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

Pullbacks of metric bundles and Cannon-Thurston maps

Swathi Krishna  and  Pranab Sardar Indian Institute of Science Education and Research, Mohali, Punjab 140306, India.
Abstract.

Metric (graph) bundles were defined by Mj and Sardar in [MS12]. In this paper, we introduce the notion of morphisms and pullbacks of metric (graph) bundles. Given a metric (graph) bundle XX over BB where XX and all the fibers are uniformly (Gromov) hyperbolic and nonelementary, and a Lipschitz quasiisometric embedding i:ABi:A\rightarrow B we show that the pullback iXi^{*}X is hyperbolic and the map i:iXXi^{*}:i^{*}X\rightarrow X admits a continuous boundary extension, i.e. the Cannon-Thurston (CT) map i:(iX)X\partial i^{*}:\partial(i^{*}X)\rightarrow\partial X. As an application of our theorem, we show that given a short exact sequence of nonelementary hyperbolic groups 1NGπQ11\rightarrow N\rightarrow G\stackrel{{\scriptstyle\pi}}{{\rightarrow}}Q\rightarrow 1 and a finitely generated quasiisometrically embedded subgroup Q1<QQ_{1}<Q, G1:=π1(Q1)G_{1}:=\pi^{-1}(Q_{1}) is hyperbolic and the inclusion G1GG_{1}\rightarrow G admits the CT map G1G\partial G_{1}\rightarrow\partial G. We then derive several interesting properties of the CT map.

1. Introduction

Given a hyperbolic group GG and a hyperbolic subgroup HH a natural question to ask is if the inclusion HGH\rightarrow G always extends continuously to HG\partial H\rightarrow\partial G (see [Bes04, Q 1.19]). This question was posed by Mahan Mitra (Mj) motivated by the seminal article of Cannon and Thurston (see [CT85]). In [CT85] the authors found the first instance of this phenomenon where HH is not quasiisometrically embedded in GG. It follows from their work that if G=π1(M)G=\pi_{1}(M) where MM is a closed hyperbolic 33-manifold fibering over a circle and H=π1(S)H=\pi_{1}(S) with SS (an orientable closed surface of genus at least 22) being the fiber, then the boundary extension HG\partial H\rightarrow\partial G exists. More generally, one may ask for a pair of (Gromov) hyperbolic metric spaces YXY\subset X if there is a continuous extension of the inclusion YXY\rightarrow X to YX\partial Y\rightarrow\partial X. Such an extension is by definition unique (see Definition 2.47) when it exists and is popularly known as the Cannon-Thurston map or ‘CT map’ for short in Geometric Group Theory. The above question of Mahan Mitra (Mj) has motivated numerous works. The reader is referred to [Mj19] for a detailed history of the problem. Although the general question for groups has been answered in the negative recently by Baker and Riley ([BR13]) there are many interesting questions to be answered in this context. In this paper, we pick up the following.

Question. Suppose 1NGπQ11\rightarrow N\rightarrow G\stackrel{{\scriptstyle\pi}}{{\rightarrow}}Q\rightarrow 1 is a short exact sequence of hyperbolic groups. Suppose Q1<QQ_{1}<Q is quasiisometrically embedded and G1=π1(Q1)G_{1}=\pi^{-1}(Q_{1}). Then does the inclusion G1<GG_{1}<G admit the CT map?

It follows by the results of [MS12] that G1G_{1} is hyperbolic (see Remark 4.4, [MS12]) so that the question makes sense. In this paper, we answer the above question affirmatively. However, we reformulate this question in terms of metric (graph) bundles as defined in [MS12] (see section 33 of this paper) and obtain the following more general result. One is referred to Lemma 2.41 and the discussion following it for the definition of barycenter map. Coarsely surjective maps are introduced in Definition 2.1(3).

Theorem 5.2. Suppose π:XB\pi:X\rightarrow B is a metric (graph) bundle such that
(1) XX is hyperbolic and
(2) all the fibers are uniformly hyperbolic and nonelementary, i.e. there are δ0\delta\geq 0 and R0R\geq 0 such that any fiber FF is δ\delta-hyperbolic and the barycenter map s3FF\partial^{3}_{s}F\rightarrow F is RR-coarsely surjective.

Suppose i:ABi:A\rightarrow B is a Lipschitz, quasiisometric embedding and πY:YA\pi_{Y}:Y\rightarrow A is the pullback bundle under ii (see Definition 3.18). Then i:YXi^{*}:Y\rightarrow X admits the CT map.

There are two main sources of examples of metric graph bundles mentioned in this paper where the above theorem can be applied. The first one is that of short exact sequences of groups.

Theorem 6.1. Suppose 1NGπQ11\rightarrow N\rightarrow G\stackrel{{\scriptstyle\pi}}{{\rightarrow}}Q\rightarrow 1 is a short exact sequence of hyperbolic groups. Suppose Q1<QQ_{1}<Q is quasiisometrically embedded and G1=π1(Q1)G_{1}=\pi^{-1}(Q_{1}). Then G1G_{1} is a hyperbolic group and the inclusion G1<GG_{1}<G admits the CT map.

We note that special cases of Theorem 5.2 and Theorem 6.1, namely when AA is a point and Q1=(1)Q_{1}=(1) respectively, were already known. See Theorem 5.3 in [MS12] and Theorem 4.3 in [Mit98a]. Another context where Theorem 5.2 applies is that of complexes of hyperbolic groups. We refer to Section 3.3.2 for relevant definitions.

Suppose 𝒴\mathcal{Y} is a finite simplicial complex and 𝔾(𝒴){\mathbb{G}}(\mathcal{Y}) is a developable complex of nonelementary hyperbolic groups over 𝒴\mathcal{Y}. Suppose that for all face σ\sigma of 𝒴{\mathcal{Y}}, GσG_{\sigma} is a nonelementary hyperbolic group and for any two faces στ\sigma\subset\tau the corresponding homomorphism GτGσG_{\tau}\rightarrow G_{\sigma} is an isomorphism onto a finite index subgroup of GσG_{\sigma}. Suppose that the fundamental group of the complex of groups, GG say, is hyperbolic. Suppose we have a good subcomplex 𝒴1𝒴{\mathcal{Y}}_{1}\subset{\mathcal{Y}} i.e. one for which the following two conditions are satisfied.
(1) The natural homomorphism π1(𝒢,𝒴1)π1(𝒢,𝒴)\pi_{1}({\mathcal{G}},{\mathcal{Y}}_{1})\rightarrow\pi_{1}({\mathcal{G}},{\mathcal{Y}}) is injective.

Let G1=π1(𝒢,𝒴1)G_{1}=\pi_{1}({\mathcal{G}},{\mathcal{Y}}_{1}). Suppose G1G_{1} and GG are both endowed with word metrics with respect to some finite generating sets. Let G^\hat{G} and G^1\hat{G}_{1} be the coned off spaces a la Farb ([Far98]), obtained by coning off all the face groups in GG and G1G_{1} respectively.
(2) Then the induced map G^1G^\hat{G}_{1}\rightarrow\hat{G} of the coned off spaces is a quasiisometric embedding. With these hypotheses we have:

Theorem 6.2. The group G1G_{1} is hyperbolic and the inclusion G1GG_{1}\rightarrow G admits the CT map.

Particularly interesting cases to which the above theorem applies are obtained in [Min11] and [MG20]. There graphs of groups are considered where all the vertex and edge groups are either surface groups ([Min11]) or free groups of rank 3\geq 3 ([MG20]) respectively.

Next, we explore properties of the Cannon-Thurston map YX\partial Y\rightarrow\partial X proved in Theorem 5.2. Suppose FF is a fiber of the bundle YY over AA. Then there is a CT map for the inclusions iF,X:FXi_{F,X}:F\rightarrow X and iF,Y:FYi_{F,Y}:F\rightarrow Y, and the map i:YXi^{*}:Y\rightarrow X. Since iF,X=iiF,Y\partial i_{F,X}=\partial i^{*}\circ\partial i_{F,Y}, if α,βF\alpha,\beta\in\partial F are identified under iF,X\partial i_{F,X} then under i\partial i^{*} the points iF,Y(α)\partial i_{F,Y}(\alpha) and iF,Y(β)\partial i_{F,Y}(\beta) are identified too. It turns out that a sort of ‘converse’ of this is also true.

Theorem 6.25. Suppose we have the hypotheses of Theorem 5.2 and also that the fibers of the bundle are proper metric spaces. Suppose γ\gamma is a (quasi)geodesic line in YY such that γ()\gamma(\infty) and γ()\gamma(-\infty) are identified by the CT map i:YX\partial i^{*}:\partial Y\rightarrow\partial X. Then πY(γ)\pi_{Y}(\gamma) is bounded. In particular, given any fiber FF of the metric bundle, γ\gamma is at a finite Hausdorff distance from a quasigeodesic line of FF.

On the other hand as an immediate application of Theorem 6.25 (in fact, see Corollary 6.26 and Proposition 6.6) we get the following:

Theorem. Suppose we have the hypotheses of Theorem 5.2 and also that the fibers of the bundle are proper metric spaces. Let FF be the fiber over a point bAb\in A. Then the CT map iF,X:FX\partial i_{F,X}:\partial F\rightarrow\partial X is surjective if and only if the CT maps iF,Yξ:FYξ\partial i_{F,Y_{\xi}}:\partial F\rightarrow\partial Y_{\xi} are surjective for all ξB\xi\in\partial B where YξY_{\xi} is the pullback of a (quasi)geodesic ray in BB asymptotic to ξ\xi.

In particular iF,Y:FY\partial i_{F,Y}:\partial F\rightarrow\partial Y is surjective if iF,X:FX\partial i_{F,X}:\partial F\rightarrow\partial X is surjective.

Following Mitra ([Mit97]) we define the Cannon-Thurston lamination X(2)(F)\partial^{(2)}_{X}(F) to be {(z1,z2)F×F:z1z2,iF,X(z1)=iF,X(z2)}\{(z_{1},z_{2})\in\partial F\times\partial F:z_{1}\neq z_{2},\,\partial i_{F,X}(z_{1})=\partial i_{F,X}(z_{2})\} and following Bowditch ([Bow13, Section 2.3]) we define for any point ξB\xi\in\partial B a subset of this lamination denoted by ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F) or simply ξ(2)(F)\partial^{(2)}_{\xi}(F) when XX is understood, where (z1,z2)ξ,X(2)(F)(z_{1},z_{2})\in\partial^{(2)}_{\xi,X}(F) if and only if iF,X(z1)=iF,X(z2)=γ~()\partial i_{F,X}(z_{1})=\partial i_{F,X}(z_{2})=\tilde{\gamma}(\infty) where γ~\tilde{\gamma} is a quasiisometric lift in XX of a (quasi)geodesic ray γ\gamma in BB converging to ξ\xi. If (z1,z2)ξ,X(2)(F)(z_{1},z_{2})\in\partial^{(2)}_{\xi,X}(F) and α\alpha is a (quasi)geodesic line in FF connecting z1,z2z_{1},z_{2} then α\alpha is referred to be a leaf of the lamination ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F). Leaves are assumed to be uniform quasigeodesics in the following theorem using Proposition 2.37.

Theorem. (See Lemma 6.17 through Lemma 6.24.) (Properties of X(2)(F)\partial^{(2)}_{X}(F))

(1)(1) X(2)(F)=ξBξ,X(2)(F)\partial^{(2)}_{X}(F)=\coprod_{\xi\in\partial B}\partial^{(2)}_{\xi,X}(F).

(2)(2) X(2)(F)\partial^{(2)}_{X}(F) and ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F) are all closed subsets of (2)F\partial^{(2)}F where (2)F={(z1,z2)F×F:z1z2}.\partial^{(2)}F=\{(z_{1},z_{2})\in\partial F\times\partial F:z_{1}\neq z_{2}\}.

(3)(3) The leaves of ξ1,X(2)(F),ξ2,X(2)(F)\partial^{(2)}_{\xi_{1},X}(F),\partial^{(2)}_{\xi_{2},X}(F) are coarsely transverse to each other for all ξ1ξ2B\xi_{1}\neq\xi_{2}\in\partial B:

Given ξ1ξ2B\xi_{1}\neq\xi_{2}\in\partial B and D>0D>0 there exists R>0R>0 such that if γi\gamma_{i} is leaf of ξi,X(2)(F)\partial^{(2)}_{\xi_{i},X}(F), i=1,2i=1,2 then γ1ND(γ2)\gamma_{1}\cap N_{D}(\gamma_{2}) has diameter less than RR.

(4)(4) If ξnξ\xi_{n}\rightarrow\xi in B\partial B and αn\alpha_{n} is a leaf of ξn,X(2)(F)\partial^{(2)}_{\xi_{n},X}(F) for all nn\in{\mathbb{N}} which converge to a geodesic line α\alpha then α\alpha is a leaf of ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F).

(5)(5) ξ,X(2)(F)=ξ,Y(2)(F)\partial^{(2)}_{\xi,X}(F)=\partial^{(2)}_{\xi,Y}(F) for all ξA\xi\in\partial A if we have the hypothesis of Theorem 5.2.

Finally, we also prove the following interesting property of the CT lamination.

Theorem 6.30. Suppose XX is a metric (graph) bundle over BB satisfying the hypotheses of Theorem 5.2 such that XX is a proper metric space. Let F=FbF=F_{b} where bBb\in B. Suppose F\partial F is not homeomorphic to a dendrite and also the CT map FX\partial F\rightarrow\partial X is surjective.

Then for all ξB\xi\in\partial B we have ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F)\neq\emptyset.

This applies in particular to the examples of short exact sequence of hyperbolic groups and the complexes of hyperbolic groups mentioned in Theorem 6.1 and Theorem 6.2 above.

Outline of the paper: In section 2 we recall basic hyperbolic geometry, Cannon-Thurston maps, etc. In section 3 we recall the basics of metric (graph) bundles and we introduce morphisms of bundles, pullbacks. Here we prove the existence of pullbacks under suitable assumptions. In section 4 we mainly recall the machinery of [MS12] and we prove a few elementary results. Section 5 is devoted to the proof of the main theorem. In section 6 we derive applications of the main result and we mention some related results.

2. Hyperbolic metric spaces

In this section, we remark on the notation and convention to be followed in the rest of the paper and we put together basic definitions and results about hyperbolic metric spaces. We begin with some basic notions from large scale geometry. Most of these are quite standard, e.g. see [Gro87], [Gd90]. We have used [MS12] where all the basic notions can be quickly found in one place.

Notation, convention and some metric space notions. One is referred to [BH99, Chapter I.1, I.3] for the definitions and basic facts about geodesic metric spaces, metric graphs and length spaces.
(0) For any set AA, IdAId_{A} will denote the identity map AAA\rightarrow A. If ABA\subset B then we denote by iA,B:ABi_{A,B}:A\rightarrow B the inclusion map of AA into BB.
(1) If xXx\in X and AXA\subset X then d(x,A)d(x,A) will denote inf{d(x,y):yA}\inf\{d(x,y):y\in A\} and will be referred to as the distance of xx from AA. For D0D\geq 0 and AXA\subset X, ND(A):={xX:d(x,a)D for someaA}N_{D}(A):=\{x\in X:\,d(x,a)\leq D\,\mbox{ for some}\,a\in A\} will be called the DD-neighborhood of AA in XX. For A,BXA,B\subset X we shall denote by d(A,B)d(A,B) the quantity inf{d(x,B):xA}\inf\{d(x,B):x\in A\} and by Hd(A,B)Hd(A,B) the quantity inf{D>0:AND(B),BND(A)}\inf\{D>0:A\subset N_{D}(B),B\subset N_{D}(A)\} and will refer to it as the Hausdorff distance of A,BA,B.
(2) If XX is a length space we consider only subspaces YXY\subset X such that the induced length metric on YY takes values in [0,)[0,\infty), or equivalently for any pair of points in YY there is a rectifiable path in XX joining them which is contained in YY. We shall refer to such subsets as rectifiably path connected. If γ\gamma is a rectifiable path in XX then l(γ)l(\gamma) will denote the length of γ\gamma.
(3) All graphs are connected for us. If XX is a metric graph then 𝒱(X){\mathcal{V}}(X) will denote the set of vertices of XX. Generally, we shall write xXx\in X to mean x𝒱(X)x\in{\mathcal{V}}(X). In metric graphs (see [BH99, Chapter I.1]) all the edges are assumed to have length 11. In a graph XX the paths are assumed to be a sequence of vertices. In other words, these are maps IXI\cap\mathbb{Z}\rightarrow X where II is a closed interval in {\mathbb{R}} with end points in {±}\mathbb{Z}\cup\{\pm\infty\}. We shall informally write this as α:IX\alpha:I\rightarrow X and sometimes refer to it as a dotted path for emphasis. Length of such a path α:IX\alpha:I\rightarrow X is defined to be l(α)=d(α(i),α(i+1))l(\alpha)=\sum d(\alpha(i),\alpha(i+1)) where the sum is taken over all ii\in{\mathbb{Z}} such that i,i+1Ii,i+1\in I. If α:[0,n]X\alpha:[0,n]\rightarrow X and β:[0,m]X\beta:[0,m]\rightarrow X are two paths with α(n)=β(0)\alpha(n)=\beta(0) then their concatenation αβ\alpha*\beta will be the path [0,m+n]X[0,m+n]\rightarrow X defined by αβ(i)=α(i)\alpha*\beta(i)=\alpha(i) if i[0,n]i\in[0,n] and αβ(j)=β(jn)\alpha*\beta(j)=\beta(j-n) if j[n,m+n]j\in[n,m+n].
(4) If XX is a geodesic metric space and x,yXx,y\in X then we shall use [x,y]X[x,y]_{X} or simply [x,y][x,y] to denote a geodesic segment joining xx to yy. This applies in particular to metric graphs. For x,y,zXx,y,z\in X we shall denote by Δxyz\Delta xyz some geodesic triangle with vertices x,y,zx,y,z.
(5) If XX is any metric space then for all AXA\subset X, diam(A)diam(A) will denote the diameter of AA.

2.1. Basic notions from large scale geometry

Suppose XX, YY are any two metric spaces and k1,ϵ0,ϵ0k\geq 1,\,\epsilon\geq 0,\epsilon^{\prime}\geq 0.

Definition 2.1.

([MS12, Definition 1.1.1])

  1. (1)

    A map ϕ:XY\phi:X\rightarrow Y is said to be metrically proper if there is an increasing function f:[0,)[0,)f:[0,\infty)\rightarrow[0,\infty) with limtf(t)=\lim_{t\rightarrow\infty}f(t)=\infty such that for any x,yXx,y\in X and R[0,)R\in[0,\infty), dY(ϕ(x),ϕ(y))Rd_{Y}(\phi(x),\phi(y))\leq R implies dX(x,y)f(R)d_{X}(x,y)\leq f(R). In this case we say that ϕ\phi is proper as measured by ff.

  2. (2)

    A subset AA of a metric space XX is said to be rr-dense in XX for some r0r\geq 0 if Nr(A)=XN_{r}(A)=X.

  3. (3)

    Suppose AA is a set. A map ϕ:AY\phi:A\rightarrow Y is said to be ϵ\epsilon-coarsely surjective if ϕ(A)\phi(A) is ϵ\epsilon-dense in YY. We will say that it is coarsely surjective if it is ϵ\epsilon-coarsely surjective for some ϵ0\epsilon\geq 0.

  4. (4)

    A map ϕ:XY\phi:X\rightarrow Y is said to be coarsely (ϵ,ϵ)(\epsilon,\epsilon^{\prime})-Lipschitz if for every x1,x2Xx_{1},x_{2}\in X, we have d(ϕ(x1),ϕ(x2))ϵd(x1,x2)+ϵd(\phi(x_{1}),\phi(x_{2}))\leq\epsilon d(x_{1},x_{2})+\epsilon^{\prime}. A coarsely (ϵ,ϵ)(\epsilon,\epsilon)-Lipschitz map will be simply called a coarsely ϵ\epsilon-Lipschitz map. A map ϕ\phi is coarsely Lipschitz if it is coarsely ϵ\epsilon-Lipschitz for some ϵ0\epsilon\geq 0.

  5. (5)

    (i)(i) A map ϕ:XY\phi:X\rightarrow Y is said to be a (k,ϵ)(k,\epsilon)-quasiisometric embedding if for every x1,x2Xx_{1},x_{2}\in X, one has

    ϵ+d(x1,x2)/kd(ϕ(x1),ϕ(x2))ϵ+kd(x1,x2).-\epsilon+d(x_{1},x_{2})/k\leq d(\phi(x_{1}),\phi(x_{2}))\leq\epsilon+kd(x_{1},x_{2}).

    A map ϕ:XY\phi:X\rightarrow Y will simply be referred to as a quasiisometric embedding if it is a (k,ϵ)(k,\epsilon)-quasiisometric embedding for some k1k\geq 1, ϵ0\epsilon\geq 0. A (k,k)(k,k)-quasiisometric embedding will be referred to as a kk-quasiisometric embedding.

    (ii)(ii) A map ϕ:XY\phi:X\rightarrow Y is a (k,ϵ)(k,\epsilon)-quasiisometry (resp. kk-quasiisometry) if it is a (k,ϵ)(k,\epsilon)-quasiisometric embedding (resp. kk-quasiisometric embedding) and moreover, it is DD-coarsely surjective for some D0D\geq 0.

    (iii)(iii) A (k,ϵ)(k,\epsilon)-quasigeodesic (resp. a kk-quasigeodesic) in a metric space XX is a (k,ϵ)(k,\epsilon)-quasiisometric embedding (resp. a kk-quasiisometric embedding) γ:IX\gamma:I\rightarrow X, where II\subseteq\mathbb{R} is an interval.

    We recall that a (1,0)(1,0)-quasigeodesic is called a geodesic.

    If I=[0,)I=[0,\infty), then γ\gamma will be called a quasigeodesic ray. If I=I={\mathbb{R}}, then we call it a quasigeodesic line. One similarly defines a geodesic ray and a geodesic line. We refer to the constant(s) kk (and ϵ\epsilon) as quasigeodesic constant(s).

    Quasigeodesics in a metric graph XX will be maps IXI\cap{\mathbb{Z}}\rightarrow X, informally written as IXI\rightarrow X where II is a closed interval in {\mathbb{R}}.

  6. (6)

    Suppose ϕ,ϕ:XY\phi,\phi^{\prime}:X\rightarrow Y are two maps and ϵ0\epsilon\geq 0.

    (i) We define d(ϕ,ϕ)d(\phi,\phi^{\prime}) to be the quantity sup{dY(ϕ(x),ϕ(x)):xX}\sup\{d_{Y}(\phi(x),\phi^{\prime}(x)):x\in X\} provided the supremum exists in {\mathbb{R}}; otherwise we write d(ϕ,ϕ)=d(\phi,\phi^{\prime})=\infty.

    (ii) A map ψ:YX\psi:Y\rightarrow X is called an ϵ\epsilon-coarse left (right) inverse of ϕ\phi if d(ψϕ,IdX)ϵd(\psi\circ\phi,Id_{X})\leq\epsilon (resp. d(ϕψ,IdY)ϵd(\phi\circ\psi,Id_{Y})\leq\epsilon).

    If ψ\psi is both an ϵ\epsilon-coarse left and right inverse then it is simply called an ϵ\epsilon-coarse inverse of ϕ\phi.

  7. (7)

    Suppose SS is any set. A map f:SXf:S\rightarrow X satisfying some properties 𝒫1,,𝒫k\mathcal{P}_{1},\cdots,\mathcal{P}_{k} will be called coarsely unique if for any other map g:SXg:S\rightarrow X with properties 𝒫1,,𝒫k\mathcal{P}_{1},\cdots,\mathcal{P}_{k} there is a constant DD such that d(f,g)Dd(f,g)\leq D.

The definition (7) above is taken from [MS12]. See the definition following Lemma 2.9 there. In places where this definition will be used the properties may not be explicitly stated but they will be clear from the context. If SS is finite then we talk about a finite subset of XX to be coarsely unique, e.g. see the remark following Lemma 2.56.

Remark on terminology: (1) All the above definitions are about certain properties of maps and in each case some parameters are involved.

(i) When the parameters are not important or they are clear from the context then we say that the map has the particular property without explicit mention of the parameters, e.g. ‘ϕ:XY\phi:X\rightarrow Y is metrically proper’ if ϕ\phi is metrically proper as measured by some function.

(ii) When we have a set of pairs of metric spaces and a map between each pair possessing the same property with the same parameters then we say that the set of maps ‘uniformly’ have the property, e.g. uniformly metrically proper, uniformly coarsely Lipschitz, uniform qi embeddings, uniform approximate nearest point projection etc.

(2) We often refer to quasiisometric embeddings as ‘qi embedding’ and quasiisometry as ‘qi’.

The following gives a characterization of quasiisometry to be used in the discussion on metric bundles.

Lemma 2.2.

([MS12, Lemma 1.1])

  1. (1)

    For every K1,K21K_{1},K_{2}\geq 1 and D0D\geq 0 there are K2.2=K2.2(K1,K2,D)K_{\ref{elem-lemma1}}=K_{\ref{elem-lemma1}}(K_{1},K_{2},D), such that the following hold.

    A K1K_{1}-coarsely Lipschitz map with a K2K_{2}-coarsely Lipschitz, DD-coarse inverse is a K2.2K_{\ref{elem-lemma1}}-quasiisometry.

  2. (2)

    Given K1K\geq 1, ϵ0\epsilon\geq 0 and R0R\geq 0 there are constants C2.2=C2.2(K,ϵ,R)C_{\ref{elem-lemma1}}=C_{\ref{elem-lemma1}}(K,\epsilon,R) and D2.2=D2.2(K,ϵ,R)D_{\ref{elem-lemma1}}=D_{\ref{elem-lemma1}}(K,\epsilon,R) such that the following holds:

    Suppose X,YX,Y are any two metric spaces and f:XYf:X\rightarrow Y is a (K,ϵ)(K,\epsilon)-quasiisometry which is RR-coarsely surjective. Then there is a (K2.2,C2.2)(K_{\ref{elem-lemma1}},C_{\ref{elem-lemma1}})-quasiisometric D2.2D_{\ref{elem-lemma1}}-coarse inverse of ff.

The following lemmas follow from simple calculations and hence we omit their proofs.

Lemma 2.3.

(1) Suppose we have a sequence of maps XfYgZX\stackrel{{\scriptstyle f}}{{\rightarrow}}Y\stackrel{{\scriptstyle g}}{{\rightarrow}}Z where f,gf,g are coarsely L1L_{1}-Lipschitz and L2L_{2}-Lipschitz respectively. Then gfg\circ f is coarsely (L1L2,L1L2+L2)(L_{1}L_{2},L_{1}L_{2}+L_{2})-Lipschitz.

(2) Suppose f:XYf:X\rightarrow Y is a (K1,ϵ1)(K_{1},\epsilon_{1})-qi embedding and g:YZg:Y\rightarrow Z is a (K2,ϵ2)(K_{2},\epsilon_{2})-qi embedding. Then gf:XZg\circ f:X\rightarrow Z is a (K1K2,K2ϵ1+ϵ2)(K_{1}K_{2},K_{2}\epsilon_{1}+\epsilon_{2})-qi embedding.

Moreover, if ff is D1D_{1}-coarsely surjective and gg is D2D_{2}-coarsely surjective then gfg\circ f is (K2D1+ϵ2+D2)(K_{2}D_{1}+\epsilon_{2}+D_{2})-coarsely surjective.

In particular, the composition of finitely many quasiisometries is a quasiisometry.

Lemma 2.4.

Suppose XX^{\prime} is any connected graph and r>0r>0. Suppose XX is another graph obtained from XX^{\prime} by introducing some new edges to XX^{\prime} where e=[v,w]e=[v,w] is an edge in XX but not in XX^{\prime} implies dX(v,w)rd_{X^{\prime}}(v,w)\leq r. Then the inclusion map XXX^{\prime}\rightarrow X is a quasiisometry.

The following lemma appears in [KS, Section 1.5] in a somewhat different form. We include a proof for the sake of completeness.

Lemma 2.5.

Let XX be any metric space, x,yXx,y\in X, γ\gamma be a (dotted) kk-quasigeodesic joining x,yx,y and α:IX\alpha:I\rightarrow X is a (dotted) coarsely LL-Lipschitz path joining x,yx,y. Suppose moreover, α\alpha is a proper embedding as measured by a function f:[0,)[0,)f:[0,\infty)\rightarrow[0,\infty) and that Hd(α,γ)DHd(\alpha,\gamma)\leq D for some D0D\geq 0. Then α\alpha is (dotted) K2.5=K2.5(k,f,D,L)K_{\ref{quasigeod criteria}}=K_{\ref{quasigeod criteria}}(k,f,D,L)-quasigeodesic in XX.

Proof.

Suppose γ\gamma is defined on an interval JJ. Let a,bIa,b\in I. Then we have

d(α(a),α(b))L|ab|+L(1)\,\,d(\alpha(a),\alpha(b))\leq L|a-b|+L\longrightarrow(1)

since α\alpha is coarsely LL-Lipschitz. Now let a,bJa^{\prime},b^{\prime}\in J be such that d(α(a),γ(a)Dd(\alpha(a),\gamma(a^{\prime})\leq D and d(α(b),γ(b))Dd(\alpha(b),\gamma(b^{\prime}))\leq D. Let R=d(α(a),α(b))R=d(\alpha(a),\alpha(b)). Then by triangle inequality d(γ(a),γ(b))2D+Rd(\gamma(a^{\prime}),\gamma(b^{\prime}))\leq 2D+R. Since γ\gamma is a kk-quasigeodesic we have k+|ab|/kd(γ(a),γ(b))2D+R-k+|a^{\prime}-b^{\prime}|/k\leq d(\gamma(a^{\prime}),\gamma(b^{\prime}))\leq 2D+R. Hence, |ab|k(2D+R)+k2|a^{\prime}-b^{\prime}|\leq k(2D+R)+k^{2}. Without loss of generality suppose aba^{\prime}\leq b^{\prime}. Consider the sequence of points a0=a,a1,,an=ba^{\prime}_{0}=a^{\prime},a^{\prime}_{1},\cdots,a^{\prime}_{n}=b^{\prime} in JJ such that ai+1=1+aia^{\prime}_{i+1}=1+a^{\prime}_{i} for 0in20\leq i\leq n-2 and anan11a^{\prime}_{n}-a^{\prime}_{n-1}\leq 1. We note that n1+k(2D+R)+k2n\leq 1+k(2D+R)+k^{2}. Let aiIa_{i}\in I be such that d(γ(ai),α(ai))Dd(\gamma(a^{\prime}_{i}),\alpha(a_{i}))\leq D, 0in0\leq i\leq n where a0=a,an=ba_{0}=a,a_{n}=b. Once again by triangle inequality we have

d(α(ai),α(ai+1))2D+d(γ(ai),γ(ai+1))2D+2kd(\alpha(a_{i}),\alpha(a_{i+1}))\leq 2D+d(\gamma(a^{\prime}_{i}),\gamma(a^{\prime}_{i+1}))\leq 2D+2k

for 0in10\leq i\leq n-1 since γ\gamma is a kk-quasigeodesic. This implies |aiai+1|f(2D+2k)|a_{i}-a_{i+1}|\leq f(2D+2k) since α\alpha is a proper embedding as measured by ff. Hence,

|ab|i=0n1|aiai+1|nf(2D+k)(1+k(2D+R)+k2)f(2D+2k).|a-b|\leq\sum_{i=0}^{n-1}|a_{i}-a_{i+1}|\leq nf(2D+k)\leq(1+k(2D+R)+k^{2})f(2D+2k).

Thus we have

1+2kD+k2k+1kf(2D+2k)|ab|R=d(α(a),α(b)).(2)\,\,-\frac{1+2kD+k^{2}}{k}+\frac{1}{kf(2D+2k)}|a-b|\leq R=d(\alpha(a),\alpha(b)).\longrightarrow(2)

Hence, by (1) and (2) we can take

K2.5=1+2D+k+kf(2D+2k)+L.K_{\ref{quasigeod criteria}}=1+2D+k+kf(2D+2k)+L.\qed

The following lemma is implicit in the proof [MS12, Proposition 2.10]. The proof of this lemma being immediate we omit it.

Lemma 2.6.

Suppose XX is a length space and YY is any metric space. Let f:XYf:X\rightarrow Y be any map. Then ff is coarsely CC-Lipschitz for some C0C\geq 0 if for all x1,x2Xx_{1},x_{2}\in X, dX(x1,x2)1d_{X}(x_{1},x_{2})\leq 1 implies dY(f(x1),f(x2))Cd_{Y}(f(x_{1}),f(x_{2}))\leq C.

Remark 1.

We spend quite some time restating some results proved in [MS12] in the generality of length spaces since the main result in our paper is about length spaces. For instance (1) the existence of pullback of metric bundles to be defined below is unclear within the category of geodesic metric spaces; and (2) we observe that for the definition of Cannon-Thurston maps the assumption of (Gromov) hyperbolic geodesic metric spaces is rather restrictive and unnecessary.

In a length metric space geodesics may not exist joining a pair of points. However, we still have the following.

Lemma 2.7.

Suppose XX is a length space. (1) Given any ϵ>0\epsilon>0, any pair of points of XX can be joined by a continuous, rectifiable, arc length parameterized path which is a (1,ϵ)(1,\epsilon)-quasigeodesic.

(2) Any pair of points of XX can be joined by a dotted 11-quasigeodesic.

Metric graph approximation to a length space

Given any length space XX, we define a metric graph YY as follows. We take the vertex set V(Y)=XV(Y)=X. We join x,yXx,y\in X by an edge (of length 11) if and only if dX(x,y)1d_{X}(x,y)\leq 1. We let ψX:X𝒱(Y)Y\psi_{X}:X\rightarrow\mathcal{V}(Y)\subset Y be the identity map. Let ϕX:YX\phi_{X}:Y\rightarrow X be defined to be the inverse of ψX\psi_{X} on 𝒱(Y){\mathcal{V}}(Y) and for any point yy in the interior of an edge ee of YY we define ϕX(y)\phi_{X}(y) to be one of the end points of the edge ee. The following hold.

Lemma 2.8.

[KS, Lemma 1.32] (1) YY is a (connected) metric graph. (2) The maps ψX\psi_{X} and ϕX|𝒱(Y)\phi_{X}|_{{\mathcal{V}}(Y)} are coarsely 11-surjective, (1,1)(1,1)-quasiisometries. (3) The map ϕX\phi_{X} is a (1,3)(1,3)-quasiisometry and it is a 11-coarse inverse of ψX\psi_{X}.

Remark 2.

We shall refer to the space YY constructed in the proof of the above lemma as the (canonical) metric graph approximation to XX. We also preserve the notations ψX\psi_{X} and ϕX\phi_{X} to be used in this context only.

Definition 2.9.

Gromov inner product: Let XX be any metric space and let p,x,yXp,x,y\in X. Then the Gromov inner product of x,yx,y with respect to pp is defined to be the number 12(d(p,x)+d(p,y)d(x,y))\frac{1}{2}(d(p,x)+d(p,y)-d(x,y)). It is denoted by (x.y)p(x.y)_{p}.

Lemma 2.10.

Suppose XX is a length space and x1,x2,x3Xx_{1},x_{2},x_{3}\in X. Let γij,i<j,1i,j3\gamma_{ij},i<j,1\leq i,j\leq 3 denote (1,1)(1,1)-quasigeodesics joining the respective pairs of points xi,xjx_{i},x_{j}. Suppose there are points w1γ23,w2γ13w_{1}\in\gamma_{23},w_{2}\in\gamma_{13} and w3γ12w_{3}\in\gamma_{12} such that d(w1,wi)Rd(w_{1},w_{i})\leq R for some R0R\geq 0, i=2,3i=2,3. Then |(x2.x3)x1d(x1,w1)|3+2R|(x_{2}.x_{3})_{x_{1}}-d(x_{1},w_{1})|\leq 3+2R.

Proof.

By triangle inequality we have |d(x2,w1)d(x2,w2)|R|d(x_{2},w_{1})-d(x_{2},w_{2})|\leq R, |d(x3,w1)d(x3,w2)|R|d(x_{3},w_{1})-d(x_{3},w_{2})|\leq R, |d(x1,w1)d(x1,wi)|R|d(x_{1},w_{1})-d(x_{1},w_{i})|\leq R, i=2,3i=2,3. Since the γij\gamma_{ij}’s are (1,1)(1,1)-quasigeodesics it is easy to see that d(x1,w3)+d(w3,x2)d(x1,x2)+3d(x_{1},w_{3})+d(w_{3},x_{2})\leq d(x_{1},x_{2})+3, d(x1,w2)+d(w2,x3)d(x1,x3)+3d(x_{1},w_{2})+d(w_{2},x_{3})\leq d(x_{1},x_{3})+3 and d(x2,w1)+d(w1,x3)d(x2,x3)+3d(x_{2},w_{1})+d(w_{1},x_{3})\leq d(x_{2},x_{3})+3. It then follows by a simple calculation that

2d(x1,w1)64Rd(x1,x2)+d(x1,x3)d(x2,x3)2d(x1,w1)+3+4R.2d(x_{1},w_{1})-6-4R\leq d(x_{1},x_{2})+d(x_{1},x_{3})-d(x_{2},x_{3})\leq 2d(x_{1},w_{1})+3+4R.

Hence, we have |(x2.x3)x1d(x1,w1)|3+2R|(x_{2}.x_{3})_{x_{1}}-d(x_{1},w_{1})|\leq 3+2R. ∎

Definition 2.11.
  1. (1)

    Suppose XX is a length space and Y1,Y2,ZY_{1},Y_{2},Z are nonempty subsets of XX. We say that ZZ coarsely disconnects Y1,Y2Y_{1},Y_{2} in XX if (i) YiZY_{i}\setminus Z\neq\emptyset, i=1,2i=1,2 and (ii) for all K1K\geq 1 there is R0R\geq 0 such that the following holds: For any yiYiy_{i}\in Y_{i}, i=1,2i=1,2 and any KK-quasigeodesic γ\gamma in XX joining y1,y2y_{1},y_{2} we have γNR(Z)\gamma\cap N_{R}(Z)\neq\emptyset.

  2. (2)

    Suppose Y,ZXY,Z\subset X, Y1,Y2YY_{1},Y_{2}\subset Y. We say that ZZ coarsely bisects YY into Y1,Y2Y_{1},Y_{2} in XX if Y=Y1Y2Y=Y_{1}\cup Y_{2} and ZZ coarsely disconnects Y1,Y2Y_{1},Y_{2} in XX.

  3. (3)

    Suppose {Xi}\{X_{i}\} is a collection of length spaces and there are nonempty sets Yi,ZiXiY_{i},Z_{i}\subset X_{i}, Yi+,YiYiY^{+}_{i},Y^{-}_{i}\subset Y_{i} such that Yi=Yi+YiY_{i}=Y^{+}_{i}\cup Y^{-}_{i}, Yi+ZiY^{+}_{i}\setminus Z_{i}\neq\emptyset, and YiZiY^{-}_{i}\setminus Z_{i}\neq\emptyset for all ii. We say that ZiZ_{i}’s uniformly coarsely bisect YiY_{i}’s into Yi+Y^{+}_{i}’s, and YiY^{-}_{i}’s if for all K1K\geq 1 there is R=R(K)0R=R(K)\geq 0 with the following property: For any ii, and for any xi+Yi+,xiYix^{+}_{i}\in Y^{+}_{i},x^{-}_{i}\in Y^{-}_{i} and any KK-quasigeodesic γiXi\gamma_{i}\subset X_{i} joining xi±x^{\pm}_{i} we have NR(Zi)γiN_{R}(Z_{i})\cap\gamma_{i}\neq\emptyset.

We note that the first part of the above definition implies Y1Y2NR(1)(Z)Y_{1}\cap Y_{2}\subset N_{R(1)}(Z). Moreover one would like to impose the condition that YiZY_{i}\setminus Z are of infinite diameter. Keeping the application we have in mind we do not assume that.

Definition 2.12.

(Approximate nearest point projection) (1) Suppose XX is any metric space, AXA\subset X, and xXx\in X. Given ϵ0\epsilon\geq 0 and yAy\in A we say that yy is an ϵ\epsilon-approximate nearest point projection of xx on AA if for all zAz\in A we have d(x,y)d(x,z)+ϵd(x,y)\leq d(x,z)+\epsilon.

(2) Suppose XX is any metric space, AXA\subset X and ϵ0\epsilon\geq 0. An ϵ\epsilon-approximate nearest point projection map f:XAf:X\rightarrow A is a map such that f(a)=af(a)=a for all aAa\in A and f(x)f(x) is an ϵ\epsilon-approximate nearest point projection of xx on AA for all xXAx\in X\setminus A.

For ϵ=0\epsilon=0 an ϵ\epsilon-approximate nearest point projection is simply referred to as a nearest point projection. A nearest point projection map from XX onto a subset AA will be denoted by PA,X:XAP_{A,X}:X\rightarrow A or simply PA:XAP_{A}:X\rightarrow A when there is no possibility of confusion.

We note that given a metric space XX and AXA\subset X a nearest point projection map XAX\rightarrow A may not be defined in general but an ϵ\epsilon-approximate nearest point projection map XAX\rightarrow A exists by axiom of choice for all ϵ>0\epsilon>0.

Lemma 2.13.

Suppose XX is a metric space and AXA\subset X. Suppose yAy\in A is an ϵ\epsilon-approximate nearest point projection of xXx\in X. Suppose α:IX\alpha:I\rightarrow X is a (1,1)(1,1)-quasigeodesic joining x,yx,y. Then yy is an (ϵ+3)(\epsilon+3)-approximate nearest point of xx^{\prime} on AA for all xαx^{\prime}\in\alpha.

Proof.

Suppose zAz\in A is any point. Then we know that d(x,y)d(x,z)+ϵd(x,y)\leq d(x,z)+\epsilon. Since α\alpha is a (1,1)(1,1)-quasigeodesic it is easy to see that d(x,x)+d(x,y)d(x,y)+3d(x,x^{\prime})+d(x^{\prime},y)\leq d(x,y)+3. Hence, d(x,x)+d(x,y)d(x,z)+3+ϵd(x,x^{\prime})+d(x^{\prime},y)\leq d(x,z)+3+\epsilon which in turn implies that d(x,y)d(x,z)d(x,x)+3+ϵd(x,z)+ϵ+3d(x^{\prime},y)\leq d(x,z)-d(x,x^{\prime})+3+\epsilon\leq d(x^{\prime},z)+\epsilon+3. Hence, yy is an (ϵ+3)(\epsilon+3)-approximate nearest point projection of xx^{\prime} on AA. ∎

Corollary 2.14.

Suppose XX is any metric space and x,y,zXx,y,z\in X. Suppose α\alpha, β\beta are (1,1)(1,1)-quasigeodesics joining x,yx,y and y,zy,z respectively. If yy is an ϵ\epsilon-approximate nearest point projection of xx on β\beta then αβ\alpha*\beta is (3,3+ϵ)(3,3+\epsilon)-quasigeodesic.

Proof.

Let xαx^{\prime}\in\alpha and yβy^{\prime}\in\beta. Let β\beta^{\prime} denote the segment of β\beta from yy to yy^{\prime}. Then yy is an ϵ\epsilon-approximate nearest point projection of xx on β\beta^{\prime} too. Hence, by the previous lemma yy is an (ϵ+3)(\epsilon+3)-approximate nearest point projection of xx^{\prime} on β\beta^{\prime}. Without loss of generality, suppose α(a)=x\alpha(a)=x^{\prime}, α(a+m)=y\alpha(a+m)=y, β(0)=y\beta(0)=y, and β(n)=y\beta(n)=y^{\prime}. Now, d(x,y)d(x,y)+ϵ+3d(x^{\prime},y)\leq d(x^{\prime},y^{\prime})+\epsilon+3. Hence d(y,y)d(x,y)+d(x,y)2d(x,y)+ϵ+3d(y,y^{\prime})\leq d(x^{\prime},y^{\prime})+d(x^{\prime},y)\leq 2d(x^{\prime},y^{\prime})+\epsilon+3. Since α,β\alpha,\beta are both (1,1)(1,1)-quasigeodesics it follows that m1d(x,y)d(x,y)+ϵ+3m-1\leq d(x^{\prime},y)\leq d(x^{\prime},y^{\prime})+\epsilon+3 and n1d(y,y)2d(x,y)+ϵ+3n-1\leq d(y,y^{\prime})\leq 2d(x^{\prime},y^{\prime})+\epsilon+3. Adding these we get m+n23d(x,y)+2ϵ+6m+n-2\leq 3d(x^{\prime},y^{\prime})+2\epsilon+6. On the other hand, d(x,y)d(x,y)+d(y,y)m+n+2d(x^{\prime},y^{\prime})\leq d(x^{\prime},y)+d(y,y^{\prime})\leq m+n+2. Putting everything together we get

13(m+n)2ϵ+83d(x,y)(m+n)+2\frac{1}{3}(m+n)-\frac{2\epsilon+8}{3}\leq d(x^{\prime},y^{\prime})\leq(m+n)+2

from which the corollary follows immediately. ∎

2.2. Rips hyperbolicity vs Gromov hyperbolicity

This subsection gives a quick introduction to some basic notions and results about hyperbolic metric spaces. One is referred to [Gro87], [Gd90], [ABC+91] for more details. The following definition of hyperbolic metric spaces is due to E. Rips and hence we refer to this as the Rips hyperbolicity.

Definition 2.15.

(1)(1) Suppose Δx1x2x3\Delta x_{1}x_{2}x_{3} is a geodesic triangle in a metric space XX and δ0\delta\geq 0, K0K\geq 0. We say that the triangle Δx1x2x3\Delta x_{1}x_{2}x_{3} is δ\delta-slim if any side of the triangle is contained in the δ\delta-neighborhood of the union of the remaining two sides.

(2)(2) Let δ0\delta\geq 0 and XX be a geodesic metric space. We say that XX is δ\delta-hyperbolic (in the sense of Rips) if all geodesic triangles in XX are δ\delta-slim.

A geodesic metric space is said to be (Rips) hyperbolic if it is δ\delta-hyperbolic in the sense of Rips for some δ0\delta\geq 0.

However, in this paper we need to deal with length spaces a lot which a priori need not be geodesic. The following definition is more relevant in that case.

Definition 2.16.

(Gromov hyperbolicity) Suppose XX is any metric space, not necessarily geodesic and δ0\delta\geq 0.

(1) Let pXp\in X. We say that the Gromov inner product on XX with respect to pp, i.e. the map X×XX\times X\rightarrow{\mathbb{R}} defined by (x,y)(x.y)p(x,y)\mapsto(x.y)_{p}, is δ\delta-hyperbolic if

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

for all x,y,zXx,y,z\in X.

(2)(2) The metric space XX is called δ\delta-hyperbolic in the sense of Gromov if the Gromov inner product on XX is δ\delta-hyperbolic with respect to any point of XX.

A metric space is called (Gromov) hyperbolic if it is δ\delta-hyperbolic in the sense of Gromov for some δ0\delta\geq 0.

However, it is a standard fact that for geodesic metric spaces the two concepts are equivalent. See [Gro87, Section 6.3C], or [BH99, Proposition 1.22, Chapter III.H] for instance. In this subsection we observe an analog of Rips hyperbolicity in Gromov hyperbolic length spaces using the next two lemmas.

The following lemma is a crucial property of Rips hyperbolic metric spaces.

Lemma 2.17.

(Stability of quasigeodesics in a Rips hyperbolic space, [Gd90]) For all δ0\delta\geq 0 and k1k\geq 1, ϵ0\epsilon\geq 0 there is a constant D2.17=D2.17(δ,k,ϵ)D_{\ref{stab-qg}}=D_{\ref{stab-qg}}(\delta,k,\epsilon) such that the following holds:

Suppose YY is a geodesic metric space δ\delta-hyperbolic in the sense of Rips. Then the Hausdorff distance between a geodesic and a (k,ϵ)(k,\epsilon)-quasigeodesic joining the same pair of end points is less than or equal to D2.17D_{\ref{stab-qg}}.

One is referred to [V0̈5, Theorem 3.18, Theorem 3.20] for a proof of the following lemma.

Lemma 2.18.

Suppose XX is a metric space which is δ\delta-hyperbolic in the sense of Gromov. If f:XYf:X\rightarrow Y is a RR-coarsely surjective, (1,C)(1,C)-quasiisometry then YY is D=D2.18(δ,R,C)D=D_{\ref{qi vs gromov hyp}}(\delta,R,C)-hyperbolic in the sense of Gromov.

Using metric graph approximations to length spaces (Lemma 2.8) and the fact that for geodesic metric spaces Gromov hyperbolicity implies Rips hyperbolicity we obtain the following three corollaries.

Corollary 2.19.

(Stability of quasigeodesics in a Gromov hyperbolic space) Given δ0,k1,ϵ0\delta\geq 0,k\geq 1,\epsilon\geq 0 there is D=D2.19(δ,k,ϵ)D=D_{\ref{cor: stab-qg}}(\delta,k,\epsilon) such that the following holds.

Suppose XX is metric space which is δ\delta-hyperbolic in the sense of Gromov. Then given (k,ϵ)(k,\epsilon)-quasigeodesics γi\gamma_{i}, i=1,2i=1,2 with the same end points we have Hd(γ1,γ2)DHd(\gamma_{1},\gamma_{2})\leq D.

Corollary 2.20.

(Analog of Rips hyperbolicity for length spaces) Suppose XX is a length space. If XX is δ\delta-hyperbolic in the sense of Gromov then for all K1K\geq 1, ϵ0\epsilon\geq 0 all (K,ϵ)(K,\epsilon)-quasigeodesic triangles in XX are D2.20=D2.20(δ,K,ϵ)D_{\ref{slim iff gromov}}=D_{\ref{slim iff gromov}}(\delta,K,\epsilon)-slim.

Conversely if all (K,ϵ)(K,\epsilon)-quasigeodesic triangles in XX are RR-slim for some R0R\geq 0 and for some sufficiently large K,ϵK,\epsilon then XX is λ2.20=λ2.20(R,K,ϵ)\lambda_{\ref{slim iff gromov}}=\lambda_{\ref{slim iff gromov}}(R,K,\epsilon)-hyperbolic in the sense of Gromov.

Slimness of triangles immediately implies slimness of polygons:

Corollary 2.21.

(Slimness of polygons) Suppose that XX is a length space. If XX is δ\delta-hyperbolic in the sense of Gromov then for all K1K\geq 1, ϵ0\epsilon\geq 0 all (K,ϵ)(K,\epsilon)-quasigeodesic nn-gons in XX are (n2)D2.20=(n2)D2.20(δ,K,ϵ)(n-2)D_{\ref{slim iff gromov}}=(n-2)D_{\ref{slim iff gromov}}(\delta,K,\epsilon)-slim.

Convention 2.22.

For the rest of the paper a δ\delta-hyperbolic (or simply hyperbolic) space will refer either to (1) a δ\delta-hyperbolic (resp. hyperbolic) space in the sense of Rips if it is a geodesic metric space or (2) a δ\delta-hyperbolic (resp. hyperbolic) space in the sense of Gromov if it is not a geodesic metric space. However, in this case the space will be assumed to be a length space. The constant δ\delta will be referred to as the hyperbolicity constant for the space involved.

2.3. Quasiconvex subspaces of hyperbolic spaces

Definition 2.23.

Let XX be a hyperbolic geodesic metric space and let AXA\subseteq X. For K0K\geq 0, we say that AA is KK-quasiconvex in XX if any geodesic with end points in AA is contained in NK(A)N_{K}(A).
If XX is a Gromov hyperbolic length space and AXA\subset X then we will say that AA is KK-quasiconvex if any (1,1)(1,1)-quasigeodesic joining a pair of points of AA is contained in NK(A)N_{K}(A).
A subset AXA\subset X is said to be quasiconvex if it is KK-quasiconvex for some K0K\geq 0.

The following lemma relates quasiconvexity with qi embedding. It is straightforward and is proved in the context of geodesic metric spaces in [KS, Chapter 1, section 1.11]. Hence we skip the proof.

Lemma 2.24.

(1) Given δ0\delta\geq 0 and k0k\geq 0 there are constants D=D(δ,k)D=D(\delta,k) and K=K(δ,k)K=K(\delta,k) such that the following holds:

Suppose XX is a δ\delta-hyperbolic metric space and AXA\subset X is kk-quasiconvex. Then ND(A)N_{D}(A) is path connected and with respect to the induced path metric on ND(A)N_{D}(A) from XX the inclusion map ND(A)XN_{D}(A)\rightarrow X is a KK-qi embedding.

(2) Suppose XX is a hyperbolic metric space and YY is a quasiconvex subset. Suppose YY is path connected and with respect to the induced path metric on YY from XX the inclusion map YXY\rightarrow X is metrically proper. Then the inclusion map is a qi embedding.

In this subsection, in a Gromov hyperbolic setting, we prove a number of results about quasiconvex sets analogous to those in [MS12, Section 1.2] which were proved in a Rips hyperbolic setting. The importance of the following lemma for this paper can be hardly exaggerated.

Lemma 2.25.

(Projection on a quasiconvex set) Let XX be a δ\delta-hyperbolic metric space, UXU\subset X is a KK-quasiconvex set and ϵ0\epsilon\geq 0. Suppose yUy\in U is an ϵ\epsilon-approximate nearest point projection of a point xXx\in X on UU. Let zUz\in U. Suppose α\alpha is a (dotted) kk-quasigeodesic joining xx to yy and β\beta is a (dotted) kk-quasigeodesic joining yy to zz. Then αβ\alpha*\beta is a (dotted) K2.25=K2.25(δ,K,k,ϵ)K_{\ref{subqc-elem}}=K_{\ref{subqc-elem}}(\delta,K,k,\epsilon)-quasigeodesic in XX.

In particular, if γ\gamma is kk-quasigeodesic joining x,zx,z then yy is contained in the D2.25(δ,K,k,ϵ)D_{\ref{subqc-elem}}(\delta,K,k,\epsilon)-neighborhood of γ\gamma.

Proof.

Without loss of generality we shall assume that XX is a δ\delta-hyperbolic length space. Suppose β1\beta_{1} is a (1,1)(1,1)-quasigeodesic in XX joining y,zy,z. Since UU is KK-quasiconvex it is clear that yy is an (ϵ+K)(\epsilon+K)-approximate nearest point projection of xx on β1\beta_{1}. Hence, if α1\alpha_{1} is a (1,1)(1,1)-quasigeodesic joining x,yx,y then α1β1\alpha_{1}*\beta_{1} is a (3,3+ϵ+K)(3,3+\epsilon+K)-quasigeodesic in XX by Corollary 2.14. By stability of quasigeodesics Hd(α,α1)D2.19(δ,k,ϵ)Hd(\alpha,\alpha_{1})\leq D_{\ref{cor: stab-qg}}(\delta,k,\epsilon), and Hd(β,β1)D2.19(δ,k,ϵ)Hd(\beta,\beta_{1})\leq D_{\ref{cor: stab-qg}}(\delta,k,\epsilon). Hence, Hd(αβ,α1β1)D2.19(δ,k,ϵ)Hd(\alpha*\beta,\alpha_{1}*\beta_{1})\leq D_{\ref{cor: stab-qg}}(\delta,k,\epsilon). By Lemma 2.5 it is enough to show now that γ=αβ\gamma=\alpha*\beta is uniformly properly embedded. Let γ1=α1β1\gamma_{1}=\alpha_{1}*\beta_{1} and R=D2.19(δ,k,ϵ)R=D_{\ref{cor: stab-qg}}(\delta,k,\epsilon). Suppose α:[0,l]X\alpha:[0,l]\rightarrow X with α(0)=x,α(l)=y\alpha(0)=x,\alpha(l)=y and β:[0,m]X\beta:[0,m]\rightarrow X with β(0)=y,β(m)=z\beta(0)=y,\beta(m)=z. Let st[0,l+m]s\leq t\in[0,l+m] and d(γ(s),γ(t))Dd(\gamma(s),\gamma(t))\leq D for some D0D\geq 0. We need find a constant D1D_{1} such that tsD1t-s\leq D_{1} where D1D_{1} depends on δ,k,K\delta,k,K and DD only. However, if s,t[0,l]s,t\in[0,l] or s,t[l,l+m]s,t\in[l,l+m] then we have k+(ts)/kD-k+(t-s)/k\leq D since both α,β\alpha,\beta are kk-quasigeodesics. Hence, in that case tsk2+kDt-s\leq k^{2}+kD. Suppose s[0,l)s\in[0,l) and t(l,m]t\in(l,m]. In this case γ(s)=α(s),γ(t)=β(tl)\gamma(s)=\alpha(s),\gamma(t)=\beta(t-l). Let xα1,yβ1x^{\prime}\in\alpha_{1},y^{\prime}\in\beta_{1} be such that d(x,γ(s))Rd(x^{\prime},\gamma(s))\leq R and d(y,γ(t))Rd(y^{\prime},\gamma(t))\leq R. Then d(x,y)2R+Dd(x^{\prime},y^{\prime})\leq 2R+D. Suppose γ1(s)=x,γ1(t)=y,γ1(u)=y\gamma_{1}(s^{\prime})=x^{\prime},\gamma_{1}(t^{\prime})=y^{\prime},\gamma_{1}(u)=y where suts^{\prime}\leq u\leq t^{\prime}. Since γ1\gamma_{1} is a (3,3+ϵ+K)(3,3+\epsilon+K)-quasigeodesic we have |st|3(3+ϵ+K)+3d(x,y)3(3+ϵ+K)+3(2R+D)|s^{\prime}-t^{\prime}|\leq 3(3+\epsilon+K)+3d(x^{\prime},y^{\prime})\leq 3(3+\epsilon+K)+3(2R+D). It follows that |su||s^{\prime}-u| and |ut||u-t^{\prime}| are both at most 3(3+ϵ+K)+3(2R+D)=9+3ϵ+3K+6R+3D3(3+\epsilon+K)+3(2R+D)=9+3\epsilon+3K+6R+3D. Hence, d(x,y),d(y,y)d(x^{\prime},y),d(y,y^{\prime}) are both at most 3(9+3ϵ+3K+6R+3D)+3+ϵ+K=30+10ϵ+10K+18R+9D=D3(9+3\epsilon+3K+6R+3D)+3+\epsilon+K=30+10\epsilon+10K+18R+9D=D^{\prime}, say. Hence, d(γ(s),y)d(\gamma(s),y), d(y,γ(t))d(y,\gamma(t)) are both at most R+DR+D^{\prime}. Since α,β\alpha,\beta are kk-quasigeodesics it follows that lsl-s and tlt-l are both at most k2+k(R+D)k^{2}+k(R+D^{\prime}). Hence, ts2(k2+k(R+D))t-s\leq 2(k^{2}+k(R+D^{\prime})). Hence, we can take D1=2k2+2kR+2kDD_{1}=2k^{2}+2kR+2kD^{\prime}. This completes the proof of the existence of K2.25K_{\ref{subqc-elem}}.

Clearly one can set D2.25(δ,K,k,ϵ)=D2.19(δ,K2.25(δ,K,k,k),K2.25(δ,K,k,k))D_{\ref{subqc-elem}}(\delta,K,k,\epsilon)=D_{\ref{cor: stab-qg}}(\delta,K_{\ref{subqc-elem}}(\delta,K,k,k),K_{\ref{subqc-elem}}(\delta,K,k,k)). ∎

Corollary 2.26.

Suppose XX is a δ\delta-hyperbolic metric space and α\alpha is a kk-quasigeodesic in XX with an end point yy. Suppose xXx\in X and yy is an ϵ\epsilon-approximate nearest point projection of xx on α\alpha. Suppose β\beta is a kk-quasigeodesic joining xx to yy. Then βα\beta*\alpha is a K2.26(δ,k,ϵ)K_{\ref{gluing quasigeodesics}}(\delta,k,\epsilon)-quasigeodesic.

Proof.

We briefly indicate the proof. One first notes by stability of quasigeodesics that images of uniform quasigeodesics are uniformly quasiconvex. Then one applies the preceding lemma. ∎

The following corollary easily follows from Lemma 2.25 and Lemma 2.13. For instance, the proof is similar to that of [MS12, Lemma 1.32].

Corollary 2.27.

(Projection on nested quasiconvex sets) Suppose XX is a δ\delta-hyperbolic metric space and VUV\subset U are two KK-quasiconvex subsets of XX. Suppose xXx\in X and x1Ux_{1}\in U, x2Vx_{2}\in V are ϵ\epsilon-approximate nearest point projection of xx on UU and VV respectively. Suppose x3x_{3} is an ϵ\epsilon-approximate nearest point projection of x1x_{1} on VV. Then d(x2,x3)D2.27(δ,K,ϵ)d(x_{2},x_{3})\leq D_{\ref{nested qc sets}}(\delta,K,\epsilon).

In particular, for any two ϵ\epsilon-approximate nearest point projections x1,x2x_{1},x_{2} of xx on UU we have d(x1,x2)D2.27(δ,K,ϵ)d(x_{1},x_{2})\leq D_{\ref{nested qc sets}}(\delta,K,\epsilon).

Corollary 2.28.

Given δ0,K0,ϵ0\delta\geq 0,K\geq 0,\epsilon\geq 0 there are constants L=L2.28(δ,K,ϵ)L=L_{\ref{cor: lip proj}}(\delta,K,\epsilon), D=D2.28(δ,K,ϵ)D=D_{\ref{cor: lip proj}}(\delta,K,\epsilon) and R=R2.28(δ,K,ϵ)R=R_{\ref{cor: lip proj}}(\delta,K,\epsilon) such that the following hold:

(1) Suppose XX is a δ\delta-hyperbolic metric space and UU is a KK-quasiconvex subset of XX. Then for all ϵ0\epsilon\geq 0 any ϵ\epsilon-approximate nearest point projection map P:XUP:X\rightarrow U is coarsely LL-Lipschitz.

(2) Suppose VV is another KK-quasiconvex subset of XX and v1,v2Vv_{1},v_{2}\in V and ui=P(vi)u_{i}=P(v_{i}), i=1,2i=1,2. If d(u1,u2)Dd(u_{1},u_{2})\geq D then u1,u2NR(V)u_{1},u_{2}\in N_{R}(V).

In particular, if the diameter of P(V)P(V) is at least DD then d(U,V)Rd(U,V)\leq R.

Proof.

(1) Suppose x,yXx,y\in X with d(x,y)1d(x,y)\leq 1. Then P(x)P(x) is an (ϵ+1)(\epsilon+1)-approximate nearest point projection of yy on UU. Hence, by Corollary 2.27 we have d(P(x),P(y))D2.27(δ,K,ϵ+1)d(P(x),P(y))\leq D_{\ref{nested qc sets}}(\delta,K,\epsilon+1). Hence, we may take L2.28(δ,K,ϵ)=D2.27(δ,K,ϵ+1)L_{\ref{cor: lip proj}}(\delta,K,\epsilon)=D_{\ref{nested qc sets}}(\delta,K,\epsilon+1) by Lemma 2.6.

(2) Consider the quadrilateral formed by (1,1)(1,1)-quasigeodesics joining the pairs (u1,u2),(u2,v2),(v2,v1)(u_{1},u_{2}),(u_{2},v_{2}),(v_{2},v_{1}) and (v1,u1)(v_{1},u_{1}). This is 2D2.20(δ,1,1)2D_{\ref{slim iff gromov}}(\delta,1,1)-slim by Corollary 2.21. Let δ=2D2.20(δ,1,1)\delta^{\prime}=2D_{\ref{slim iff gromov}}(\delta,1,1). Suppose no point of the side v1v2v_{1}v_{2} is contained in a δ\delta^{\prime}-neighborhood of the side u1u2u_{1}u_{2}. Then there are two points say x1,x2v1v2x_{1},x_{2}\in v_{1}v_{2} such that xiNδ(uivi)x_{i}\in N_{\delta^{\prime}}(u_{i}v_{i}), i=1,2i=1,2 and d(x1,x2)2d(x_{1},x_{2})\leq 2. Hence there are points yiuiviy_{i}\in u_{i}v_{i}, i=1,2i=1,2 such that d(y1,y2)2+2δd(y_{1},y_{2})\leq 2+2\delta^{\prime}. However, uiu_{i} is an (ϵ+3)(\epsilon+3)-approximate nearest point projection of yiy_{i} on UU by Lemma 2.13. Hence, by the first part of the Corollary 2.28 we have d(u1,u2)L2.28(δ,K,ϵ+3)+(2+2δ)L2.28(δ,K,ϵ+3)d(u_{1},u_{2})\leq L_{\ref{cor: lip proj}}(\delta,K,\epsilon+3)+(2+2\delta^{\prime})L_{\ref{cor: lip proj}}(\delta,K,\epsilon+3). Hence, if the diameter of P(V)P(V) is bigger than D=L2.28(δ,K,ϵ+3)+(2+2δ)L2.28(δ,K,ϵ+3)D=L_{\ref{cor: lip proj}}(\delta,K,\epsilon+3)+(2+2\delta^{\prime})L_{\ref{cor: lip proj}}(\delta,K,\epsilon+3) then there is a point xv1v2x\in v_{1}v_{2} and yu1u2y\in u_{1}u_{2} such that d(x,y)δd(x,y)\leq\delta^{\prime}. Since UU is KK-quasiconvex we have thus xNK+δ(U)x\in N_{K+\delta^{\prime}}(U). Thus we may choose R=K+δR=K+\delta^{\prime}. ∎

The second part of the above corollary is implied in Lemma 1.35 of [MS12] too. The next lemma roughly says that the nearest point projection of a quasigeodesic on a quasiconvex set is close to a quasigeodesic.

Lemma 2.29.

Given K0,R0,δ0K\geq 0,R\geq 0,\delta\geq 0 there is a constant D=D2.29(R,K,δ)D=D_{\ref{trivial lemma}}(R,K,\delta) such that the following holds:
Suppose XX is a δ\delta-hyperbolic metric space and AA is a KK-quasiconvex subset of XX. Suppose x,yXx,y\in X and x¯,y¯A\bar{x},\bar{y}\in A respectively are their 11-approximate nearest point projections on AA. Let [x,y],[x¯,y¯][x,y],[\bar{x},\bar{y}] denote 11-quasigeodesics in XX joining x,yx,y and x¯,y¯\bar{x},\bar{y} respectively. Suppose z[x,y]z\in[x,y] and z¯\bar{z} is a 11-approximate nearest point projection of zz on AA and d(z,z¯)Rd(z,\bar{z})\leq R. Then d(z,[x¯,y¯])Dd(z,[\bar{x},\bar{y}])\leq D.

Proof.

By Corollary 2.21 quadrilaterals in XX formed by 11-quasigeodesics are 2D2.20(δ,1,1)2D_{\ref{slim iff gromov}}(\delta,1,1)-slim. Hence, there is z[x,x¯][x¯,y¯][y,y¯]z^{\prime}\in[x,\bar{x}]\cup[\bar{x},\bar{y}]\cup[y,\bar{y}] such that d(z,z)2D2.20(δ,1,1)d(z,z^{\prime})\leq 2D_{\ref{slim iff gromov}}(\delta,1,1). If z[x¯,y¯]z^{\prime}\in[\bar{x},\bar{y}] then we are done. Suppose not. Without loss of generality let us assume that z[x,x¯]z^{\prime}\in[x,\bar{x}]. Then d(z,A)d(z,z)+d(z,A)2D2.20(δ,1,1)+Rd(z^{\prime},A)\leq d(z,z^{\prime})+d(z,A)\leq 2D_{\ref{slim iff gromov}}(\delta,1,1)+R. Since x¯\bar{x} is a 11-approximate nearest point projection of xx on AA, x¯\bar{x} is a 44-approximate nearest point projection of zz^{\prime} on AA by Lemma 2.13. Hence, by Corollary 2.28, d(x¯,z¯)L2.28(δ,K,4)d(z,z)L2.28(δ,K,4)(2D2.20(δ,1,1)+R)d(\bar{x},\bar{z})\leq L_{\ref{cor: lip proj}}(\delta,K,4)d(z^{\prime},z)\leq L_{\ref{cor: lip proj}}(\delta,K,4)(2D_{\ref{slim iff gromov}}(\delta,1,1)+R). But d(z,z¯)Rd(z,\bar{z})\leq R. Hence, d(z,x¯)R+L2.28(δ,K,4)(2D2.20(δ,1,1)+R)d(z,\bar{x})\leq R+L_{\ref{cor: lip proj}}(\delta,K,4)(2D_{\ref{slim iff gromov}}(\delta,1,1)+R). Thus we can take D2.29(R,K,δ)=max{2D2.20(δ,1,1),R+L2.28(δ,K,4)(2D2.20(δ,1,1)+R)}D_{\ref{trivial lemma}}(R,K,\delta)=\max\{2D_{\ref{slim iff gromov}}(\delta,1,1),R+L_{\ref{cor: lip proj}}(\delta,K,4)(2D_{\ref{slim iff gromov}}(\delta,1,1)+R)\}. ∎

The following lemma asserts that quasiconvexity and nearest point projections are preserved under qi embeddings.

Lemma 2.30.

Suppose XX is a δ\delta-hyperbolic metric graph and YXY\subset X is a connected sub-graph such that the inclusion (Y,dY)(X,dX)(Y,d_{Y})\rightarrow(X,d_{X}) is a kk-qi embedding. Suppose AYA\subset Y is KK-quasiconvex in YY. Then the following holds.

(1) AA is K2.30(δ,k,K)K_{\ref{qc in subspace}}(\delta,k,K)-quasiconvex in XX.

(2) For any xYx\in Y if x1,x2Ax_{1},x_{2}\in A are the nearest point projections of xx on AA in YY and XX respectively then dY(x1,x2)D2.30(δ,k,K)d_{Y}(x_{1},x_{2})\leq D_{\ref{qc in subspace}}(\delta,k,K).

Proof.

(1) Suppose x,yAx,y\in A and let α,β\alpha,\beta be geodesics joining x,yx,y in YY and XX respectively. Since, YY is kk-qi embedded α\alpha is a (k,k)(k,k)-quasigeodesic in XX by Lemma 2.3. Hence, by stability of quasigeodesics Hd(α,β)D2.17(δ,k,k)Hd(\alpha,\beta)\leq D_{\ref{stab-qg}}(\delta,k,k). However, AA being KK-quasiconvex in YY, αNK(A)\alpha\subset N_{K}(A) in YY and hence in XX as well. Thus βNK+D2.17(δ,k,k)(A)\beta\subset N_{K+D_{\ref{stab-qg}}(\delta,k,k)}(A) in XX. Hence, we can take K2.30(δ,k,K)=K+D2.17(δ,k,k)K_{\ref{qc in subspace}}(\delta,k,K)=K+D_{\ref{stab-qg}}(\delta,k,k).

(2) Suppose K1=K2.30(δ,k,K)K_{1}=K_{\ref{qc in subspace}}(\delta,k,K). Then x2ND([x,x1]X)x_{2}\in N_{D}([x,x_{1}]_{X}) in XX where D=D2.25(δ,K1,1,1)D=D_{\ref{subqc-elem}}(\delta,K_{1},1,1). We have Hd([x,x1]Y,[x,x1]X)D2.17(δ,k,k)Hd([x,x_{1}]_{Y},[x,x_{1}]_{X})\leq D_{\ref{stab-qg}}(\delta,k,k) by stability of quasigeodesics. Thus there is a point x2[x,x1]Yx^{\prime}_{2}\in[x,x_{1}]_{Y} such that dX(x2,x2)D+D2.17(δ,k,k)=D1d_{X}(x_{2},x^{\prime}_{2})\leq D+D_{\ref{stab-qg}}(\delta,k,k)=D_{1}, say. Then dY(x2,x2)k(D1+k)d_{Y}(x_{2},x^{\prime}_{2})\leq k(D_{1}+k) since YY is kk-qi embedded in XX. Since x1x_{1} is a nearest point projection of xx on AA in YY, it is also a nearest point projection of x2x^{\prime}_{2} on AA in YY. Hence, dY(x2,x1)dY(x2,x2)k(D1+k)d_{Y}(x^{\prime}_{2},x_{1})\leq d_{Y}(x^{\prime}_{2},x_{2})\leq k(D_{1}+k). Hence, dY(x1,x2)2k(D1+k)d_{Y}(x_{1},x_{2})\leq 2k(D_{1}+k) by triangle inequality. Thus we can take D2.30(δ,k,K)=2k(D1+k)D_{\ref{qc in subspace}}(\delta,k,K)=2k(D_{1}+k). ∎

Definition 2.31.

Suppose XX is a δ\delta-hyperbolic metric space and A,BA,B are two quasiconvex subsets. Let R>0R>0. We say that A,BA,B are mutually RR-cobounded, or simply RR-cobounded, if the set of all 11-approximate nearest point projections of the points of AA on BB has a diameter at most RR and vice versa.

When the constant RR is understood or is not important we just say that A,BA,B are cobounded.

The following corollary is an immediate consequence of Corollary 2.28(2).

Corollary 2.32.

([MS12, Lemma 1.35]) Given δ0,k0\delta\geq 0,k\geq 0 there are constants D=D2.32(δ,k)D=D_{\ref{cobounded cor}}(\delta,k) and R=R2.32(δ,k)R=R_{\ref{cobounded cor}}(\delta,k) such that the following holds.

Suppose XX is a δ\delta-hyperbolic metric space and A,BXA,B\subset X are two kk-quasiconvex subsets. If d(A,B)Dd(A,B)\geq D then A,BA,B are mutually RR-cobounded.

The following proposition and its proof are motivated by an analogous result due to Hamenstadt ([Ham05, Lemma 3.5]). See also [MS12, Corollary 1.52]. Before we state the proposition let us explain the set-up.

(P 0)(P\,0) Suppose XX is a δ\delta-hyperbolic metric graph and YXY\subset X is a KK-quasiconvex subgraph, for some δ0,K0\delta\geq 0,K\geq 0. Suppose II is an interval in {\mathbb{R}} with end points in {,}\mathbb{Z}\cup\{\infty,-\infty\} and Π:YI\Pi:Y\rightarrow I is a map such that IΠ(Y)I\cap{\mathbb{Z}}\subset\Pi(Y). Let Yi:=Π1(i)Y_{i}:=\Pi^{-1}(i) for all iIi\in I\cap\mathbb{Z} and Yij=Π1([i,j])Y_{ij}=\Pi^{-1}([i,j]) for all i,jIi,j\in I\cap\mathbb{Z} with i<ji<j such that the following hold.

(P 1)(P\,1) All the sets YiY_{i} and YijY_{ij}, i,jIi,j\in I, i<ji<j are KK-quasiconvex in XX.

(P 2)(P\,2) YiY_{i} uniformly coarsely bisects YY into Yi:=Π1((,i]I)Y^{-}_{i}:=\Pi^{-1}((-\infty,i]\cap I) and Yi+:=Π1([i,)I)Y^{+}_{i}:=\Pi^{-1}([i,\infty)\cap I) for all iIi\in I. Let R0R\geq 0 be such that any geodesic in YY joining Yi+Y^{+}_{i} and YiY^{-}_{i} passes through NR(Yi)N_{R}(Y_{i}) for all iIi\in I\cap{\mathbb{Z}}.

(P 3)(P\,3) d(Yii+1,Yjj+1)>2K+1d(Y_{ii+1},Y_{jj+1})>2K+1 for all i,jIi,j\in I if j+1Ij+1\in I and i+1<ji+1<j.

(P 4)(P\,4) There is D0D\geq 0 such that the sets YiY_{i} and YjY_{j} are DD-cobounded in XX for all i,jIi,j\in I\cap{\mathbb{Z}} with i<ji<j unless j=i+1j=i+1 and i,ji,j are the end points of II.

The proposition below is about a description of uniform quasigeodesics in XX joining points of YY.

Proposition 2.33.

Given δ0,K0,D0\delta\geq 0,K\geq 0,D\geq 0, λ1\lambda\geq 1, ϵ1\epsilon\geq 1 and R0R\geq 0 there are λ=λ2.33(δ,K,D,λ,ϵ,R)1\lambda^{\prime}=\lambda_{\ref{hamenstadt}}(\delta,K,D,\lambda,\epsilon,R)\geq 1 and μ2.33=μ2.33(δ,K,D,ϵ,R)0\mu_{\ref{hamenstadt}}=\mu_{\ref{hamenstadt}}(\delta,K,D,\epsilon,R)\geq 0 such that the following holds.

Suppose we have the aforementioned hypotheses (P 0)(P\,0), (P 1)(P\,1), (P 2)(P\,2), (P 3)(P\,3) and (P 4)(P\,4). Suppose m,nIm,n\in I\cap{\mathbb{Z}} and yYm,yYny\in Y_{m},y^{\prime}\in Y_{n}. Suppose yiYy_{i}\in Y, minm\leq i\leq n are defined as follows: ym=yy_{m}=y, yi+1y_{i+1} is an ϵ\epsilon-approximate nearest point projection of yiy_{i} on Yi+1Y_{i+1} for min1m\leq i\leq n-1. Suppose αiYii+1\alpha_{i}\subset Y_{ii+1} is a λ\lambda-quasigeodesic in XX joining yiy_{i} and yi+1y_{i+1}, min1m\leq i\leq n-1 and β\beta is a λ\lambda-quasigeodesic joining yny_{n} and yy^{\prime}.

Then the concatenation of the all the αi\alpha_{i}’s and β\beta is a λ\lambda^{\prime}-quasigeodesic in XX joining y,yy,y^{\prime}. Moreover, each yiy_{i} is an μ2.33\mu_{\ref{hamenstadt}}-approximate nearest point projection of yy on YiY_{i} for m+2inm+2\leq i\leq n.

Proof.

The proof is broken into the following three claims. In course of the proof we shall denote the concatenation of the αi\alpha_{i}’s and β\beta by α\alpha.

Claim 1: Suppose xYix\in Y^{-}_{i} for some ii. Let x¯\bar{x} be an ϵ\epsilon-approximate nearest point projection of xx on YiY_{i}. Then x¯\bar{x} is an ϵ\epsilon^{\prime}-approximate nearest point projection of xx on Yi+Y^{+}_{i} where ϵ\epsilon^{\prime} depends only on ϵ\epsilon and the parameters δ,D,K\delta,D,K and RR.

Proof of Claim 1: Suppose xx^{\prime} is a 11-approximate nearest point projection of xx on Yi+Y^{+}_{i}. Since Yi+Y^{+}_{i} is KK-quasiconvex [x,x][x,x¯][x,x^{\prime}]*[x^{\prime},\bar{x}] is a K2.25(δ,K,1,1)K_{\ref{subqc-elem}}(\delta,K,1,1)-quasigeodesic by Lemma 2.25. Let k1=K2.25(δ,K,1,1)k_{1}=K_{\ref{subqc-elem}}(\delta,K,1,1). Then by stability of quasigeodesics there is a point z[x,x¯]z\in[x,\bar{x}] such that d(x,z)D2.17(δ,k1)=D1d(x^{\prime},z)\leq D_{\ref{stab-qg}}(\delta,k_{1})=D_{1}, say. We claim that zz is uniformly close to YiY_{i}. Since YiY^{-}_{i} is KK-quasiconvex there is a point wYiw\in Y^{-}_{i} such that d(z,w)Kd(z,w)\leq K. It follows that d(w,x)D1+Kd(w,x^{\prime})\leq D_{1}+K. By (P 2)(P\,2), there is a point z1[w,x]z_{1}\in[w,x^{\prime}] such that d(z1,Yi)Rd(z_{1},Y_{i})\leq R. Since, d(z1,w)d(w,x)D1+Kd(z_{1},w)\leq d(w,x^{\prime})\leq D_{1}+K and d(w,z)Kd(w,z)\leq K it follows by triangle inequality that d(z,Yi)2K+D1+Rd(z,Y_{i})\leq 2K+D_{1}+R. Now, by Lemma 2.13 x¯\bar{x} is an (ϵ+3)(\epsilon+3)-approximate nearest point projection of zz on YiY_{i}. Hence, d(x,x¯)d(x,z)+d(z,x¯)D1+ϵ+3+d(z,Yi)d(x^{\prime},\bar{x})\leq d(x^{\prime},z)+d(z,\bar{x})\leq D_{1}+\epsilon+3+d(z,Y_{i}). It follows that ϵ=3+ϵ+2K+2D1+R\epsilon^{\prime}=3+\epsilon+2K+2D_{1}+R works.

Note: We shall use D1D_{1} again in the proof of Claim 33 to denote the same constant as in the proof of Claim 11 above.

Claim 2. Next we claim that for all m+2in1m+2\leq i\leq n-1 there is uniformly bounded set AiYiA_{i}\subset Y_{i} such that ϵ\epsilon-nearest point projection of any point of YjY^{-}_{j}, j<ij<i on YiY_{i} is contained in AiA_{i}.

Proof of Claim 2: Consider any YiY_{i}, m+2in1m+2\leq i\leq n-1. Let BiYiB_{i}\subset Y_{i} be the set of all 11-approximate nearest point projections of points of Yi1Y_{i-1} on YiY_{i} in XX. Then the diameter of BiB_{i} is at most DD by (P 4)(P\,4). Suppose xYjx\in Y^{-}_{j}, j<ij<i. Let x1,x2x_{1},x_{2} be respectively ϵ\epsilon-approximate nearest point projections of xx on Yi1Y_{i-1} and YiY_{i} respectively. Let x3x_{3} be an ϵ\epsilon-nearest point projection of x1x_{1} on YiY_{i}. Now, by Step 1 x1x_{1} is an ϵ\epsilon^{\prime}-approximate nearest point projection of xx on Yi1+Y^{+}_{i-1} and x2x_{2}, x3x_{3} are ϵ\epsilon^{\prime}-approximate nearest point projection of xx and x1x_{1} respectively on Yi+Y^{+}_{i}. Therefore, by the first part of Corollary 2.27 we have d(x2,x3)D2.27(δ,K,ϵ)d(x_{2},x_{3})\leq D_{\ref{nested qc sets}}(\delta,K,\epsilon^{\prime}). However, if x1Bix^{\prime}_{1}\in B_{i} is a 11-approximate nearest point projection of x1x_{1} on YiY_{i} then by the second part of the Corollary 2.27 we have d(x3,Bi)d(x3,x1)D2.27(δ,K,ϵ)d(x_{3},B_{i})\leq d(x_{3},x^{\prime}_{1})\leq D_{\ref{nested qc sets}}(\delta,K,\epsilon) since ϵ1\epsilon\geq 1. Hence, d(x2,Bi)2D2.27(δ,K,ϵ)d(x_{2},B_{i})\leq 2D_{\ref{nested qc sets}}(\delta,K,\epsilon). Therefore, we can take Ai=N2D2.27(δ,K,ϵ)(Bi)YiA_{i}=N_{2D_{\ref{nested qc sets}}(\delta,K,\epsilon)}(B_{i})\cap Y_{i}.

Let r=supm+2in1{diam(Ai)}.r=\sup_{m+2\leq i\leq n-1}\{diam(A_{i})\}. We note that rD+2D2.27(δ,K,ϵ)r\leq D+2D_{\ref{nested qc sets}}(\delta,K,\epsilon).

Claim 3. Finally we claim that (1) α\alpha is contained in a uniformly small neighborhood of a geodesic joining y,yy,y^{\prime} and (2) α\alpha is uniformly properly embedded in XX.

We note that the proposition follows from Claim 3 using Lemma 2.5.

Proof of Claim 3: Suppose x,xαx,x^{\prime}\in\alpha, Π(x)<Π(x)\Pi(x)<\Pi(x^{\prime}). Choose smallest k,lk,l such that xαYkk+1,xαYll+1x\in\alpha\cap Y_{kk+1},x^{\prime}\in\alpha\cap Y_{ll+1}, where mklnm\leq k\leq l\leq n. Let γ\gamma be a geodesic in XX joining x,xx,x^{\prime}.

(1) It is enough to show that the segment of α\alpha joining xx to xx^{\prime} is contained in a uniformly small neighborhood of γ\gamma. Hence, without loss of generality k<lk<l. Due to Corollary 2.21 it is enough to prove that the points yiy_{i}, k+1il1k+1\leq i\leq l-1 are contained in a uniformly small neighborhood of γ\gamma in order to show that the segment of α\alpha joining xx to xx^{\prime} is contained in a uniformly small neighborhood of γ\gamma. (We note that the path αn1β\alpha_{n-1}*\beta is a D2.25(δ,K,λ,ϵ)D_{\ref{subqc-elem}}(\delta,K,\lambda,\epsilon)-quasigeodesic joining yn1y_{n-1} and yy^{\prime}.) For this first we note that xx is on αk\alpha_{k}. Let γk\gamma_{k} be a geodesic joining yk,yk+1y_{k},y_{k+1}. Then by stability of quasigeodesics there is a point x1γkx_{1}\in\gamma_{k} such that d(x1,x)D2.19(δ,λ,λ)d(x_{1},x)\leq D_{\ref{cor: stab-qg}}(\delta,\lambda,\lambda). Since yk+1y_{k+1} is an ϵ\epsilon-approximate nearest point projection of yky_{k} on Yk+1Y_{k+1}, by Lemma 2.13 yk+1y_{k+1} is an (ϵ+3)(\epsilon+3)-approximate nearest point projection of x1x_{1} on Yk+1Y_{k+1}. Hence, yk+1y_{k+1} is an (ϵ+3+D2.19(δ,λ,λ))(\epsilon+3+D_{\ref{cor: stab-qg}}(\delta,\lambda,\lambda))-approximate nearest point projection of x1x_{1} on Yk+1Y_{k+1}. Let ϵ1=ϵ+3+D2.19(δ,λ,λ)\epsilon_{1}=\epsilon+3+D_{\ref{cor: stab-qg}}(\delta,\lambda,\lambda). By Step 1 yk+1y_{k+1} is an ϵ1\epsilon^{\prime}_{1}-nearest point projection of xx on Yk+1+Y^{+}_{k+1} where ϵ1=3+ϵ1+2D1++2K+R\epsilon^{\prime}_{1}=3+\epsilon_{1}+2D_{1}++2K+R. Now the concatenation of a geodesic joining yk+1y_{k+1} to xx^{\prime} with the segment of α\alpha from xx to yk+1y_{k+1} is a uniform quasigeodesic by Lemma 2.25. Thus by Corollary 2.19 yk+1y_{k+1} is uniformly close to γ\gamma. On the other hand by Step 2 yiy_{i} is an (ϵ+r)(\epsilon+r)-approximate nearest point projection of xx on YiY_{i} and hence an (ϵ+r)(\epsilon+r)^{\prime}-approximate nearest point projection on Yi+Y^{+}_{i} for all k+2il1k+2\leq i\leq l-1. Hence, again by Lemma 2.25 and Corollary 2.19 yiy_{i} is within a uniformly small neighborhood of γ\gamma. This proves (1).

(2) Suppose L=sup{d(yi,γ):k+1il1}L=\sup\{d(y_{i},\gamma):k+1\leq i\leq l-1\}. Suppose x,xαx,x^{\prime}\in\alpha as above with d(x,x)Nd(x,x^{\prime})\leq N. Once again, without loss of generality k<lk<l. We claim that lk+Nl\leq k+N. To see this consider two adjacent vertices vi,vi+1v_{i},v_{i+1} on γ\gamma. If viNK(Yss+1)v_{i}\in N_{K}(Y_{ss+1}) and vi+1NK(Ytt+1)v_{i+1}\in N_{K}(Y_{tt+1}) with s<ts<t then by the (P 3)(P\,3) we have t=s+1t=s+1. The claim follows from this. Suppose α(sk)=x\alpha(s_{k})=x, α(si)=yi\alpha(s_{i})=y_{i} for k+1il1k+1\leq i\leq l-1 and α(sl)=x\alpha(s_{l})=x^{\prime}. We note that d(α(si),α(si+1))N+2Ld(\alpha(s_{i}),\alpha(s_{i+1}))\leq N+2L for kil1k\leq i\leq l-1. Since lkNl-k\leq N and since the segments of α\alpha joining α(si),α(si+1)\alpha(s_{i}),\alpha(s_{i+1}), kil1k\leq i\leq l-1 are uniform quasigeodesics we are done.

For the second part of the proposition we have already noticed that yiy_{i} is an (ϵ+r)(\epsilon+r)-approximate nearest point projection of any point YjY^{-}_{j}, in particular of yy, on YiY_{i} for all j<ij<i, m+2in1m+2\leq i\leq n-1. On the other hand, yn1y_{n-1} is an (ϵ+r)=(ϵ+r+3+2D1+2K+R)(\epsilon+r)^{\prime}=(\epsilon+r+3+2D_{1}+2K+R)-approximate nearest point projection of yy on Yn1+Y^{+}_{n-1}. Hence, by Corollary 2.27 if yny^{\prime}_{n} is a 11-approximate point projection of yy on YnYn1+Y_{n}\subset Y^{+}_{n-1} then d(yn,yn)D2.27(δ,K,(ϵ+r))d(y^{\prime}_{n},y_{n})\leq D_{\ref{nested qc sets}}(\delta,K,(\epsilon+r)^{\prime}). Thus yny_{n} is an (1+D2.27(δ,K,(ϵ+r)))(1+D_{\ref{nested qc sets}}(\delta,K,(\epsilon+r)^{\prime}))-approximate nearest point projection of yy on YnY_{n}. ∎

Lemma 2.34.

Given δ0,k1,ϵ0\delta\geq 0,k\geq 1,\epsilon\geq 0 there is a constant D=D2.34(δ,k,ϵ)D=D_{\ref{gromov product meaning}}(\delta,k,\epsilon) such that the following is true.

Suppose XX is a δ\delta-hyperbolic metric space. Suppose x1,x2,pXx_{1},x_{2},p\in X and α\alpha is a (k,ϵ)(k,\epsilon)-quasigeodesic in XX joining x1,x2x_{1},x_{2}. Then |(x1.x2)pd(p,α)|D|(x_{1}.x_{2})_{p}-d(p,\alpha)|\leq D.

Proof.

Without loss generality we shall assume that XX is a length space δ\delta-hyperbolic in the sense of Gromov. Let wαw\in\alpha be a 11-approximate nearest point projection of pp on α\alpha. Let β1,β2\beta_{1},\beta_{2} be (1,1)(1,1)-quasigeodesics joining the pairs of points (x1,p),(x2,p)(x_{1},p),(x_{2},p) respectively. Let γ\gamma be a (1,1)(1,1)-quasigeodesic joining p,wp,w and let α\alpha^{\prime} be a (1,1)(1,1)-quasigeodesic joining x1,x2x_{1},x_{2}. Let C=D2.19(δ,k,ϵ+1)C=D_{\ref{cor: stab-qg}}(\delta,k,\epsilon+1). Now, by Corollary 2.19 Hd(α,α)CHd(\alpha,\alpha^{\prime})\leq C and α\alpha is CC-quasiconvex. Let α1\alpha_{1} be the portion of α\alpha from x1x_{1} to ww and let α2\alpha_{2} be the portion of α\alpha from ww to x2x_{2}. Then α1γ\alpha_{1}*\gamma, α2γ\alpha_{2}*\gamma are K=K2.25(δ,C,k+ϵ,k+ϵ)K=K_{\ref{subqc-elem}}(\delta,C,k+\epsilon,k+\epsilon)-quasigeodesics. Hence by Corollary 2.19 Hd(βi,αiγ)D2.19(δ,K,K)Hd(\beta_{i},\alpha_{i}*\gamma)\leq D_{\ref{cor: stab-qg}}(\delta,K,K). Let wiβiw_{i}\in\beta_{i} be such that d(w,wi)D2.19(δ,K,K)d(w,w_{i})\leq D_{\ref{cor: stab-qg}}(\delta,K,K). Since Hd(α,α)CHd(\alpha,\alpha^{\prime})\leq C, there is a point wαw^{\prime}\in\alpha^{\prime} such that d(w,w)Cd(w,w^{\prime})\leq C. Hence, d(w,wi)C+D2.19(δ,K,K)=Rd(w^{\prime},w_{i})\leq C+D_{\ref{cor: stab-qg}}(\delta,K,K)=R, say. Now by Lemma 2.10 |(x1.x2)pd(p,w)|3+2R|(x_{1}.x_{2})_{p}-d(p,w^{\prime})|\leq 3+2R. It follows that |(x1.x2)pd(p,w)|3+2R+C|(x_{1}.x_{2})_{p}-d(p,w)|\leq 3+2R+C. Since ww is a 11-approximate nearest point projection of pp on α\alpha we have for all zαz\in\alpha, d(p,w)d(p,z)+1d(p,w)\leq d(p,z)+1. Thus |d(p,α)d(p,w)|1|d(p,\alpha)-d(p,w)|\leq 1. Hence, |(x1.x2)pd(p,α)|4+2R+C|(x_{1}.x_{2})_{p}-d(p,\alpha)|\leq 4+2R+C. ∎

2.4. Boundaries of hyperbolic spaces and CT maps

Given a hyperbolic metric space, there are the following three standard ways to define a boundary. Some of the results in this subsection are mentioned without proof. One may refer to [BH99] and [ABC+91] for details.

Definition 2.35.
  1. (1)

    Geodesic boundary. Suppose XX is a (geodesic) hyperbolic metric space. Let 𝒢\mathcal{G} denote the set of all geodesic rays in XX. The geodesic boundary X\partial X of XX is defined to be 𝒢/\mathcal{G}/\sim where \sim is the equivalence relation on 𝒢\mathcal{G} defined by setting αβ\alpha\sim\beta iff Hd(α,β)<Hd(\alpha,\beta)<\infty.

  2. (2)

    Quasigeodesic boundary. Suppose XX is a hyperbolic metric space in the sense of Gromov. Let 𝒬\mathcal{Q} be the set of all quasigeodesic rays in XX. Then the quasigeodesic boundary qX\partial_{q}X is defined to be 𝒬/\mathcal{Q}/\sim where \sim is defined as above.

  3. (3)

    Gromov boundary or sequential boundary. Suppose XX is a hyperbolic metric space in the sense of Gromov and pXp\in X. Let 𝒮\mathcal{S} be the set of all sequences {xn}\{x_{n}\} in XX such that limi,j(xi.xj)p=\lim_{i,j\rightarrow\infty}(x_{i}.x_{j})_{p}=\infty. All such sequences are said to converge to infinity. On 𝒮\mathcal{S} we define an equivalence relation where {xn}{yn}\{x_{n}\}\sim\{y_{n}\} if and only if limi,j(xi.yj)p=\lim_{i,j\rightarrow\infty}(x_{i}.y_{j})_{p}=\infty for some (any) base point pXp\in X. The Gromov boundary or the sequential boundary sX\partial_{s}X of XX, as a set, is defined to be 𝒮/\mathcal{S}/\sim.

Notation and convention. (1) The equivalence class of a geodesic ray or a quasigeodesic ray α\alpha in X\partial X or qX\partial_{q}X is denoted by α()\alpha(\infty). It is customary to fix a base point and require that all the rays start from there to define X\partial X and qX\partial_{q}X but it is not essential.

(2) If α\alpha is a (quasi)geodesic ray with α(0)=x\alpha(0)=x, α()=ξ\alpha(\infty)=\xi then we say that α\alpha joins xx to ξ\xi. We use [x,ξ)[x,\xi) to denote any (quasi)geodesic ray joining xx to ξ\xi when the parametrization of the (quasi)geodesic ray is not important or is understood.

(3) If α\alpha is a quasigeodesic line with α()=ξ1,α()=ξ2qX\alpha(\infty)=\xi_{1},\alpha(-\infty)=\xi_{2}\in\partial_{q}X then we say that α\alpha joins ξ1,ξ2\xi_{1},\xi_{2}. We denote by (ξ1,ξ2)(\xi_{1},\xi_{2}) any quasigeodesic line joining ξ1,ξ2\xi_{1},\xi_{2} when the parameters of the quasigeodesic are understood.

(4) If ξ=[{xn}]sX\xi=[\{x_{n}\}]\in\partial_{s}X then we write xnξx_{n}\rightarrow\xi or ξ=limnxn\xi=\lim_{n\rightarrow\infty}x_{n} and say that the sequence {xn}\{x_{n}\} converges to ξ\xi.

(5) We shall denote by X^\widehat{X} the set XsXX\cup\partial_{s}X.

The following lemma and proposition summarizes all the basic properties of the boundary of hyperbolic spaces that we will need in this paper.

Lemma 2.36.

([DK18, Theorem 11.108]) Let XX, YY be hyperbolic metric spaces.
(1) Given a qi embedding ϕ:XY\phi:X\rightarrow Y we have an injective map ϕ:sXsY\partial\phi:\partial_{s}X\rightarrow\partial_{s}Y.

(2) (i) If XϕYψZX\stackrel{{\scriptstyle\phi}}{{\rightarrow}}Y\stackrel{{\scriptstyle\psi}}{{\rightarrow}}Z are qi embeddings then (ψϕ)=ψϕ\partial(\psi\circ\phi)=\partial\psi\circ\partial\phi

(ii) (IdX)\partial(Id_{X}) is the identity map on sX\partial_{s}X.

(iii) A qi induces a bijective boundary map.

The following proposition relates the three definitions of boundaries.

Proposition 2.37.

(1) For any metric space XX the inclusion 𝒢𝒬\mathcal{G}\rightarrow\mathcal{Q} induces an injective map XqX\partial X\rightarrow\partial_{q}X.

(2) Given a quasigeodesic ray α\alpha, limnα(n)\lim_{n\rightarrow\infty}\alpha(n) is well defined and αβ\alpha\sim\beta implies limnα(n)=limnβ(n)\lim_{n\rightarrow\infty}\alpha(n)=\lim_{n\rightarrow\infty}\beta(n). This induces an injective map qXsX\partial_{q}X\rightarrow\partial_{s}X.

(3) If XX is a proper geodesic hyperbolic metric space then the map XqX\partial X\rightarrow\partial_{q}X is a bijection.

(4) The map qXsX\partial_{q}X\rightarrow\partial_{s}X is a bijection for all Gromov hyperbolic length spaces.

In fact, given δ0\delta\geq 0 there is a constant k2.37=k2.37(δ)k_{\ref{visibility}}=k_{\ref{visibility}}(\delta) such that given any δ\delta-hyperbolic length space XX, any pair of points x,yX^x,y\in\widehat{X} can be joined by a k2.37k_{\ref{visibility}}-quasigeodesic.

Proof.

(1), (2), (3) are standard. See [BH99, Chapter III.H] for instance.

(4) is proved for geodesic metric spaces in Section 2 of [MS12]. See Lemma 2.4 there. The same result for a general length space then is a simple consequence of the existence of a metric graph approximation of a length space and the preceding lemma. ∎

Lemma 2.38.

(Ideal triangles are slim) Suppose XX is a δ\delta-hyperbolic metric space in the sense of Rips or Gromov. Suppose x,y,zX^x,y,z\in\widehat{X} and we have three kk-quasigeodesics joining each pair of points from {x,y,z}\{x,y,z\}. Then the triangle is R=R2.38(δ,k)R=R_{\ref{ideal triangles are slim}}(\delta,k)-slim.

In particular, if γ1,γ2\gamma_{1},\gamma_{2} are two kk-quasigeodesic rays with γ1(0)=γ2(0)\gamma_{1}(0)=\gamma_{2}(0) and γ1()=γ2()\gamma_{1}(\infty)=\gamma_{2}(\infty) then Hd(γ1,γ2)RHd(\gamma_{1},\gamma_{2})\leq R.

The proof of the above lemma is pretty standard and hence we omit it. However, slimness of ideal triangles immediately implies slimness of ideal polygons:

Corollary 2.39.

(Ideal polygons are slim) Suppose XX is a δ\delta-hyperbolic metric space in the sense of Rips or Gromov. Suppose x1,x2,,xnX^x_{1},x_{2},\ldots,x_{n}\in\widehat{X} are nn points and we have nn kk-quasigeodesics joining pairs of points (x1,x2),(x2,x3),,(xn1,xn)(x_{1},x_{2}),(x_{2},x_{3}),\ldots,(x_{n-1},x_{n}) and (xn,x1)(x_{n},x_{1}). Then this nn-gon is R=R2.39(δ,k,n)R=R_{\ref{ideal polygons are slim}}(\delta,k,n)-slim, i.e. every side is contained in RR-neighborhood of the union of the remaining n1n-1 sides.

The following lemma gives a geometric interpretation for sequential boundary in terms of quasigeodesics.

Lemma 2.40.

Let xXx\in X be any point. Suppose {xn}\{x_{n}\} is any sequence of points in XX and βm,n\beta_{m,n} is a kk-quasigeodesic joining xmx_{m} to xnx_{n} for all m,nm,n\in{\mathbb{N}}. Suppose αn\alpha_{n} is a kk-quasigeodesic joining xx to xnx_{n}. Then

(1) {xn}𝒮\{x_{n}\}\in\mathcal{S} if and only if limm,nd(x,βm,n)=\lim_{m,n\rightarrow\infty}d(x,\beta_{m,n})=\infty if and only if there is a constant DD such that for all M>0M>0 there is N>0N>0 with Hd(αmB(x;M),αnB(x;M))DHd(\alpha_{m}\cap B(x;M),\alpha_{n}\cap B(x;M))\leq D for all m,nNm,n\geq N.

(2) Suppose moreover ξsX\xi\in\partial_{s}X and γn\gamma_{n} is a kk-quasigeodesic in XX joining xnx_{n} to ξ\xi for all nn\in{\mathbb{N}} and α\alpha is a kk-quasigeodesic joining xx to ξ\xi.

Then xnξx_{n}\rightarrow\xi if and only if d(x,γn)d(x,\gamma_{n})\rightarrow\infty iff there is constant D>0D>0 such that for all M>0M>0 there is N>0N>0 with Hd(αB(x;M),αnB(x;M))DHd(\alpha\cap B(x;M),\alpha_{n}\cap B(x;M))\leq D for all nNn\geq N.

We skip the proof of this lemma. In fact, the first statement of the lemma is an easy consequence of Lemma 2.34 and stability of quasigeodesics. The second statement is a simple consequence of Lemma 2.34, stability of quasigeodesics and the Lemma 2.38.

The following lemma is proved in section 2 of [MS12] (see Lemma 2.7 and Lemma 2.9 there) for hyperbolic geodesic metric spaces. The same statements are true for length spaces too. To prove it for length spaces one just takes a metric graph approximation. Since the proof is straightforward we omit it.

Lemma 2.41.

(Barycenters of ideal triangles) Given δ0\delta\geq 0 there is r00r_{0}\geq 0 such that for any δ\delta-hyperbolic length space XX, any three distinct points x,y,zX^x,y,z\in\widehat{X} and any three k2.37(δ)k_{\ref{visibility}}(\delta)-quasigeodesics joining x,y,zx,y,z in pairs there is a point x0Xx_{0}\in X such that Nr0(x0)N_{r_{0}}(x_{0}) intersects all the three quasigeodesics.

We refer to a point with this property to be a barycenter of the ideal triangle Δxyz\Delta xyz. There is a constant L0L_{0} such that if x0,x1x_{0},x_{1} are two barycenters of Δxyz\Delta xyz then d(x0,x1)L0d(x_{0},x_{1})\leq L_{0}.

Thus we have a coarsely well-defined map s3XX\partial^{3}_{s}X\rightarrow X. We shall refer to this map as the barycenter map. It is a standard fact that for a non-elementary hyperbolic group GG if XX is a Cayley graph of GG then the barycenter map s3XX\partial^{3}_{s}X\rightarrow X is coarsely surjective and vice versa. If XX is a hyperbolic metric space such that the barycenter map for XX is coarsely surjective then XX will be called a nonelementary hyperbolic space. In section 44 and 55 we deal with spaces with this property.

The following lemma is clear. For instance, we can apply the proof of [MS12, Lemma 2.9].

Lemma 2.42.

Barycenter maps being coarsely surjective is a qi invariant property among hyperbolic length spaces.

2.4.1. Topology on sX\partial_{s}X and Cannon-Thurston maps

Definition 2.43.

(1) If {ξn}\{\xi_{n}\} is a sequence of points in sX\partial_{s}X, we say that {ξn}\{\xi_{n}\} converges to ξsX\xi\in\partial_{s}X if the following holds: Suppose ξn=[{xkn}k]\xi_{n}=[\{x^{n}_{k}\}_{k}] and ξ=[{xk}]\xi=[\{x_{k}\}]. Then limn(lim infi,j(xi.xjn)p)=\lim_{n\rightarrow\infty}(\liminf_{i,j\rightarrow\infty}(x_{i}.x^{n}_{j})_{p})=\infty.

(2) A subset AsXA\subset\partial_{s}X is said to be closed if for any sequence {ξn}\{\xi_{n}\} in AA, ξnξ\xi_{n}\rightarrow\xi implies ξA\xi\in A.

The definition of convergence that we have stated here is equivalent to the one stated in [ABC+91]. Moreover, that the convergence mentioned above is well-defined follows from [ABC+91] and hence we skip it. The next two lemmas give a geometric meaning of the convergence.

Lemma 2.44.

Given k1k\geq 1 and δ0\delta\geq 0 there are constants D=D2.44(k,δ)D=D_{\ref{lem: lim defn}}(k,\delta), L=L2.44(k,δ)L=L_{\ref{lem: lim defn}}(k,\delta) and r=r2.44(k,δ)r=r_{\ref{lem: lim defn}}(k,\delta) with the following properties:

Suppose α,β\alpha,\beta are two kk-quasigeodesic rays starting from a point xXx\in X such that α()β()\alpha(\infty)\neq\beta(\infty) and γ\gamma is a kk-quasigeodesic line joining α()\alpha(\infty) and β()\beta(\infty). Then the following hold:

(1) There exists NN\in{\mathbb{N}} such that |(α(m).β(n))xd(x,γ)|D|(\alpha(m).\beta(n))_{x}-d(x,\gamma)|\leq D for all m,nNm,n\geq N.

In particular, |lim infm,n(α(m).β(n))xd(x,γ)|D|\liminf_{m,n\rightarrow\infty}(\alpha(m).\beta(n))_{x}-d(x,\gamma)|\leq D.

(2) Suppose R=d(x,γ)R=d(x,\gamma) then Hd(αB(x;Rr),βB(x;Rr))LHd(\alpha\cap B(x;R-r),\beta\cap B(x;R-r))\leq L.

Proof.

(1) Since α()β()\alpha(\infty)\neq\beta(\infty) by Lemma 2.38 there is NN\in{\mathbb{N}} such that for all m,nNm,n\geq N, α(m)NR2.38(γ)\alpha(m)\in N_{R_{\ref{ideal triangles are slim}}}(\gamma) and β(n)NR2.38(γ)\beta(n)\in N_{R_{\ref{ideal triangles are slim}}}(\gamma). Let xm,ynγx_{m},y_{n}\in\gamma be such that d(xm,α(m))R2.38d(x_{m},\alpha(m))\leq R_{\ref{ideal triangles are slim}} and d(yn,β(n))R2.38d(y_{n},\beta(n))\leq R_{\ref{ideal triangles are slim}}. Then by joining xm,α(m)x_{m},\alpha(m) and yn,β(n)y_{n},\beta(n) and applying Corollary 2.21 we see that Hausdorff distance between any (1,1)(1,1)-quasigeodesic joining α(m),β(n)\alpha(m),\beta(n), say cm,nc_{m,n} and the portion of γ\gamma between xm,ynx_{m},y_{n} is at most R2.38+2D2.20(δ,k,k)R_{\ref{ideal triangles are slim}}+2D_{\ref{slim iff gromov}}(\delta,k,k). It is clear that for large enough NN, d(x,γ)d(x,\gamma) is the same as the distance of xx and the segment of γ\gamma between xm,ynx_{m},y_{n} if m,nNm,n\geq N. Thus for such m,nm,n we have |d(x,cm,n)d(x,γ)|R2.38+2D2.20(δ,k,k)|d(x,c_{m,n})-d(x,\gamma)|\leq R_{\ref{ideal triangles are slim}}+2D_{\ref{slim iff gromov}}(\delta,k,k). But by Lemma 2.34, |(α(m).β(n))xd(x,cm,n|D2.34(δ,k,k)|(\alpha(m).\beta(n))_{x}-d(x,c_{m,n}|\leq D_{\ref{gromov product meaning}}(\delta,k,k). Hence, |(α(m).β(n))xd(x,γ)|R2.38+2D2.20(δ,k,k)+D2.34(δ,k,k)|(\alpha(m).\beta(n))_{x}-d(x,\gamma)|\leq R_{\ref{ideal triangles are slim}}+2D_{\ref{slim iff gromov}}(\delta,k,k)+D_{\ref{gromov product meaning}}(\delta,k,k) for all large m,nm,n.

(2) To see this we take a 11-approximate nearest point projection, say zz, of xx on γ\gamma. Let xzxz denote a 11-quasigeodesic joining x,zx,z. Then by Corollary 2.26 concatenation of xzxz and the portions of γ\gamma joining zz to γ(±)\gamma(\pm\infty) respectively are both K2.26(δ,k,k)K_{\ref{gluing quasigeodesics}}(\delta,k,k)-quasigeodesics. Call them α\alpha^{\prime} and β\beta^{\prime} respectively. Note that α()=α()\alpha(\infty)=\alpha^{\prime}(\infty) and β()=β()\beta(\infty)=\beta^{\prime}(\infty). Let K=max{k,K2.26(δ,k,ϵ)}K=\max\{k,K_{\ref{gluing quasigeodesics}}(\delta,k,\epsilon)\}. Then by the last part of Lemma 2.38 it follows that zNr(α)Nr(β)z\in N_{r}(\alpha)\cap N_{r}(\beta) where r=R2.38(δ,K)r=R_{\ref{ideal triangles are slim}}(\delta,K). Suppose xα,yβx^{\prime}\in\alpha,y^{\prime}\in\beta are such that d(z,x)rd(z,x^{\prime})\leq r and d(y,z)rd(y^{\prime},z)\leq r. By Lemma 2.20 the Hausdorff distance between xzxz and the portions of α\alpha from xx to xx^{\prime} and the portion of β\beta from xx to yy^{\prime} are each at most D2.20(δ,k,k)+rD_{\ref{slim iff gromov}}(\delta,k,k)+r. Thus these segments of α\alpha and β\beta are at a Hausdorff distance at most L=2D2.20(δ,k,k)+2rL=2D_{\ref{slim iff gromov}}(\delta,k,k)+2r from each other. This completes the proof. ∎

Lemma 2.45.

Let xXx\in X be any point. Suppose {ξn}\{\xi_{n}\} is any sequence of points in sX\partial_{s}X. Suppose βm,n\beta_{m,n} is a kk-quasigeodesic line joining ξm\xi_{m} to ξn\xi_{n} for all m,nm,n\in{\mathbb{N}} and αn\alpha_{n} is a kk-quasigeodesic ray joining xx to ξn\xi_{n} for all nn\in{\mathbb{N}}. Then

(1) limm,nd(x,βm,n)=\lim_{m,n\rightarrow\infty}d(x,\beta_{m,n})=\infty iff there is a constant D=D(k,δ)D=D(k,\delta) such that for all M>0M>0 there is N>0N>0 with Hd(αmB(x;M),αnB(x;M))DHd(\alpha_{m}\cap B(x;M),\alpha_{n}\cap B(x;M))\leq D for all m,nNm,n\geq N and in this case {ξn}\{\xi_{n}\} converges to some point of sX\partial_{s}X.

(2) Suppose moreover ξsX\xi\in\partial_{s}X, γn\gamma_{n} is a kk-quasigeodesic ray in XX joining ξn\xi_{n} to ξ\xi for all nn, and α\alpha is a kk-quasigeodesic ray joining xx to ξ\xi. Then ξnξ\xi_{n}\rightarrow\xi iff d(x,γn)d(x,\gamma_{n})\rightarrow\infty iff there is constant D=D(k,δ)D^{\prime}=D^{\prime}(k,\delta) such that for all M>0M>0 there is N>0N>0 with Hd(αB(x;M),αnB(x;M))DHd(\alpha\cap B(x;M),\alpha_{n}\cap B(x;M))\leq D for all nNn\geq N. In this case limm,nd(x,βm,n)=\lim_{m,n\rightarrow\infty}d(x,\beta_{m,n})=\infty.

Proof.

(1) The ‘iff’ part is an immediate consequence of Lemma 2.44. We prove the last part. Let nin_{i} be an increasing sequence in {\mathbb{N}} such that for all m,nnim,n\geq n_{i} we have Hd(αmB(x;i),αnB(x;i))DHd(\alpha_{m}\cap B(x;i),\alpha_{n}\cap B(x;i))\leq D. Let yiy_{i} be a point of αniB(x;i)\alpha_{n_{i}}\cap B(x;i) such that d(x,yi)+1sup{d(x,y):xαniB(x;i)}d(x,y_{i})+1\geq sup\{d(x,y):x\in\alpha_{n_{i}}\cap B(x;i)\}. We claim that yiy_{i} converges to a point of sX\partial_{s}X. Clearly d(x,yi)d(x,y_{i})\rightarrow\infty. Given iji\leq j\in{\mathbb{N}} we have d(yi,αn)Dd(y_{i},\alpha_{n})\leq D and d(yj,αn)Dd(y_{j},\alpha_{n})\leq D for all nnjn\geq n_{j}. By slimness of polygons we see that any (1,1)(1,1)-quasigeodesic joining yi,yjy_{i},y_{j} is uniformly close to αn\alpha_{n}. It follows that limi,j(yi.yj)x=\lim_{i,j\rightarrow\infty}(y_{i}.y_{j})_{x}=\infty. Let ξ=[{yn}]\xi=[\{y_{n}\}]. It is clear that ξnξ\xi_{n}\rightarrow\xi.

(2) Both iff statements are immediate from Lemma 2.44. The last part follows from slimness of ideal triangle since d(x,γn)d(x,\gamma_{n})\rightarrow\infty. ∎

Corollary 2.46.

Suppose {xn}\{x_{n}\} is a sequence of points in X^\widehat{X} such that {xn}X\{x_{n}\}\subset X or {xn}sX\{x_{n}\}\subset\partial_{s}X. Suppose xnξsXx_{n}\rightarrow\xi\in\partial_{s}X and γn\gamma_{n} is a kk-quasigeodesic joining xnx_{n} to ξ\xi for each nn. Let ynγny_{n}\in\gamma_{n} such that d(x,yn)d(x,y_{n})\rightarrow\infty. Then limnyn=ξ\lim_{n\rightarrow\infty}y_{n}=\xi.

Definition 2.47.

(Cannon-Thurston map, [Mit98b]) If f:YXf:Y\rightarrow X is any map of hyperbolic metric spaces then we say that the Cannon-Thurston or the CT map exists for ff or that ff admits the CT map if ff gives rise to a continuous map f:YsXs\partial{f}:\partial{Y}_{s}\rightarrow\partial{X}_{s} in the following sense:

Given any ξsY\xi\in\partial_{s}Y and any sequence of points {yn}\{y_{n}\} in YY converging to ξ\xi, the sequence {f(yn)}\{f(y_{n})\} converges to a definite point of sX\partial_{s}X independent of the {yn}\{y_{n}\} and the resulting map f:sYsX\partial{f}:\partial_{s}{Y}\rightarrow\partial_{s}{X} is continuous.

Generally, one assumes that the map ff is a proper embedding but for the sake of the definition it is unnecessary. We note that the CT map is unique when it exists. The following lemma gives a sufficient condition for the existence of CT maps.

Lemma 2.48.

(Mitra’s criterion, [Mit98b, Lemma 2.1]) Suppose XX, YY are geodesic hyperbolic metric spaces and f:YXf:Y\rightarrow X is a metrically proper map. Then ff admits the CT map if the following holds:

()(\ast) Let y0Yy_{0}\in Y. There exists a function τ:00\tau:{\mathbb{R}}_{\geq 0}\rightarrow{\mathbb{R}}_{\geq 0}, with the property that τ(n)\tau(n)\rightarrow\infty as nn\rightarrow\infty such that for all geodesic segments [y1,y2]Y[y_{1},y_{2}]_{Y} in YY lying outside the nn-ball around y0Yy_{0}\in Y, any geodesic segment [f(y1),f(y2)]X[f(y_{1}),f(y_{2})]_{X} in XX joining the pair of points f(y1),f(y2)f(y_{1}),f(y_{2}) lies outside the τ(n)\tau(n)-ball around f(y0)Xf(y_{0})\in X.

Remark 3.

(1) The main set of examples where Lemma 2.48 applies comes from taking YY to be a rectifiably path connected subspace of a hyperbolic space XX with induced length metric and the map ff is assumed to be the inclusion map. One also considers the orbit map GXG\rightarrow X where GG is a hyperbolic group acting properly by isometries on a hyperbolic metric space XX. In these examples, the map ff is coarsely Lipschitz as well as metrically proper. The proof of the lemma by Mitra also assumes that XX, YY are proper geodesic metric spaces and Mitra considered the geodesic boundaries. However, these conditions are not necessary as the following lemma and examples show.

(2) The proof of Lemma 2.48 by Mitra only checks that the map is a well-defined extension of ff rather than it is continuous. However, with very little effort the condition ()(\ast) can be shown to be sufficient for the well-definedness as well as the continuity of the CT map.

(3) One can easily check that the condition ()(\ast) is also necessary provided X,YX,Y are proper hyperbolic spaces and ff is coarsely Lipschitz and metrically proper.

The following lemma is the main tool for the proof of our theorem of Cannon-Thurston map. We shall refer to this as Mitra’s lemma.

Lemma 2.49.

Suppose X,YX,Y are length spaces hyperbolic in the sense of Gromov, and f:YXf:Y\rightarrow X is any map. Let pYp\in Y.

()(\ast\ast) Suppose for all N>0N>0 there is M=M(N)>0M=M(N)>0 such that NN\rightarrow\infty implies MM\rightarrow\infty with the following property: For any y1,y2Yy_{1},y_{2}\in Y, any (1,1)(1,1)-quasigeodesic α\alpha in YY joining y1,y2y_{1},y_{2} and any (1,1)(1,1)-quasigeodesic β\beta in XX joining f(y1),f(y2)f(y_{1}),f(y_{2}), B(p,N)α=B(p,N)\cap\alpha=\emptyset implies B(f(p),M)β=B(f(p),M)\cap\beta=\emptyset.

Then the CT map exists for f:YXf:Y\rightarrow X.

Proof.

Suppose {yn}\{y_{n}\} is any sequence in YY. Suppose αi,j\alpha_{i,j} is a (1,1)(1,1)-quasigeodesic in YY joining yi,yjy_{i},y_{j} and suppose γi,j\gamma_{i,j} is a (1,1)(1,1)-quasigeodesic in XX joining f(yi),f(yj)f(y_{i}),f(y_{j}). Then by Lemma 2.34 limi,j(yi.yj)p=\lim_{i,j\rightarrow\infty}(y_{i}.y_{j})_{p}=\infty if and only if limi,jd(p,αi,j)=\lim_{i,j\rightarrow\infty}d(p,\alpha_{i,j})=\infty and limi,j(f(yi).f(yj))f(p)=\lim_{i,j\rightarrow\infty}(f(y_{i}).f(y_{j}))_{f(p)}=\infty if and only if limi,jdX(f(p),γi,j)=\lim_{i,j\rightarrow\infty}d_{X}(f(p),\gamma_{i,j})=\infty. On the other hand by ()(\ast\ast) limi,jd(p,αi,j)=\lim_{i,j\rightarrow\infty}d(p,\alpha_{i,j})=\infty implies limi,jdX(f(p),γi,j)=\lim_{i,j\rightarrow\infty}d_{X}(f(p),\gamma_{i,j})=\infty. Thus {yn}\{y_{n}\} converges to a point of sY\partial_{s}Y implies {f(yn)}\{f(y_{n})\} converges to a point of sX\partial_{s}X. The same argument shows that if {yn}\{y_{n}\} and {zn}\{z_{n}\} are two sequences in YY representing the same point of sY\partial_{s}Y then {f(yn)}\{f(y_{n})\} and {f(zn)}\{f(z_{n})\} also represent the same point of sX\partial_{s}X. Thus we have a well-defined map f:sYsX\partial f:\partial_{s}Y\rightarrow\partial_{s}X.

Now we prove the continuity of the map. We need to show that if ξnξ\xi_{n}\rightarrow\xi in sY\partial_{s}Y then f(ξn)f(ξ)\partial f(\xi_{n})\rightarrow\partial f(\xi). Suppose ξn\xi_{n} is represented by the class of {ykn}k\{y^{n}_{k}\}_{k} and ξ\xi is the equivalence class of {yk}\{y_{k}\}. Then

limn(lim infi,j(yin.yj)p)=.\lim_{n\rightarrow\infty}(\liminf_{i,j\rightarrow\infty}(y^{n}_{i}.y_{j})_{p})=\infty.

By Lemma 2.34 then we have

limn(lim infi,jd(p,αi,jn)=\lim_{n\rightarrow\infty}(\liminf_{i,j\rightarrow\infty}d(p,\alpha^{n}_{i,j})=\infty

for any (1,ϵ)(1,\epsilon)-quasigeodesic αi,jn\alpha^{n}_{i,j} in YY joining yiny^{n}_{i} and yjy_{j}. By ()(\ast) then we have

limn(lim infi,jd(f(p),γi,jn)=\lim_{n\rightarrow\infty}(\liminf_{i,j\rightarrow\infty}d(f(p),\gamma^{n}_{i,j})=\infty

where γi,jn\gamma^{n}_{i,j} is any (1,ϵ)(1,\epsilon)-quasigeodesic in XX joining f(yin),f(yj)f(y^{n}_{i}),f(y_{j}). This in turn implies by Lemma 2.34 that

limn(lim infi,j(f(yin).f(yj))f(p))=.\lim_{n\rightarrow\infty}(\liminf_{i,j\rightarrow\infty}(f(y^{n}_{i}).f(y_{j}))_{f(p)})=\infty.

Therefore, f(ξn)f(ξ)\partial f(\xi_{n})\rightarrow\partial f(\xi) as was required. ∎

Examples and remarks:

  1. (1)

    Suppose f:00f:{\mathbb{R}}_{\geq 0}\rightarrow{\mathbb{R}}_{\geq 0} is the function f(x)=ex1f(x)=e^{x}-1. Then ff is not coarsely Lipschitz but ff admits the CT map.

  2. (2)

    One can easily cook up an example along the line of the above example where metric properness is also violated but the CT map exists as we see in the example below. We will see another interesting example in Corollary 6.10.

  3. (3)

    The condition ()(\ast) in the above lemma is also not necessary in general for the existence of the CT map. Here is an example in which both metric properness and ()(\ast) fail to hold but nevertheless the CT map exists. Suppose XX is a tree built in two steps. First we have a star, i.e. a tree with one central vertex on which end points of finite intervals are glued where the lengths of the intervals are unbounded. Then two distinct rays are glued to each vertex of the star other than the central vertex. Suppose YY is obtained by collapsing the central star in XX to a point and ff is the quotient map. Then clearly the CT map exists but ()(\ast) is violated.

The following lemma is very standard and hence we skip mentioning its proof.

Lemma 2.50.

(Functoriality of CT maps) (1) Suppose X,Y,ZX,Y,Z are hyperbolic metric spaces and f:XYf:X\rightarrow Y and g:YZg:Y\rightarrow Z admit the CT maps. Then so does gfg\circ f and (gf)=gf\partial(g\circ f)=\partial g\circ\partial f.

(2) If i:XXi:X\rightarrow X is the identity map then it admits the CT map i\partial i which is the identity map on sX\partial_{s}X

(3) If two maps f,h:XYf,h:X\rightarrow Y are at a finite distance admitting the CT maps then they induce the same CT map.

(4) Suppose f:XYf:X\rightarrow Y is a qi embedding of hyperbolic length spaces. Then ff admits the CT map f:sXsY\partial f:\partial_{s}X\rightarrow\partial_{s}Y which is a homeomorphism onto the image.

If ff is a quasiisometry then f\partial f is a homeomorphism. In particular, the action by left multiplication of a hyperbolic group GG on itself induces an action of GG on G\partial G by homeomorphisms.

2.4.2. Limit sets

Definition 2.51.

Suppose XX is a hyperbolic metric space and AXA\subset X. Then the limit set of AA in XX is the set ΛX(A)={limnansX:{an}is a sequence inA}\Lambda_{X}(A)=\{\lim_{n\rightarrow\infty}a_{n}\in\partial_{s}X:\{a_{n}\}\,\mbox{is a sequence in}\,A\}.

When XX is understood then the limit set of AXA\subset X will be denoted simply by Λ(A)\Lambda(A). In this subsection, we collect some basic results on limit sets that we need in Section 6 of the paper. In each case, we briefly indicate the proofs for the sake of completeness. The following is straightforward.

Lemma 2.52.

Suppose XX is a hyperbolic metric space and A,BXA,B\subset X with Hd(A,B)<Hd(A,B)<\infty. Then Λ(A)=Λ(B)\Lambda(A)=\Lambda(B).

Lemma 2.53.

Suppose XX is a hyperbolic metric space and YXY\subset X. Suppose ZYZ\subset Y coarsely bisects YY in XX into Y1,Y2Y_{1},Y_{2} where ZY1Y2Z\subset Y_{1}\cap Y_{2}. Then Λ(Y1)Λ(Y2)=Λ(Z)\Lambda(Y_{1})\cap\Lambda(Y_{2})=\Lambda(Z).

Proof.

This is a straightforward consequence of Lemma 2.34. ∎

Lemma 2.54.

Suppose XX is a δ\delta-hyperbolic metric space and AXA\subset X is λ\lambda-quasiconvex. Suppose ξΛ(A)\xi\in\Lambda(A) and γ\gamma is a KK-quasigeodesic ray converging to ξ\xi. Then there are NN\in{\mathbb{N}} and D=D2.54(δ,λ,K)>0D=D_{\ref{limset lem}}(\delta,\lambda,K)>0 such that γ(n)ND(A)\gamma(n)\in N_{D}(A) for all nNn\geq N.

Proof.

Rather than explicitly computing the constants we indicate how to obtain them. Suppose {xn}\{x_{n}\} is a sequence in AA such that xnξx_{n}\rightarrow\xi. Let y1γy_{1}\in\gamma be a 11-approximate nearest point projection of x1x_{1} on γ\gamma. Let α1\alpha_{1} denote a (1,1)(1,1)-quasigeodesic joining x1,y1x_{1},y_{1}. Then the concatenation, say γ1\gamma_{1}, of α1\alpha_{1} and the segment of γ\gamma from y1y_{1} to ξ\xi is a uniform quasigeodesic by Corollary 2.26. For all m>1m>1, let ymy_{m} denote a 11-approximate nearest point projection of xmx_{m} on γ1\gamma_{1}. Then ymy_{m} is contained in γ1\gamma_{1} for all large mm. However, once again by Corollary 2.26 the concatenation of the portion of γ1\gamma_{1} between x1,ymx_{1},y_{m} and a 11-quasigeodesic joining xm,ymx_{m},y_{m} is a uniform quasigeodesic. Now it follows by stability of quasigeodesics that the segment of γ1\gamma_{1} between y1,ymy_{1},y_{m} is contained in a uniformly small neighborhood of AA since AA is quasiconvex. ∎

Lemma 2.55.

Suppose X,YX,Y are hyperbolic metric spaces, and f:YXf:Y\rightarrow X is any metrically proper map. Suppose that the CT map exists for ff. Then we have Λ(f(Y))=f(Y)\Lambda(f(Y))=\partial f(\partial Y) in each of the following cases:

(1) YY is a proper metric space.

(2) ff is a qi embedding.

Proof.

(1) It is clear that f(Y)Λ(f(Y))\partial f(\partial Y)\subset\Lambda(f(Y)). Suppose yny_{n} is any sequence such that f(yn)ξf(y_{n})\rightarrow\xi for some ξsX\xi\in\partial_{s}X. Since ff is proper {yn}\{y_{n}\} is an unbounded sequence. Since YY is a proper length space it is a geodesic metric space by Hopf-Rinow theorem (see [BH99], Proposition 3.7, Chapter I.3). Now it is a standard fact that any unbounded sequence in a proper geodesic metric space has a subsequence converging to a point of the Gromov boundary of the space. Since YY is proper, we have a subsequence {ynk}\{y_{n_{k}}\} of {yn}\{y_{n}\} such that ynkηy_{n_{k}}\rightarrow\eta for some ηsY\eta\in\partial_{s}Y. It is clear that f(η)=ξ\partial f(\eta)=\xi. Hence Λ(f(Y))f(Y)\Lambda(f(Y))\subset\partial f(\partial Y).

(2) Let yYy\in Y and x=f(y)x=f(y). Suppose {yn}\{y_{n}\} is a sequence of points in YY such that limm,n(f(ym).f(yn))x=\lim_{m,n\rightarrow\infty}(f(y_{m}).f(y_{n}))_{x}=\infty and η=[{f(ym)}]\eta=[\{f(y_{m})\}]. Then by Lemma 2.34 for any 11-quasigeodesic βm,n\beta_{m,n} in XX joining f(ym),f(yn)f(y_{m}),f(y_{n}) for all m,nm,n\in{\mathbb{N}}, we have limm,ndX(x,βm,n)=\lim_{m,n\rightarrow\infty}d_{X}(x,\beta_{m,n})=\infty. Since ff is a qi embedding if αm,n\alpha_{m,n} is a 11-quasigeodesic in YY joining ym,yny_{m},y_{n} for all m,nm,n\in{\mathbb{N}} then f(αm,n)f(\alpha_{m,n}) are uniform quasigeodesics in XX. Hence, by stability of quasigeodesics in XX we have Hd(f(αm,n),βm,n)<DHd(f(\alpha_{m,n}),\beta_{m,n})<D for some constant D0D\geq 0. Thus limm,ndX(x,f(αm,n))=\lim_{m,n\rightarrow\infty}d_{X}(x,f(\alpha_{m,n}))=\infty. Since ff is a qi embedding and x=f(y)x=f(y) it follows that limm,ndY(y,αm,n)=\lim_{m,n\rightarrow\infty}d_{Y}(y,\alpha_{m,n})=\infty. Therefore, limm,n(ym.yn)y=\lim_{m,n\rightarrow\infty}(y_{m}.y_{n})_{y}=\infty again by Lemma 2.34. Hence, if ξ=[{yn}]\xi=[\{y_{n}\}] then f(ξ)=η\partial f(\xi)=\eta. ∎

Lemma 2.56.

(Projection of boundary points on quasiconvex sets) Given δ0\delta\geq 0 and k0k\geq 0 there is a constant R=R2.56(δ,k)R=R_{\ref{qc last lemma}}(\delta,k) such that the following holds:

Suppose XX is a δ\delta-hyperbolic metric space, AXA\subset X is kk-quasiconvex and ξXΛ(A)\xi\in\partial X\setminus\Lambda(A). Then there is a point xAx\in A with the following property: Suppose {xn}\{x_{n}\} is any sequence where xnξx_{n}\rightarrow\xi. Then there is an N>0N>0 such that for all nNn\geq N we have PA(xn)AB(x,R)P_{A}(x_{n})\in A\cap B(x,R).

Proof.

Suppose {xn},{yn}\{x_{n}\},\{y_{n}\} are two sequences in XX such that xnξx_{n}\rightarrow\xi and ynξy_{n}\rightarrow\xi. Let αm,n\alpha_{m,n} be a 11-quasigeodesic in XX joining xm,ynx_{m},y_{n} for all m,nm,n\in{\mathbb{N}}. Let PA:XAP_{A}:X\rightarrow A be a 11-approximate nearest point projection on AA.

Claim: There is a constant R0>0R_{0}>0 depending only on δ\delta and kk and there is N>0N>0 such that diam(PA(αm,n))R0diam(P_{A}(\alpha_{m,n}))\leq R_{0} for all m,nNm,n\geq N.

We first note that limm,nd(A,αm,n)=\lim_{m,n\rightarrow\infty}d(A,\alpha_{m,n})=\infty. In fact, if this is not the case then there is r>0r>0 such that for all N>0N>0 there are mN,nNNm_{N},n_{N}\geq N with d(A,αmN,nN)rd(A,\alpha_{m_{N},n_{N}})\leq r. In that case let aNAa_{N}\in A be such that d(aN,αmN,nN)rd(a_{N},\alpha_{m_{N},n_{N}})\leq r. It is then clear that aNξa_{N}\rightarrow\xi by Lemma 2.40(1), contradicting the hypothesis that ξΛ(A)\xi\not\in\Lambda(A). By stability of quasigeodesics, any 11-quasigeodesic is uniformly quasiconvex in XX and AA is given to be kk-quasiconvex. Hence, by Corollary 2.32 there are constants D0,R0D_{0},R_{0} such that d(A,αm,n)>D0d(A,\alpha_{m,n})>D_{0} implies that diam(PA(αm,n))R0diam(P_{A}(\alpha_{m,n}))\leq R_{0}. Since, limm,nd(A,αm,n)=\lim_{m,n\rightarrow\infty}d(A,\alpha_{m,n})=\infty there is N>0N>0 such that d(A,αm,n)>D0d(A,\alpha_{m,n})>D_{0} for all m,nNm,n\geq N. This proves the existence of NN and R0R_{0}.

Now, by specializing the claim to the case {xn}={yn}\{x_{n}\}=\{y_{n}\} we have N0>0N_{0}>0 such that if βm,n\beta_{m,n} is a 11-quasigeodesic joining xm,xnx_{m},x_{n} then diam(PA(βm,n))R0diam(P_{A}(\beta_{m,n}))\leq R_{0} for all m,nN0m,n\geq N_{0}. Let x=PA(xN0)x=P_{A}(x_{N_{0}}). Now, given any sequence {xn}\{x^{\prime}_{n}\} in XX with xnξx^{\prime}_{n}\rightarrow\xi by the claim there is M>0M>0 such that for all m,nMm,n\geq M, d(PA(xm),PA(xn))R0d(P_{A}(x_{m}),P_{A}(x^{\prime}_{n}))\leq R_{0}. Hence, if N=max{N0,M}N=\max\{N_{0},M\} then d(x,PA(xn))d(x,PA(xN))+d(PA(xN),PA(xn))2R0d(x,P_{A}(x^{\prime}_{n}))\leq d(x,P_{A}(x_{N}))+d(P_{A}(x_{N}),P_{A}(x^{\prime}_{n}))\leq 2R_{0}. Thus we can take R=2R0R=2R_{0}.∎

Since the point xAx\in A in the above lemma is coarsely unique we shall call any such point to be the nearest point projection of ξ\xi on AA and we shall denote it by PA(ξ)P_{A}(\xi).

3. Metric bundles

In this section, we recall necessary definitions and some elementary properties of the primary objects of study in this paper namely, metric bundles and metric graph bundles from [MS12]. We make a minor modification (see Definition 3.2) to the definition of a metric bundle but use the same definition of metric graph bundles as in [MS12].

3.1. Basic definitions and properties.

Definition 3.1.

(Metric bundles [MS12, Definition 1.2]) Suppose (X,d)(X,d) and (B,dB)(B,d_{B}) are geodesic metric spaces; let c1c\geq 1 and let η:[0,)[0,)\eta:[0,\infty)\rightarrow[0,\infty) be a function. We say that XX is an (η,c)(\eta,c)- metric bundle over BB if there is a surjective 11-Lipschitz map π:XB\pi:X\rightarrow B such that the following conditions hold:
(1) For each point zBz\in B, Fz:=π1(z)F_{z}:=\pi^{-1}(z) is a geodesic metric space with respect to the path metric dzd_{z} induced from XX. The inclusion maps i:(Fz,dz)Xi:(F_{z},d_{z})\rightarrow X are uniformly metrically proper as measured by η\eta.
(2) Suppose z1,z2Bz_{1},z_{2}\in B, dB(z1,z2)1d_{B}(z_{1},z_{2})\leq 1 and let γ\gamma be a geodesic in BB joining them.
Then for any point zγz\in\gamma and xFzx\in F_{z} there is a path γ~:[0,1]π1(γ)X\tilde{\gamma}:[0,1]\rightarrow\pi^{-1}(\gamma)\subset X of length at most cc such that γ~(0)Fz1\tilde{\gamma}(0)\in F_{z_{1}}, γ~(1)Fz2\tilde{\gamma}(1)\in F_{z_{2}} and xγ~x\in\tilde{\gamma}.

If XX is a metric bundle over BB in the above sense then we shall refer to it as a geodesic metric bundle in this paper. However, the above definition seems a little restrictive. Therefore, we propose the following.

Definition 3.2.

(Length metric bundles) Suppose (X,d)(X,d) and (B,dB)(B,d_{B}) are length spaces, c1c\geq 1 and we have a function η:[0,)[0,)\eta:[0,\infty)\rightarrow[0,\infty). We say that XX is an (η,c)(\eta,c)- length metric bundle over BB if there is a surjective 11-Lipschitz map π:XB\pi:X\rightarrow B such that the following conditions hold:
(1) For each point zBz\in B, Fz:=π1(z)F_{z}:=\pi^{-1}(z) is a length space with respect to the path metric dzd_{z} induced from XX. The inclusion maps i:(Fz,dz)Xi:(F_{z},d_{z})\rightarrow X are uniformly metrically proper as measured by η\eta.
(2) Suppose z1,z2Bz_{1},z_{2}\in B, and let γ\gamma be a path of length at most 11 in BB joining them.
Then for any point zγz\in\gamma and xFzx\in F_{z} there is a path γ~:[0,1]π1(γ)X\tilde{\gamma}:[0,1]\rightarrow\pi^{-1}(\gamma)\subset X of length at most cc such that γ~(0)Fz1\tilde{\gamma}(0)\in F_{z_{1}}, γ~(1)Fz2\tilde{\gamma}(1)\in F_{z_{2}} and xγ~x\in\tilde{\gamma}.

Given length spaces XX and BB we will say that XX is a length metric bundle over BB if XX is an (η,c)(\eta,c)-length metric bundle over BB in the above sense for some function η:++\eta:{\mathbb{R}}^{+}\rightarrow{\mathbb{R}}^{+} and some constant c1c\geq 1.

Convention 3.3.

From now on whenever we speak of a metric bundle we mean a length metric bundle.

Definition 3.4.

(Metric graph bundles [MS12, Definition 1.5]) Suppose XX and BB are metric graphs. Let η:[0,)[0,)\eta:[0,\infty)\rightarrow[0,\infty) be a function. We say that XX is an η\eta-metric graph bundle over BB if there exists a surjective simplicial map π:XB\pi:X\rightarrow B such that:
1.1. For each b𝒱(B)b\in\mathcal{V}(B), Fb:=π1(b)F_{b}:=\pi^{-1}(b) is a connected subgraph of XX and the inclusion maps i:FbXi:F_{b}\rightarrow X are uniformly metrically proper as measured by η\eta for the path metrics dbd_{b} induced on FbF_{b}.
2.2. Suppose b1,b2𝒱(B)b_{1},b_{2}\in\mathcal{V}(B) are adjacent vertices. Then each vertex x1x_{1} of Fb1F_{b_{1}} is connected by an edge with a vertex in Fb2F_{b_{2}}.

Remark 4.

Since the map π\pi is simplicial it follows that it is 11-Lipschitz.

For a metric (graph) bundle the spaces (Fz,dz)(F_{z},d_{z}), zBz\in B will be referred to as fibers and the dzd_{z}-distance between two points in FzF_{z} will be referred to as their fiber distance. A geodesic in FzF_{z} will be called a fiber geodesic. The spaces XX and BB will be referred to as the total space and the base space of the bundle respectively. By a statement of the form ‘XX is a metric bundle (resp. metric graph bundle)’ we will mean that it is the total space of a metric bundle (resp. metric graph bundle).

Most of the results proved for geodesic metric bundles in [MS12] have their analogs for length metric bundles. We explicitly prove this phenomenon or provide sufficient arguments for all the results needed for our purpose.

Convention 3.5.

Very often in a lemma, proposition, corollary, or a theorem we shall omit explicit mention of some of the parameters on which a constant may depend if the parameters are understood.

Definition 3.6.

Suppose π:XB\pi:X\rightarrow B is a metric (graph) bundle.

(1) Suppose ABA\subset B and k1k\geq 1. A kk-qi section over AA is a kk-qi embedding s:AXs:A\rightarrow X (resp. s:𝒱(A)Xs:{\mathcal{V}}(A)\rightarrow X)such that πs=IdA\pi\circ s=Id_{A} (resp. πs=Id𝒱(A)\pi\circ s=Id_{{\mathcal{V}}(A)}) where AA has the restricted metric from BB and IdAId_{A} (resp. Id𝒱(A)Id_{{\mathcal{V}}(A)}) denotes the identity map on AAA\rightarrow A (resp. 𝒱(A)𝒱(A){\mathcal{V}}(A)\rightarrow{\mathcal{V}}(A)).

(2) Given any metric space (resp. graph) ZZ and any qi embedding f:ZBf:Z\rightarrow B (resp. f:𝒱(Z)𝒱(B)f:{\mathcal{V}}(Z)\rightarrow{\mathcal{V}}(B)) a kk-qi lift of ff is a kk-qi embedding f~:ZX\tilde{f}:Z\rightarrow X (resp. f~:𝒱(Z)𝒱(X)\tilde{f}:{\mathcal{V}}(Z)\rightarrow{\mathcal{V}}(X)) such that πf~=f\pi\circ\tilde{f}=f.

Convention 3.7.

(1) Most of the time we shall refer to the image of a qi section (or a qi lift) to be the qi section (resp. the qi lift).
(2) Suppose γ:IB\gamma:I\rightarrow B is a (quasi)geodesic and γ~\tilde{\gamma} is a qi lift of γ\gamma. Let b=γ(t)b=\gamma(t) for some tIt\in I. Then we will denote γ~(t)\tilde{\gamma}(t) by γ~(b)\tilde{\gamma}(b) also.
(3) In the context of a metric graph bundle (X,B,π)(X,B,\pi), when we talk about a point in XX, BB or a fiber, we mean that the point is a vertex in the corresponding space.

The following lemma is immediate from the definition of a metric (graph) bundle. Hence we briefly indicate its proof.

Lemma 3.8.

( Path lifting lemma) Suppose π:XB\pi:X\rightarrow B is an (η,c)(\eta,c)-metric bundle or an η\eta-metric graph bundle.

  1. (1)

    Suppose b1,b2Bb_{1},b_{2}\in B. Suppose γ:[0,L]B\gamma:[0,L]\rightarrow B is a continuous, rectifiable, arc length parameterized path (resp. an edge path) in BB joining b1b_{1} to b2b_{2}. Given any xFb1x\in F_{b_{1}} there is a path γ~\tilde{\gamma} in π1(γ)\pi^{-1}(\gamma) such that l(γ~)(L+1)cl(\tilde{\gamma})\leq(L+1)c (resp l(γ~)=L)l(\tilde{\gamma})=L) joining xx to some point of Fb2F_{b_{2}}.

    In particular, in case XX is a metric graph bundle over BB any geodesic γ\gamma of BB can be lifted to a geodesic starting from any given point of π1(γ)\pi^{-1}(\gamma).

  2. (2)

    For any k1k\geq 1 and ϵ0\epsilon\geq 0, any dotted (k,ϵ)(k,\epsilon)-quasigeodesic β:[m,n]B\beta:[m,n]\rightarrow B has a lift β~\tilde{\beta} starting from any point of Fβ(m)F_{\beta(m)} such that the following hold, where we assume c=1c=1 for metric graph bundles.

    For all i,j[m,n]i,j\in[m,n] we have

    ϵ+1k|ij|dX(β~(i),β~(j))c(k+ϵ+1)|ij|.-\epsilon+\frac{1}{k}|i-j|\leq d_{X}(\tilde{\beta}(i),\tilde{\beta}(j))\leq c\cdot(k+\epsilon+1)|i-j|.

    In particular it is a c(k+ϵ+1)c\cdot(k+\epsilon+1)-qi lift of β\beta. Also we have

    l(β~)ck(k+ϵ+1)(ϵ+dB(b1,b2)).l(\tilde{\beta})\leq ck(k+\epsilon+1)(\epsilon+d_{B}(b_{1},b_{2})).
Proof.

(1) We fix a sequence of points 0=t0,t1,,tn=L0=t_{0},t_{1},\cdots,t_{n}=L in [0,L][0,L] such that l(γ|[ti,ti+1])=1l(\gamma|_{[t_{i},t_{i+1}]})=1 for 0i<n10\leq i<n-1 and l(γ|[tn1,tn])1l(\gamma|_{[t_{n-1},t_{n}]})\leq 1 for the metric bundle case. For the metric graph bundle γ(ti)\gamma(t_{i}) are the consecutive vertices on γ\gamma, 0iL=n0\leq i\leq L=n. Now given any x=:x0Ft0x=:x_{0}\in F_{t_{0}} we can inductively construct a sequence of points xiFtix_{i}\in F_{t_{i}}, 0in0\leq i\leq n and a sequence of paths αi\alpha_{i} of length at most cc (resp. an edge) joining xix_{i} to xi+1x_{i+1} for 0in10\leq i\leq n-1. Concatenation of these paths gives a candidate for γ~\tilde{\gamma}.

The second statement for metric graph bundles follow because π:XB\pi:X\rightarrow B is a 11-Lipschitz map.

(2) We construct a lift β~\tilde{\beta} of β\beta starting from any point xFβ(m)x\in F_{\beta(m)} inductively as follows. We know that dB(β(i),β(i+1))k+ϵd_{B}(\beta(i),\beta(i+1))\leq k+\epsilon. Let βi\beta_{i} be a path in BB joining β(i)\beta(i) to β(i+1)\beta(i+1) which is of length at most k+ϵ+1k+\epsilon+1 for min1m\leq i\leq n-1. We can then find a sequence of paths of length at most (k+ϵ+1)c(k+\epsilon+1)\cdot c in π1(βi)\pi^{-1}(\beta_{i}) (where c=1c=1 for metric graph bundle) min1m\leq i\leq n-1 using the first part of the lemma such that βm\beta_{m} starts at xx and βi+1\beta_{i+1} starts at the end point of βi\beta_{i} for m+1in1m+1\leq i\leq n-1. Let xix_{i} be the starting point of βi\beta_{i} for min1m\leq i\leq n-1 and let xnx_{n} be the end point of βn1\beta_{n-1}. Then we define β~\tilde{\beta} by setting β~(i)=xi\tilde{\beta}(i)=x_{i}, minm\leq i\leq n.

Clearly dX(β~(i),β~(j))c(k+ϵ+1)|ij|d_{X}(\tilde{\beta}(i),\tilde{\beta}(j))\leq c\cdot(k+\epsilon+1)|i-j|. Also, dB(πβ~(i),πβ~(j))=dB(β(i),β(j))dX(β~(i),β~(j))d_{B}(\pi\circ\tilde{\beta}(i),\pi\circ\tilde{\beta}(j))=d_{B}(\beta(i),\beta(j))\leq d_{X}(\tilde{\beta}(i),\tilde{\beta}(j)) since π\pi is 11-Lipschitz. Since β\beta is a dotted (k,ϵ)(k,\epsilon) quasigeodesic, we have ϵ+1k|ij|dB(β(i),β(j))-\epsilon+\frac{1}{k}|i-j|\leq d_{B}(\beta(i),\beta(j)). This proves that

ϵ+1k|ij|dX(β~(i),β~(j))c(k+ϵ+1)|ij|.-\epsilon+\frac{1}{k}|i-j|\leq d_{X}(\tilde{\beta}(i),\tilde{\beta}(j))\leq c\cdot(k+\epsilon+1)|i-j|.

For the last part of (2) we see that

l(β~)=i=mn1dX(β~(i),β~(i+1))i=mn1c(k+ϵ+1)=(nm)c(k+ϵ+1).l(\tilde{\beta})=\sum_{i=m}^{n-1}d_{X}(\tilde{\beta}(i),\tilde{\beta}(i+1))\leq\sum_{i=m}^{n-1}c\cdot(k+\epsilon+1)=(n-m)c\cdot(k+\epsilon+1).

On the other hand since β\beta is a (k,ϵ)(k,\epsilon)-quasigeodesic we have ϵ+1k(nm)dB(b1,b2)-\epsilon+\frac{1}{k}(n-m)\leq d_{B}(b_{1},b_{2}). The conclusion immediately follows from these two inequalities. ∎

The following corollary follows from the proof of Proposition 2.10 of [MS12]. We include it for the sake of completeness.

Corollary 3.9.

Given any metric (graph) bundle π:XB\pi:X\rightarrow B and b1,b2Bb_{1},b_{2}\in B we can define a map ϕ:Fb1Fb2\phi:F_{b_{1}}\rightarrow F_{b_{2}} such that dX(x,ϕ(x))3c+3cdB(b1,b2)d_{X}(x,\phi(x))\leq 3c+3cd_{B}(b_{1},b_{2}) (resp. d(x,ϕ(x))=dB(b1,b2)d(x,\phi(x))=d_{B}(b_{1},b_{2})) for all xFb1x\in F_{b_{1}}.

Proof.

The statement about the metric graph bundle is trivially true by Lemma 3.8 (1). For the metric bundle case, fix a dotted 11-quasigeodesic γ\gamma joining b1b_{1} to b2b_{2}. Then for all xFb1x\in F_{b_{1}} fix for once and all a dotted lift γ~\tilde{\gamma} as constructed in the proof of the Lemma 3.8 which starts from xx and set ϕ(x)=γ~(b2)\phi(x)=\tilde{\gamma}(b_{2}). The statement then follows from Lemma 3.8(2). ∎

Remark 5.

For all b1,b2Bb_{1},b_{2}\in B any map f:Fb1Fb2f:F_{b_{1}}\rightarrow F_{b_{2}} such that dX(x,f(x))Dd_{X}(x,f(x))\leq D for some constant DD independent of xx will be referred to as a fiber identification map.

The proof of the first part of the following lemma is immediate from Corollary 3.9 whereas the next two parts essentially follow from the proof of Proposition 2.10 of [MS12]. Hence we skip the proofs.

Lemma 3.10.

Suppose π:XB\pi:X\rightarrow B is an (η,c)(\eta,c)-metric bundle or an η\eta-metric graph bundle and R0R\geq 0. Suppose b1,b2Bb_{1},b_{2}\in B. The we have the following.

  1. (1)

    Hd(Fb1,Fb2)3c+3cdB(b1,b2)Hd(F_{b_{1}},F_{b_{2}})\leq 3c+3cd_{B}(b_{1},b_{2}) (resp. Hd(Fb1,Fb2)=dB(b1,b2)Hd(F_{b_{1}},F_{b_{2}})=d_{B}(b_{1},b_{2})).

  2. (2)

    Suppose ϕb1b2:Fb1Fb2\phi_{b_{1}b_{2}}:F_{b_{1}}\rightarrow F_{b_{2}} is a map such that for all xFb1x\in F_{b_{1}}, d(x,ϕb1b2(x))Rd(x,\phi_{b_{1}b_{2}}(x))\leq R for all xFb1x\in F_{b_{1}}.

    Then ϕb1b2\phi_{b_{1}b_{2}} is a K3.10=K3.10(R)K_{\ref{fibers qi}}=K_{\ref{fibers qi}}(R)-quasiisometry which is D3.10D_{\ref{fibers qi}}-surjective.

  3. (3)

    If ψb1b2:Fb1Fb2\psi_{b_{1}b_{2}}:F_{b_{1}}\rightarrow F_{b_{2}} is any other map such that d(x,ψb1b2(x))Rd(x,\psi_{b_{1}b_{2}}(x))\leq R^{\prime} for all xFb1x\in F_{b_{1}} then d(ϕb1b2,ψb1b2)η(R+R)d(\phi_{b_{1}b_{2}},\psi_{b_{1}b_{2}})\leq\eta(R+R^{\prime}).

    In particular, the maps ϕb1b2\phi_{b_{1}b_{2}} are coarsely unique (see Definition 2.1(7)).

In this lemma, we deliberately suppress the dependence of K3.10K_{\ref{fibers qi}} on the parameter(s) of the bundle.

Corollary 3.11.

Suppose π:XB\pi:X\rightarrow B is a metric (graph) bundle and b1,b2Bb_{1},b_{2}\in B (resp. b1,b2𝒱(B)b_{1},b_{2}\in{\mathcal{V}}(B)) such that dB(b1,b2)Rd_{B}(b_{1},b_{2})\leq R. Suppose ϕb1b2:Fb1Fb2\phi_{b_{1}b_{2}}:F_{b_{1}}\rightarrow F_{b_{2}} is a fiber identification map as constructed in the proof of Corollary 3.9. Then ϕb1b2\phi_{b_{1}b_{2}} is a K3.11=K3.11(R)K_{\ref{fibers unif qi}}=K_{\ref{fibers unif qi}}(R)-quasiisometry.

Proof.

By Corollary 3.9 dX(x,ϕb1b2(x))3c+3cdB(b1,b2)3c+3cRd_{X}(x,\phi_{b_{1}b_{2}}(x))\leq 3c+3cd_{B}(b_{1},b_{2})\leq 3c+3cR for all xFb1x\in F_{b_{1}} (resp. dX(x,ϕb1b2(x))=dB(b1,b2)Rd_{X}(x,\phi_{b_{1}b_{2}}(x))=d_{B}(b_{1},b_{2})\leq R for all x𝒱(B)x\in{\mathcal{V}}(B)). Hence by Lemma 3.10(2) ϕb1b2\phi_{b_{1}b_{2}} is K3.11=K3.10(3c+3cR)K_{\ref{fibers unif qi}}=K_{\ref{fibers qi}}(3c+3cR)-qi for the metric bundle and K3.11=K3.10(R)K_{\ref{fibers unif qi}}=K_{\ref{fibers qi}}(R)-qi for the metric graph bundle case. ∎

The following corollary is proved as a simple consequence of Lemma 3.10 and Corollary 3.9. (See Corollary 1.14, and Corollary 1.16 of [MS12].) Therefore, we skip the proof of it.

Corollary 3.12.

(Bounded flaring condition) For all kk\in\mathbb{R}, k1k\geq 1 there is a function μk:\mu_{k}:{\mathbb{N}}\rightarrow{\mathbb{N}} such that the following holds:

Suppose π:XB\pi:X\rightarrow B is an (η,c)(\eta,c)-metric bundle or an η\eta-metric graph bundle. Let γB\gamma\subset B be a dotted (1,1)(1,1)-quasigeodesic (resp. a geodesic) joining b1,b2Bb_{1},b_{2}\in B, and let γ1~\tilde{\gamma_{1}}, γ2~\tilde{\gamma_{2}} be two kk-qi lifts of γ\gamma in XX. Suppose γ~i(b1)=xiFb1\tilde{\gamma}_{i}(b_{1})=x_{i}\in F_{b_{1}} and γ~i(b2)=yiFb2\tilde{\gamma}_{i}(b_{2})=y_{i}\in F_{b_{2}}, i=1,2i=1,2.

Then

db2(y1,y2)μk(N)max{db1(x1,x2),1}.d_{b_{2}}(y_{1},y_{2})\leq\mu_{k}(N)\mbox{max}\{d_{b_{1}}(x_{1},x_{2}),1\}.

if dB(b1,b2)Nd_{B}(b_{1},b_{2})\leq N.

In the rest of the paper, we will summarize the conclusion of Corollary 3.12 by saying that a metric (graph) bundle satisfies the bounded flaring condition.

Remark 6.

(Metric bundles in the literature) Metric (graph) bundles appear in several places in other people’s work. In [Bow13, Section 2.1] Bowditch defines stacks of (hyperbolic) spaces which can easily be shown to be quasiisometric to metric graph bundles over an interval in {\mathbb{R}}. Conversely, a metric (graph) bundle whose base is an interval in {\mathbb{R}} is clearly a stack of spaces as per [Bow13, Section 2.1]. In [Why10] Whyte defines coarse bundles which are also quasiisometric to metric graph bundles but with additional restrictions.

3.2. Some natural constructions of metric bundles

In this section, we discuss a few general constructions that produce metric (graph) bundles.

Definition 3.13.

(1)(1) (Metric bundle morphisms) Suppose (Xi,Bi,πi)(X_{i},B_{i},\pi_{i}), i=1,2i=1,2 are metric bundles. A morphism from (X1,B1,π1)(X_{1},B_{1},\pi_{1}) to (X2,B2,π2)(X_{2},B_{2},\pi_{2}) (or simply from X1X_{1} to X2X_{2} when there is no possibility of confusion) consists of a pair of coarsely LL-Lipschitz maps f:X1X2f:X_{1}\rightarrow X_{2} and g:B1B2g:B_{1}\rightarrow B_{2} for some L0L\geq 0 such that π2f=gπ1\pi_{2}\circ f=g\circ\pi_{1}, i.e. the following diagram (Figure 1) is commutative.

X1X_{1}X2X_{2}B1B_{1}B2B_{2}ffπ1\pi_{1}π2\pi_{2}gg
Figure 1.

(2)(2)(Metric graph bundle morphisms) Suppose (Xi,Bi,πi)(X_{i},B_{i},\pi_{i}), i=1,2i=1,2 are metric graph bundles. A morphism from (X1,B1,π1)(X_{1},B_{1},\pi_{1}) to (X2,B2,π2)(X_{2},B_{2},\pi_{2}) (or simply from X1X_{1} to X2X_{2} when there is no possibility of confusion) consists of a pair of coarsely LL-Lipschitz maps f:𝒱(X1)𝒱(X2)f:{\mathcal{V}}(X_{1})\rightarrow{\mathcal{V}}(X_{2}) and g:𝒱(B1)𝒱(B2)g:{\mathcal{V}}(B_{1})\rightarrow{\mathcal{V}}(B_{2}) for some L0L\geq 0 such that π2f=gπ1\pi_{2}\circ f=g\circ\pi_{1}.

(3)(3) (Isomorphisms) A morphism (f,g)(f,g) from a metric (graph) bundle (X1,B1,π1)(X_{1},B_{1},\pi_{1}) to a metric (graph) bundle (X2,B2,π2)(X_{2},B_{2},\pi_{2}) is called an isomorphism if there is a morphism (f,g)(f^{\prime},g^{\prime}) from (X2,B2,π2)(X_{2},B_{2},\pi_{2}) to (X1,B1,π1)(X_{1},B_{1},\pi_{1}) such that ff^{\prime} is a coarse inverse of ff and gg^{\prime} is a coarse inverse of gg.

We note that for any morphism (f,g)(f,g) from a metric (graph) bundle (X1,B1,π1)(X_{1},B_{1},\pi_{1}) to a metric (graph) bundle (X2,B2,π2)(X_{2},B_{2},\pi_{2}) we have f(π11(b))π21(g(b))f(\pi^{-1}_{1}(b))\subset\pi^{-1}_{2}(g(b)) for all bB1b\in B_{1}. We will denote by fb:π11(b)π21(g(b))f_{b}:\pi^{-1}_{1}(b)\rightarrow\pi^{-1}_{2}(g(b)) the restriction of ff to π11(b)\pi^{-1}_{1}(b) for all bB1b\in B_{1}. We shall refer to these maps as the fiber maps of the morphisms. We also note that in the case of metric graph bundles coarse Lipschitzness is equivalent to Lipschitzness.

Lemma 3.14.

Given k1,K1k\geq 1,K\geq 1 and L0L\geq 0 there are constants L3.14,K3.14L_{\ref{metric bundle map}},K_{\ref{metric bundle map}} such that the following hold.

Suppose (f,g)(f,g) is a morphism of metric (graph) bundles as in the definition above. Then the following hold:

(1) For all bB1b\in B_{1} the map fb:π11(b)π21(g(b))f_{b}:\pi^{-1}_{1}(b)\rightarrow\pi^{-1}_{2}(g(b)) is coarsely L3.14L_{\ref{metric bundle map}}-Lipschitz with respect to the induced length metric on the fibers.

(2) Suppose γ:IB1\gamma:I\rightarrow B_{1} is a dotted (1,1)(1,1)-quasigeodesic (or simply a geodesic in the case of a metric graph bundle) and suppose γ~\tilde{\gamma} is a kk-qi lift of γ\gamma. If gg is a KK-qi embedding then fγ~f\circ\tilde{\gamma} is a K3.14=K3.14(k,K,L)K_{\ref{metric bundle map}}=K_{\ref{metric bundle map}}(k,K,L)-qi lift of gγg\circ\gamma.

Proof.

We shall check the lemma only for the metric bundle case because for metric graph bundles the proofs are similar and in fact easier.

Suppose πi:XiBi\pi_{i}:X_{i}\rightarrow B_{i}, i=1,2i=1,2 are (ηi,ci)(\eta_{i},c_{i})-metric bundles.

(1) Let bB1b\in B_{1} and x,yπ11(b)x,y\in\pi^{-1}_{1}(b) be such that db(x,y)1d_{b}(x,y)\leq 1. Since ff is coarsely LL-Lipschitz, dX2(f(x),f(y))L+LdX1(x,y)L+Ldb(x,y)2Ld_{X_{2}}(f(x),f(y))\leq L+Ld_{X_{1}}(x,y)\leq L+Ld_{b}(x,y)\leq 2L. Now, the fibers of π2\pi_{2} are uniformly properly embedded as measured by η2\eta_{2}. Hence, dg(b)(f(x),f(y))η2(2L)d_{g(b)}(f(x),f(y))\leq\eta_{2}(2L). Therefore, by Lemma 2.6 the fiber map fb:π11(b)π21(g(b))f_{b}:\pi^{-1}_{1}(b)\rightarrow\pi^{-1}_{2}(g(b)) is η2(2L)\eta_{2}(2L)-coarsely Lipschitz. Hence, L3.14=η2(2L)L_{\ref{metric bundle map}}=\eta_{2}(2L) will do.

(2) Let γ2=gγ\gamma_{2}=g\circ\gamma and γ~2=fγ~\tilde{\gamma}_{2}=f\circ\tilde{\gamma}. Then clearly, π2γ~2=γ2\pi_{2}\circ\tilde{\gamma}_{2}=\gamma_{2} whence γ~2\tilde{\gamma}_{2} is a lift of γ2\gamma_{2}. By Lemma 2.3(1) γ~2=fγ~\tilde{\gamma}_{2}=f\circ\tilde{\gamma} is coarsely (kL,kL+L)(kL,kL+L)-Lipschitz. Hence, for all s,tIs,t\in I we have

dX2(γ~2(s),γ~2(t))kL|st|+(kL+L).d_{X_{2}}(\tilde{\gamma}_{2}(s),\tilde{\gamma}_{2}(t))\leq kL|s-t|+(kL+L).

On the other hand, for s,tIs,t\in I we have

dX2(γ~2(s),γ~2(t))dB2(π2γ~2(s),π2γ~2(t))=dB2(γ2(s),γ2(t)).d_{X_{2}}(\tilde{\gamma}_{2}(s),\tilde{\gamma}_{2}(t))\geq d_{B_{2}}(\pi_{2}\circ\tilde{\gamma}_{2}(s),\pi_{2}\circ\tilde{\gamma}_{2}(t))=d_{B_{2}}(\gamma_{2}(s),\gamma_{2}(t)).

However, by Lemma 2.3(2) γ2=gγ\gamma_{2}=g\circ\gamma is a (K,2K)(K,2K)-qi embedding. Hence, we have

dX2(γ~2(s),γ~2(t))dB2(γ2(s),γ2(t))2K+1K|st|.d_{X_{2}}(\tilde{\gamma}_{2}(s),\tilde{\gamma}_{2}(t))\geq d_{B_{2}}(\gamma_{2}(s),\gamma_{2}(t))\geq-2K+\frac{1}{K}|s-t|.

Therefore, it follows that γ~2\tilde{\gamma}_{2} is a K3.14=max{2K,kL+L}K_{\ref{metric bundle map}}=\max\{2K,kL+L\}-qi lift of γ2\gamma_{2}. ∎

The following theorem characterizes isomorphisms of metric (graph) bundles.

Theorem 3.15.

If (f,g)(f,g) is an isomorphism of metric (graph) bundles as in the above definition then the maps f,gf,g are quasiisometries and all the fiber maps are uniform quasiisometries.

Conversely, if the map gg is a qi and the fiber maps are uniform qi then (f,g)(f,g) is an isomorphism.

Proof.

We shall prove the theorem in the case of a metric bundle only. The proof in the case of a metric graph bundle is very similar and hence we skip it.

If (f,g)(f,g) is an isomorphism then f,gf,g are qi by Lemma 2.2(1). We need to show that the fiber maps are quasiisometries.

Suppose (f,g)(f^{\prime},g^{\prime}) is a coarse inverse of (f,g)(f,g) such that dX2(ff(x2),x2)Rd_{X_{2}}(f\circ f^{\prime}(x_{2}),x_{2})\leq R and dX1(ff(x1),x1)Rd_{X_{1}}(f^{\prime}\circ f(x_{1}),x_{1})\leq R for all x1X1x_{1}\in X_{1} and x2X2x_{2}\in X_{2}. It follows that for all b1B1,b2B2b_{1}\in B_{1},b_{2}\in B_{2} we have dB1(b1,gg(b1))Rd_{B_{1}}(b_{1},g^{\prime}\circ g(b_{1}))\leq R and dB2(b2,gg(b2))Rd_{B_{2}}(b_{2},g\circ g^{\prime}(b_{2}))\leq R since the maps π1,π2\pi_{1},\pi_{2} are 11-Lipschitz. Suppose f,gf^{\prime},g^{\prime} are coarsely LL^{\prime}-Lipschitz. Let L1=η2(2L)L_{1}=\eta_{2}(2L) and L2=η1(2L)L_{2}=\eta_{1}(2L^{\prime}). Then for all uB1u\in B_{1}, fu:π11(u)π21(g(u))f_{u}:\pi^{-1}_{1}(u)\rightarrow\pi^{-1}_{2}(g(u)) is coarsely L1L_{1}-Lipschitz and for all vB2v\in B_{2}, fv:π21(v)π11(g(v))f^{\prime}_{v}:\pi^{-1}_{2}(v)\rightarrow\pi^{-1}_{1}(g^{\prime}(v)) is coarsely L2L_{2}-Lipschitz by Lemma 3.14(1).

Let bB1b\in B_{1}. To show that fb:π11(b)π21(g(b)f_{b}:\pi^{-1}_{1}(b)\rightarrow\pi^{-1}_{2}(g(b) is a uniform quasiisometry, it is enough by Lemma 2.2(1) to find a uniformly coarsely Lipschitz map π21(g(b))π11(b)\pi^{-1}_{2}(g(b))\rightarrow\pi^{-1}_{1}(b) which is uniform coarse inverse of fbf_{b}. We already know that fg(b)f^{\prime}_{g(b)} is L2L_{2}-coarsely Lipschitz. Let b1=gg(b)b_{1}=g^{\prime}\circ g(b). We also noted that dB1(b,b1)Rd_{B_{1}}(b,b_{1})\leq R. Hence, it follows by Corollary 3.9 and Corollary 3.11 that we have a K3.10(R)K_{\ref{fibers qi}}(R)-qi ϕb1b:π11(b1)π11(b)\phi_{b_{1}b}:\pi^{-1}_{1}(b_{1})\rightarrow\pi^{-1}_{1}(b) such that dX1(x,ϕb1b(x))3c1+3c1Rd_{X_{1}}(x,\phi_{b_{1}b}(x))\leq 3{c_{1}}+3{c_{1}}R for all xπ11(b1)x\in\pi^{-1}_{1}(b_{1}). Let h=ϕb1bfg(b)h=\phi_{b_{1}b}\circ f^{\prime}_{g(b)}. We claim that hh is a uniformly coarsely Lipschitz, uniform coarse inverse of fbf_{b}. Since fg(b)f^{\prime}_{g(b)} is L2L_{2}-coarsely Lipschitz and clearly ϕb1b\phi_{b_{1}b} is K3.10(R)K_{\ref{fibers qi}}(R)-coarsely Lipschitz, it follows by Lemma 2.3(1) that hh is (L2K3.10(R)+K3.10(R))(L_{2}K_{\ref{fibers qi}}(R)+K_{\ref{fibers qi}}(R))-coarsely Lipschitz.

Moreover, for all xπ11(b)x\in\pi^{-1}_{1}(b) we have dX1(x,hfb(x))dX1(x,fg(b)fb(x))+dX1(fg(b)fb(x),hfb(x))R+3c1+3c1Rd_{X_{1}}(x,h\circ f_{b}(x))\leq d_{X_{1}}(x,f^{\prime}_{g(b)}\circ f_{b}(x))+d_{X_{1}}(f^{\prime}_{g(b)}\circ f_{b}(x),h\circ f_{b}(x))\leq R+3{c_{1}}+3{c_{1}}R. Hence, db(x,hfb(x))η1(R+3c1+3c1R)d_{b}(x,h\circ f_{b}(x))\leq\eta_{1}(R+3{c_{1}}+3{c_{1}}R). Let yπ21(g(b))y\in\pi^{-1}_{2}(g(b)). Then

dX2(y,fbh(y))=dX2(y,fϕb1bf(y))dX2(y,ff(y))+dX2(ff(y),fϕb1bf(y))d_{X_{2}}(y,f_{b}\circ h(y))=d_{X_{2}}(y,f\circ\phi_{b_{1}b}\circ f^{\prime}(y))\leq d_{X_{2}}(y,f\circ f^{\prime}(y))+d_{X_{2}}(f\circ f^{\prime}(y),f\circ\phi_{b_{1}b}\circ f^{\prime}(y))
R+L(3c1+3c1R)+L\leq R+L(3{c_{1}}+3{c_{1}}R)+L

since dX1(f(y),ϕb1bf(y))3c1+3c1Rd_{X_{1}}(f^{\prime}(y),\phi_{b_{1}b}\circ f^{\prime}(y))\leq 3{c_{1}}+3{c_{1}}R. Hence, dg(b)(y,fbh(y))η2(R+L(3c1+3c1R)+L)d_{g(b)}(y,f_{b}\circ h(y))\leq\eta_{2}(R+L(3{c_{1}}+3{c_{1}}R)+L). Hence by Lemma 2.2(1) fbf_{b} is a uniform qi.

Conversely, suppose all the fiber maps of the morphism (f,g)(f,g) are (λ,ϵ)(\lambda,\epsilon)-qi which are RR-coarsely surjective and gg is a (λ1,ϵ1)(\lambda_{1},\epsilon_{1})-qi which is R1R_{1}-surjective. Let gg^{\prime} be a coarsely (K,C)(K,C)-quasiisometric, DD-coarse inverse of gg where K=K2.2(λ1,ϵ1,R1)K=K_{\ref{elem-lemma1}}(\lambda_{1},\epsilon_{1},R_{1}), C=C2.2(λ1,ϵ1,R1)C=C_{\ref{elem-lemma1}}(\lambda_{1},\epsilon_{1},R_{1}) and D=D2.2(λ1,ϵ1,R1)D=D_{\ref{elem-lemma1}}(\lambda_{1},\epsilon_{1},R_{1}). For all uB1u\in B_{1} let f¯u\bar{f}_{u} be a D1D_{1}-coarse inverse of fu:FuFg(u)f_{u}:F_{u}\rightarrow F_{g(u)}. We will define a map f:X2X1f^{\prime}:X_{2}\rightarrow X_{1} such that (f,g)(f^{\prime},g^{\prime}) is morphism from X2X_{2} to X1X_{1} and ff^{\prime} is a coarse inverse of ff as follows.

For all uB2u\in B_{2} we define fu:FuFg(u)f^{\prime}_{u}:F_{u}\rightarrow F_{g^{\prime}(u)} as the composition f¯g(u)ϕug(g(u))\bar{f}_{g^{\prime}(u)}\circ\phi_{ug(g^{\prime}(u))} where ϕug(g(u))\phi_{ug(g^{\prime}(u))} is a fiber identification map as constructed in the proof of Corollary 3.9. Collectively this defines ff^{\prime}. Now we shall check that ff^{\prime} satisfies the desired properties.

(i) We first check that (f,g)(f^{\prime},g^{\prime}) is a morphism. It is clear from the definition that π1f=gπ2\pi_{1}\circ f^{\prime}=g^{\prime}\circ\pi_{2}. Hence we will be done by showing that ff^{\prime} is coarsely Lipschitz. By Lemma 2.6 it is enough to show that for all u2,v2B2u_{2},v_{2}\in B_{2} and xFu2,yFv2x\in F_{u_{2}},y\in F_{v_{2}} with dX2(x,y)1d_{X_{2}}(x,y)\leq 1, dX1(f(x),f(y))d_{X_{1}}(f^{\prime}(x),f^{\prime}(y)) is uniformly small. We note that dB2(u2,v2)1d_{B_{2}}(u_{2},v_{2})\leq 1. Let u1=g(u2)u_{1}=g^{\prime}(u_{2}) and v1=g(v2)v_{1}=g^{\prime}(v_{2}). Then dB1(u1,v1)K+Cd_{B_{1}}(u_{1},v_{1})\leq K+C, dB2(u2,g(u1))Dd_{B_{2}}(u_{2},g(u_{1}))\leq D and dB2(v2,g(v1))Dd_{B_{2}}(v_{2},g(v_{1}))\leq D. This means dX2(x,ϕu2g(u1)(x))3Dc2+3c2d_{X_{2}}(x,\phi_{u_{2}g(u_{1})}(x))\leq 3Dc_{2}+3c_{2} and dX2(y,ϕv2g(v1))3Dc2+3c2d_{X_{2}}(y,\phi_{v_{2}g(v_{1})})\leq 3Dc_{2}+3c_{2} by Lemma 3.8 and Corollary 3.9. Hence, dX2(ϕu2g(u1)(x),ϕv2g(v1)(y))1+6c2+6Dc2d_{X_{2}}(\phi_{u_{2}g(u_{1})}(x),\phi_{v_{2}g(v_{1})}(y))\leq 1+6c_{2}+6Dc_{2}. Let x2=ϕu2g(u1)(x)x_{2}=\phi_{u_{2}g(u_{1})}(x), y2=ϕv2g(v1)(y)y_{2}=\phi_{v_{2}g(v_{1})}(y), x1=f(x2)=f¯g(u1)(x2)x_{1}=f^{\prime}(x_{2})=\bar{f}_{g(u_{1})}(x_{2}) and y1=f(y2)=f¯g(v1)(y2)y_{1}=f^{\prime}(y_{2})=\bar{f}_{g(v_{1})}(y_{2}). Therefore, dX2(x2,y2)1+6c2+6Dc2=R2d_{X_{2}}(x_{2},y_{2})\leq 1+6c_{2}+6Dc_{2}=R_{2}, say and we want to show that dX1(x1,y1)d_{X_{1}}(x_{1},y_{1}) is uniformly small. Let x2=f(x1)=fu1(x1),y2=f(y1)=fv1(y1)x^{\prime}_{2}=f(x_{1})=f_{u_{1}}(x_{1}),y^{\prime}_{2}=f(y_{1})=f_{v_{1}}(y_{1}). Then dX2(x2,x2)D1d_{X_{2}}(x_{2},x^{\prime}_{2})\leq D_{1} and dX2(y2,y2)D1d_{X_{2}}(y_{2},y^{\prime}_{2})\leq D_{1}. Hence, dX2(x2,y2)R2+2D1d_{X_{2}}(x^{\prime}_{2},y^{\prime}_{2})\leq R_{2}+2D_{1}. Since dB1(u1,v1)K+Cd_{B_{1}}(u_{1},v_{1})\leq K+C there is a point y1Fu1y^{\prime}_{1}\in F_{u_{1}} such that dX1(x1,y1)(K+C)c1+c1d_{X_{1}}(x_{1},y^{\prime}_{1})\leq(K+C)c_{1}+c_{1}. Hence, dX2(x2,f(y1))((K+C)c1+c1).L+Ld_{X_{2}}(x^{\prime}_{2},f(y^{\prime}_{1}))\leq((K+C)c_{1}+c_{1}).L+L. Hence, dX2(f(y1),y2)dX2(f(y1),x2)+dX2(x2,y2)((K+C)c1+c1).L+L+2D1+R2d_{X_{2}}(f(y^{\prime}_{1}),y^{\prime}_{2})\leq d_{X_{2}}(f(y^{\prime}_{1}),x^{\prime}_{2})+d_{X_{2}}(x^{\prime}_{2},y^{\prime}_{2})\leq((K+C)c_{1}+c_{1}).L+L+2D_{1}+R_{2}. This implies that dv2(f(y1),f(y1))η2(((K+C)c1+c1).L+L+2D1+R2)=D2d_{v_{2}}(f(y^{\prime}_{1}),f(y_{1}))\leq\eta_{2}(((K+C)c_{1}+c_{1}).L+L+2D_{1}+R_{2})=D_{2}, say. Since fv1f_{v_{1}} is a (λ,ϵ)(\lambda,\epsilon)-qi we have ϵ+1λdv1(y1,y1)D2-\epsilon+\frac{1}{\lambda}d_{v_{1}}(y_{1},y^{\prime}_{1})\leq D_{2}. Hence, dv1(y1,y1)(ϵ+D2)λd_{v_{1}}(y_{1},y^{\prime}_{1})\leq(\epsilon+D_{2})\lambda. Thus, dX1(x1,y1)dX1(x1,y1)+dX1(y1,y1)(K+C)c1+c1+(ϵ+D2)λd_{X_{1}}(x_{1},y_{1})\leq d_{X_{1}}(x_{1},y^{\prime}_{1})+d_{X_{1}}(y^{\prime}_{1},y_{1})\leq(K+C)c_{1}+c_{1}+(\epsilon+D_{2})\lambda.

(ii) We already know that gg^{\prime} is a coarse inverse of gg. Hence we will be done by checking that ff^{\prime} is a coarse inverse of ff. We will check only that d(ff,IdX1)<d(f^{\prime}\circ f,Id_{X_{1}})<\infty leaving the proof of d(ff,IdX2)<d(f\circ f^{\prime},Id_{X_{2}})<\infty for the reader. Suppose bB1b\in B_{1} and xπ11(b)x\in\pi^{-1}_{1}(b). Then f(f(x))=f¯gg(b)ϕg(b)gg(g(b))fb(x)f^{\prime}(f(x))=\bar{f}_{g^{\prime}\circ g(b)}\circ\phi_{g(b)g\circ g^{\prime}(g(b))}\circ f_{b}(x). We want to show that dX1(x,f(f(x)))d_{X_{1}}(x,f^{\prime}(f(x))) is uniformly small. Let h=fgg(b)f¯gg(b)h=f_{g^{\prime}\circ g(b)}\circ\bar{f}_{g^{\prime}\circ g(b)}. Then dX2(f(x),f(f(f(x))))=dX2(fb(x),hϕg(b)gg(g(b))fb(x))dX2(fb(x),ϕg(b)gg(g(b))(fb(x)))+dX2(ϕg(b)gg(g(b))(fb(x)),hϕg(b)gg(g(b))fb(x))d_{X_{2}}(f(x),f(f^{\prime}(f(x))))=d_{X_{2}}(f_{b}(x),h\circ\phi_{g(b)g\circ g^{\prime}(g(b))}\circ f_{b}(x))\leq d_{X_{2}}(f_{b}(x),\phi_{g(b)g\circ g^{\prime}(g(b))}(f_{b}(x)))+d_{X_{2}}(\phi_{g(b)g\circ g^{\prime}(g(b))}(f_{b}(x)),h\circ\phi_{g(b)g\circ g^{\prime}(g(b))}\circ f_{b}(x)). Now since, d(gg,IdB2)Dd(g\circ g^{\prime},Id_{B_{2}})\leq D, dX2(fb(x),ϕg(b)gg(g(b))(fb(x)))3Dc2+3c2d_{X_{2}}(f_{b}(x),\phi_{g(b)g\circ g^{\prime}(g(b))}(f_{b}(x)))\leq 3Dc_{2}+3c_{2}. Since d(h,IdFg(g(g(b))))D1d(h,Id_{F_{g(g^{\prime}(g(b)))}})\leq D_{1} we have dX2(ϕg(b)gg(g(b))(fb(x)),hϕg(b)gg(g(b))fb(x))D1d_{X_{2}}(\phi_{g(b)g\circ g^{\prime}(g(b))}(f_{b}(x)),h\circ\phi_{g(b)g\circ g^{\prime}(g(b))}\circ f_{b}(x))\leq D_{1}. Thus dX2(f(x),f(f(f(x))))3Dc2+3c2+D1d_{X_{2}}(f(x),f(f^{\prime}(f(x))))\leq 3Dc_{2}+3c_{2}+D_{1}. Hence, it is enough to show that ff is a proper embedding. Here is how this is proved. Suppose b,bBb,b^{\prime}\in B, xπ11(b)x\in\pi^{-1}_{1}(b) and xπ11(b)x^{\prime}\in\pi^{-1}_{1}(b^{\prime}). Suppose dX2(f(x),f(x))Nd_{X_{2}}(f(x),f(x^{\prime}))\leq N for some N0N\geq 0. This implies dB2(g(b),g(b))=dB2(π2f(x),π2f(x))Nd_{B_{2}}(g(b),g(b^{\prime}))=d_{B_{2}}(\pi_{2}\circ f(x),\pi_{2}\circ f(x^{\prime}))\leq N. Since gg is a (λ1,ϵ1)(\lambda_{1},\epsilon_{1})-qi we have ϵ1+dB1(b,b)/λ1N-\epsilon_{1}+d_{B_{1}}(b,b^{\prime})/\lambda_{1}\leq N, i.e. dB1(b,b)(N+ϵ1)λ1=N1d_{B_{1}}(b,b^{\prime})\leq(N+\epsilon_{1})\lambda_{1}=N_{1}, say. Hence by Corollary 3.9 there is a point x′′π11(b)x^{\prime\prime}\in\pi^{-1}_{1}(b^{\prime}) such that dX1(x,x′′)3N1c1+3c1d_{X_{1}}(x,x^{\prime\prime})\leq 3N_{1}c_{1}+3c_{1}. Since ff is coarsely LL-Lipschitz we have dX2(f(x),f(x′′))L(3N1c1+3c1)+Ld_{X_{2}}(f(x),f(x^{\prime\prime}))\leq L(3N_{1}c_{1}+3c_{1})+L. It follows that d(f(x),f(x′′))d(f(x),f(x))+d(f(x),f(x′′))N+L(3N1c1+3c1)+L=N2d(f(x^{\prime}),f(x^{\prime\prime}))\leq d(f(x^{\prime}),f(x))+d(f(x),f(x^{\prime\prime}))\leq N+L(3N_{1}c_{1}+3c_{1})+L=N_{2}, say. Hence, dg(b)(f(x),f(x′′))η2(N2)d_{g(b^{\prime})}(f(x^{\prime}),f(x^{\prime\prime}))\leq\eta_{2}(N_{2}). Since fbf_{b^{\prime}} is a (λ,ϵ)(\lambda,\epsilon)-qi we have dX1(x,x′′)db(x,x′′)λ(ϵ+η2(N2))d_{X_{1}}(x^{\prime},x^{\prime\prime})\leq d_{b^{\prime}}(x^{\prime},x^{\prime\prime})\leq\lambda(\epsilon+\eta_{2}(N_{2})). Hence, dX1(x,x)dX1(x,x′′)+dX1(x,x′′)3N1c1+3c1+λ(ϵ+η2(N2))d_{X_{1}}(x,x^{\prime})\leq d_{X_{1}}(x,x^{\prime\prime})+d_{X_{1}}(x^{\prime},x^{\prime\prime})\leq 3N_{1}c_{1}+3c_{1}+\lambda(\epsilon+\eta_{2}(N_{2})). This completes the proof. ∎

Definition 3.16.

(Subbundle) Suppose (Xi,B,πi)(X_{i},B,\pi_{i}), i=1,2i=1,2 are metric (graph) bundles with the same base space BB. We say that (X1,B,π1)(X_{1},B,\pi_{1}) is subbundle of (X2,B,π2)(X_{2},B,\pi_{2}) or simply X1X_{1} is a subbundle of X2X_{2} if there is a metric (graph) bundle morphism (f,g)(f,g) from (X1,B,π1)(X_{1},B,\pi_{1}) to (X2,B,π2)(X_{2},B,\pi_{2}) such that all the fiber maps fbf_{b}, bBb\in B are uniform qi embeddings and gg is the identity map on BB (resp. on 𝒱(B){\mathcal{V}}(B)).

The most important example of a subbundle that concerns us is that of ladders which we discuss in a later section. The following gives another way to construct a metric (graph) bundle. We omit the proof since it is immediate.

Lemma 3.17.

(Restriction bundle) Suppose π:XB\pi:X\rightarrow B is a metric (graph) bundle and ABA\subset B is a connected subset such that any pair of points in AA can be joined by a path of finite length in AA (resp. AA is a connected subgraph). Then the restriction of π\pi to Y=π1(A)Y=\pi^{-1}(A) gives a metric (graph) bundle with the same parameters as that of π:XB\pi:X\rightarrow B where AA and YY are given the induced length metrics from BB and XX respectively.

Moreover, if f:YXf:Y\rightarrow X and g:ABg:A\rightarrow B are the inclusion maps then (f,g):(Y,A)(X,B)(f,g):(Y,A)\rightarrow(X,B) is a morphism of metric (graph) bundles.

Definition 3.18.

(1)(1) (Pullback of a metric bundle) Given a metric bundle (X,B,π)(X,B,\pi) and a coarsely Lipschitz map g:B1Bg:B_{1}\rightarrow B a pullback of (X,B,π)(X,B,\pi) under gg is a metric bundle (X1,B1,π1)(X_{1},B_{1},\pi_{1}) together with a morphism (f:X1X,g:B1B)(f:X_{1}\rightarrow X,g:B_{1}\rightarrow B) such that the following universal property holds: Suppose π2:YB1\pi_{2}:Y\rightarrow B_{1} is another metric bundle and (fY,g)(f^{Y},g) is a morphism from YY to XX. Then there is a coarsely unique morphism (f,IdB1)(f^{\prime},Id_{B_{1}}) from YY to X1X_{1} making the following diagram commutative.

X1X_{1}XXB1B_{1}BBYYffπ1\pi_{1}ggπ\piπ2\pi_{2}fYf^{Y}ff^{\prime}
Figure 2.

(2)(2) (Pullback of a metric graph bundle) In the case of a metric graph bundle, the diagram is replaced by one where we have the vertex sets instead of the whole spaces.

The following lemma follows by a standard argument.

Lemma 3.19.

Suppose we have a metric bundle (X,B,π)(X,B,\pi) and a coarsely Lipschitz map g:B1Bg:B_{1}\rightarrow B for which there are two pullbacks i.e. metric bundles (Xi,B1,πi)(X_{i},B_{1},\pi_{i}) together with a morphisms (fi:XiX,g:B1B)(f_{i}:X_{i}\rightarrow X,g:B_{1}\rightarrow B), i=1,2i=1,2 satisfying the universal property of the Definition 3.18. Then there is a coarsely unique metric (graph) bundle isomorphism from X1X_{1} to X2X_{2}.

With the above lemma in mind, in the context of Definition 3.18, we say that f:X1Xf:X_{1}\rightarrow X is the pullback of XX under g:B1Bg:B_{1}\rightarrow B or simply X1X_{1} is the pullback of XX under gg when all the other maps are understood.

Lemma 3.20.

Given L0L\geq 0 and functions ϕ1,ϕ2:[0,)[0,)\phi_{1},\phi_{2}:[0,\infty)\rightarrow[0,\infty) there is a function ϕ:[0,)[0,)\phi:[0,\infty)\rightarrow[0,\infty) such that the following hold:

Suppose we have the following commutative diagram of maps between metric spaces satisfying the properties (1)-(3) below.

X1X_{1}XXB1B_{1}YYffπ1\pi_{1}π2\pi_{2}fYf^{Y}ff^{\prime}

(1) All the maps (except possibly ff^{\prime}) are coarsely LL-Lipschitz.

(2) If dB1(b,b)Nd_{B_{1}}(b,b^{\prime})\leq N then Hd(π11(b),π11(b))ϕ1(N)Hd(\pi^{-1}_{1}(b),\pi^{-1}_{1}(b^{\prime}))\leq\phi_{1}(N) for all b,bB1b,b^{\prime}\in B_{1} and N[0,)N\in[0,\infty).

(3) The restrictions of ff on the fibers of π1\pi_{1} are uniformly properly embedded as measured by ϕ2\phi_{2}.

Then dY(y,y)Rd_{Y}(y,y^{\prime})\leq R implies dX1(f(y),f(y))ϕ(R)d_{X_{1}}(f^{\prime}(y),f^{\prime}(y^{\prime}))\leq\phi(R) for all y,yYy^{\prime},y\in Y and R[0,)R\in[0,\infty). In particular, if YY is a length space or the vertex set of a connected metric graph with restricted metric then ff^{\prime} is coarsely ϕ(1)\phi(1)-Lipschitz.

Moreover, ff^{\prime} is coarsely unique, i.e. there is a constant D>0D>0 such that if f′′:YX1f^{\prime\prime}:Y\rightarrow X_{1} is another map making the above diagram commutative then d(f,f′′)Dd(f^{\prime},f^{\prime\prime})\leq D.

Proof.

Suppose y,yYy,y^{\prime}\in Y with dY(y,y)Rd_{Y}(y,y^{\prime})\leq R. Let x=f(y),x=f(y)x=f^{\prime}(y),x^{\prime}=f^{\prime}(y^{\prime}). Then dB1(π1(x),π1(x))=dB1(π2(y),π2(y))LR+Ld_{B_{1}}(\pi_{1}(x),\pi_{1}(x^{\prime}))=d_{B_{1}}(\pi_{2}(y),\pi_{2}(y^{\prime}))\leq LR+L. Let b=π2(y),b=π2(y)b=\pi_{2}(y),b^{\prime}=\pi_{2}(y^{\prime}). Then Hd(π11(b),π11(b))ϕ1(LR+L)=R1Hd(\pi^{-1}_{1}(b),\pi^{-1}_{1}(b^{\prime}))\leq\phi_{1}(LR+L)=R_{1}, say. Let x1π11(b)x^{\prime}_{1}\in\pi^{-1}_{1}(b^{\prime}) be such that dX1(x,x1)R1d_{X_{1}}(x,x^{\prime}_{1})\leq R_{1}. Then dX(f(x),f(x1))LR1+Ld_{X}(f(x),f(x^{\prime}_{1}))\leq LR_{1}+L. On the other hand dX(f(x),f(x))=dX(fY(y),fY(y))LR+Ld_{X}(f(x),f(x^{\prime}))=d_{X}(f^{Y}(y),f^{Y}(y^{\prime}))\leq LR+L. By triangle inequality, we have dX(f(x),f(x1))LR+L+LR1+L=2L+RL+R1Ld_{X}(f(x^{\prime}),f(x^{\prime}_{1}))\leq LR+L+LR_{1}+L=2L+RL+R_{1}L. Hence, by the hypothesis (3) of the lemma dX1(x,x1)ϕ2(2L+RL+R1L)d_{X_{1}}(x^{\prime},x^{\prime}_{1})\leq\phi_{2}(2L+RL+R_{1}L). Thus dX1(x,x)dX1(x,x1)+dX1(x,x1)R1+ϕ2(2L+RL+R1L)d_{X_{1}}(x,x^{\prime})\leq d_{X_{1}}(x,x^{\prime}_{1})+d_{X_{1}}(x^{\prime},x^{\prime}_{1})\leq R_{1}+\phi_{2}(2L+RL+R_{1}L). Hence, we may choose ϕ(t)=ϕ1(Lt+L)+ϕ2(2L+tL+Lϕ1(Lt+L))\phi(t)=\phi_{1}(Lt+L)+\phi_{2}(2L+tL+L\phi_{1}(Lt+L)).

In case YY is a length space or the vertex set of a connected metric graph it follows by Lemma 2.6 that ff^{\prime} is coarsely ϕ(1)\phi(1)-Lipschitz.

Lastly, suppose f′′:YX1f^{\prime\prime}:Y\rightarrow X_{1} is another map making the diagram commutative. In particular we have fY=ff=ff′′f^{Y}=f\circ f^{\prime}=f\circ f^{\prime\prime}. Hence for all yYy\in Y we have f(f(y))=f(f′′(y))f(f^{\prime}(y))=f(f^{\prime\prime}(y)). Since π1(f(y))=π1(f′′(y))=π2(y)\pi_{1}(f^{\prime}(y))=\pi_{1}(f^{\prime\prime}(y))=\pi_{2}(y) by the hypothesis (3) of the lemma it follows that dX1(f(y),f′′(y))ϕ2(0)d_{X_{1}}(f^{\prime}(y),f^{\prime\prime}(y))\leq\phi_{2}(0). Hence d(f,f′′)ϕ2(0)d(f^{\prime},f^{\prime\prime})\leq\phi_{2}(0). ∎

Remark 7.

We note that the condition (2) of the lemma above holds in case π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} is a metric (graph) bundle.

Proposition 3.21.

(Pullbacks of metric bundles) Suppose (X,B,π)(X,B,\pi) is a metric bundle and g:B1Bg:B_{1}\rightarrow B is a Lipschitz map. Then there is a pullback.

More precisely the following hold: Suppose X1X_{1} is the set theoretic pullback with the induced length metric from X×B1X\times B_{1} and let π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} be the projection on the second coordinate and let f:X1Xf:X_{1}\rightarrow X be the projection on the first coordinate. Then (1) π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} is metric bundle and ff is a coarsely Lipschitz map so that (f,g)(f,g) is a morphism from X1X_{1} to XX. (2) f:X1Xf:X_{1}\rightarrow X is the metric bundle pullback of XX under gg. (3) All the fiber maps fb:π11(b)π1(g(b))f_{b}:\pi^{-1}_{1}(b)\rightarrow\pi^{-1}(g(b)), bB1b\in B_{1} are isometries with respect to induced length metrics from X1X_{1} and XX respectively.

Proof.

By definition X1={(x,t)X×B1:g(t)=π(x)}X_{1}=\{(x,t)\in X\times B_{1}:g(t)=\pi(x)\}. We put on it the induced length metric from X×B1X\times B_{1}. Let π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} be the restriction of the projection map X×B1B1X\times B_{1}\rightarrow B_{1} to X1X_{1}. We first show that X1X_{1} is a length space. Suppose gg is LL-Lipschitz. Let (x,s),(y,t)X1(x,s),(y,t)\in X_{1}. Let α\alpha be a rectifiable path joining s,ts,t in B1B_{1}. Then gαg\circ\alpha is a rectifiable path in BB of length at most l(α)Ll(\alpha)L. By Lemma 3.8 and Corollary 3.9 this path can be lifted to a rectifiable path in XX starting from xx and ending at some point say zz in FtF_{t} such that the length of the path is at most 3c+3cLl(α)3c+3cLl(\alpha). By construction this lift is contained in X1X_{1}. Finally we can join (y,t),(z,t)(y,t),(z,t) by a rectifiable path in FtF_{t}. This show that (x,s)(x,s) and (y,t)(y,t) can be joined in X1X_{1} by a rectifiable path. This proves that X1X_{1} is a length space. Now, since π11(t)=π1(g(t))\pi^{-1}_{1}(t)=\pi^{-1}(g(t)) is uniformly properly embedded in XX for all tB1t\in B_{1} and XX is properly embedded in X×B1X\times B_{1}, π11(t)\pi^{-1}_{1}(t) is uniformly properly embedded in X1X_{1} for all tB1t\in B_{1}. The same argument also shows that any path in B1B_{1} of length at most 11 can be lifted to a path of length at most 3c+3cL3c+3cL verifying the condition 2 of metric bundles.

Hence (X1,B1,π1)(X_{1},B_{1},\pi_{1}) is a metric bundle. Let f:X1Xf:X_{1}\rightarrow X be the restriction of the projection map X×B1XX\times B_{1}\rightarrow X to X1X_{1}. Clearly f:X1Xf:X_{1}\rightarrow X is a morphism of metric bundles. Finally, we check the universal property. If there is a metric bundle π2:YB1\pi_{2}:Y\rightarrow B_{1} and a morphism (fY,g)(f^{Y},g) from YY to XX then there is a map f:YX1f^{\prime}:Y\rightarrow X_{1} making the diagram 2 commutative since we are working with the set theoretic pullback. That ff^{\prime} is a coarsely unique, coarsely Lipschitz map now follows from Lemma 3.20. In fact, condition (2) of that lemma follows from Lemma 3.10(1) since π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} is a metric bundle and (3) follows because fibers of metric bundles are uniformly properly embedded and in this case the restriction of ff, π11(b)π1(g(b))X\pi^{-1}_{1}(b)\rightarrow\pi^{-1}(g(b))\subset X is an isometry with respect to the induced path metric on π11(b)\pi^{-1}_{1}(b) and π1(g(b))\pi^{-1}(g(b)) for all bB1b\in B_{1}. ∎

Corollary 3.22.

Suppose (X,B,π)(X,B,\pi) is a metric bundle and g:B1Bg:B_{1}\rightarrow B is a Lipschitz map. Suppose π2:X2B1\pi_{2}:X_{2}\rightarrow B_{1} is an arbitrary metric bundle and (f2:X2X,g)(f_{2}:X_{2}\rightarrow X,g) is a morphism of metric bundles. If X2X_{2} is the pullback of XX under gg and f2:X2Xf_{2}:X_{2}\rightarrow X is the pullback map then for all bB1b\in B_{1} the fiber map (f2)b:π21(b)π1(g(b))(f_{2})_{b}:\pi^{-1}_{2}(b)\rightarrow\pi^{-1}(g(b)) is a uniform quasiisometry with respect to the induced length metrics on the fibers of π2\pi_{2} and π\pi respectively.

Proof.

Suppose X1X_{1} is the pullback of XX under gg as constructed in the proof of the proposition above. Then the fiber maps fb:π11(b)π1(g(b))f_{b}:\pi^{-1}_{1}(b)\rightarrow\pi^{-1}(g(b)) are isometries with respect to the induced metrics on the fibers of π1\pi_{1} and π\pi respectively. On the other hand by Lemma 3.19 there is a coarsely unique metric bundle isomorphism (h,Id)(h,Id) from X2X_{2} to X1X_{1} making the diagram 3 below commutative.

X1X_{1}XXB1B_{1}BBX2X_{2}ffπ1\pi_{1}ggπ\piπ2\pi_{2}f2f_{2}hh
Figure 3.

Now, by Theorem 3.15 the fiber maps hb:π21(b)π11(b)h_{b}:\pi^{-1}_{2}(b)\rightarrow\pi^{-1}_{1}(b) are uniform quasiisometries with respect to the induced length metrics on the fibers of π2\pi_{2} and π1\pi_{1} respectively. Since (f2)b=fbhb(f_{2})_{b}=f_{b}\circ h_{b} for all bB1b\in B_{1} are done by Lemma 2.3(2). ∎

Example 1.

Suppose (X,B,π)(X,B,\pi) is a metric bundle and B1BB_{1}\subset B which is path connected and such that with respect to the path metric induced from BB, B1B_{1} is a length space. Let X1=π1(B1)X_{1}=\pi^{-1}(B_{1}) be endowed with the induced path metric from XX. Let π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} be the restriction of π\pi to X1X_{1}. Let g:B1Bg:B_{1}\rightarrow B and f:X1Xf:X_{1}\rightarrow X be the inclusion maps. It is clear that (X1,B1,π1)(X_{1},B_{1},\pi_{1}) is a metric bundle and also that X1X_{1} is the pullback of gg.

Remark 8.

The notion of morphisms of metric bundles was implicit in the work of Whyte([Why10]). Along the line of [Why10], one can define a more general notion of metric bundles by relaxing the hypothesis of length spaces. In that category of spaces, pullbacks should exist under any coarsely Lipschitz maps. However, we do not delve into it here.

Proposition 3.23.

(Pullbacks for metric graph bundles) Suppose (X,B,π)(X,B,\pi) is an η\eta-metric graph bundle, B1B_{1} is a metric graph and g:𝒱(B1)𝒱(B)g:{\mathcal{V}}(B_{1})\rightarrow{\mathcal{V}}(B) is a coarsely LL-Lipschitz map for some constant L1L\geq 1. Then there is a pullback π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} of gg such that all the fiber maps fb:π11(b)π1(g(b))f_{b}:\pi^{-1}_{1}(b)\rightarrow\pi^{-1}(g(b)), b𝒱(B1)b\in{\mathcal{V}}(B_{1}) are isometries with respect to induced length metrics from X1X_{1} and XX respectively.

Proof.

The proof is a little long. Hence we break this into steps for the sake of clarity.

Step 1. Construction of X1X_{1} and π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} and f:𝒱(X1)𝒱(X)f:{\mathcal{V}}(X_{1})\rightarrow{\mathcal{V}}(X). We first construct a metric graph X1X_{1}, a candidate for the total space of the bundle. The vertex set of X1X_{1} is the disjoint union of the vertex sets of π1(g(b))\pi^{-1}(g(b)), b𝒱(B1)b\in{\mathcal{V}}(B_{1}). There are two types of edges. First of all for all b𝒱(B1)b\in{\mathcal{V}}(B_{1}), we take all the edges appearing in π1(g(b))\pi^{-1}(g(b)). In other words, the full subgraph π1(g(b))\pi^{-1}(g(b)) is contained in X1X_{1}. Let us denote that by FbF_{b}. For all adjacent vertices s,tB1s,t\in B_{1} we introduce some other edges with one end point in FsF_{s} and the other in FtF_{t}. We note that Fs,FtX1F_{s},F_{t}\subset X_{1} are identical copies of Fg(s)F_{g(s)} and Fg(t)F_{g(t)} respectively. Let fs:FsFg(s)f_{s}:F_{s}\rightarrow F_{g(s)} denote this identification. Let ee be an edge joining s,ts,t and let α\alpha be a geodesic in BB joining g(s),g(t)g(s),g(t). Now for each xFsx\in F_{s} we lift the path α\alpha starting from fs(x)f_{s}(x) isometrically by Lemma 3.8(1) to say α~\tilde{\alpha}. For each such lift we join xx by an edge to yV(Ft)y\in V(F_{t}) if and only if ft(y)=α~(g(t))f_{t}(y)=\tilde{\alpha}(g(t)). This completes the construction of X1X_{1}. We note that dB(g(s),g(t))2Ld_{B}(g(s),g(t))\leq 2L and hence l(α~)2Ll(\tilde{\alpha})\leq 2L too. Now we define f:𝒱(X1)𝒱(X)f:{\mathcal{V}}(X_{1})\rightarrow{\mathcal{V}}(X) by setting f(x)=fπ1(x)(x)f(x)=f_{\pi_{1}(x)}(x) for all x𝒱(X1)x\in{\mathcal{V}}(X_{1}). It is clear that this map is 2L2L-Lipschitz.

Step 2. π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} is a metric graph bundle and (f,g)(f,g) is a morphism. We need to verify that the fibers are uniformly properly embedded in X1X_{1} so that X1X_{1} is a metric graph bundle. Suppose x,yFsx,y\in F_{s} and dX1(x,y)Dd_{X_{1}}(x,y)\leq D. Let α\alpha be a (dotted) geodesic in X1X_{1} joining x,yx,y. Then fαf\circ\alpha is a (dotted) path of length at most 2LD2LD. Thus dX(f(x),f(y))2LDd_{X}(f(x),f(y))\leq 2LD. Since XX is an η\eta-metric graph bundle dg(s)(f(x),f(y))η(2LD)d_{g(s)}(f(x),f(y))\leq\eta(2LD). Since ff is an isometry when restricted to FsF_{s} we have ds(x,y)η(2LD)d_{s}(x,y)\leq\eta(2LD). This proves that X1X_{1} is a metric graph bundle over B1B_{1}.

On the other hand, ff is 2L2L-Lipschitz by step 1 and gg is coarsely LL-Lipschitz by hypothesis. It is also clear that πf=gπ1\pi\circ f=g\circ\pi_{1} by the definition of ff. Thus (f,g)(f,g) is a morphism of metric graph bundles from X1X_{1} to XX.

𝒱(X1){\mathcal{V}}(X_{1})𝒱(X){\mathcal{V}}(X)𝒱(B1){\mathcal{V}}(B_{1})𝒱(B){\mathcal{V}}(B)𝒱(Y){\mathcal{V}}(Y)ffπ1\pi_{1}ggπ\piπ2\pi_{2}fYf^{Y}ff^{\prime}
Figure 4.

Step 3. X1X_{1} is a pullback. Now we check that X1X_{1} is a pullback of XX under gg. Suppose π2:YB1\pi_{2}:Y\rightarrow B_{1} is a metric graph bundle and (fY,g)(f^{Y},g) is a morphism of metric graph bundles from YY to XX where fYf^{Y} is coarsely L1L_{1}-Lipschitz We need to find a coarsely unique, coarsely Lipschitz map f:𝒱(Y)𝒱(X1)f^{\prime}:{\mathcal{V}}(Y)\rightarrow{\mathcal{V}}(X_{1}) such that (f,Id)(f^{\prime},Id) is a morphism from YY to X1X_{1} and the whole diagram 4 is commutative where Id:𝒱(B1)𝒱(B1)Id:{\mathcal{V}}(B_{1})\rightarrow{\mathcal{V}}(B_{1}) is the identity map.

The map ff^{\prime}: For all s𝒱(B1)s\in{\mathcal{V}}(B_{1}) we define ff^{\prime} on 𝒱(π21(s)){\mathcal{V}}(\pi^{-1}_{2}(s)) as the composition fs1fsYf^{-1}_{s}\circ f^{Y}_{s}. Collectively these maps define ff^{\prime}. It is clear that ff^{\prime} makes the whole diagram above commutative.

The rest of the argument follows from Lemma 3.20. In fact, condition (2) of that lemma follows from Lemma 3.10(1) since π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} is a metric graph bundle and (3) follows because fibers of metric graph bundles are uniformly properly embedded and in this case the restriction of ff, π11(b)π1(g(b))X\pi^{-1}_{1}(b)\rightarrow\pi^{-1}(g(b))\subset X is an isometry with respect to the induced path metric on π11(b)\pi^{-1}_{1}(b) and π1(g(b))\pi^{-1}(g(b)) for all b𝒱(B1)b\in{\mathcal{V}}(B_{1}). ∎

The corollary below follows immediately from the proof of the above proposition.

Corollary 3.24.

Suppose π:XB\pi:X\rightarrow B is a metric graph bundle. Suppose AA is a connected subgraph of BB. Let g:ABg:A\rightarrow B denote the inclusion map. Let XA=π1(A)X_{A}=\pi^{-1}(A), πA\pi_{A} be the restriction of π\pi and let f:XAXf:X_{A}\rightarrow X denote the inclusion map. Then XAX_{A} is the pullback of XX under gg.

The proof of the following corollary is similar to that of Corollary 3.22 and hence we omit the proof.

Corollary 3.25.

Suppose (X,B,π)(X,B,\pi) is a metric graph bundle and g:𝒱(B1)𝒱(B)g:{\mathcal{V}}(B_{1})\rightarrow{\mathcal{V}}(B) is a coarsely Lipschitz map. Suppose π2:X2B1\pi_{2}:X_{2}\rightarrow B_{1} is an arbitrary metric graph bundle and (f2:𝒱(X2)𝒱(X),g)(f_{2}:{\mathcal{V}}(X_{2})\rightarrow{\mathcal{V}}(X),g) is a morphism of metric graph bundles. If X2X_{2} is the pullback of XX under gg and f2:𝒱(X2)𝒱(X)f_{2}:{\mathcal{V}}(X_{2})\rightarrow{\mathcal{V}}(X) is the pullback map then for all b𝒱(B1)b\in{\mathcal{V}}(B_{1}) the fiber map (f2)b:𝒱(π21(b))𝒱(π1(g(b)))(f_{2})_{b}:{\mathcal{V}}(\pi^{-1}_{2}(b))\rightarrow{\mathcal{V}}(\pi^{-1}(g(b))) is a uniform quasiisometry with respect to the induced length metrics on the fibers of π2\pi_{2} and π\pi respectively.

3.3. Some examples

In this section we discuss in detail two main sources of examples for metric graph bundles to which the main theorem of this paper will be applied.

3.3.1. Short exact sequence of groups

Example 2.

Given a short exact sequence of finitely generated groups

1NGπQ11\rightarrow N\rightarrow G\stackrel{{\scriptstyle\pi}}{{\rightarrow}}Q\rightarrow 1

we have a naturally associated metric graph bundle. This is the main motivating example of metric graph bundles. We recall the definition from [MS12, Example 1.8] with a minor modification.

Suppose H<QH<Q is a finitely generated subgroup. Let G1=π1(H)G_{1}=\pi^{-1}(H). We fix a generating set SNS_{N} of NN, a generating set SSNS\supseteq S_{N} of GG such that SS contains a generating set S1S_{1} of G1G_{1}, SNS1S_{N}\subset S_{1} and NS=SNN\cap S=S_{N}. Let SQ=π(S){1}S_{Q}=\pi(S)\setminus\{1\} and SH=π(S1){1}S_{H}=\pi(S_{1})\setminus\{1\}. Then we have a metric graph bundle π:Γ(G,S)Γ(Q,SQ)\pi:\Gamma(G,S)\rightarrow\Gamma(Q,S_{Q}). Clearly Γ(H,SH)\Gamma(H,S_{H}) is a subgraph of Γ(Q,SQ)\Gamma(Q,S_{Q}) and Γ(G1,S1)=π1(Γ(H,SH))\Gamma(G_{1},S_{1})=\pi^{-1}(\Gamma(H,S_{H})). Hence, by Corollary 3.24 it follows that Γ(G1,S1)\Gamma(G_{1},S_{1}) is the pullback of Γ(G,S)\Gamma(G,S) under the inclusion Γ(H,SH)Γ(Q,SQ)\Gamma(H,S_{H})\hookrightarrow\Gamma(Q,S_{Q}).

3.3.2. Complexes of groups

For this example, we refer to [Hae92]. Suppose 𝒴\mathcal{Y} is a finite simplicial complex and G(𝒴)\mbox{{\bf G}}(\mathcal{Y}) is a developable complex of groups defined over 𝒴\mathcal{Y}. (See [Hae92, Definition 2.2].) For any face σ\sigma of 𝒴\mathcal{Y}, let KσK_{\sigma} be a K(Gσ,1)K(G_{\sigma},1)-space. Then by [Hae92, Theorem 3.4.1] there is a complex of spaces p:𝒳𝒴p:\mathcal{X}\rightarrow\mathcal{Y} (compare with good complexes of spaces due to Corson [Cor92]) which is a cellular aspherical realization (see [Hae92, Definition 3.3.4]) of the complex of groups G(𝒴)\mbox{{\bf G}}(\mathcal{Y}) such that inverse image under pp of the barycenter of each face σ\sigma is KσK_{\sigma}. It follows from the construction of 𝒳\mathcal{X} that there is a continuous section ss of p:𝒳𝒴p:\mathcal{X}\rightarrow\mathcal{Y} over the 11-skeleton 𝒴(1)\mathcal{Y}^{(1)} of 𝒴\mathcal{Y}. We fix a maximal tree of s(𝒴(1))s(\mathcal{Y}^{(1)}) and a base vertex v0𝒴(0)v_{0}\in\mathcal{Y}^{(0)} in it. Let G=π1(𝒳,s(v0))G=\pi_{1}(\mathcal{X},s(v_{0})). Thus for any v𝒴(0)v\in\mathcal{Y}^{(0)} we have a natural injective homomorphism π1(Xv,s(v))G\pi_{1}(X_{v},s(v))\rightarrow G. We identify the image of the same with GvG_{v}. Next following Corson [Cor92] we take the universal cover π𝒳:𝒳~𝒳\pi_{{\mathcal{X}}}:\tilde{\mathcal{X}}\rightarrow\mathcal{X}. We put a CW complex structure on 𝒳~\tilde{\mathcal{X}} in the standard way so that π𝒳\pi_{{\mathcal{X}}} is a cellular map. Then for all y𝒴y\in\mathcal{Y}, we collapse each connected component of (pπ𝒳)1(y)(p\circ\pi_{{\mathcal{X}}})^{-1}(y) to a point. Suppose {\mathcal{B}} is the quotient complex thus obtained and let q:𝒳~q:\tilde{{\mathcal{X}}}\rightarrow{\mathcal{B}} be the quotient map. Then we note that there is a cellular map π¯𝒳:𝒴\bar{\pi}_{{\mathcal{X}}}:{\mathcal{B}}\rightarrow{\mathcal{Y}} making the following diagram commutative.

𝒳~\tilde{{\mathcal{X}}}{\mathcal{B}}𝒳{\mathcal{X}}𝒴{\mathcal{Y}}qqπ𝒳\pi_{{\mathcal{X}}}π¯𝒳\bar{\pi}_{{\mathcal{X}}}pp
Figure 5.

Now for our purpose, we shall also assume that all the face groups GσG_{\sigma} are finitely generated, the 0-skeleton of each KσK_{\sigma} is a point xσx_{\sigma}, the 11-skeleton is a wedge of finitely many circles and the developable complex of groups satisfies the qi condition as defined below.

Definition 3.26.

Suppose we have a developable complex of groups (𝒢,𝒴)({\mathcal{G}},{\mathcal{Y}}).

(1) We say that it satisfies the qi condition if for any faces στ\sigma\subset\tau of 𝒴{\mathcal{Y}} the corresponding homomorphism GτGσG_{\tau}\rightarrow G_{\sigma} is an isomorphism onto a finite index subgroup of GσG_{\sigma}.

(2) If all the face groups of GσG_{\sigma} satisfies a group theoretic property 𝒫\mathcal{P} then we shall say that (𝒢,𝒴)({\mathcal{G}},{\mathcal{Y}}) is a developable complex of groups with property 𝒫\mathcal{P}.

For instance, we shall work in section 6 with the developable complexes of nonelementary hyperbolic groups.

However, we now aim to associate to the complex of groups a metric graph bundle as follows. Let X=(pπ𝒳)1(𝒴(1))(1)X^{\prime}=(p\circ\pi_{{\mathcal{X}}})^{-1}({\mathcal{Y}}^{(1)})^{(1)} and B=π¯𝒳1(𝒴(1))(1)B=\bar{\pi}_{{\mathcal{X}}}^{-1}({\mathcal{Y}}^{(1)})^{(1)} where we denote by 𝒵(1)\mathcal{Z}^{(1)} the 11-skeleton of any CW complex 𝒵\mathcal{Z}. Now we construct a metric graph bundle π:XB\pi:X\rightarrow B as follows. For all v(0)v\in{\mathcal{B}}^{(0)} let Fv:=q1(v)(1)F_{v}:=q^{-1}(v)^{(1)}. Suppose v,w(0)v,w\in{\mathcal{B}}^{(0)} are connected by an edge ee. We look at the subcomplex 𝒳~[v,w]=q1([v,w])\tilde{{\mathcal{X}}}_{[v,w]}=q^{-1}([v,w]). Let v0=π¯𝒳(v),w0=π¯𝒳(w)v_{0}=\bar{\pi}_{{\mathcal{X}}}(v),w_{0}=\bar{\pi}_{{\mathcal{X}}}(w) and e0=π¯𝒳(e)e_{0}=\bar{\pi}_{{\mathcal{X}}}(e). Then 𝒳~[v,w]π𝒳1(p1([v0,w0])\tilde{{\mathcal{X}}}_{[v,w]}\subset\pi^{-1}_{{\mathcal{X}}}(p^{-1}([v_{0},w_{0}]). However, we recall from Haefliger [Hae92] how p1([v0,w0])𝒳p^{-1}([v_{0},w_{0}])\subset{\mathcal{X}} is built from the spaces the Kv0,Kw0K_{v_{0}},K_{w_{0}} and Ke0K_{e_{0}}. There are injective homomorphisms Ge0Gv0,Ge0Gw0G_{e_{0}}\rightarrow G_{v_{0}},G_{e_{0}}\rightarrow G_{w_{0}}. We choose cellular maps f0:Ke0Kv0,f1:Ke0Kw0f_{0}:K_{e_{0}}\rightarrow K_{v_{0}},f_{1}:K_{e_{0}}\rightarrow K_{w_{0}} such that the induced maps in the fundamental groups are those group homomorphisms. Then one glues Ke0×[0,1]K_{e_{0}}\times[0,1] to Kv0Kw0K_{v_{0}}\bigsqcup K_{w_{0}} by gluing Ke0×{0}K_{e_{0}}\times\{0\} to Kv0K_{v_{0}} and Ke0×{1}K_{e_{0}}\times\{1\} to Kw0K_{w_{0}} using the maps f0,f1f_{0},f_{1} respectively. Let m0m_{0} be the midpoint of xe0×[0,1]p1([v0,w0])x_{e_{0}}\times[0,1]\subset p^{-1}([v_{0},w_{0}]) and let mem\in e be the midpoint of ee. Then through any aq1(m)(0)a\in q^{-1}(m)^{(0)} we lift xe0×[0,1]x_{e_{0}}\times[0,1]. The lift is a 11-cell joining avq1(v)(0)a_{v}\in q^{-1}(v)^{(0)} to awq1(w)(0)a_{w}\in q^{-1}(w)^{(0)}. Let us denote the map aava\mapsto a_{v} by fe,vf_{e,v} and the map aawa\mapsto a_{w} by fe,wf_{e,w}

Lemma 3.27.

(1) The map fe,v:q1(m)(0)q1(v)(0)f_{e,v}:q^{-1}(m)^{(0)}\rightarrow q^{-1}(v)^{(0)} is uniformly coarsely surjective with respect to the graph metric on q1(m)(0),q1(v)(0)q^{-1}(m)^{(0)},q^{-1}(v)^{(0)} coming from q1(m)(1),q1(v)(1)q^{-1}(m)^{(1)},q^{-1}(v)^{(1)} respectively.

(2) Similar statement holds for fe,wf_{e,w}.

Proof.

We will only prove (1) as the proof of (2) is similar. The group Gv<GG_{v}<G is isomorphic to Gv0G_{v_{0}} and q1(v)q^{-1}(v) is a universal cover of Kv0K_{v_{0}} since the complex of groups is developable. The groups GvG_{v} acts properly discontinuously with quotient Kv0K_{v_{0}}. Since the action is cellular the action of GvG_{v} on q1(v)(1)q^{-1}(v)^{(1)} is simply transitive. Similarly the action of GmG_{m} is simply transitive on q1(m)(1)q^{-1}(m)^{(1)}. We note that Gm<GvG_{m}<G_{v} and the map fe,vf_{e,v} is equivariant. It is also clear that [Gv:Gm]=[Gv0:Ge0][G_{v}:G_{m}]=[G_{v_{0}}:G_{e_{0}}]. Finally we note that q1(v)(1)q^{-1}(v)^{(1)} is naturally isometric to a Cayley graph of GvG_{v} when q1(v)(1)q^{-1}(v)^{(1)} is given graph metric where each edge has length 11. The lemma is immediate from this. ∎

Let R>0R>0 be such that fe,vf_{e,v} is coarsely RR-surjective for all 0-cell vv and 11-cell ee of {\mathcal{B}} where ee is incident on vv. Then we construct a graph XX from XX^{\prime} by introducing new edges as follows. Given v,w(0)v,w\in{\mathcal{B}}^{(0)} connected by an edge ee we join all xq1(v)x\in q^{-1}(v) to yq1(w)y\in q^{-1}(w) by an edge if there is aq1(me)(0)a\in q^{-1}(m_{e})^{(0)} such that d(x,fe,v(a))Rd(x,f_{e,v}(a))\leq R, d(y,fe,w(a))Rd(y,f_{e,w}(a))\leq R where the distances are taken in the respective 11-skeletons of q1(v)q^{-1}(v) and q1(w)q^{-1}(w).

Proposition 3.28.

Suppose we identify GG as the group of deck transformation on the covering map π𝒳:𝒳~𝒳\pi_{{\mathcal{X}}}:\tilde{{\mathcal{X}}}\rightarrow{\mathcal{X}}. Then we have the following:

(1) GG acts on XX and on BB through simplicial maps. The map qq is GG-equivariant.

(2) The GG-action is proper and cofinite on XX but it is only cofinite on BB. Also B/GB/G is isomorphic to 𝒴(1){\mathcal{Y}}^{(1)}.

(3) For all v𝒴(0)v\in{\mathcal{Y}}^{(0)} and v~π¯𝒳1(v)\tilde{v}\in\bar{\pi}_{{\mathcal{X}}}^{-1}(v), Gv~G_{\tilde{v}} is a conjugate of GvG_{v} in GG.

(4) The action of Gv~G_{\tilde{v}} on Xv~=q1(v~)X_{\tilde{v}}=q^{-1}(\tilde{v}) is proper and cocompact. In fact the action on V(Xv~)V(X_{\tilde{v}}) is transitive and on E(Xv~)E(X_{\tilde{v}}) is cofinite. In particular if the GvG_{v} is hyperbolic for all v𝒴(0)v\in\mathcal{Y}^{(0)} then for all v𝒴(0)v\in\mathcal{Y}^{(0)} and v~π¯𝒳1(v)\tilde{v}\in\bar{\pi}_{{\mathcal{X}}}^{-1}(v), Xv~X_{\tilde{v}} is uniformly hyperbolic.

(5) π:XB\pi:X\rightarrow B is a metric graph bundle.

Proof.

The group GG acts through deck transformations of the covering map π𝒳:𝒳~𝒳\pi_{{\mathcal{X}}}:\tilde{{\mathcal{X}}}\rightarrow{\mathcal{X}}. Hence it follows that GG permutes the connected components of (pπ𝒳)1(y)(p\circ\pi_{{\mathcal{X}}})^{-1}(y) for all y𝒴y\in{\mathcal{Y}}. The action is also simplicial. Hence, (1) follows from this. For (2) we note that the action of GG on XX^{\prime} is proper and cofinite. On the other hand, the inclusion map XXX^{\prime}\rightarrow X is a GG-equivariant quasiisometry by Lemma 2.4. Hence the GG-action on XX is proper and cofinite. Clearly, B/GB/G is isomorphic to 𝒴(1){\mathcal{Y}}^{(1)} whence the GG-action on BB is cofinite. (3) is a consequence of a basic covering space argument using the GG-equivariance of the map qq. In (4) the properness follows from the properness of the action of GG on XX^{\prime}. Cocompactness is due to the fact that X/GX^{\prime}/G is finite. The second part also follows from the nature of K(Gv,1)K(G_{v},1) used to construct 𝒳{\mathcal{X}}, where v=π¯𝒳(v~)v=\bar{\pi}_{{\mathcal{X}}}(\tilde{v}). The last part follows from the second by Milnor-Scwarz lemma. What remains is to prove (5). For all v~V(B)\tilde{v}\in V(B), let Xv~=π1(v~)X_{\tilde{v}}=\pi^{-1}(\tilde{v}). Since B/GB/G is finite and the map qq is GG-equivariant the Xv~X_{\tilde{v}}’s are uniformly properly embedded in XX^{\prime} iff for all wV(B/G)w\in V(B/G) there is one w~π¯𝒳1(w)\tilde{w}\in\bar{\pi}_{{\mathcal{X}}}^{-1}(w) such that Xw~X_{\tilde{w}} is uniformly properly embedded in XX^{\prime}. However, each inclusion Xv~XX_{\tilde{v}}\rightarrow X^{\prime} is Gv~G_{\tilde{v}}-equivariant, the Gv~G_{\tilde{v}} action on Xv~X_{\tilde{v}} is proper and cocompact and Gv~G_{\tilde{v}} is a finitely generated subgroup of GG. Since each finitely generated subgroup of a finitely generated group is uniformly properly embedded it follows that Xv~X_{\tilde{v}} is properly embedded in XX^{\prime}. Since XX^{\prime} is quasiisometric to XX, it follows that Xv~X_{\tilde{v}}’s are properly embedded in XX. This verifies property (1) of metric graph bundles. Property (2) follows from Lemma 3.27 and the construction of the new edges. ∎

Subcomplexes of groups

In the above set-up we now assume further that we have a connected subcomplex 𝒴1𝒴{\mathcal{Y}}_{1}\subset{\mathcal{Y}}. Let 𝒳1=p1(𝒴1){\mathcal{X}}_{1}=p^{-1}({\mathcal{Y}}_{1}). We shall assume that the base point x0𝒳x_{0}\in{\mathcal{X}} is contained in 𝒳1{\mathcal{X}}_{1} and a maximal tree of s(𝒴1(1))s({\mathcal{Y}}^{(1)}_{1}) is chosen so that it is contained in the chosen maximal tree of s(𝒴(1))s({\mathcal{Y}}^{(1)}). Suppose the inclusion 𝒳1𝒳{\mathcal{X}}_{1}\rightarrow{\mathcal{X}} is π1\pi_{1}-injective. Then the restriction G(𝒴1)\mbox{{\bf G}}({\mathcal{Y}}_{1}) of G(𝒴)\mbox{{\bf G}}({\mathcal{Y}}) to 𝒴1{\mathcal{Y}}_{1} is a developable complex of groups by [BH99, Corollary 2.15]. Let G1=π1(𝒳1,x0)G_{1}=\pi_{1}({\mathcal{X}}_{1},x_{0}). However, 𝒳1𝒴1{\mathcal{X}}_{1}\rightarrow{\mathcal{Y}}_{1} is a complex of spaces which is a cellular aspherical realization of the complex of groups G(𝒴1)\mbox{{\bf G}}(\mathcal{Y}_{1}). Hence, we can build a metric graph bundle π1:X1B1\pi_{1}:X_{1}\rightarrow B_{1} as described in Proposition 3.28.

In fact fixing a point x~0π𝒳1(x0)\tilde{x}_{0}\in\pi^{-1}_{{\mathcal{X}}}(x_{0}) we may identify GG as the group of deck transformations on 𝒳~\tilde{{\mathcal{X}}}. Then G1G_{1} stabilizes the connected component of π1(𝒳1)\pi^{-1}({\mathcal{X}}_{1}) containing x~0\tilde{x}_{0}. Since 𝒳1𝒳{\mathcal{X}}_{1}\rightarrow{\mathcal{X}} is π1\pi_{1}-injective this connected component, say 𝒳~1\tilde{{\mathcal{X}}}_{1}, is a universal cover of 𝒳1{\mathcal{X}}_{1}. We set B1=q(𝒳~1)BB_{1}=q(\tilde{{\mathcal{X}}}_{1})\cap B and X1=π1(B1)X_{1}=\pi^{-1}(B_{1}). The following proposition records these in a nutshell.

Proposition 3.29.

Suppose 𝒴{\mathcal{Y}} is a finite connected simplicial complex and G(𝒴)\mbox{{\bf G}}({\mathcal{Y}}) is a developable complex of groups with qi condition and with fundamental group GG and suppose 𝒴1{\mathcal{Y}}_{1} is a connected subcomplex of 𝒴{\mathcal{Y}}. Suppose G1G_{1} is the fundamental group of G(𝒴1)\mbox{{\bf G}}({\mathcal{Y}}_{1}). Suppose the inclusion G(𝒴1)G(𝒴)\mbox{{\bf G}}({\mathcal{Y}}_{1})\rightarrow\mbox{{\bf G}}({\mathcal{Y}}) induces injective homomorphism G1GG_{1}\rightarrow G.

Then there is a metric graph bundle π:XB\pi:X\rightarrow B, a connected subgraph B1BB_{1}\subset B such that the following hold:

(1) GG acts on XX and on BB through simplicial maps. The map π\pi is GG-equivariant. The action is proper and cofinite on XX but it is only cofinite on BB. Also, there is a simplicial GG-equivariant map B𝒴(1)B\rightarrow{\mathcal{Y}}^{(1)} with trivial action on 𝒴(1){\mathcal{Y}}^{(1)} inducing an isomorphism of graphs B/G𝒴(1)B/G\rightarrow{\mathcal{Y}}^{(1)}. The group Gb<GG_{b}<G is a conjugate of Gb¯G_{\bar{b}} in GG where b¯\bar{b} is the image of bb under the map B𝒴(1)B\rightarrow{\mathcal{Y}}^{(1)}. Also the GbG_{b}-action on FbF_{b} is proper and cofinite for all bV(B)b\in V(B).

(2) Let X1=π1(B1)X_{1}=\pi^{-1}(B_{1}). Then G1G_{1} stabilizes X1X_{1} and the G1G_{1}-action on X1X_{1} is proper and cofinite. Also the restriction of the map B/G𝒴(1)B/G\rightarrow{\mathcal{Y}}^{(1)} to B1/G1B_{1}/G_{1} is an isomorphism of graphs B1/G1𝒴1(1)B_{1}/G_{1}\rightarrow{\mathcal{Y}}^{(1)}_{1}.

Later on we shall work with rather special subcomplexes of groups as defined below.

Definition 3.30.

Suppose 𝒴{\mathcal{Y}} is a finite connected simplicial complex and (𝒢,𝒴)({\mathcal{G}},{\mathcal{Y}}) is a developable complexes of groups with qi condition over 𝒴{\mathcal{Y}}. We shall call a connected subcomplex 𝒴1𝒴{\mathcal{Y}}_{1}\subset{\mathcal{Y}} a good subcomplex if the following hold:

(1) The induced natural homomorphism π1(𝒢,𝒴1)π1(𝒢,𝒴)\pi_{1}({\mathcal{G}},{\mathcal{Y}}_{1})\rightarrow\pi_{1}({\mathcal{G}},{\mathcal{Y}}) is injective. Suppose the image is G1G_{1}.

(2) If π:XB\pi:X\rightarrow B is a metric graph bundle obtained as in Proposition 3.28 from (𝒢,𝒴)({\mathcal{G}},{\mathcal{Y}}) and B1BB_{1}\subset B is as in Proposition 3.29. Then the inclusion B1BB_{1}\subset B is a qi embedding.

We note that XX is quasiisometric to GG and X1X_{1} is quasiisometric to G1G_{1}. Thus it follows that BB is quasiisometric to the ‘coned-off’ space a la Farb([Far98]) obtained from GG by coning off the cosets of the various face groups of (𝒢,𝒴)({\mathcal{G}},{\mathcal{Y}}). Similarly B1B_{1} is obtained by coning off various cosets of the face groups of (𝒢,𝒴1)({\mathcal{G}},{\mathcal{Y}}_{1}). Thus condition (2) of the above definition is intrinsic and independent of the particular metric graph bundle obtained from (𝒢,𝒴)({\mathcal{G}},{\mathcal{Y}}).

4. Geometry of metric bundles

In this section, we recall some results from [MS12] and also add a few of our own which are going to be useful for the proof of our main theorem in the next section. Especially some of the results which were stated for geodesic metric spaces in [MS12] but whose proofs require little adjustments to hold true for length spaces are mentioned here.

4.1. Metric graph bundles arising from metric bundles

An analogue of the following result is proved in [MS12](see Lemma 1.17 through Lemma 1.21 in [MS12]). We give an independent and relatively simpler proof here. We also construct an approximating metric graph bundle morphism starting with a given metric bundle morphism. However, one disadvantage of our construction is that the metric graphs so obtained are never proper.

Proposition 4.1.

Suppose π:XB\pi^{\prime}:X^{\prime}\rightarrow B^{\prime} is an (η,c)(\eta,c)-metric bundle. Then there is a metric graph bundle π:XB\pi:X\rightarrow B along with quasiisometries ψB:BB\psi_{B}:B^{\prime}\rightarrow B and ψX:XX\psi_{X}:X^{\prime}\rightarrow X such that (1) πψX=ψBπ\pi\circ\psi_{X}=\psi_{B}\circ\pi^{\prime} and (2) for all bBb\in B^{\prime} the map ψX\psi_{X} restricted to π1(b)\pi^{\prime-1}(b) is a (1,1)(1,1)-quasiisometry onto π1(ψB(b))\pi^{-1}(\psi_{B}(b)).

Moreover, the maps ψX,ψB\psi_{X},\psi_{B} have coarse inverses ϕX\phi_{X}, ϕB\phi_{B} respectively making the following diagram commutative:

XX^{\prime}XXBB^{\prime}BBψX\psi_{X}ϕX\phi_{X}π\pi^{\prime}π\piψB\psi_{B}ϕB\phi_{B}
Figure 6.
Proof.

(1) For the proof we use the construction of Lemma 2.8. We shall briefly recall the construction of the spaces. We define 𝒱(B)=B{\mathcal{V}}(B)=B^{\prime} and s,t𝒱(B)s,t\in{\mathcal{V}}(B) are connected by an edge if and only if sts\neq t and dB(s,t)1d_{B^{\prime}}(s,t)\leq 1. This defines the graph. We also have a natural map ψB:BB\psi_{B}:B^{\prime}\rightarrow B which is just the inclusion map when BB^{\prime} is identified with the vertex set of BB. To define XX, we take 𝒱(X)=X{\mathcal{V}}(X)=X^{\prime}. Edges are of two types.

Type 1 edges: For all sBs\in B^{\prime}, x,yπ1(s)x,y\in\pi^{\prime-1}(s) are connected by an edge if and only if ds(x,y)1d_{s}(x,y)\leq 1.

Type 2 edges: If stBs\neq t\in B^{\prime}, xπ1(s)x\in\pi^{\prime-1}(s) and yπ1(t)y\in\pi^{\prime-1}(t) then x,yx,y are connected by an edge if and only if dB(s,t)1d_{B^{\prime}}(s,t)\leq 1 and dX(x,y)cd_{X^{\prime}}(x,y)\leq c.

The map ψX:XX\psi_{X}:X^{\prime}\rightarrow X is defined as before to be the inclusion map. By Lemma 2.8 ψB\psi_{B} is a qi. We also note that πψX=ψBπ\pi\circ\psi_{X}=\psi_{B}\circ\pi^{\prime}. We need to verify that ψX\psi_{X} is a qi. For that, it is enough to produce Lipschitz coarse inverses ϕX\phi_{X}, ϕB\phi_{B} as claimed in the second part of the proposition and then apply Lemma 2.2 since it is clear that ψX\psi_{X} is 11-Lipschitz. We first choose a coarse inverse ϕB\phi_{B} of ψB\psi_{B} as follows. On 𝒱(B){\mathcal{V}}(B) it is simply the identity map. The interior of each edge is then sent to one of its end points. The map ϕX\phi_{X} on 𝒱(X){\mathcal{V}}(X) is also defined as the identity map. The interior of a type 1 edge is sent to one of its end points. Then interior of each type 2 edge e=[x,y]e=[x,y] is sent to one of the end points xx or yy according as the edge π(e)\pi(e) is mapped by ϕB\phi_{B} to π(x)\pi(x) or π(y)\pi(y) respectively. It follows that the diagram in Figure 6 commutes. We just need to check that ϕX\phi_{X} is coarsely Lipschitz, since ϕB,ϕX\phi_{B},\phi_{X} are inverses of ψB,ψX\psi_{B},\psi_{X} respectively on a 11-dense subset, they will be coarse inverse automatically. However, by Lemma 2.6 it is enough to show that edges are mapped to small diameter sets. This is again clear. In fact, the image of an edge has diameter at most cc. This proves the first part of the proposition.

(2) This is immediate from the definition of ψX\psi_{X} and the construction in Lemma 2.8.

(3) Finally, we need to check that (X,B,π)(X,B,\pi) is a metric graph bundle. Let sBs\in B and x,yπ1(s)x,y\in\pi^{-1}(s) such that dX(x,y)Md_{X}(x,y)\leq M for some M>0M>0. Since ϕX\phi_{X} is a quasiisometry, dX(x,y)Md_{X^{\prime}}(x,y)\leq M^{\prime}, where M>0M^{\prime}>0 depends on MM and ϕX\phi_{X}. Since π1(ϕB(s))\pi^{\prime-1}(\phi_{B}(s)) is properly embedded in XX^{\prime} as measured by η\eta, we have dϕB(s)(x,y)η(M)d_{\phi_{B}(s)}(x,y)\leq\eta(M^{\prime}). Now, using the above fact that π1(ϕB(s))\pi^{\prime-1}(\phi_{B}(s)) is (1,1)(1,1)-quasiisometric to π1(s)\pi^{-1}(s), we have ds(x,y)η(M)+1d_{s}(x,y)\leq\eta(M^{\prime})+1. Hence, π1(s)\pi^{-1}(s) is uniformly properly embedded in XX. Next we check the condition (2)(2) of Definition 3.4. Suppose s,t𝒱(B)s,t\in{\mathcal{V}}(B) are adjacent vertices. Then, dB(s,t)1d_{B^{\prime}}(s,t)\leq 1. Let α\alpha be a path in BB^{\prime} joining s,ts,t with lB(α)1l_{B^{\prime}}(\alpha)\leq 1. Then, for any xπ1(s)x\in\pi^{\prime-1}(s), α\alpha can be lifted to a path of length at most cc, joining xx to some yπ1(t)y\in\pi^{\prime-1}(t). Then there exists an edge joining xx and yy in XX, which is a lift of the edge joining ss and tt in BB.∎

Remark 9.

We shall refer to the metric graph bundle XX obtained from XX^{\prime} as the canonical metric graph bundle associated to the bundle XX. Since we are working with length metric spaces some of the machinery of [MS12] may not apply directly. The above proposition then comes to the rescue. We sometimes modify our definitions suitably to make things work. Consequently, all the results proved for metric graph bundles have their close analogs in metric bundles. We shall make this precise for instance in Proposition 4.3 and Definition 4.5.

Approximating a metric bundle morphism

Suppose π:XB\pi^{\prime}:X^{\prime}\rightarrow B^{\prime} is a metric bundle and g:ABg:A^{\prime}\rightarrow B^{\prime} is a Lipschitz map. Suppose YY^{\prime} is the pullback of the bundle under the map gg as constructed in the proof of Proposition 3.21, i.e. YY^{\prime} is also the set theoretic pullback. Let gπ:YAg^{*}\pi^{\prime}:Y^{\prime}\rightarrow A^{\prime} be the corresponding bundle projection map and f:YXf:Y^{\prime}\rightarrow X^{\prime} be the pullback map. Suppose we use the recipe of the above proposition to construct metric graph bundles πX:XB\pi_{X}:X\rightarrow B, πY:YA\pi_{Y}:Y\rightarrow A with quasiisometries ψA:AA\psi_{A}:A^{\prime}\rightarrow A, ψB:BB\psi_{B}:B^{\prime}\rightarrow B, ψY:YY\psi_{Y}:Y^{\prime}\rightarrow Y and ψX:XX\psi_{X}:X^{\prime}\rightarrow X such that πYψY=ψAgπ\pi_{Y}\circ\psi_{Y}=\psi_{A}\circ g^{*}\pi^{\prime} and πXψX=ψBπ\pi_{X}\circ\psi_{X}=\psi_{B}\circ\pi^{\prime}.

Suppose ϕX,ϕB,ϕY,ϕA\phi_{X},\phi_{B},\phi_{Y},\phi_{A} are the coarse inverses (as constructed in the proposition above) of ψX\psi_{X}, ψB\psi_{B}, ψY\psi_{Y}, and ψA\psi_{A} respectively. We then have a commutative diagram:

YYYY^{\prime}XX^{\prime}XXAAAA^{\prime}BB^{\prime}BBψY\psi_{Y}ϕY\phi_{Y}ffψX\psi_{X}ϕX\phi_{X}πY\pi_{Y}gπg^{*}\pi^{\prime}π\pi^{\prime}πX\pi_{X}ψY\psi_{Y}ψA\psi_{A}ϕA\phi_{A}ggψB\psi_{B}ϕB\phi_{B}
Figure 7.

Let f¯,g¯\bar{f},\bar{g} denote the restrictions of ψXfϕY\psi_{X}\circ f\circ\phi_{Y} and ψBgϕA\psi_{B}\circ g\circ\phi_{A} on the vertex sets of YY and AA respectively.

Proposition 4.2.

(1) The pair of maps (f¯,g¯)(\bar{f},\bar{g}) gives a morphism of metric graph bundles from YY to XX.

Moreover, if YY^{\prime} is the pullback of XX^{\prime} under gg and ff is the pullback map then YY is the pullback of XX under g¯\bar{g} and f¯\bar{f} is the pullback map.

(2) In case, X,YX^{\prime},Y^{\prime} are hyperbolic then ff admits the CT map if and only if so does f¯\bar{f}.

Proof.

(1) Since all the maps in consideration, i.e. ψX,f,ϕY,ψB,g,ϕA\psi_{X},f,\phi_{Y},\psi_{B},g,\phi_{A} are coarsely Lipschitz the maps f¯,g¯\bar{f},\bar{g} are also coarsely Lipschitz by Lemma 2.3(1). It also follows that πXf¯=g¯πY\pi_{X}\circ\bar{f}=\bar{g}\circ\pi_{Y}. Thus (f¯,g¯)(\bar{f},\bar{g}) is a morphism.

YYYY^{\prime}XX^{\prime}XXAAAA^{\prime}BB^{\prime}BBY1Y_{1}ψY\psi_{Y}ϕY\phi_{Y}ffψX\psi_{X}ϕX\phi_{X}πY\pi_{Y}gπg^{*}\pi^{\prime}π\pi^{\prime}πX\pi_{X}ψY\psi_{Y}ψA\psi_{A}ϕA\phi_{A}ggψB\psi_{B}ϕB\phi_{B}f1f_{1}π1\pi_{1}f2f_{2}
Figure 8.

Suppose YY^{\prime} is a the pullback of XX^{\prime} under gg. To show that YY is the pullback of XX we need to verify the universal property. Suppose π1:Y1A\pi_{1}:Y_{1}\rightarrow A is any metric bundle and f1:𝒱(Y1)𝒱(X)f_{1}:{\mathcal{V}}(Y_{1})\rightarrow{\mathcal{V}}(X) is a coarsely Lipschitz map such that the pair (f1,g¯)(f_{1},\bar{g}) is a morphism of metric graph bundles from Y1Y_{1} to XX. We note that π(ϕXf1)=g(ϕAπ1)\pi^{\prime}\circ(\phi_{X}\circ f_{1})=g\circ(\phi_{A}\circ\pi_{1}). Since YY^{\prime} is a set theoretic pullback there is a unique map f2:𝒱(Y1)Yf_{2}:{\mathcal{V}}(Y_{1})\rightarrow Y^{\prime} making the whole diagram below commutative.

Now, by Lemma 2.3(1) the maps ϕXf1\phi_{X}\circ f_{1} and ϕAπ1\phi_{A}\circ\pi_{1} are coarsely Lipschitz. Hence, it follows by Lemma 3.20 and Remark 7 that the map f2f_{2} is coarsely Lipschitz. Let h=ψYf2h=\psi_{Y}\circ f_{2}. Then hh is coarsely Lipschitz by Lemma 2.3(1) and we have f¯h=f1\bar{f}\circ h=f_{1} and πYh=π1\pi_{Y}\circ h=\pi_{1}. Hence, (h,IdA)(h,Id_{A}) is a morphism from Y1Y_{1} to YY. Finally coarse uniqueness of hh follows from Lemma 3.20.

(2) This is a simple application of Lemma 2.50. ∎

4.2. Metric bundles with hyperbolic fibers

For the rest of this section we shall assume that all our metric (graph) bundles π:XB\pi:X\rightarrow B have the following property:

(\ast) Each of the fibers FbF_{b} , bBb\in B (resp. b𝒱(B)b\in\mathcal{V}(B)) is a δ\delta^{\prime}-hyperbolic metric space with respect to the path metric dbd_{b} induced from XX.

We will refer to this by saying that the metric (graph) bundle has uniformly hyperbolic fibers. Moreover, the following property is crucial for the existence of (global) qi sections.

(\ast\ast) There is N0N\geq 0 such that for all bBb\in B the barycenter map ϕb:3FbFb\phi_{b}:\partial^{3}F_{b}\rightarrow F_{b} is coarsely NN-surjective. (Recall that barycenter maps were defined right after Lemma 2.41.)

Proposition 4.3.

([MS12, Proposition 2.10, Proposition 2.12]) Global qi sections for metric (graph) bundles: For all δ,c0,N0\delta^{\prime},c\geq 0,N\geq 0 and η:[0,)[0,)\eta:[0,\infty)\rightarrow[0,\infty) there exists K0=K0(c,η,δ,N)K_{0}=K_{0}(c,\eta,\delta^{\prime},N) such that the following holds.

Suppose p:XBp:X^{\prime}\rightarrow B^{\prime} is an (η,c)(\eta,c)-metric bundle or an η\eta-metric graph bundle satisfying ()(\ast) and ()(\ast\ast). Then there is a K0K_{0}-qi section over BB^{\prime} through each point of XX^{\prime} (where we assume c=1c=1 for the metric graph bundle).

Convention 4.4.

(1) With the notation of Proposition 4.1, we note that for any qi section Σ{\Sigma} in XX over BB, ϕX(Σ)=Σ\phi_{X}({\Sigma})={\Sigma} since ϕX\phi_{X} is the identity map when restricted to 𝒱(X){\mathcal{V}}(X). We shall refer to it as a qi section of the metric graph bundle transported to the metric bundle.

(2) Whenever we talk about a KK-qi section in a metric bundle we shall mean that it is the transport of a KK-qi section contained in the associated canonical metric graph bundle.

Definition 4.5.

([MS12, Definition 2.13]) Suppose Σ1\Sigma_{1} and Σ2\Sigma_{2} are two KK-qi sections of the metric graph bundle XX. For each b𝒱(B)b\in\mathcal{V}(B) we join the points Σ1Fb{\Sigma}_{1}\cap F_{b}, Σ2Fb{\Sigma}_{2}\cap F_{b} by a geodesic in FbF_{b}. We denote the union of these geodesics by 𝕃(Σ1,Σ2){\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}), and call it a KK-ladder (formed by the sections Σ1{\Sigma}_{1} and Σ2{\Sigma}_{2}).

For a metric bundle by a ladder, we will mean one transported from the canonical metric graph bundle associated to it (by the canonical map ϕX\phi_{X} as in Proposition 4.1.)

The following are the most crucial properties of a ladder summarized from [MS12].

Proposition 4.6.

Given K0K\geq 0, δ0\delta\geq 0 there are C=C4.6(K)0C=C_{\ref{ladders are qi embedded}}(K)\geq 0, R=R4.6(K)0R=R_{\ref{ladders are qi embedded}}(K)\geq 0 and K4.6(δ,K)0K_{\ref{ladders are qi embedded}}(\delta,K)\geq 0 such that the following holds:

Suppose π:XB\pi:X\rightarrow B is an η\eta-metric graph bundle satisfying ()(\ast). Suppose Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are two KK-qi sections in XX and 𝕃=𝕃(Σ1,Σ2){\mathbb{L}}={\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}) is the ladder formed by them. Then the following hold.

(1) (Ladders are coarse Lipschitz retracts) There is a coarsely CC-Lipschitz retraction π𝕃:X𝕃\pi_{{\mathbb{L}}}:X\rightarrow{\mathbb{L}} defined as follows:

For all xXx\in X we define π𝕃(x)\pi_{{\mathbb{L}}}(x) to be a nearest point projection of xx in Fπ(x)F_{\pi(x)} on 𝕃Fπ(x){\mathbb{L}}\cap F_{\pi(x)}.

(2) Given a kk-qi section γ\gamma in XX over a geodesic in BB, π𝕃(γ)\pi_{{\mathbb{L}}}(\gamma) is a (C+2kC)(C+2kC)-qi section in XX contained in 𝕃{\mathbb{L}} over the same geodesic in BB.

(3) (QI sections in ladders) If XX also satisfies ()(\ast\ast) then through any point of 𝕃{\mathbb{L}} there is (1+2K)C(1+2K)C-qi section contained in 𝕃{\mathbb{L}}.

(4) (Quasiconvexity of ladders) The RR-neighborhood of 𝕃{\mathbb{L}} is (i) connected and (ii) uniformly qi embedded in XX.

In particular if XX is δ\delta-hyperbolic then 𝕃{\mathbb{L}} is K4.6(δ,K)K_{\ref{ladders are qi embedded}}(\delta,K)-quasiconvex in XX.

Proof.

(1) is stated as Theorem 3.2 in [MS12]. (2), (3) are immediate from (1) or one can refer to Lemma 3.1 of [MS12]. (4) is proved in Lemma 3.6 in [MS12] assuming ()(\ast\ast). However, we briefly indicate the argument here without assuming ()(\ast\ast).

4(i) Suppose b,bBb,b^{\prime}\in B, dB(b,b)=1d_{B}(b,b^{\prime})=1. Let x𝕃Fbx\in{\mathbb{L}}\cap F_{b}. Then there is a point xFbx^{\prime}\in F_{b^{\prime}} such that d(x,x)=1d(x,x^{\prime})=1. Hence, d(π𝕃(x),π𝕃(x))=d(x,π𝕃(x))2Cd(\pi_{{\mathbb{L}}}(x),\pi_{{\mathbb{L}}}(x^{\prime}))=d(x,\pi_{{\mathbb{L}}}(x^{\prime}))\leq 2C. If we define R=2CR=2C then clearly the RR-neighborhood of 𝕃{\mathbb{L}} is connected.

4(ii) We first claim that the NR(𝕃)=YN_{R}({\mathbb{L}})=Y say, is also properly embedded in XX. Suppose x,yYx^{\prime},y^{\prime}\in Y with dX(x,y)Nd_{X}(x^{\prime},y^{\prime})\leq N. Let x,y𝕃x,y\in{\mathbb{L}} be such that d(x,x)R,d(y,y)Rd(x,x^{\prime})\leq R,d(y,y^{\prime})\leq R. Then d(x,y)2R+Nd(x,y)\leq 2R+N. Hence, dB(π(x),π(y))2R+Nd_{B}(\pi(x),\pi(y))\leq 2R+N. Let α\alpha be a geodesic in BB joining π(x),π(y)\pi(x),\pi(y). Then by Lemma 3.8 there is a geodesic lift α~\tilde{\alpha} of α\alpha starting from xx. It follows that for all adjacent vertices b1,b2αb_{1},b_{2}\in\alpha we have d(π𝕃(α~)(b1),π𝕃(α~)(b2))2Cd(\pi_{{\mathbb{L}}}(\tilde{\alpha})(b_{1}),\pi_{{\mathbb{L}}}(\tilde{\alpha})(b_{2}))\leq 2C. Hence, the length of π𝕃(α~)\pi_{{\mathbb{L}}}(\tilde{\alpha}) is at most 2C(2R+N)2C(2R+N). Hence, d(y,π𝕃(α~(π(y)))d(x,y)+d(x,π𝕃(α~(π(y))))2R+N+l(π𝕃(α~))2R+N+2C(2R+N)d(y,\pi_{{\mathbb{L}}}(\tilde{\alpha}(\pi(y)))\leq d(x,y)+d(x,\pi_{{\mathbb{L}}}(\tilde{\alpha}(\pi(y))))\leq 2R+N+l(\pi_{{\mathbb{L}}}(\tilde{\alpha}))\leq 2R+N+2C(2R+N). Hence, dπ(y)(y,π𝕃(α~(π(y))))η(2R+N+4CR+2CN)d_{\pi(y)}(y,\pi_{{\mathbb{L}}}(\tilde{\alpha}(\pi(y))))\leq\eta(2R+N+4CR+2CN). Since π𝕃(α~)Y\pi_{{\mathbb{L}}}(\tilde{\alpha})\subset Y, dY(x,y)dπ(y)(y,π𝕃(α~(π(y))))+l(π𝕃(α~))η(2R+N+4CR+2CN)+4CR+2CNd_{Y}(x,y)\leq d_{\pi(y)}(y,\pi_{{\mathbb{L}}}(\tilde{\alpha}(\pi(y))))+l(\pi_{{\mathbb{L}}}(\tilde{\alpha}))\leq\eta(2R+N+4CR+2CN)+4CR+2CN. Hence, dY(x,y)4CR+2CN+η(2R+N+4CR+2CN)d_{Y}(x^{\prime},y^{\prime})\leq 4CR+2CN+\eta(2R+N+4CR+2CN).

Finally we prove the qi embedding. Let f(N)=η(2R+N+4CR+2CN)+4CR+2CNf(N)=\eta(2R+N+4CR+2CN)+4CR+2CN for all NN\in{\mathbb{N}}. Given x,y𝕃x,y\in{\mathbb{L}}, dX(x,y)=nd_{X}(x,y)=n and a geodesic γ:[0,n]X\gamma:[0,n]\rightarrow X joining them. By the proof of (4)(i) we have dY(π𝕃(γ(i)),π𝕃(γ(i+1))f(2C)d_{Y}(\pi_{{\mathbb{L}}}(\gamma(i)),\pi_{{\mathbb{L}}}(\gamma(i+1))\leq f(2C) for all 0in10\leq i\leq n-1 whence d𝕃(x,y)nf(2C)=f(2C)dX(x,y)d_{{\mathbb{L}}}(x,y)\leq nf(2C)=f(2C)d_{X}(x,y). Clearly dX(x,y)d𝕃(x,y)d_{X}(x,y)\leq d_{{\mathbb{L}}}(x,y). This proves the qi embedded part.

It follows that for all x,y𝕃x,y\in{\mathbb{L}} a geodesic joining x,yx,y in YY is a (f(2C),0)(f(2C),0)-quasigeodesic in XX. Since XX is δ\delta-hyperbolic stability of quasigeodesics implies that 𝕃{\mathbb{L}} is uniformly quasiconvex. In fact, we can take K4.6(δ,K)=R+D2.19(δ,f(2C),0)K_{\ref{ladders are qi embedded}}(\delta,K)=R+D_{\ref{cor: stab-qg}}(\delta,f(2C),0). ∎

Remark 10.

Part (3) and (4) are clearly also true for metric bundles which satisfy the properties ()(\ast) and ()(\ast\ast).

The following corollary is immediate.

Corollary 4.7.

(Ladders form subbundles) Suppose π:XB\pi:X\rightarrow B is an η\eta-metric graph bundle satisfying ()(\ast) and ()(\ast\ast). Let C,RC,R be as in the previous proposition. Suppose 𝕃=𝕃(Σ1,Σ2){\mathbb{L}}={\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}) is a KK-ladder. Consider the metric graph ZZ obtained from 𝕃{\mathbb{L}} by introducing some extra edges as follows: Suppose b,bBb,b^{\prime}\in B are adjacent vertices then for all x𝕃Fbx\in{\mathbb{L}}\cap F_{b}, x𝕃Fbx^{\prime}\in{\mathbb{L}}\cap F_{b^{\prime}} we join x,xx,x^{\prime} by an edge if and only if dX(x,x)C+2KCd_{X}(x,x^{\prime})\leq C+2KC. Let πZ:ZB\pi_{Z}:Z\rightarrow B be the simplicial map such that π=πZ\pi=\pi_{Z} on 𝒱(Z){\mathcal{V}}(Z) and the extra edges are mapped isometrically to edges of BB.

Then ZZ is a metric graph bundle and the natural map ZXZ\rightarrow X gives a subbundle of XX which is also a (uniform) qi onto NR(𝕃)N_{R}({\mathbb{L}}) and hence a (uniform) qi embedding in XX.

In the next section of the paper, we will exclusively deal with bundles π:XB\pi:X\rightarrow B which are hyperbolic satisfying ()(\ast) and ()(\ast\ast) and we will need to understand geodesics in XX. Since ladders are quasiconvex we look for quasigeodesics contained in ladders. The lemma below is the last technical piece of information needed for that purpose. However, we need the following definitions for stating the lemma.

Definition 4.8.

Suppose XX is a metric graph bundle over BB and suppose Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are any two qi sections.

(1) Neck of ladders ([MS12, Definition 2.16]). Suppose R0R\geq 0. Then the set UR(Σ1,Σ2)={bB:db(Σ1Fb,Σ2Fb)R}U_{R}({\Sigma}_{1},{\Sigma}_{2})=\{b\in B:\,d_{b}({\Sigma}_{1}\cap F_{b},{\Sigma}_{2}\cap F_{b})\leq R\} is called the RR-neck of the ladder 𝕃(Σ1,Σ2){\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}).

For a metric bundle the RR-neck of a ladder will be defined to be the one transported from the canonical metric graph bundle associated to it, i.e. the image under ϕB\phi_{B}.

(2) Girth of ladders ([MS12, Definition 2.15]). The quantity min{db(Σ1Fb,Σ2Fb):bB}\min\{d_{b}({\Sigma}_{1}\cap F_{b},{\Sigma}_{2}\cap F_{b}):b\in B\} is called the girth of the ladder 𝕃(Σ1,Σ2){\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}) and it will be denoted by dh(Σ1,Σ2)d_{h}({\Sigma}_{1},{\Sigma}_{2}).

Definition 4.9.

([MS12, Definition 1.12])(Flaring for metric graph bundles) Suppose π:XB\pi:X\rightarrow B is a metric graph bundle. We say that it satisfies a flaring condition if for all k1k\geq 1, there exist νk>1\nu_{k}>1 and nk,Mkn_{k},M_{k}\in\mathbb{N} such that the following holds:
Let γ:[nk,nk]B\gamma:[-n_{k},n_{k}]\rightarrow B be a geodesic and let γ1~\tilde{\gamma_{1}} and γ2~\tilde{\gamma_{2}} be two kk-qi lifts of γ\gamma in XX. If dγ(0)(γ1~(0),γ2~(0))Mkd_{\gamma(0)}(\tilde{\gamma_{1}}(0),\tilde{\gamma_{2}}(0))\geq M_{k}, then we have

 νk.dγ(0)(γ1~(0),γ2~(0))max{dγ(nk)(γ1~(nk),γ2~(nk)),dγ(nk)(γ1~(nk),γ2~(nk))}.\mbox{ {\small$\nu_{k}.d_{\gamma(0)}(\tilde{\gamma_{1}}(0),\tilde{\gamma_{2}}(0))\leq\mbox{max}\{d_{\gamma(n_{k})}(\tilde{\gamma_{1}}(n_{k}),\tilde{\gamma_{2}}(n_{k})),d_{\gamma(-n_{k})}(\tilde{\gamma_{1}}(-n_{k}),\tilde{\gamma_{2}}(-n_{k}))\}$}}.

We note that existence of flaring in a metric graph bundle implies the existence of three functions νk,nk,Mk\nu_{k},n_{k},M_{k} of kk with the said property in the above definition. This is independent of the hypotheses about metric graph bundles and the conditions ()(\ast) and ()(\ast\ast) mentioned in the beginning of this subsection. This notion is motivated from the hallway flaring condition of Bestvina-Feighn ([BF92]).

Definition 4.10.

(Flaring for metric bundles) We shall say that a metric bundle π:XB\pi:X\rightarrow B satisfies a (νk,Mk,nk)(\nu_{k},M_{k},n_{k})-flaring condition if the canonical metric graph bundle associated to it satisfies a (νk,Mk,nk)(\nu_{k},M_{k},n_{k})-flaring condition.

Remark 11.

(1) Since the base for a metric bundle need not be a geodesic metric space, it is not reasonable to use [MS12, Definition 1.12] of flaring for metric bundles. However, one can formulate analogous flaring of qi sections over uniform quasigeodesics in the base and then show that this is indeed equivalent to Definition 4.10. Since this discussion is not directly related to the rest of the paper we move it to the end of the paper and we include it as an appendix. See Lemma A.5 and Lemma A.6.

(2) This definition of flaring for metric bundles is equivalent to [MS12, Definition 1.12] in the case of geodesic metric bundles. In fact it follows from Lemma A.5 and Lemma A.6 that a geodesic metric bundle satisfies flaring as per [MS12, Definition 1.12] iff the canonical metric graph bundle associated to it also satisfies flaring.

The following lemma will be crucial for the next section of the paper.

Lemma 4.11.

(Quasiconvexity of necks of ladders, [MS12, Lemma 2.18]) Let XX be an η\eta-metric graph bundle over BB satisfying (νk,Mk,nk)(\nu_{k},M_{k},n_{k})-flaring condition for all k1k\geq 1. Then for all c11c_{1}\geq 1 and R>1R>1 there are constants D4.11=D4.11(c1,R)D_{\ref{qc-level-set-new}}=D_{\ref{qc-level-set-new}}(c_{1},R) and K4.11=K4.11(c1)K_{\ref{qc-level-set-new}}=K_{\ref{qc-level-set-new}}(c_{1}) such that the following holds:
Suppose Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are two c1c_{1}-qi sections of BB in XX and let Lmax{Mc1,dh(Σ1,Σ2)}L\geq max\{M_{c_{1}},d_{h}({\Sigma}_{1},{\Sigma}_{2})\}.

  1. (1)

    Let γ:[t0,t1]B\gamma:[t_{0},t_{1}]\rightarrow B be a geodesic, t0,t1t_{0},t_{1}\in\mathbb{Z}, such that
    a) dγ(t0)(Σ1Fγ(t0),Σ2Fγ(t0))=LRd_{\gamma(t_{0})}({\Sigma}_{1}\cap F_{\gamma(t_{0})},{\Sigma}_{2}\cap F_{\gamma(t_{0})})=LR.
    b) γ(t1)UL:=UL(Σ1,Σ2)\gamma(t_{1})\in U_{L}:=U_{L}({\Sigma}_{1},{\Sigma}_{2}) but for all t[t0,t1)t\in[t_{0},t_{1})\cap\mathbb{Z}, γ(t)UL\gamma(t)\not\in U_{L}.
    Then the length of γ\gamma is at most D4.11(c1,R)D_{\ref{qc-level-set-new}}(c_{1},R).

  2. (2)

    For any b1,b2ULb_{1},b_{2}\in U_{L} and any geodesic [b1,b2][b_{1},b_{2}] joining them in BB, we have [b1,b2]NK4.11(UL)[b_{1},b_{2}]\subset N_{K_{\ref{qc-level-set-new}}}(U_{L}). In particular, if BB is hyperbolic then ULU_{L} is K4.11K_{\ref{qc-level-set-new}}-quasiconvex in BB.

  3. (3)

    If dh(Σ1,Σ2)Mc1d_{h}({\Sigma}_{1},{\Sigma}_{2})\geq M_{c_{1}} then the diameter of the set ULU_{L} is at most D4.11=D4.11(c1,L)D^{\prime}_{\ref{qc-level-set-new}}=D^{\prime}_{\ref{qc-level-set-new}}(c_{1},L).

Part (2) of the above lemma is slightly different from that of [MS12, Lemma 2.18] but the proof there actually showed this. However, ladders with short necks to which Lemma 4.11 applies are given a special name:

Definition 4.12.

(Small girth ladders) Given two KK-qi sections Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} in a metric graph bundle satisfying a flaring condition the ladder 𝕃(Σ1,Σ2){\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}) is called a small girth ladder if UL(Σ1,Σ2)U_{L}({\Sigma}_{1},{\Sigma}_{2})\neq\emptyset where L=MKL=M_{K}.

Remark 12.

Suppose XBX^{\prime}\rightarrow B^{\prime} is a metric bundle and XBX\rightarrow B is the canonical metric graph bundle associated to it. Suppose a flaring condition holds for XX. This is the case for instance when XX or equivalently XX^{\prime} is hyperbolic. In such a case, a small girth ladder in XX^{\prime} for us will be, by definition, the transport of a small girth ladder from XX under ϕX\phi_{X} (as in Proposition 4.1).

We end this section with two simple lemmas. We note that flaring condition is not needed for these to hold.

Lemma 4.13.

Given D0,K1D\geq 0,K\geq 1 there is R=R4.13(D,K)R=R_{\ref{distance from qi section}}(D,K) such that the following holds.

Suppose Σ{\Sigma} is KK-qi section in XX and xXx\in X. Let b=π(x)b=\pi(x). Then d(x,Σ)Dd(x,{\Sigma})\geq D if db(x,ΣFb)Rd_{b}(x,{\Sigma}\cap F_{b})\geq R.

Proof.

Suppose yΣy\in{\Sigma} a nearest point from xx. Let αΣ\alpha\subset{\Sigma} be the lift of a geodesic [b,π(y)][b,\pi(y)] joining bb to π(y)\pi(y) joining yy to ΣFb{\Sigma}\cap F_{b}. We note that dB(b,π(y))d(x,y)d_{B}(b,\pi(y))\leq d(x,y). Hence, d(y,α(b))Kd(x,y)+Kd(y,\alpha(b))\leq Kd(x,y)+K. Therefore, d(x,α(b))d(x,y)+d(y,α(b))(K+1)d(x,y)+Kd(x,\alpha(b))\leq d(x,y)+d(y,\alpha(b))\leq(K+1)d(x,y)+K. This implies d(x,y)1K+1d(x,α(b))d(x,y)\geq\frac{1}{K+1}d(x,\alpha(b)) since all distances are integers in this case. Now fibers of XX are properly embedded as measured by η\eta. Thus if db(x,α(b))η((K+1)D)d_{b}(x,\alpha(b))\geq\eta((K+1)D) then d(x,y)Dd(x,y)\geq D. Hence, we can take R=η(KD+D)R=\eta(KD+D). ∎

The corollary below gives a relation between the girth of a ladder 𝕃(Σ1,Σ2){\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}) and d(Σ1,Σ2)d({\Sigma}_{1},{\Sigma}_{2}).

Corollary 4.14.

Given D0,K1D\geq 0,K\geq 1 there is an R=R4.14(D,K)R=R_{\ref{girth vs section distance}}(D,K) such that the following holds.
Suppose Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are two KK-qi sections in XX. Then d(Σ1,Σ2)Dd({\Sigma}_{1},{\Sigma}_{2})\geq D if UR(Σ1,Σ2)=U_{R}({\Sigma}_{1},{\Sigma}_{2})=\emptyset.

The next lemma is a generalization of Lemma 4.13. Nevertheless we keep both of them since they are used many times in the next section.

Lemma 4.15.

Given K,DK,D there is R=R4.15(K,D)R=R_{\ref{distance from ladder}}(K,D) such that the following holds.

Suppose Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are two KK-qi sections in XX and 𝕃=𝕃(Σ1,Σ2){\mathbb{L}}={\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}). Suppose xXx\in X and π(x)=b\pi(x)=b. Then d(x,𝕃)Dd(x,{\mathbb{L}})\geq D if db(x,𝕃Fb)Rd_{b}(x,{\mathbb{L}}\cap F_{b})\geq R.

Proof.

Suppose y𝕃y\in{\mathbb{L}} is a nearest point from xx. Let α\alpha be a geodesic lift of any geodesic [b,π(y)][b,\pi(y)] joining bb to π(y)\pi(y) such that α\alpha joins yy to FbF_{b}. Now π𝕃(α)\pi_{{\mathbb{L}}}(\alpha) is a 2C2C-qi lift of [b,π(y)][b,\pi(y)] where C=C4.6(K)C=C_{\ref{ladders are qi embedded}}(K). Thus d(y,π𝕃(α)(b))2CdB(b,π(y)+2C2Cd(x,y)+2Cd(y,\pi_{{\mathbb{L}}}(\alpha)(b))\leq 2Cd_{B}(b,\pi(y)+2C\leq 2Cd(x,y)+2C. Hence, d(x,𝕃Fb)d(x,y)+d(y,π𝕃(α)(b))(2C+1)d(x,y)+2Cd(x,{\mathbb{L}}\cap F_{b})\leq d(x,y)+d(y,\pi_{{\mathbb{L}}}(\alpha)(b))\leq(2C+1)d(x,y)+2C. Therefore, d(x,y)12C+1d(x,𝕃Fb)d(x,y)\geq\frac{1}{2C+1}d(x,{\mathbb{L}}\cap F_{b}). Hence, we can take R=η((2C+1)D)R=\eta((2C+1)D). ∎

5. Cannon-Thurston maps for pull-back bundles

In this section, we prove the main result of the paper. Here is the set-up. From now on we suppose that π:XB\pi:X\rightarrow B is an (η,c)(\eta,c)-metric bundle or an η\eta-metric graph bundle satisfying the following hypotheses.

  • (H1)

    BB is a δ0\delta_{0}-hyperbolic metric space.

  • (H2)

    Each of the fibers FbF_{b}, bBb\in B is a δ0\delta_{0}-hyperbolic metric space with respect to the path metric induced from XX.

  • (H3)

    The barycenter maps 3FbFb\partial^{3}F_{b}\rightarrow F_{b}, bBb\in B (resp. b𝒱(B)b\in\mathcal{V}(B)) are N0N_{0}-coarsely surjective for some constant N0N_{0}.

  • (H4)

    The (νk,Mk,nk)(\nu_{k},M_{k},n_{k})-flaring condition is satisfied for all k1k\geq 1.

The following theorem is the main result of [MS12]:

Theorem 5.1.

([MS12, Theorem 4.3 and Proposition 5.8]) If π:XB\pi:X\rightarrow B is a geodesic metric bundle or a metric graph bundle satisfying H1,H2,H3H1,H2,H3 then XX is a hyperbolic metric space if and only if XX satisfies a flaring condition.

5.1. Proof of the main theorem

We are now ready to state and prove the main theorem of the paper.

Theorem 5.2.

(Main Theorem) Suppose π:XB\pi:X\rightarrow B is a metric (graph) bundle satisfying the hypotheses H1, H2, H3, and H4. Suppose g:ABg:A\rightarrow B is a Lipschitz kk-qi embedding and suppose p:YAp:Y\rightarrow A is the pullback bundle. Let f:YXf:Y\rightarrow X be the pullback map.

Then YY is a hyperbolic metric space and the CT map exists for f:YXf:Y\rightarrow X.

Proof.

We first note that XX is hyperbolic. This follows from Theorem 5.1 if XX is a metric graph bundle (or a geodesic metric bundle). In case XX is a (length) metric bundle one may first pass to the canonical metric graph bundle associated to it, and then verify the hypotheses of Theorem 5.1 for it. In fact, if any metric bundle satisfies (H1), (H2), and (H3) then the canonical metric graph bundle associated to it also has these properties with possibly different values of the respective parameters. Flaring condition (H4) follows from Definition 4.10. It then follows that the metric graph bundle is hyperbolic. Consequently, XX is hyperbolic by Proposition 4.1. We shall assume that XX is δ\delta-hyperbolic. We begin with the following reductions: (1) It is enough to prove the theorem only for metric graph bundles. Indeed this follows from Proposition 4.2(2). So for the rest of the proof we shall assume that π:XB\pi:X\rightarrow B is a metric graph bundle satisfying (H1), (H2), (H3), (H4).

Since we work with graphs from now, for the rest of the section by hyperbolicity we shall mean Rips hyperbolicity.

(2) We may moreover assume that AA is a connected subgraph, g:ABg:A\rightarrow B is the inclusion map and YY is the restriction bundle for that inclusion. In particular, f:YXf:Y\rightarrow X is the inclusion map and Y=π1(A)Y=\pi^{-1}(A).

Since g:ABg:A\rightarrow B is a kk-qi embedding and BB is δ0\delta_{0}-hyperbolic, g(A)g(A) is D2.19(δ0,k,k)D_{\ref{cor: stab-qg}}(\delta_{0},k,k)-quasiconvex in BB. Let AA^{\prime} be the D2.19(δ0,k,k)D_{\ref{cor: stab-qg}}(\delta_{0},k,k)-neighborhood of g(A)g(A) in BB. Then clearly AA^{\prime} is connected subgraph of BB and g:AAg:A\rightarrow A^{\prime} is a quasiisometry with respect to the induced path metric on AA^{\prime} from BB. Clearly AA^{\prime} is (1,4D2.19(δ0,k,k))(1,4D_{\ref{cor: stab-qg}}(\delta_{0},k,k))-qi embedded. Let π:X=π1(A)A\pi^{\prime}:X^{\prime}=\pi^{-1}(A^{\prime})\rightarrow A^{\prime} be the restriction of π\pi on XX^{\prime}. Then π:XA\pi^{\prime}:X^{\prime}\rightarrow A^{\prime} is a metric graph bundle by Lemma 3.17. Also, we note that (f,g):YX(f,g):Y\rightarrow X^{\prime} is a morphism of metric graph bundles. By Corollary 3.25 the fiber maps of the morphism f:YXf:Y\rightarrow X^{\prime} are uniform quasiisometries and hence by Theorem 3.15 we see that f:YXf:Y\rightarrow X^{\prime} is an isomorphism of metric graph bundles. Since (Rips) hyperbolicity of graphs is a qi invariant, we are reduced to proving hyperbolicity of XX^{\prime} and also by Lemma 2.50(1) we are reduced to proving the existence of the CT map for the inclusion XXX^{\prime}\rightarrow X.

Hyperbolicity of YY
YY
is hyperbolic by Remark 4.44.4 of [MS12]. In fact, by Theorem 5.1 it is enough to check that flaring holds for the bundle YAY\rightarrow A. This is a consequence of flaring of the bundle π:XB\pi:X\rightarrow B and bounded flaring.

Remark 13.

(1) The sole purpose of (H3)(H3) is to have global uniform qi sections through every point of XX which is guaranteed by Proposition 4.3. For the rest of this section, we shall also assume the following.

(H3\,{}^{\prime}) Through any point of XX there is a global K0K_{0}-qi section.

(2) Clearly YY is an η\eta-metric graph bundle over AA satisfying H2, H3. We shall assume that AA is δ0\delta^{\prime}_{0}-hyperbolic. We shall also assume the bundle YY satisfies a (νk,Mk,nk)(\nu^{\prime}_{k},M^{\prime}_{k},n^{\prime}_{k})-flaring condition for all k1k\geq 1.

Existence of CT map
Outline of the proof: To prove the existence of the CT map we use Lemma 2.49. The different steps used in the proof are as follows. (1) Given y,yYy,y^{\prime}\in Y first we define a uniform quasigeodesic c(y,y)c(y,y^{\prime}) in XX joining y,yy,y^{\prime}. This is extracted from [MS12]. (2) In the next step we modify c(y,y)c(y,y^{\prime}) to obtain a path c¯(y,y)\bar{c}(y,y^{\prime}) in YY. (3) We then check that these paths are uniform quasigeodesics in YY. (4) Finally we verify the condition of Lemma 2.49 for the paths c(y,y)c(y,y^{\prime}) and c¯(y,y)\bar{c}(y,y^{\prime}). Since X,YX,Y are hyperbolic metric spaces, stability of quasigeodesics and Lemma 2.49 finishes the proof. To maintain modularity of the arguments we state intermediate observations as lemma, proposition etc.

Remark 14.

Although we assumed that y,yYy,y^{\prime}\in Y as is necessary for our proof, c(y,y)c(y,y^{\prime}) as defined below is a uniform quasigeodesic for all y,yXy,y^{\prime}\in X as it will follow from the proof.

However, we would like to note that description of uniform quasigeodesics in a metric graph bundle with the above properties H1-H4 is already contained in [MS12], e.g. see Proposition 3.4, and Proposition 3.14 of [MS12]. We make it more explicit with the help of Proposition 2.33.

Step 1: Descriptions of the uniform quasigeodesic c(y,y)c(y,y^{\prime}).
The description of the paths and the proof that they are uniform quasigeodesics in XX is broken up into three further substeps.

Step 1(a): Choosing a ladder containing y,yy,y^{\prime}. We begin by choosing any two K0K_{0}-qi sections Σ,Σ{\Sigma},{\Sigma}^{\prime} in XX containing y,yy,y^{\prime} respectively. Let 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) be the ladder formed by them. Throughout Step 1 we shall work with these qi sections and ladder. The path c(y,y)c(y,y^{\prime}) that we shall construct in Step 1(c) will be contained in this ladder.

Step 1(b): Decomposition of the ladder into small girth ladders.
We next choose finitely many qi sections in 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) after [MS12, Proposition 3.14] in a way suitable for using Proposition 2.33. This requires a little preparation. We start with the following.

Lemma 5.3.

For all K1K\geq 1 there is D5.3(K)D_{\ref{lem: ladders cobdd}}(K) such that the following holds in XX.

Suppose Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are two KK-qi sections and dh(Σ1,Σ2)MKd_{h}({\Sigma}_{1},{\Sigma}_{2})\geq M_{K}. Then Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are D5.3(K)D_{\ref{lem: ladders cobdd}}(K)-cobounded.

Proof.

We note that Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are K=D2.19(δ,K,K)K^{\prime}=D_{\ref{cor: stab-qg}}(\delta,K,K)-quasiconvex in XX. Suppose P:XΣ1P:X\rightarrow{\Sigma}_{1} is an 11-approximate nearest point projection map and the diameter of P(Σ2)P({\Sigma}_{2}) is bigger than D=D2.28(δ,K,1)D=D_{\ref{cor: lip proj}}(\delta,K^{\prime},1). Then d(Σ1,Σ2)R=R2.28(δ,K,1)d({\Sigma}_{1},{\Sigma}_{2})\leq R=R_{\ref{cor: lip proj}}(\delta,K^{\prime},1). If xΣ2x\in{\Sigma}_{2} such that d(x,Σ1)Rd(x,{\Sigma}_{1})\leq R and b=π(x)b=\pi(x) then db(x,Σ1Fb)R4.13(R,K)=R¯d_{b}(x,{\Sigma}_{1}\cap F_{b})\leq R_{\ref{distance from qi section}}(R,K^{\prime})=\bar{R}, say. Hence, π(P(Σ2))UR¯(Σ1,Σ2)\pi(P({\Sigma}_{2}))\subset U_{\bar{R}}({\Sigma}_{1},{\Sigma}_{2}). However, by Lemma 4.11 the diameter of UR¯(Σ1,Σ2)U_{\bar{R}}({\Sigma}_{1},{\Sigma}_{2}) is at most D4.11(K,R¯)D^{\prime}_{\ref{qc-level-set-new}}(K^{\prime},\bar{R}). It follows that the diameter of P(Σ2)P({\Sigma}_{2}) is at most K+KD4.11(K,R¯)K+KD^{\prime}_{\ref{qc-level-set-new}}(K^{\prime},\bar{R}). Hence we may choose D5.3(K)=max{D2.28(δ,K,1),K+KD4.11(K,R¯)}D_{\ref{lem: ladders cobdd}}(K)=\max\{D_{\ref{cor: lip proj}}(\delta,K^{\prime},1),K+KD^{\prime}_{\ref{qc-level-set-new}}(K^{\prime},\bar{R})\}. ∎

Lemma 5.4.

Suppose Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are two KK-qi sections and Σ𝕃(Σ1,Σ2){\Sigma}\subset{\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}) is KK-qi section. Then Σ{\Sigma} coarsely uniformly bisects 𝕃(Σ1,Σ2){\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}) into the subladders 𝕃(Σ1,Σ){\mathbb{L}}({\Sigma}_{1},{\Sigma}) and 𝕃(Σ,Σ2){\mathbb{L}}({\Sigma},{\Sigma}_{2}).

Proof.

First of all any ladder formed by KK-qi sections is K4.6(δ,K)K_{\ref{ladders are qi embedded}}(\delta,K)-quasiconvex. Let K=K4.6(δ,K)K^{\prime}=K_{\ref{ladders are qi embedded}}(\delta,K). Let k1k\geq 1, and xiΣix_{i}\in{\Sigma}_{i}, i=1,2i=1,2 be any points. Let γx1x2:IX\gamma_{x_{1}x_{2}}:I\rightarrow X be a kk-quasigeodesic joining them where II is an interval. Then there are points t1,t2It_{1},t_{2}\in I with |t1t2|1|t_{1}-t_{2}|\leq 1 such that γx1x2(t1)NK(𝕃(Σ1,Σ))\gamma_{x_{1}x_{2}}(t_{1})\in N_{K^{\prime}}({\mathbb{L}}({\Sigma}_{1},{\Sigma})) and γx1x2(t2)NK(𝕃(Σ,Σ2))\gamma_{x_{1}x_{2}}(t_{2})\in N_{K^{\prime}}({\mathbb{L}}({\Sigma},{\Sigma}_{2})). Let y1𝕃(Σ1,Σ)y_{1}\in{\mathbb{L}}({\Sigma}_{1},{\Sigma}) and y2𝕃(Σ,Σ2)y_{2}\in{\mathbb{L}}({\Sigma},{\Sigma}_{2}) be such that d(yi,γx1x2(ti))Kd(y_{i},\gamma_{x_{1}x_{2}}(t_{i}))\leq K^{\prime}, i=1,2i=1,2. We note that d(γx1x2(t1),γx1x2(t2))2kd(\gamma_{x_{1}x_{2}}(t_{1}),\gamma_{x_{1}x_{2}}(t_{2}))\leq 2k. Hence, d(y1,y2)2K+2kd(y_{1},y_{2})\leq 2K^{\prime}+2k. Let b=π(y1)b=\pi(y_{1}). Then db(y1,𝕃(Σ,Σ2)Fb)R4.15(K,2K+2k)d_{b}(y_{1},{\mathbb{L}}({\Sigma},{\Sigma}_{2})\cap F_{b})\leq R_{\ref{distance from ladder}}(K,2K^{\prime}+2k). This implies db(y1,ΣFb)R4.15(K,2K+2k)d_{b}(y_{1},{\Sigma}\cap F_{b})\leq R_{\ref{distance from ladder}}(K,2K^{\prime}+2k). Thus d(γx1x2(t1),Σ)K+R4.15(K,2K+2k)d(\gamma_{x_{1}x_{2}}(t_{1}),{\Sigma})\leq K^{\prime}+R_{\ref{distance from ladder}}(K,2K^{\prime}+2k). This proves the lemma. ∎

Lemma 5.5.

If 𝒬{\mathcal{Q}} is a KK-qi section in XX then 𝒬Y{\mathcal{Q}}\cap Y is a K5.5(K)K_{\ref{retricting sections}}(K)-qi section of AA in YY.

Proof.

Suppose s:BXs:B\rightarrow X is the KK-qi embedding such that s(B)=𝒬s(B)={\mathcal{Q}}. Let ss also denote the restriction on AA. Since the bundle map YAY\rightarrow A is 11-Lipschitz we have dA(u,v)dY(s(u),s(v))d_{A}(u,v)\leq d_{Y}(s(u),s(v)) for all u,vAu,v\in A. Thus it is enough to show that s:AYs:A\rightarrow Y is uniformly coarsely Lipschitz. Suppose u,vAu,v\in A are adjacent vertices. Then dX(s(u),s(v))2Kd_{X}(s(u),s(v))\leq 2K. Now, there is a vertex xFvx\in F_{v} adjacent to s(u)Fus(u)\in F_{u}. Hence, dX(s(v),x)1+2Kd_{X}(s(v),x)\leq 1+2K. Therefore, dv(s(v),x)η(1+2K)d_{v}(s(v),x)\leq\eta(1+2K). Hence, dY(s(u),s(v))1+η(1+2K)d_{Y}(s(u),s(v))\leq 1+\eta(1+2K). It follows that for all u,vAu,v\in A we have dY(s(u),s(v))(1+η(1+2K))dA(u,v)d_{Y}(s(u),s(v))\leq(1+\eta(1+2K))d_{A}(u,v). Hence, we can take K5.5(K)=1+η(1+2K)K_{\ref{retricting sections}}(K)=1+\eta(1+2K). ∎

The following corollary is proved exactly as Lemma 5.3. Hence we omit the proof.

Corollary 5.6.

For all K1K\geq 1 there is D5.6(K)0D_{\ref{Y ladders cobdd}}(K)\geq 0 such that the following holds.

Suppose Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are two KK-qi sections in XX and dh(Σ1,Σ2)MKd_{h}({\Sigma}_{1},{\Sigma}_{2})\geq M_{K}. Then Σ1Y,Σ2Y{\Sigma}_{1}\cap Y,{\Sigma}_{2}\cap Y are D5.6(K)D_{\ref{Y ladders cobdd}}(K)-cobounded in YY.

Before describing the decomposition of ladders the following conclusions and notation on qi sections and ladders will be useful to record.

Convention 5.7.

(C0) We recall that AA is kk-qi embedded in BB. We let k0=D2.17(δ0,k,k)k_{0}=D_{\ref{stab-qg}}(\delta_{0},k,k) so that AA is k0k_{0}-quasiconvex in BB. Finally we assume that YY is δ\delta^{\prime} hyperbolic.

(C1) Let Ki+1=(1+2K0)C4.6(Ki)K_{i+1}=(1+2K_{0})C_{\ref{ladders are qi embedded}}(K_{i}) for all ii\in{\mathbb{N}} where K0K_{0} is as in (H33^{\prime}). Therefore, through any point of a KiK_{i}-ladder in XX, there is a Ki+1K_{i+1}-qi section contained in the ladder. Let Ki=K5.5(Ki)K^{\prime}_{i}=K_{\ref{retricting sections}}(K_{i}).

(C2) We let λi=max{D2.19(δ,Ki,Ki),K4.6(δ,Ki),D2.19(δ,Ki,Ki),K4.6(δ,Ki)}\lambda_{i}=\max\{D_{\ref{cor: stab-qg}}(\delta,K_{i},K_{i}),K_{\ref{ladders are qi embedded}}(\delta,K_{i}),D_{\ref{cor: stab-qg}}(\delta^{\prime},K^{\prime}_{i},K^{\prime}_{i}),K_{\ref{ladders are qi embedded}}(\delta^{\prime},K^{\prime}_{i})\} so that any KiK_{i}-qi section 𝒬X{\mathcal{Q}}\subset X and any ladder 𝕃X{\mathbb{L}}\subset X formed by two KiK_{i}-qi sections in XX are λi\lambda_{i}-quasiconvex in XX and moreover 𝒬Y{\mathcal{Q}}\cap Y and 𝕃Y{\mathbb{L}}\cap Y are λi\lambda_{i}-quasiconvex in YY.

(C3) If Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are two KiK_{i}-qi sections in XX and dh(Σ1,Σ2)MKid_{h}({\Sigma}_{1},{\Sigma}_{2})\geq M_{K_{i}} then they are DiD_{i}-cobounded in XX, as are Σ1Y,Σ2Y{\Sigma}_{1}\cap Y,{\Sigma}_{2}\cap Y in YY where Di=max{D5.3(Ki),D5.6(Ki)}D_{i}=\max\{D_{\ref{lem: ladders cobdd}}(K_{i}),D_{\ref{Y ladders cobdd}}(K_{i})\}.

(C4) For each pair of KiK_{i}-qi sections Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} in XX with dh(Σ1,Σ2)>ri=max{R4.14(2λi+1,Ki),R4.14(2λi+1,Ki)}d_{h}({\Sigma}_{1},{\Sigma}_{2})>r_{i}=\max\{R_{\ref{girth vs section distance}}(2\lambda_{i}+1,K_{i}),R_{\ref{girth vs section distance}}(2\lambda_{i}+1,K^{\prime}_{i})\} we have dX(Σ1,Σ2)>2λi+1d_{X}({\Sigma}_{1},{\Sigma}_{2})>2\lambda_{i}+1 and dY(Σ1Y,Σ2Y)>2λi+1d_{Y}({\Sigma}_{1}\cap Y,{\Sigma}_{2}\cap Y)>2\lambda_{i}+1.

The following proposition is extracted from Proposition 3.14 of [MS12]. The various parts of this proposition are contained in the different steps of the proof of [MS12, Proposition 3.14].

Let us fix a point b0Ab_{0}\in A once and for all. Suppose α:[0,l]Fb0𝕃(Σ,Σ)\alpha:[0,l]\rightarrow F_{b_{0}}\cap{\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) is an isometry such that α(0)=ΣFb0\alpha(0)={\Sigma}\cap F_{b_{0}} and ΣFb0=α(l){\Sigma}^{\prime}\cap F_{b_{0}}=\alpha(l).

Proposition 5.8.

(See [MS12, Corollary 3.13 and Proposition 3.14]) There is a constant L0L_{0} such that for all LL0L\geq L_{0} there is a partition 0=t0<t1<<tn=l0=t_{0}<t_{1}<\cdots<t_{n}=l of [0,l][0,l] and K1K_{1}-qi sections Σi{\Sigma}_{i} passing through α(ti)\alpha(t_{i}), 0in0\leq i\leq n inside 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) such that the following holds.

  1. (1)

    Σ0=Σ,Σn=Σ{\Sigma}_{0}={\Sigma},{\Sigma}_{n}={\Sigma}^{\prime}.

  2. (2)

    For 0in20\leq i\leq n-2, Σi+1𝕃(Σi,Σ){\Sigma}_{i+1}\subset{\mathbb{L}}({\Sigma}_{i},{\Sigma}^{\prime}).

  3. (3)

    For 0in20\leq i\leq n-2 either (I) dh(Σi,Σi+1)=Ld_{h}({\Sigma}_{i},{\Sigma}_{i+1})=L, or (II) dh(Σi,Σi+1)>Ld_{h}({\Sigma}_{i},{\Sigma}_{i+1})>L and there is a K2K_{2}-qi section Σi{\Sigma}^{\prime}_{i} through α(ti+11)\alpha(t_{i+1}-1) inside 𝕃(Σi,Σi+1){\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1}) such that dh(Σi,Σi)<C+CLd_{h}({\Sigma}_{i},{\Sigma}^{\prime}_{i})<C+CL where C=C4.6(K1)C=C_{\ref{ladders are qi embedded}}(K_{1}).

  4. (4)

    dh(Σn1,Σn)Ld_{h}({\Sigma}_{n-1},{\Sigma}_{n})\leq L.

However, we will need a slightly different decomposition of 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) than what is described here. It is derived as the following corollary to the Proposition 5.8.

Convention 5.9.

We shall fix L=L0+MK3+r3L=L_{0}+M_{K_{3}}+r_{3} and denote it by R0R_{0} for the rest of the paper. Also we shall define R1=C+CR0R_{1}=C+CR_{0} where C=C4.6(K1)C=C_{\ref{ladders are qi embedded}}(K_{1}). Thus we have the following.

Corollary 5.10.

(Decomposition of 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime})) There is a partition 0=t0<t1<<tn=l0=t_{0}<t_{1}<\cdots<t_{n}=l of [0,l][0,l] and K1K_{1}-qi sections Σi{\Sigma}_{i} passing through α(ti)\alpha(t_{i}), 0in0\leq i\leq n inside 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) such that the following holds.

  1. (1)

    Σ0=Σ,Σn=Σ{\Sigma}_{0}={\Sigma},{\Sigma}_{n}={\Sigma}^{\prime}.

  2. (2)

    For 0in20\leq i\leq n-2, Σi+1𝕃(Σi,Σ){\Sigma}_{i+1}\subset{\mathbb{L}}({\Sigma}_{i},{\Sigma}^{\prime}).

  3. (3)

    For 0in20\leq i\leq n-2 either (I) dh(Σi,Σi+1)=R0d_{h}({\Sigma}_{i},{\Sigma}_{i+1})=R_{0}, or (II) dh(Σi,Σi+1)>R0d_{h}({\Sigma}_{i},{\Sigma}_{i+1})>R_{0} and there is a K2K_{2}-qi section Σi{\Sigma}^{\prime}_{i} through α(ti+11)\alpha(t_{i+1}-1) inside 𝕃(Σi,Σi+1){\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1}) such that dh(Σi,Σi)<R1d_{h}({\Sigma}_{i},{\Sigma}^{\prime}_{i})<R_{1}.

    In either case dX(Σi,Σi+1)>2λ1+1d_{X}({\Sigma}_{i},{\Sigma}_{i+1})>2\lambda_{1}+1 and Σi,Σi+1{\Sigma}_{i},{\Sigma}_{i+1} are D1D_{1}-cobounded in XX.

  4. (4)

    dh(Σn1,Σn)R0d_{h}({\Sigma}_{n-1},{\Sigma}_{n})\leq R_{0}.

We note that the second part of (3) follows from (C1), (C2), (C3) above. However, a subladder 𝕃(Σi,Σi+1){\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1}) of 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) will be referred to as a type (I) subladder or a type (II) subladder according as dh(Σi,Σi+1)=R0d_{h}({\Sigma}_{i},{\Sigma}_{i+1})=R_{0} or dh(Σi,Σi+1)>R0d_{h}({\Sigma}_{i},{\Sigma}_{i+1})>R_{0} respectively.

Remark 15.

(1) We note that by the choice of R0,R1R_{0},R_{1} it follows that dY(ΣiY,Σi+1Y)>2λ1+1d_{Y}({\Sigma}_{i}\cap Y,{\Sigma}_{i+1}\cap Y)>2\lambda_{1}+1 and ΣiY,Σi+1Y{\Sigma}_{i}\cap Y,{\Sigma}_{i+1}\cap Y are D1D_{1}-cobounded in YY for 0in20\leq i\leq n-2.

(2) We shall use Σi{\Sigma}_{i} to mean qi sections in 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) exactly as in the corollary above for the rest of this section.

(3) Finally we note that Σn,Σn1{\Sigma}_{n},{\Sigma}_{n-1} need not be cobounded in general and the same remark applies to ΣnY,Σn1Y{\Sigma}_{n}\cap Y,{\Sigma}_{n-1}\cap Y.

Lemma 5.11.

Let Π:𝕃(Σ,Σ)[0,n]\Pi:{\mathbb{L}}({\Sigma},{\Sigma}^{\prime})\rightarrow[0,n] be any map that sends Σi{\Sigma}_{i} to i[0,n]i\in[0,n]\cap\mathbb{Z} and sends any point of 𝕃(Σi,Σi+1){ΣiΣi+1}{\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1})\setminus\{{\Sigma}_{i}\cup{\Sigma}_{i+1}\} to a point in (i,i+1)(i,i+1). Then the hypotheses of Proposition 2.33 are verified for both Π\Pi and its restriction 𝕃(Σ,Σ)Y[0,n]{\mathbb{L}}({\Sigma},{\Sigma}^{\prime})\cap Y\rightarrow[0,n].

Proof.

For both Π\Pi and its restriction to 𝕃(Σ,Σ)Y{\mathbb{L}}({\Sigma},{\Sigma}^{\prime})\cap Y, (P 0),(P 1)(P\,0),(P\,1) follow from (C2), (P 2)(P\,2) follows from Lemma 5.4, and (P 3)(P\,3) follows from (C4). (P 4)(P\,4) for Π\Pi follows from (C3) and for the restriction of Π\Pi to 𝕃(Σ,Σ)Y{\mathbb{L}}({\Sigma},{\Sigma}^{\prime})\cap Y from Remark 15(1).∎

Step 1(c): Joining y,yy,y^{\prime} inside 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime}). We now inductively define a finite sequence of points yiΣiy_{i}\in{\Sigma}_{i}, 0in+10\leq i\leq n+1 with y0=y,yn+1=yy_{0}=y,y_{n+1}=y^{\prime} such that each yiy_{i}, 1in1\leq i\leq n, is a uniform approximate nearest point projection of yi1y_{i-1} on Σi{\Sigma}_{i} in XX. We also define uniform quasigeodesics γi\gamma_{i} in XX joining yi,yi+1y_{i},y_{i+1}. The concatenation of these γi\gamma_{i}’s then forms a uniform quasigeodesic in XX joining y,yy,y^{\prime} by Proposition 2.33 and Lemma 5.11.

We define γn\gamma_{n} to be the lift of [π(yn),π(yn+1)][\pi(y_{n}),\pi(y_{n+1})] in Σ{\Sigma}^{\prime}.

Suppose y0,,yiy_{0},\ldots,y_{i} and γ0,,γi1\gamma_{0},\ldots,\gamma_{i-1} are already constructed, 0in20\leq i\leq n-2. We next explain how to define yi+1y_{i+1} and γi\gamma_{i}.

Case I. Suppose 𝕃i=𝕃(Σi,Σi+1){\mathbb{L}}_{i}={\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1}) is of type (I) i.e. dh(Σi,Σi+1)=R0d_{h}({\Sigma}_{i},{\Sigma}_{i+1})=R_{0} or i=n1i=n-1. Then, UR0(Σi,Σi+1)U_{R_{0}}({\Sigma}_{i},{\Sigma}_{i+1}) is non-empty. Let uiu_{i} be a nearest point projection of π(yi)\pi(y_{i}) on UR0(Σi,Σi+1)U_{R_{0}}({\Sigma}_{i},{\Sigma}_{i+1}). We define yi+1=Σi+1Fuiy_{i+1}={\Sigma}_{i+1}\cap F_{u_{i}}. Let αi\alpha_{i} be the lift of [π(yi),ui][\pi(y_{i}),u_{i}] in Σi{\Sigma}_{i}, and let σi\sigma_{i} be the subsegment of Fui𝕃iF_{u_{i}}\cap{\mathbb{L}}_{i} joining αi(ui)\alpha_{i}(u_{i}) and yi+1y_{i+1}. We define γi\gamma_{i} to be the concatenation of αi\alpha_{i} and σi\sigma_{i}. Then clearly γi\gamma_{i} is a (K1+R0)(K_{1}+R_{0})-quasigeodesic in XX. That yi+1y_{i+1} is a uniform approximate nearest point projection of yiy_{i} on Σi+1{\Sigma}_{i+1} follows from the following lemma.

Lemma 5.12.

Given K1K\geq 1 and RMKR\geq M_{K} there are constants ϵ5.12(K,R)\epsilon_{\ref{step 1(c)-1}}(K,R) and ϵ5.12(K,R)\epsilon^{\prime}_{\ref{step 1(c)-1}}(K,R) such that the following holds.

Suppose 𝒬1,𝒬2{\mathcal{Q}}_{1},{\mathcal{Q}}_{2} are two KK-qi sections and dh(𝒬1,𝒬2)Rd_{h}({\mathcal{Q}}_{1},{\mathcal{Q}}_{2})\leq R. Let x𝒬1x\in{\mathcal{Q}}_{1} and let U=UR(𝒬1,𝒬2)U=U_{R}({\mathcal{Q}}_{1},{\mathcal{Q}}_{2}). Suppose bb is a nearest point projection of π(x)\pi(x) on UU. Then 𝒬2Fb{\mathcal{Q}}_{2}\cap F_{b} is ϵ5.12(K,R)\epsilon_{\ref{step 1(c)-1}}(K,R)-approximate nearest point projection of xx on 𝒬2{\mathcal{Q}}_{2}.

If dh(𝒬1,𝒬2)MKd_{h}({\mathcal{Q}}_{1},{\mathcal{Q}}_{2})\geq M_{K} then for any bUb^{\prime}\in U the point 𝒬2Fb{\mathcal{Q}}_{2}\cap F_{b^{\prime}} is an ϵ5.12(K,R)\epsilon^{\prime}_{\ref{step 1(c)-1}}(K,R)-approximate nearest point projection of any point of 𝒬1{\mathcal{Q}}_{1} on 𝒬2{\mathcal{Q}}_{2}.

This lemma follows from Corollary 1.40 and Proposition 3.4 of [MS12] given that ladders are quasiconvex. However, we give an independent proof using the hyperbolicity of XX.

Proof.

Suppose x¯\bar{x} is a nearest point projection of xx on 𝒬2{\mathcal{Q}}_{2} and let x=𝒬2Fbx^{\prime}={\mathcal{Q}}_{2}\cap F_{b}. Let γxx\gamma_{xx^{\prime}} be the concatenation of the lift in 𝒬1{\mathcal{Q}}_{1} of any geodesic in BB joining π(x)\pi(x) to bb and any geodesic in FbF_{b} joining 𝒬1Fb{\mathcal{Q}}_{1}\cap F_{b} to 𝒬2Fb{\mathcal{Q}}_{2}\cap F_{b}. Clearly it is a (K+R)(K+R)-quasigeodesic in XX. Also by Lemma 2.25 the concatenation of any 11-quasigeodesics joining x,x¯x,\bar{x} and x¯,x\bar{x},x^{\prime} is a K2.25(δ,K,1,0)K_{\ref{subqc-elem}}(\delta,K,1,0)-quasigeodesic. Hence, by stability of quasigeodesics we have x¯ND(γi)\bar{x}\in N_{D}(\gamma_{i}) where D=D2.19(δ,K,K)D=D_{\ref{cor: stab-qg}}(\delta,K^{\prime},K^{\prime}) and K=max{K+R,K2.25(δ,K,1,0)}K^{\prime}=\max\{K+R,K_{\ref{subqc-elem}}(\delta,K,1,0)\}. This implies there is a point zγxxz\in\gamma_{xx^{\prime}} such that d(z,x¯)Dd(z,\bar{x})\leq D. If zFbγxxz\in F_{b}\cap\gamma_{xx^{\prime}} then d(x¯,x)D+Rd(\bar{x},x^{\prime})\leq D+R and hence xx^{\prime} is a (D+R)(D+R)-approximate nearest point projection of xx on 𝒬2{\mathcal{Q}}_{2}.

Suppose z𝒬1γxxz\in{\mathcal{Q}}_{1}\cap\gamma_{xx^{\prime}}. Then dπ(z)(z,𝒬2Fπ(z))R4.13(D,K)d_{\pi(z)}(z,{\mathcal{Q}}_{2}\cap F_{\pi(z)})\leq R_{\ref{distance from qi section}}(D,K). Hence, by Lemma 4.11 we have dB(π(z),b)D4.11(K,R)d_{B}(\pi(z),b)\leq D_{\ref{qc-level-set-new}}(K,R^{\prime}) where R=R4.13(D,K)/RR^{\prime}=R_{\ref{distance from qi section}}(D,K)/R. Therefore, d(x¯,x)d(x¯,z)+d(z,𝒬1Fb)+d(𝒬1Fb,x)D+(K+KD4.11(K,R))+Rd(\bar{x},x^{\prime})\leq d(\bar{x},z)+d(z,{\mathcal{Q}}_{1}\cap F_{b})+d({\mathcal{Q}}_{1}\cap F_{b},x^{\prime})\leq D+(K+KD_{\ref{qc-level-set-new}}(K,R^{\prime}))+R. Hence in this case xx^{\prime} is a (D+K+KD4.11(K,R)+R)(D+K+KD_{\ref{qc-level-set-new}}(K,R^{\prime})+R)-approximate nearest point projection of xx on 𝒬2{\mathcal{Q}}_{2}. We may set ϵ5.12(K,R)=D+K+KD4.11(K,R)+R\epsilon_{\ref{step 1(c)-1}}(K,R)=D+K+KD_{\ref{qc-level-set-new}}(K,R^{\prime})+R.

For the last part, we note that the diameter of UU is at most D4.11(K,R)D^{\prime}_{\ref{qc-level-set-new}}(K,R). Thus clearly ϵ5.12(K,R)=ϵ5.12(K,R)+K+KD4.11(K,R)\epsilon^{\prime}_{\ref{step 1(c)-1}}(K,R)=\epsilon_{\ref{step 1(c)-1}}(K,R)+K+KD^{\prime}_{\ref{qc-level-set-new}}(K,R) works. ∎

Case II. Suppose 𝕃i=𝕃(Σi,Σi+1){\mathbb{L}}_{i}={\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1}) is of type (II), i.e. dh(Σi,Σi+1)>R0d_{h}({\Sigma}_{i},{\Sigma}_{i+1})>R_{0}. In this case there exists a K2K_{2}-qi section Σi{\Sigma}^{\prime}_{i} inside 𝕃i=𝕃(Σi,Σi+1){\mathbb{L}}_{i}={\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1}) passing through α(ti+11)\alpha(t_{i+1}-1) such that dh(Σi,Σi)R1d_{h}({\Sigma}_{i},{\Sigma}^{\prime}_{i})\leq R_{1}. We thus use Case (I) twice as follows. First we project yiy_{i} on Σi{\Sigma}^{\prime}_{i}. Suppose the projection is yiy^{\prime}_{i}. Then we project yiy^{\prime}_{i} on Σi+1{\Sigma}_{i+1} which we call yi+1y_{i+1} and so on. Here are the details involved.

Let viv_{i} be a nearest point projection of π(yi)\pi(y_{i}) on UR1(Σi,Σi)U_{R_{1}}({\Sigma}_{i},{\Sigma}^{\prime}_{i}) and let wiw_{i} be a nearest point projection viv_{i} on UR1(Σi,Σi+1)U_{R_{1}}({\Sigma}^{\prime}_{i},{\Sigma}_{i+1}). Then yi+1=Σi+1Fwiy_{i+1}={\Sigma}_{i+1}\cap F_{w_{i}}. In this case we let αi\alpha_{i} denote the lift of [π(yi),vi][\pi(y_{i}),v_{i}] in Σi{\Sigma}_{i} and let βi\beta_{i} denote the lift of [vi,wi][v_{i},w_{i}] in Σi{\Sigma}^{\prime}_{i}. Then γi\gamma_{i} is the concatenation of the paths αi\alpha_{i}, [ΣiFvi,ΣiFvi]vi[{\Sigma}_{i}\cap F_{v_{i}},{\Sigma}^{\prime}_{i}\cap F_{v_{i}}]_{v_{i}}, βi\beta_{i} and [ΣiFwi,Σi+1Fwi]wi[{\Sigma}^{\prime}_{i}\cap F_{w_{i}},{\Sigma}_{i+1}\cap F_{w_{i}}]_{w_{i}}. That yi+1y_{i+1} is a uniform approximate nearest point projection of yiy_{i} on Σi+1{\Sigma}_{i+1} and that γi\gamma_{i} is a uniform quasigeodesic follow immediately from Lemma 5.12 and the last part of Proposition 2.33.

Remark 16.

We note that 𝕃(Σ,Σ)Y{\mathbb{L}}({\Sigma},{\Sigma}^{\prime})\cap Y is a ladder in YY formed by the qi sections ΣY{\Sigma}\cap Y and ΣY{\Sigma}^{\prime}\cap Y defined over AA. However, in this case the subladders 𝕃(Σi,Σi+1)Y{\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1})\cap Y may not be of type (I) or (II). Therefore, we cannot directly use the above procedure to construct a uniform quasigeodesic in YY joining y,yy,y^{\prime}.

Step 2: Modification of the path c(y,y)c(y,y^{\prime}).

In this step we shall construct a path c¯(y,y)\bar{c}(y,y^{\prime}) in YY joining y,yy,y^{\prime} by modifying c(y,y)c(y,y^{\prime}). For 0in0\leq i\leq n, let bib_{i} be a nearest point projection of π(yi)\pi(y_{i}) on AA and let y¯i=FbiΣi\bar{y}_{i}=F_{b_{i}}\cap{\Sigma}_{i}. We define a path γ¯iY\bar{\gamma}_{i}\subset Y joining the points y¯i,y¯i+1\bar{y}_{i},\bar{y}_{i+1} for 0in0\leq i\leq n. Finally the path c¯(y,y)\bar{c}(y,y^{\prime}) is defined to be the concatenation of these paths. The path γ¯n\bar{\gamma}_{n} is the lift of [π(yn+1),π(y¯n)]A[\pi(y_{n+1}),\pi(\bar{y}_{n})]_{A} in ΣY{\Sigma}^{\prime}\cap Y. The definition of γ¯i\bar{\gamma}_{i}, for 0in10\leq i\leq n-1, depends on the type of the subladder 𝕃i=𝕃(Σi,Σi+1){\mathbb{L}}_{i}={\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1}) given by Corollary 5.10(3).

Case 2(I): Suppose 𝕃i{\mathbb{L}}_{i} is of type (I) or i=n1i=n-1. Let α¯i\bar{\alpha}_{i} denote the lift of [bi,bi+1]A[b_{i},b_{i+1}]_{A} in Σi{\Sigma}_{i} starting at y¯i\bar{y}_{i}. The path γ¯i\bar{\gamma}_{i} is defined to be the concatenation of α¯i\bar{\alpha}_{i} and the fiber geodesic Fbi+1𝕃(Σi,Σi+1)F_{b_{i+1}}\cap{\mathbb{L}}({\Sigma}_{i},{\Sigma}_{i+1}).

Case 2(II): Suppose 𝕃i{\mathbb{L}}_{i} is of type (II). In this case, we apply Case 2(I) to each of the subladders 𝕃(Σi,Σi){\mathbb{L}}({\Sigma}_{i},{\Sigma}^{\prime}_{i}) and 𝕃(Σi,Σi+1){\mathbb{L}}({\Sigma}^{\prime}_{i},{\Sigma}_{i+1}). Let yiy^{\prime}_{i} be as defined in step 1(c). Let biAb^{\prime}_{i}\in A be a nearest point projection π(yi)\pi(y^{\prime}_{i}) on AA and y¯i=π1(bi)Σi\bar{y}^{\prime}_{i}=\pi^{-1}({b}^{\prime}_{i})\cap{\Sigma}^{\prime}_{i}. Next we connect y¯i,y¯i\bar{y}_{i},\bar{y}^{\prime}_{i} and y¯i,y¯i+1\bar{y}^{\prime}_{i},\bar{y}_{i+1} as in Case 2(I) inside the ladders 𝕃(ΣiY,ΣiY){\mathbb{L}}({\Sigma}_{i}\cap Y,{\Sigma}^{\prime}_{i}\cap Y) and 𝕃(ΣiY,Σi+1Y){\mathbb{L}}({\Sigma}^{\prime}_{i}\cap Y,{\Sigma}_{i+1}\cap Y) respectively. We shall denote by α¯i\bar{\alpha}_{i} and β¯i\bar{\beta}_{i} the lift of [bi,bi]A[b_{i},b^{\prime}_{i}]_{A} in ΣiY{\Sigma}_{i}\cap Y and [bi,bi+1]A[b^{\prime}_{i},b_{i+1}]_{A} in ΣiY{\Sigma}^{\prime}_{i}\cap Y respectively. The concatenation of the paths α¯i\bar{\alpha}_{i}, [ΣiFbi,ΣiFbi]bi𝕃(Σ,Σ)[{\Sigma}_{i}\cap F_{b^{\prime}_{i}},{\Sigma}^{\prime}_{i}\cap F_{b^{\prime}_{i}}]_{b^{\prime}_{i}}\subset{\mathbb{L}}({\Sigma},{\Sigma}^{\prime}), β¯i\bar{\beta}_{i} and [ΣiFbi+1,Σi+1Fbi+1]bi+1𝕃(Σ,Σ)[{\Sigma}^{\prime}_{i}\cap F_{b_{i+1}},{\Sigma}_{i+1}\cap F_{b_{i+1}}]_{b_{i+1}}\subset{\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) is defined to be γ¯i\bar{\gamma}_{i}.

Step 3: Proving that c¯(y,y)\bar{c}(y,y^{\prime}) is a uniform quasigeodesic in YY. To show that c¯(y,y)\bar{c}(y,y^{\prime}) is a quasigeodesic it is enough, by Proposition 2.33, to show that the paths γ¯i\bar{\gamma}_{i} are all uniform quasigeodesics in YY and that for 0in10\leq i\leq n-1, y¯i+1\bar{y}_{i+1} is an approximate nearest point projection of y¯i\bar{y}_{i} in Σi+1Y{\Sigma}_{i+1}\cap Y. The proof of this is broken into three cases depending on the type of the ladder 𝕃i{\mathbb{L}}_{i}. We start with the following lemma as a preparation for the proof.

The lemma below is true for any metric bundle that satisfies the hypotheses (H1)-(H4), (H33^{\prime}) although we are stating it for XX only. For instance, it is true for YY too.

Lemma 5.13.

Suppose bBb\in B, x,yFbx,y\in F_{b}. Suppose for all KK0K\geq K_{0} and RMKR\geq M_{K} there is a constant D=D(K,R)0D=D(K,R)\geq 0 such that for all x,y[x,y]bx^{\prime},y^{\prime}\in[x,y]_{b} and any two KK-qi sections 𝒬1\mathcal{Q}_{1} and 𝒬2\mathcal{Q}_{2} in XX passing through x,yx^{\prime},y^{\prime} respectively, either UR(𝒬1,𝒬2)=U_{R}(\mathcal{Q}_{1},\mathcal{Q}_{2})=\emptyset or dB(b,UR(𝒬1,𝒬2))Dd_{B}(b,U_{R}(\mathcal{Q}_{1},\mathcal{Q}_{2}))\leq D. Then the following hold:

  1. (1)

    [x,y]b[x,y]_{b} is a λ5.13\lambda_{\ref{Y-path 1.1}}-quasigeodesic in XX where λ5.13\lambda_{\ref{Y-path 1.1}} depends on the function DD (and the parameters of the metric bundle).

  2. (2)

    If 𝒬\mathcal{Q} and 𝒬\mathcal{Q}^{\prime} are two KK-qi sections passing through x,yx,y respectively then xx is a uniform approximate nearest point projection of yy on 𝒬{\mathcal{Q}} and yy is a uniform approximate nearest point projection of xx on 𝒬{\mathcal{Q}}^{\prime}.

Proof.

(1) Since the arc length parametrization of [x,y]b[x,y]_{b} is a uniform proper embedding, by Lemma 2.5 it is enough to show that [x,y]b[x,y]_{b} is uniformly close to a geodesic in XX joining x,yx,y.

Claim: Suppose Σx,Σy{\Sigma}_{x},{\Sigma}_{y} are two K0K_{0}-qi sections passing through x,yx,y respectively. Given any z[x,y]bz\in[x,y]_{b} and any K1K_{1}-qi section Σz{\Sigma}_{z} passing through zz contained in the ladder 𝕃(Σx,Σy){\mathbb{L}}({\Sigma}_{x},{\Sigma}_{y}) the nearest point projection of xx on Σz{\Sigma}_{z} is uniformly close to zz.

We note that once the claim is proved then applying Proposition 2.33 to the ladder 𝕃(Σx,Σy)=𝕃(Σx,Σz)𝕃(Σz,Σy){\mathbb{L}}({\Sigma}_{x},{\Sigma}_{y})={\mathbb{L}}({\Sigma}_{x},{\Sigma}_{z})\cup{\mathbb{L}}({\Sigma}_{z},{\Sigma}_{y}) it follows that zz is uniformly close to a geodesic joining x,yx,y. From this (1) follows immediately.

Proof of the claim: First suppose UMK1(Σx,Σz)U_{M_{K_{1}}}({\Sigma}_{x},{\Sigma}_{z})\neq\emptyset. Then we can find a uniform approximate nearest point projection of xx on Σz{\Sigma}_{z} using Step 1(c), Case I and Lemma 5.12 above which is uniformly close to zz by hypothesis.

Now suppose UMK1(Σx,Σz)=U_{M_{K_{1}}}({\Sigma}_{x},{\Sigma}_{z})=\emptyset. Let αzx:[0,l]Fb\alpha_{zx}:[0,l]\rightarrow F_{b} be the unit speed parametrization of the geodesic 𝕃(Σx,Σz)Fb{\mathbb{L}}({\Sigma}_{x},{\Sigma}_{z})\cap F_{b} joining zz to xx. By Corollary 5.10 there is a K2K_{2}-qi section Σz{\Sigma}_{z^{\prime}} contained in the ladder 𝕃(Σx,Σz){\mathbb{L}}({\Sigma}_{x},{\Sigma}_{z}) passing through z=αzx(t)z^{\prime}=\alpha_{zx}(t) for some t[0,l]t\in[0,l] such that 𝕃(Σz,Σz){\mathbb{L}}({\Sigma}_{z},{\Sigma}_{z^{\prime}}) is a K2K_{2}-ladder of type (I) or (II). Let xx^{\prime} be a nearest point projection of xx on Σz{\Sigma}_{z^{\prime}}. By the last part of Proposition 2.33 applied to 𝕃(Σx,Σz){\mathbb{L}}({\Sigma}_{x},{\Sigma}_{z}), it is enough to find a uniform approximate nearest point projection of xx^{\prime} on Σz{\Sigma}_{z} which is also uniformly close to zz. However, in this case Σz,Σz{\Sigma}_{z},{\Sigma}_{z^{\prime}} are D2D_{2}-cobounded. Hence it is enough to find a uniform approximate nearest point projection of zz^{\prime} on Σz{\Sigma}_{z} which is uniformly close to zz. The proof of this is broken into two cases as follows.

(I) Suppose dh(Σz,Σz)=R0d_{h}({\Sigma}_{z},{\Sigma}_{z^{\prime}})=R_{0}. By the last part of Lemma 5.12 if vUR0(Σz,Σz)v\in U_{R_{0}}({\Sigma}_{z},{\Sigma}_{z^{\prime}}) then FvΣzF_{v}\cap{\Sigma}_{z} is a uniform approximate nearest point projection of any point of Σz{\Sigma}_{z^{\prime}}. Since dA(b,v)d_{A}(b,v) is uniformly small by hypothesis, d(z,FvΣz)d(z,F_{v}\cap{\Sigma}_{z}) is also uniformly small.

(II) Suppose dh(Σz,Σz)>R0d_{h}({\Sigma}_{z},{\Sigma}_{z^{\prime}})>R_{0}. Then there is a K3K_{3}-qi section Σz′′{\Sigma}_{z^{\prime\prime}} in 𝕃(Σz,Σz){\mathbb{L}}({\Sigma}_{z},{\Sigma}_{z^{\prime}}) passing through z′′=αzx(t1)z^{\prime\prime}=\alpha_{zx}(t-1) such that UR0(Σz,Σz′′)U_{R_{0}}({\Sigma}_{z},{\Sigma}_{z^{\prime\prime}})\neq\emptyset. Let vv^{\prime} be a nearest point projection of bb on UR0(Σz,Σz′′)U_{R_{0}}({\Sigma}_{z},{\Sigma}_{z^{\prime\prime}}). Then by hypothesis d(b,v)d(b,v^{\prime}) is uniformly small whence d(z,FvΣz)d(z,F_{v^{\prime}}\cap{\Sigma}_{z}) is uniformly small. Also by Lemma 5.12 the point ΣzFv{\Sigma}_{z}\cap F_{v^{\prime}} is a uniform approximate nearest point projection of z′′z^{\prime\prime} on Σz{\Sigma}_{z}. It follows that zz is a uniform approximate nearest point projection of z′′z^{\prime\prime} on Σz{\Sigma}_{z}. Finally, since d(z,z′′)1d(z^{\prime},z^{\prime\prime})\leq 1, zz is a uniform approximate nearest point projection of zz^{\prime}.

(2) We shall prove only the first statement since the proof of the second would be an exact copy. Suppose x1𝒬x_{1}\in{\mathcal{Q}} is a nearest point projection of yy on 𝒬{\mathcal{Q}}. Consider the KK-qi section over [b,π(x1)][b,\pi(x_{1})] contained in 𝒬{\mathcal{Q}}. This is a KK-quasigeodesic of XX joining x,x1x,x_{1}. Since 𝒬{\mathcal{Q}} is a KK-qi section, by stability of quasigeodesics it is D2.17(δ,K,K)D_{\ref{stab-qg}}(\delta,K,K)-quasiconvex in XX. Hence by Lemma 2.25 the concatenation of this quasigeodesic with a geodesic in XX joining yy to x1x_{1} is a K2.25(δ,K~,K,0)K_{\ref{subqc-elem}}(\delta,\tilde{K},K,0)-quasigeodesic where K~=D2.17(δ,K,K)\tilde{K}=D_{\ref{stab-qg}}(\delta,K,K). Let k=max{K~,λ5.13}k^{\prime}=\max\{\tilde{K},\lambda_{\ref{Y-path 1.1}}\}. Since [x,y]b[x,y]_{b} is a λ5.13\lambda_{\ref{Y-path 1.1}}-quasigeodesic, by stability of quasigeodesics we have x1N2D([x,y]b)x_{1}\in N_{2D^{\prime}}([x,y]_{b}) where D=D2.17(δ,k,k)D^{\prime}=D_{\ref{stab-qg}}(\delta,k^{\prime},k^{\prime}). Suppose z[x,y]bz\in[x,y]_{b} be such that d(x1,z)2Dd(x_{1},z)\leq 2D^{\prime}. Then dB(π(x1),π(z))=dB(π(x1),b)2Dd_{B}(\pi(x_{1}),\pi(z))=d_{B}(\pi(x_{1}),b)\leq 2D^{\prime}. Hence, d(x,x1)K+2DKd(x,x_{1})\leq K+2D^{\prime}K. Thus xx is a (K+2DK)(K+2D^{\prime}K)-approximate nearest point projection of yy on 𝒬{\mathcal{Q}}. ∎

Remark 17.

The proof of the first part of the above lemma uses the hypothesis for KK3K\leq K_{3} only whereas the proof of the second part follows directly from the statement of the first part and is independent of the hypotheses of the lemma.

The following lemma is actually a trivial consequence of flaring ( Lemma 4.13) and it is going to be used in the next two lemmas following it.

Lemma 5.14.

Given R0,K,K1R\geq 0,K,K^{\prime}\geq 1 and RMKR^{\prime}\geq M_{K^{\prime}} there is a constant R5.14(R,R,K,K)R_{\ref{Y-path 1.2}}(R,R^{\prime},K,K^{\prime}) and D5.14(R,R,K,K)D_{\ref{Y-path 1.2}}(R,R^{\prime},K,K^{\prime}) such that the following holds.

Suppose uBu\in B and PA(u)=bP_{A}(u)=b, where PA:BAP_{A}:B\rightarrow A is a nearest point projection map. Suppose x,yFbx,y\in F_{b} and let γx,γy\gamma_{x},\gamma_{y} be two KK-qi sections over [u,b][u,b]. Let 𝒬1,𝒬2{\mathcal{Q}}_{1},{\mathcal{Q}}_{2} be two KK^{\prime}-qi sections over AA in YY and U=UR(𝒬1,𝒬2)U=U_{R^{\prime}}({\mathcal{Q}}_{1},{\mathcal{Q}}_{2}). If du(γx(u),γy(u))Rd_{u}(\gamma_{x}(u),\gamma_{y}(u))\leq R and UU\neq\emptyset, then db(x,y)R5.14(R,R,K,K)d_{b}(x,y)\leq R_{\ref{Y-path 1.2}}(R,R^{\prime},K,K^{\prime}) and dA(b,U)D5.14(R,R,K,K)d_{A}(b,U)\leq D_{\ref{Y-path 1.2}}(R,R^{\prime},K,K^{\prime}).

Proof.

Suppose UU\neq\emptyset and du(γx(u),γy(y))Rd_{u}(\gamma_{x}(u),\gamma_{y}(y))\leq R. Let bUMK(𝒬1,𝒬2)b^{\prime}\in U_{M_{K^{\prime}}}({\mathcal{Q}}_{1},{\mathcal{Q}}_{2}) be any point and let [b,b][b,b^{\prime}] denote a geodesic in AA joining b,bb,b^{\prime}. Then the concatenation [u,b][b,b][u,b]*[b,b^{\prime}] is a K2.25(δ0,k0,k,0)K_{\ref{subqc-elem}}(\delta_{0},k_{0},k,0)-quasigeodesic in BB by Lemma 2.25 since AA is kk-qi embedded and k0k_{0}-quasiconvex. Concatenation of γx,γy\gamma_{x},\gamma_{y} with the qi sections over [b,b][b,b^{\prime}] contained in 𝒬1,𝒬2{\mathcal{Q}}_{1},{\mathcal{Q}}_{2} respectively defines max{K,K}\max\{K,K^{\prime}\}-qi sections over [u,b][b,b][u,b]*[b,b^{\prime}] passing through x,yx,y respectively. Let k=K2.25(δ0,k0,k,0)k^{\prime}=K_{\ref{subqc-elem}}(\delta_{0},k_{0},k,0) and k′′=max{K,K}k^{\prime\prime}=\max\{K,K^{\prime}\}. Then by Lemma 2.3 these qi sections are (kk′′,k′′k+k′′)(k^{\prime}k^{\prime\prime},k^{\prime\prime}k^{\prime}+k^{\prime\prime})-quasigeodesics in XX. Since XX is δ\delta-hyperbolic and d(γx(u),γy(u))Rd(\gamma_{x}(u),\gamma_{y}(u))\leq R and d(𝒬1Fb,𝒬2Fb)Rd({\mathcal{Q}}_{1}\cap F_{b^{\prime}},{\mathcal{Q}}_{2}\cap F_{b^{\prime}})\leq R^{\prime}, by Corollary 2.21 xx is contained in the D:=(R+R+2D2.20(δ,kk′′,kk′′+k′′))D^{\prime}:=(R+R^{\prime}+2D_{\ref{slim iff gromov}}(\delta,k^{\prime}k^{\prime\prime},k^{\prime}k^{\prime\prime}+k^{\prime\prime}))-neighborhood of the qi section over [u,b][b,b][u,b]*[b,b^{\prime}] passing through yy. Applying Lemma 4.13 to the restriction bundles over [u,b][u,b] and [b,b][b,b^{\prime}] we have db(x,y)R1d_{b}(x,y)\leq R^{\prime}_{1} where R1=R4.13(D,K)R^{\prime}_{1}=R_{\ref{distance from qi section}}(D^{\prime},K). Hence, we can take R5.14(R,R,K,K)=R1R_{\ref{Y-path 1.2}}(R,R^{\prime},K,K^{\prime})=R^{\prime}_{1}. Finally by Lemma 4.11 dA(b,U)D4.11(K,R1/MK)d_{A}(b,U)\leq D_{\ref{qc-level-set-new}}(K^{\prime},R^{\prime}_{1}/M_{K^{\prime}}). This completes the proof by taking D5.14(R,R,K,K)=D4.11(K,R1/MK)D_{\ref{Y-path 1.2}}(R,R^{\prime},K,K^{\prime})=D_{\ref{qc-level-set-new}}(K^{\prime},R^{\prime}_{1}/M_{K^{\prime}}). ∎

We recall that the paths c¯(y,y)\bar{c}(y,y^{\prime}) were constructed from c(y,y)c(y,y^{\prime}) by replacing parts of c(y,y)c(y,y^{\prime}) by some fiber geodesic segments. The main aim of the following three lemmas is to proving that these fiber geodesic segments are uniform quasigeodesics in YY. Depending on how the corresponding subladders of XX intersect YY we have three scenarios and hence we divided the proof into three lemmas.

Lemma 5.15.

Given KK0K\geq K_{0} and RMKR\geq M_{K} there are constants K5.15=K5.15(K,R)K_{\ref{Y-path 1.3}}=K_{\ref{Y-path 1.3}}(K,R), ϵ5.15=ϵ5.15(K,R)\epsilon_{\ref{Y-path 1.3}}=\epsilon_{\ref{Y-path 1.3}}(K,R) and D5.15=D5.15(K,R)D_{\ref{Y-path 1.3}}=D_{\ref{Y-path 1.3}}(K,R) such that the following holds.

Suppose 𝒬,𝒬{\mathcal{Q}},{\mathcal{Q}}^{\prime} are two KK-qi sections in XX and dh(𝒬,𝒬)Rd_{h}({\mathcal{Q}},{\mathcal{Q}}^{\prime})\leq R in XX. Let U=UR(𝒬,𝒬)U=U_{R}({\mathcal{Q}},{\mathcal{Q}}^{\prime}). Suppose dh(𝒬Y,𝒬Y)Rd_{h}({\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y)\geq R in YY. Then the following hold.

  1. (1)

    The projection of UU on AA is of diameter at most D5.15D_{\ref{Y-path 1.3}}.

  2. (2)

    For any bPA(U)b\in P_{A}(U), Fb𝕃(𝒬,𝒬)F_{b}\cap{\mathbb{L}}({\mathcal{Q}},{\mathcal{Q}}^{\prime}) is a K5.15K_{\ref{Y-path 1.3}}-quasigeodesic in YY; moreover, Fb𝒬F_{b}\cap{\mathcal{Q}} is an ϵ5.15\epsilon_{\ref{Y-path 1.3}}-approximate nearest point projection of any point of 𝒬{\mathcal{Q}}^{\prime} on 𝒬{\mathcal{Q}} and vice versa.

Proof.

(1) We know that AA is k0k_{0}-quasiconvex in BB. By Lemma 4.11 UU is K4.11(K)K_{\ref{qc-level-set-new}}(K)-quasiconvex in BB. Let λ=max{k0,K4.11(K)}\lambda^{\prime}=\max\{k_{0},K_{\ref{qc-level-set-new}}(K)\}. Suppose PA:BAP_{A}:B\rightarrow A is a nearest point projection map and a,aPA(U)a,a^{\prime}\in P_{A}(U) with dB(a,a)D2.28(δ,λ,0)d_{B}(a,a^{\prime})\geq D_{\ref{cor: lip proj}}(\delta,\lambda^{\prime},0). Then there are u,uUu,u^{\prime}\in U such that dB(a,u)R2.28(δ,λ,0)d_{B}(a,u)\leq R_{\ref{cor: lip proj}}(\delta,\lambda^{\prime},0) and dB(a,u)R2.28(δ,λ,0)d_{B}(a^{\prime},u^{\prime})\leq R_{\ref{cor: lip proj}}(\delta,\lambda^{\prime},0). Let D=R2.28(δ,λ,0)D=R_{\ref{cor: lip proj}}(\delta,\lambda^{\prime},0). We know du(𝒬Fu,𝒬Fu)Rd_{u}({\mathcal{Q}}\cap F_{u},{\mathcal{Q}}^{\prime}\cap F_{u})\leq R. Hence by the bounded flaring condition we have da(𝒬Fa,𝒬Fa)μK(D)Rd_{a}({\mathcal{Q}}\cap F_{a},{\mathcal{Q}}^{\prime}\cap F_{a})\leq\mu_{K}(D)R. Similarly da(𝒬Fa,𝒬Fa)μK(D)Rd_{a^{\prime}}({\mathcal{Q}}\cap F_{a^{\prime}},{\mathcal{Q}}^{\prime}\cap F_{a^{\prime}})\leq\mu_{K}(D)R. Let R1=μK(D)RR_{1}=\mu_{K}(D)R. Thus, a,aUR1(𝒬Y,𝒬Y)a,a^{\prime}\in U_{R_{1}}({\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y). Since R1MKR_{1}\geq M_{K}, by Lemma 4.11 we have diam(UR1(𝒬Y,𝒬Y))D4.11(K,R1)diam(U_{R_{1}}({\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y))\leq D_{\ref{qc-level-set-new}}(K,R_{1}). This proves (1). In fact, we can take D5.15=max{D2.28(δ,λ,0),D4.11(K,R1)}D_{\ref{Y-path 1.3}}=\max\{D_{\ref{cor: lip proj}}(\delta,\lambda^{\prime},0),D_{\ref{qc-level-set-new}}(K,R_{1})\}.

We derive (2) from Lemma 5.13 as follows. Let uUu\in U be such that PA(u)=bP_{A}(u)=b and let x,yFb𝕃(𝒬,𝒬)x,y\in F_{b}\cap{\mathbb{L}}({\mathcal{Q}},{\mathcal{Q}}^{\prime}). Suppose 𝒬1,𝒬1{\mathcal{Q}}_{1},{\mathcal{Q}}^{\prime}_{1} are two KK^{\prime}-qi sections in YY passing through x,yx,y respectively and U=UMK(𝒬1,𝒬1)U^{\prime}=U_{M_{K^{\prime}}}({\mathcal{Q}}_{1},{\mathcal{Q}}^{\prime}_{1}). Suppose UU^{\prime}\neq\emptyset. Consider the restriction ZZ of the bundle XX on [u,b]B[u,b]\subset B. In this bundle 𝒬Z,𝒬Z{\mathcal{Q}}\cap Z,{\mathcal{Q}}^{\prime}\cap Z are KK-qi sections. By Proposition 4.6(3) there are (1+2K0)C4.6(K)(1+2K_{0})C_{\ref{ladders are qi embedded}}(K)-qi sections over ubub contained in the ladder 𝕃(𝒬Z,𝒬Z){\mathbb{L}}({\mathcal{Q}}\cap Z,{\mathcal{Q}}^{\prime}\cap Z) passing through x,yx,y. Call them γx,γy\gamma_{x},\gamma_{y} respectively. We note that d(γx(u),γy(u))Rd(\gamma_{x}(u),\gamma_{y}(u))\leq R. Now applying Lemma 5.14 we know that dB(b,U)d_{B}(b,U^{\prime}) is uniformly small. This verifies the hypothesis of Lemma 5.13. Thus 𝒬Fb{\mathcal{Q}}\cap F_{b} is a uniform approximate nearest point projection of 𝒬Fb{\mathcal{Q}}^{\prime}\cap F_{b} on 𝒬{\mathcal{Q}}. Since dh(𝒬Y,𝒬Y)RMKd_{h}({\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y)\geq R\geq M_{K} the qi sections 𝒬Y,𝒬Y{\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y are uniformly cobounded by Lemma 5.3. This shows that 𝒬Fb{\mathcal{Q}}\cap F_{b} is a uniform approximate nearest point projection of any point of 𝒬{\mathcal{Q}}^{\prime} on 𝒬{\mathcal{Q}}. That 𝒬Fb{\mathcal{Q}}^{\prime}\cap F_{b} is a uniform approximate nearest point projection of any point of 𝒬{\mathcal{Q}} on 𝒬{\mathcal{Q}}^{\prime} is similar and hence we skip it. ∎

Lemma 5.16.

Given D0D\geq 0, KK0K\geq K_{0} and RMKR\geq M_{K} there are constants K5.16=K5.16(D,K,R)K_{\ref{Y-path 1.4}}=K_{\ref{Y-path 1.4}}(D,K,R) ϵ5.16=ϵ5.16(D,K,R)\epsilon_{\ref{Y-path 1.4}}=\epsilon_{\ref{Y-path 1.4}}(D,K,R) and D5.16=D5.16(D,K,R)D_{\ref{Y-path 1.4}}=D_{\ref{Y-path 1.4}}(D,K,R) such that the following holds.

Suppose 𝒬,𝒬{\mathcal{Q}},{\mathcal{Q}}^{\prime} are two KK-qi sections in XX and dh(𝒬,𝒬)Rd_{h}({\mathcal{Q}},{\mathcal{Q}}^{\prime})\leq R in XX. Let U=UR(𝒬,𝒬)U=U_{R}({\mathcal{Q}},{\mathcal{Q}}^{\prime}). Suppose UU\neq\emptyset and diam(U)Ddiam(U)\leq D. Then the following holds.

  1. (1)

    diam(PA(U))D5.16diam(P_{A}(U))\leq D_{\ref{Y-path 1.4}}.

  2. (2)

    For any bPA(U)b\in P_{A}(U), Fb𝕃(𝒬,𝒬)F_{b}\cap{\mathbb{L}}({\mathcal{Q}},{\mathcal{Q}}^{\prime}) is a K5.16K_{\ref{Y-path 1.4}}-quasigeodesic in YY.

  3. (3)

    Fb𝒬F_{b}\cap{\mathcal{Q}} is an ϵ5.16\epsilon_{\ref{Y-path 1.4}}-approximate nearest point projection of any point of 𝒬{\mathcal{Q}}^{\prime} on 𝒬{\mathcal{Q}} and vice versa.

Proof.

(1) Since BB is δ0\delta_{0}-hyperbolic and AA is k0k_{0}-quasiconvex in BB any nearest point projection map PA:BAP_{A}:B\rightarrow A is coarsely L:=L2.28(δ0,k0,0)L:=L_{\ref{cor: lip proj}}(\delta_{0},k_{0},0)-Lipschitz. Hence, diam(PA(U))L+DLdiam(P_{A}(U))\leq L+DL.

We can derive (2), (3) from Lemma 5.13 and the hypotheses of Lemma 5.13 can be verified using Lemma 5.14. The proof is an exact copy of the proof of Lemma 5.15(2),(3). Hence we omit it. The only part that requires explanation is why 𝒬Y{\mathcal{Q}}\cap Y, 𝒬Y{\mathcal{Q}}^{\prime}\cap Y are uniformly cobounded in YY. If dh(𝒬Y,𝒬Y)>Rd_{h}({\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y)>R then we are done by Lemma 5.3. Suppose this is not the case. Then by the hypothesis diam(UR(𝒬Y,𝒬Y))k(k+D)diam(U_{R}({\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y))\leq k(k+D) since AA is kk-qi embedded in BB. Then we are done by the first part of Lemma 5.12. ∎

Lemma 5.17.

Given KK0K\geq K_{0} and RMKR\geq M_{K} there is a constant D5.17=D5.17(K,R)D_{\ref{Y-path 1.5}}=D_{\ref{Y-path 1.5}}(K,R) such that the following holds.

Suppose 𝒬,𝒬{\mathcal{Q}},{\mathcal{Q}}^{\prime} are two KK-qi sections in XX and dh(𝒬Y,𝒬Y)Rd_{h}({\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y)\leq R. Let U=UR(𝒬,𝒬)U=U_{R}({\mathcal{Q}},{\mathcal{Q}}^{\prime}). Then the following holds. For any bPA(U)b\in P_{A}(U), db(𝒬Fb,𝒬Fb)D5.17d_{b}({\mathcal{Q}}\cap F_{b},{\mathcal{Q}}^{\prime}\cap F_{b})\leq D_{\ref{Y-path 1.5}}.

Proof.

Suppose uUu\in U and PA(u)=bP_{A}(u)=b. If uAu\in A then b=ub=u and db(𝒬Fb,𝒬Fb)Rd_{b}({\mathcal{Q}}\cap F_{b},{\mathcal{Q}}^{\prime}\cap F_{b})\leq R. Suppose uAu\not\in A. We note that UY=U(𝒬Y,𝒬Y)U\cap Y=U({\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y)\neq\emptyset. Let vU(𝒬Y,𝒬Y)v\in U({\mathcal{Q}}\cap Y,{\mathcal{Q}}^{\prime}\cap Y). Then by Lemma 2.25 [u,b][b,v][u,b]*[b,v] is a K2.25(δ0,k0,1,0)K_{\ref{subqc-elem}}(\delta_{0},k_{0},1,0)-quasigeodesic in BB. Since UU is K4.11(K)K_{\ref{qc-level-set-new}}(K)-quasiconvex in BB. Let k=K2.25(δ0,k0,1,0)k^{\prime}=K_{\ref{subqc-elem}}(\delta_{0},k_{0},1,0). Hence, by Lemma 2.17, bND(U)b\in N_{D}(U) where D=D2.17(δ0,k,k)+K4.11(K)D=D_{\ref{stab-qg}}(\delta_{0},k^{\prime},k^{\prime})+K_{\ref{qc-level-set-new}}(K). Finally by the bounded flaring db(𝒬Fb,𝒬Fb)Rmax{1,μK(D)}d_{b}({\mathcal{Q}}\cap F_{b},{\mathcal{Q}}^{\prime}\cap F_{b})\leq R\max\{1,\mu_{K}(D)\}. Hence we can take D5.17=Rmax{1,μK(D)}D_{\ref{Y-path 1.5}}=R\max\{1,\mu_{K}(D)\}. ∎

Finally, we are ready to finish the proof of step 3.

Lemma 5.18.

For 0in10\leq i\leq n-1 we have the following.

  1. (1)

    y¯i+1\bar{y}_{i+1} is a uniform approximate nearest point projection of y¯i\bar{y}_{i} on Σi+1Y{\Sigma}_{i+1}\cap Y.

  2. (2)

    γ¯i\bar{\gamma}_{i} is a uniform quasigeodesic in YY.

Proof.

The proof is broken into three cases depending on the type of 𝕃i{\mathbb{L}}_{i}.

Case 1: in2i\leq n-2 and 𝕃i{\mathbb{L}}_{i} is of type (I): By Corollary 4.11 UR0(Σi,Σi+1)U_{R_{0}}({\Sigma}_{i},{\Sigma}_{i+1}) has uniformly small diameter. Hence by Lemma 5.16(2) [ΣiFbi+1,Σi+1Fbi+1]bi+1[{\Sigma}_{i}\cap F_{b_{i+1}},{\Sigma}_{i+1}\cap F_{b_{i+1}}]_{b_{i+1}} is a uniform quasigeodesic in YY. By the part (3) of the same lemma Σi+1Fbi+1{\Sigma}_{i+1}\cap F_{b_{i+1}} is a uniform approximate nearest point projection of ΣiFbi+1{\Sigma}_{i}\cap F_{b_{i+1}} on Σi+1Y{\Sigma}_{i+1}\cap Y and ΣiFbi+1{\Sigma}_{i}\cap F_{b_{i+1}} is a uniform approximate nearest point projection of Σi+1Fbi+1{\Sigma}_{i+1}\cap F_{b_{i+1}} on ΣiY{\Sigma}_{i}\cap Y in YY. Hence the second part of the lemma follows, in this case, by Lemma 2.25.

Case 2: in2i\leq n-2 and 𝕃i{\mathbb{L}}_{i} is of type (II): Suppose 𝕃i{\mathbb{L}}_{i} is a ladder of type (II). In this case, it is enough, by Proposition 2.33, to show the following two statements (2)(2^{\prime}) and (2′′)(2^{\prime\prime}):

(2)(2^{\prime}): y¯i\bar{y}^{\prime}_{i} is a uniform approximate nearest point projection of y¯i\bar{y}_{i} on ΣiY{\Sigma}^{\prime}_{i}\cap Y in YY and the concatenation of α¯i\bar{\alpha}_{i} and the fiber geodesic [ΣiFbi,ΣiFbi]bi[{\Sigma}_{i}\cap F_{b^{\prime}_{i}},{\Sigma}^{\prime}_{i}\cap F_{b^{\prime}_{i}}]_{b^{\prime}_{i}} is a uniform quasigeodesic in YY.

We know that dh(Σi,Σi)R1d_{h}({\Sigma}_{i},{\Sigma}^{\prime}_{i})\leq R_{1}. Depending on the nature of dh(ΣiY,ΣiY)d_{h}({\Sigma}_{i}\cap Y,{\Sigma}^{\prime}_{i}\cap Y) the proof of (2)(2^{\prime}) is broken into the following two cases.

Case (2)(i)(2^{\prime})(i): Suppose dh(ΣiY,ΣiY)R1d_{h}({\Sigma}_{i}\cap Y,{\Sigma}^{\prime}_{i}\cap Y)\leq R_{1}. In this case dbi(ΣiFbi,ΣiFbi)d_{b^{\prime}_{i}}({\Sigma}_{i}\cap F_{b^{\prime}_{i}},{\Sigma}^{\prime}_{i}\cap F_{b^{\prime}_{i}}) is uniformly small by Lemma 5.17. By Lemma 5.12 if bi′′b^{\prime\prime}_{i} is a nearest point projection of π(y¯i)\pi(\bar{y}_{i}) on UR1(ΣiY,ΣiY)U_{R_{1}}({\Sigma}_{i}\cap Y,{\Sigma}^{\prime}_{i}\cap Y) then Fbi′′ΣiF_{b^{\prime\prime}_{i}}\cap{\Sigma}^{\prime}_{i} is a uniform approximate nearest point projection of y¯i\bar{y}_{i} on ΣiY{\Sigma}^{\prime}_{i}\cap Y in YY. Thus it is enough to show that dB(bi′′,bi)d_{B}(b^{\prime\prime}_{i},b^{\prime}_{i}) uniformly bounded to prove that y¯i\bar{y}^{\prime}_{i} is a uniform approximate nearest point projection of y¯i\bar{y}_{i} on ΣiY{\Sigma}^{\prime}_{i}\cap Y in YY. Then since ΣiY{\Sigma}_{i}\cap Y is K1K^{\prime}_{1}-qi section in YY and dbi(ΣiFbi,ΣiFbi)d_{b^{\prime}_{i}}({\Sigma}_{i}\cap F_{b^{\prime}_{i}},{\Sigma}^{\prime}_{i}\cap F_{b^{\prime}_{i}}) is uniformly small it will follow that the concatenation of α¯i\bar{\alpha}_{i} and the fiber geodesic [ΣiFbi,ΣiFbi]bi[{\Sigma}_{i}\cap F_{b^{\prime}_{i}},{\Sigma}^{\prime}_{i}\cap F_{b^{\prime}_{i}}]_{b^{\prime}_{i}} is a uniform quasigeodesic in YY.

That dB(bi′′,bi)d_{B}(b^{\prime\prime}_{i},b^{\prime}_{i}) uniformly bounded is proved as follows. Let U=UR1(Σi,Σi)U=U_{R_{1}}({\Sigma}_{i},{\Sigma}^{\prime}_{i}), V=UA=UR1(ΣiY,ΣiY)V=U\cap A=U_{R_{1}}({\Sigma}_{i}\cap Y,{\Sigma}^{\prime}_{i}\cap Y). Since BB is δ0\delta_{0}-hyperbolic, AA is kk-qi embedded in BB and VV is λ2\lambda_{2}-quasiconvex in AA, VV is K2.30(δ0,k,λ2)K_{\ref{qc in subspace}}(\delta_{0},k,\lambda_{2})-quasiconvex in BB. Let k=max{λ2,k0,K4.11(K2),K2.30(δ0,k,λ2)}k^{\prime}=\max\{\lambda_{2},k_{0},K_{\ref{qc-level-set-new}}(K_{2}),K_{\ref{qc in subspace}}(\delta_{0},k,\lambda_{2})\}. Then A,U,VA,U,V are all kk^{\prime}-quasiconvex in BB. By the definitions of yiy_{i}’s we know that π(yi)\pi(y^{\prime}_{i}) is the nearest point projection of π(yi)\pi(y_{i}) on UU. Let b¯i\bar{b}^{\prime}_{i} be a nearest point projection of π(yi)\pi(y^{\prime}_{i}) on VV. Also bi=π(y¯i)b^{\prime}_{i}=\pi(\bar{y}^{\prime}_{i}) is the nearest point projection of π(yi)\pi(y^{\prime}_{i}) on AA. On the other hand, bi=π(y¯i)b_{i}=\pi(\bar{y}_{i}) is a nearest point projection of π(yi)\pi(y_{i}) on AA and bi′′b^{\prime\prime}_{i} is the nearest point projection of bib_{i} on VV. Therefore, dB(bi′′,b¯i)2D2.27(δ0,k,0)d_{B}(b^{\prime\prime}_{i},\bar{b}^{\prime}_{i})\leq 2D_{\ref{nested qc sets}}(\delta_{0},k^{\prime},0) by Corollary 2.27.

Now, by Lemma 5.17 dbi(ΣiFbi,ΣiFbi)D5.17(K2,R1)d_{b^{\prime}_{i}}({\Sigma}_{i}\cap F_{b^{\prime}_{i}},{\Sigma}^{\prime}_{i}\cap F_{b^{\prime}_{i}})\leq D_{\ref{Y-path 1.5}}(K_{2},R_{1}). Hence, by Lemma 4.11 dA(bi,V)D4.11(K2,D5.17(K2,R1)/R1)=D1d_{A}(b^{\prime}_{i},V)\leq D_{\ref{qc-level-set-new}}(K_{2},D_{\ref{Y-path 1.5}}(K_{2},R_{1})/R_{1})=D_{1}, say. Let vVv\in V be such that dA(bi,v)D1d_{A}(b^{\prime}_{i},v)\leq D_{1}. Then dB(bi,v)kD1+kd_{B}(b^{\prime}_{i},v)\leq kD_{1}+k. Hence, Hd([π(yi),bi]B,[π(yi),v]B)δ0+k+kD1Hd([\pi(y^{\prime}_{i}),b^{\prime}_{i}]_{B},[\pi(y^{\prime}_{i}),v]_{B})\leq\delta_{0}+k+kD_{1}. However, the concatenation [π(yi),b¯i]B[b¯i,v]B[\pi(y^{\prime}_{i}),\bar{b}^{\prime}_{i}]_{B}*[\bar{b}^{\prime}_{i},v]_{B} is a K2.25(δ0,k,1,0)K_{\ref{subqc-elem}}(\delta_{0},k^{\prime},1,0)-quasigeodesic. Hence, there is a point w[π(yi),v]Bw\in[\pi(y^{\prime}_{i}),v]_{B} such that dB(w,b¯i)D2.17(δ0,K2.25(δ0,k,1,0),K2.25(δ0,k,1,0))=D2d_{B}(w,\bar{b}^{\prime}_{i})\leq D_{\ref{stab-qg}}(\delta_{0},K_{\ref{subqc-elem}}(\delta_{0},k^{\prime},1,0),K_{\ref{subqc-elem}}(\delta_{0},k^{\prime},1,0))=D_{2}, say. Thus there is a point w[π(yi),bi]w^{\prime}\in[\pi(y^{\prime}_{i}),b^{\prime}_{i}] such that dB(w,b¯i)D2+δ0+k+kD1=D3d_{B}(w^{\prime},\bar{b}^{\prime}_{i})\leq D_{2}+\delta_{0}+k+kD_{1}=D_{3}, say. But bib^{\prime}_{i} is a nearest point projection of π(yi)\pi(y^{\prime}_{i}) on AA and b¯iVA\bar{b}^{\prime}_{i}\in V\subset A. Thus dB(w,bi)D3d_{B}(w^{\prime},b^{\prime}_{i})\leq D_{3}. Thus dB(b¯i,bi)2D3d_{B}(\bar{b}^{\prime}_{i},b^{\prime}_{i})\leq 2D_{3}. Hence, dB(bi,bi′′)dB(bi′′,b¯i)+dB(b¯i,bi)2D2.27(δ0,k,0)+2D3d_{B}(b^{\prime}_{i},b^{\prime\prime}_{i})\leq d_{B}(b^{\prime\prime}_{i},\bar{b}^{\prime}_{i})+d_{B}(\bar{b}^{\prime}_{i},b^{\prime}_{i})\leq 2D_{\ref{nested qc sets}}(\delta_{0},k^{\prime},0)+2D_{3}.

Case (2)(ii)(2^{\prime})(ii): Suppose dh(ΣiY,ΣiY)R1d_{h}({\Sigma}_{i}\cap Y,{\Sigma}^{\prime}_{i}\cap Y)\geq R_{1}. In this case Lemma 5.15 and Lemma 2.25 do the job.

(2′′)(2^{\prime\prime}): y¯i+1\bar{y}_{i+1} is a uniform approximate nearest point projection of y¯i\bar{y}^{\prime}_{i} on Σi+1Y{\Sigma}_{i+1}\cap Y in YY and the concatenation of β¯\bar{\beta} and the fiber geodesic [ΣiFbi+1,Σi+1Fbi+1]bi+1[{\Sigma}^{\prime}_{i}\cap F_{b_{i+1}},{\Sigma}_{i+1}\cap F_{b_{i+1}}]_{b_{i+1}} is a uniform quasigeodesic joining y¯i\bar{y}^{\prime}_{i} to y¯i+1\bar{y}_{i+1} in YY.

In this case dh(ΣiY,Σi+1Y)1d_{h}({\Sigma}^{\prime}_{i}\cap Y,{\Sigma}_{i+1}\cap Y)\leq 1 hence we are done as in Case (2)(i)(2^{\prime})(i).

Case 3: i=n1i=n-1: The proof of this case is also analogous to that of the proof of Case (2)(i)(2^{\prime})(i) since dh(Σn1,Σn)R0d_{h}({\Sigma}_{n-1},{\Sigma}_{n})\leq R_{0}. ∎

Remark 18.

The conclusion of Lemma 5.16 is subsumed by Lemma 5.15 and Lemma 5.17. But we still keep Lemma 5.16 for the sake of ease of explanation.

Thus by Lemma 5.11 and Lemma 5.18, we have proved the following.

Proposition 5.19.

Let x,yYx,y\in Y and let Σ{\Sigma} and Σ{\Sigma}^{\prime} be two K0K_{0}-qi sections in XX through xx and yy respectively. Let c(x,y)c(x,y) be a uniform quasigeodesic in XX joining xx and yy which is contained in 𝕃(Σ,Σ){\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) as constructed in step 1(c). Then the corresponding modified path c~(x,y)\tilde{c}(x,y), as constructed in step 2, is a uniform quasigeodesic in YY.

Step 4. Verification of the hypothesis of Lemma 2.49.

Lemma 5.20.

(Proper embedding of the pullback YY) The pullback YY is metrically properly embedded in XX. In fact, the distortion function for YY is the composition of a linear function with η\eta, the common distortion function for all the fibers of the bundle XX.

Proof.

As was done in the proof of the main theorem, we shall assume that gg is the inclusion map and Y=π1(A)Y=\pi^{-1}(A) and pp is the restriction of π\pi. Let x,yYx,y\in Y such that dX(x,y)Md_{X}(x,y)\leq M. Let π(x)=b1\pi(x)=b_{1} and π(y)=b2\pi(y)=b_{2}. Then, dB(b1,b2)Md_{B}(b_{1},b_{2})\leq M and hence dA(b1,b2)k+kMd_{A}(b_{1},b_{2})\leq k+kM. Let [b1,b2]A[b_{1},b_{2}]_{A} be a geodesic joining b1b_{1} and b2b_{2} in AA. This is a quasigeodesic in BB. By Lemma 3.8, there exists an isometric section γ\gamma over [b1,b2]A[b_{1},b_{2}]_{A}, through xx in YY. Clearly, γ\gamma is a qi lift in XX, say kk^{\prime}-qi lift. We have, lX(γ)k(kM+k)+k=:D(M)l_{X}(\gamma)\leq k^{\prime}(kM+k)+k^{\prime}=:D(M). The concatenation of γ\gamma and the fiber geodesic [γFb2,y]Fb2[\gamma\cap F_{b_{2}},y]_{F_{b_{2}}} is a path, denoted by α\alpha, joining xx and yy in XX. So,

dX(γFb2,y)dX(γFb2,x)+dX(x,y)lX(γ)+dX(x,y)D(M)+M.d_{X}(\gamma\cap F_{b_{2}},y)\leq d_{X}(\gamma\cap F_{b_{2}},x)+d_{X}(x,y)\leq l_{X}(\gamma)+d_{X}(x,y)\leq D(M)+M.

Now, since Fb2F_{b_{2}} is uniformly properly embedded as measured by η\eta, we have, db2(γFb2,y)η(D(M)+M)d_{b_{2}}(\gamma\cap F_{b_{2}},y)\leq\eta(D(M)+M). Now, α\alpha lies in YY and lY(γ)kM+kl_{Y}({\gamma})\leq kM+k. Then,

dY(x,y)lY(α)lY(γ)+dY(γFb2,y)kM+k+db2(γ~Fb2,y).d_{Y}(x,y)\leq l_{Y}(\alpha)\leq l_{Y}({\gamma})+d_{Y}({\gamma}\cap F_{b_{2}},y)\leq kM+k+d_{b_{2}}(\tilde{\gamma}\cap F_{b_{2}},y).

Therefore, dY(x,y)kM+k+η(D(M)+M)d_{Y}(x,y)\leq kM+k+\eta(D(M)+M). Setting η0(M):=kM+k+η(D(M)+M)\eta_{0}(M):=kM+k+\eta(D(M)+M), we have the following: for all x,yYx,y\in Y, d(x,y)Md(x,y)\leq M implies dY(x,y)η0(M)d_{Y}(x,y)\leq\eta_{0}(M). ∎

We recall that we fixed a vertex b0Ab_{0}\in A to define the paths c(y,y)c(y,y^{\prime}) in the last step. Let y0Fb0y_{0}\in F_{b_{0}}. However, the following lemma completes the proof of Theorem 5.2.

Lemma 5.21.

Given D>0D>0, there is D1>0D_{1}>0 such that the following holds.

If dX(y0,c(y,y))Dd_{X}(y_{0},c(y,y^{\prime}))\leq D then dY(y0,c¯(y,y))D1d_{Y}(y_{0},\bar{c}(y,y^{\prime}))\leq D_{1}.

Proof.

Let xc(y,y)x\in c(y,y^{\prime}) be such that dX(y0,x)Dd_{X}(y_{0},x)\leq D. This implies that dB(π(x),b0)Dd_{B}(\pi(x),b_{0})\leq D. We recall that the path c(y,y)c(y,y^{\prime}) is a concatenation of γj\gamma_{j}, j=0,1,,nj=0,1,\cdots,n. Suppose xγix\in\gamma_{i}, 0in0\leq i\leq n. We claim that there is a point of γ¯i\bar{\gamma}_{i} uniformly close to y0y_{0}. Now, γi\gamma_{i} is either a lift of geodesic segments of BB in a K2K_{2}-qi section Σi{\Sigma}_{i} or possibly Σi{\Sigma}^{\prime}_{i} or it is the concatenation of such a lift and a fiber geodesic of length at most R1R_{1}. Let 𝒬{\mathcal{Q}} denote the corresponding qi section and suppose c(y,y)𝒬c(y,y^{\prime})\cap{\mathcal{Q}} joins the points z𝒬z\in{\mathcal{Q}} to w𝒬w\in{\mathcal{Q}}. If i=ni=n then γi\gamma_{i} is a qi lift of [π(z),π(w)]B[\pi(z),\pi(w)]_{B} in 𝒬{\mathcal{Q}} joining z,wz,w. Otherwise there is a fiber geodesic σc(y,y)Fπ(w)\sigma\subset c(y,y^{\prime})\cap F_{\pi(w)} connecting 𝒬{\mathcal{Q}} to the next qi section 𝒬{\mathcal{Q}}^{\prime}, say. Then both the points zz and 𝒬σ{\mathcal{Q}}^{\prime}\cap\sigma are one of the yiy_{i}’s or yjy^{\prime}_{j}’s. Let z=𝒬σz^{\prime}={\mathcal{Q}}^{\prime}\cap\sigma and b=π(z)b^{\prime}=\pi(z^{\prime}). Let bb be the nearest point projection of π(x)\pi(x) on AA. It follows that dB(π(x),b)Dd_{B}(\pi(x),b)\leq D.

Suppose xσx\in\sigma. By the definition of c¯(y,y)\bar{c}(y,y^{\prime}) we have 𝒬Fbc¯(y,y){\mathcal{Q}}^{\prime}\cap F_{b}\in\bar{c}(y,y^{\prime}). However, dB(b,b0)dB(b,π(x))+dB(b0,π(x))2Dd_{B}(b,b_{0})\leq d_{B}(b,\pi(x))+d_{B}(b_{0},\pi(x))\leq 2D. Since AA is kk-qi embedded in BB we have dA(b,b0)k+2Dkd_{A}(b,b_{0})\leq k+2Dk. Hence, dY(𝒬Fb0,𝒬Fb)K2+(k+2Dk).K2d_{Y}({\mathcal{Q}}^{\prime}\cap F_{b_{0}},{\mathcal{Q}}^{\prime}\cap F_{b})\leq K_{2}+(k+2Dk).K_{2}. On the other hand in this case π(x)=b\pi(x)=b^{\prime} and db(z,x)R1d_{b^{\prime}}(z^{\prime},x)\leq R_{1}. Hence, dX(z,y0)R1+Dd_{X}(z^{\prime},y_{0})\leq R_{1}+D. Thus dX(y0,𝒬Fb0)dX(y0,x)+dX(x,z)+dX(z,𝒬Fb0)D+R1+K2+DK2d_{X}(y_{0},{\mathcal{Q}}^{\prime}\cap F_{b_{0}})\leq d_{X}(y_{0},x)+d_{X}(x,z^{\prime})+d_{X}(z^{\prime},{\mathcal{Q}}^{\prime}\cap F_{b_{0}})\leq D+R_{1}+K_{2}+DK_{2} since dB(b,b0)2Dd_{B}(b,b_{0})\leq 2D. Hence, dY(y0,𝒬Fb0)db0(y0,𝒬Fb0)η(D+R1+K2+DK2)d_{Y}(y_{0},{\mathcal{Q}}^{\prime}\cap F_{b_{0}})\leq d_{b_{0}}(y_{0},{\mathcal{Q}}^{\prime}\cap F_{b_{0}})\leq\eta(D+R_{1}+K_{2}+DK_{2}). Thus dY(y0,𝒬Fb)dY(y0,𝒬Fb0)+dY(𝒬Fb,𝒬Fb0)dY(y0,𝒬Fb0)+K2+K2dA(b,b0)η(D+R1+K2+DK2)+K2+(k+2Dk)K2d_{Y}(y_{0},{\mathcal{Q}}^{\prime}\cap F_{b})\leq d_{Y}(y_{0},{\mathcal{Q}}^{\prime}\cap F_{b_{0}})+d_{Y}({\mathcal{Q}}^{\prime}\cap F_{b},{\mathcal{Q}}^{\prime}\cap F_{b_{0}})\leq d_{Y}(y_{0},{\mathcal{Q}}^{\prime}\cap F_{b_{0}})+K_{2}+K_{2}d_{A}(b,b_{0})\leq\eta(D+R_{1}+K_{2}+DK_{2})+K_{2}+(k+2Dk)K_{2}. Hence, in this case dY(y0,c¯(y,y))(1+k+2Dk)K2+η(D+R1+K2+DK2)d_{Y}(y_{0},\bar{c}(y,y^{\prime}))\leq(1+k+2Dk)K_{2}+\eta(D+R_{1}+K_{2}+DK_{2}).

Otherwise suppose xx is contained in the lift of [π(z),π(w)]B[\pi(z),\pi(w)]_{B} in 𝒬{\mathcal{Q}}. We note that π(x)[π(z),π(w)]B\pi(x)\in[\pi(z),\pi(w)]_{B} and dB(π(x),A)Dd_{B}(\pi(x),A)\leq D. Now AA is k0k_{0}-quasiconvex in BB. Hence, by Lemma 2.29 we have dB(π(x),[π(z)¯,π(w)¯]B)D2.29(D,k0,δ)d_{B}(\pi(x),[\overline{\pi(z)},\overline{\pi(w)}]_{B})\leq D_{\ref{trivial lemma}}(D,k_{0},\delta). where π(z)¯\overline{\pi(z)}, π(w)¯\overline{\pi(w)} are nearest point projections of π(z),π(w)\pi(z),\pi(w) respectively on AA. Since AA is kk-qi embedded in BB by stability of quasigeodesics Hd([π(z)¯,π(w)¯]B,[π(z)¯,π(w)¯]A)D2.17(δ,k,k)Hd([\overline{\pi(z)},\overline{\pi(w)}]_{B},[\overline{\pi(z)},\overline{\pi(w)}]_{A})\leq D_{\ref{stab-qg}}(\delta,k,k). Hence, dB(π(x),[π(z)¯,π(w)¯]A)D2.29(D,k0,δ)+D2.17(δ,k,k)d_{B}(\pi(x),[\overline{\pi(z)},\overline{\pi(w)}]_{A})\leq D_{\ref{trivial lemma}}(D,k_{0},\delta)+D_{\ref{stab-qg}}(\delta,k,k). Let α\alpha be the lift of [π(z)¯,π(w)¯]A[\overline{\pi(z)},\overline{\pi(w)}]_{A} in 𝒬{\mathcal{Q}}. Then αc¯(y,y)\alpha\subset\bar{c}(y,y). On the other hand, dX(x,α)K2+K2dB(π(x),[π(z)¯,π(w)¯]A)K2+K2(D2.29(D,k0,δ)+D2.17(δ,k,k))=D1d_{X}(x,\alpha)\leq K_{2}+K_{2}d_{B}(\pi(x),[\overline{\pi(z)},\overline{\pi(w)}]_{A})\leq K_{2}+K_{2}(D_{\ref{trivial lemma}}(D,k_{0},\delta)+D_{\ref{stab-qg}}(\delta,k,k))=D_{1}, say. Hence, dX(y0,α)dX(y0,x)+dX(x,α)D+D1d_{X}(y_{0},\alpha)\leq d_{X}(y_{0},x)+d_{X}(x,\alpha)\leq D+D_{1}. This implies that dY(y0,α)d_{Y}(y_{0},\alpha) is also bounded by a function of DD and the other parameters of the metric graph bundles XX and YY, by Lemma 5.20. ∎

5.2. An example

For the convenience of the reader, we briefly illustrate a special case of our main theorem where B=,A=(,0]B={\mathbb{R}},A=(-\infty,0]. This discussion will also be used in the proof of the last proposition of the next section. We shall assume b0=0b_{0}=0 here.

As in the proof of Lemma 5.21 suppose 𝒬,𝒬{\mathcal{Q}},{\mathcal{Q}}^{\prime} are two qi sections among the various Σi,Σj{\Sigma}_{i},{\Sigma}^{\prime}_{j}’s and let w𝒬,z,w𝒬w^{\prime}\in{\mathcal{Q}}^{\prime},z,w\in{\mathcal{Q}} are points of c(y,y)c(y,y^{\prime}) where π(w)=π(w)\pi(w^{\prime})=\pi(w), dπ(w)(w,w)R1d_{\pi(w)}(w,w^{\prime})\leq R_{1} and the concatenation of the lift say α\alpha, of [π(z),π(w)][\pi(z),\pi(w)] in 𝒬{\mathcal{Q}} and the vertical geodesic segment, say σ\sigma, in Fπ(w)F_{\pi(w)} is a part of c(y,y)c(y,y^{\prime}). Following are the possibilities.

Case 1. If w,zYc(y,y)w^{\prime},z\in Y\cap c(y,y^{\prime}) then ασY\alpha*\sigma\subset Y and it is the corresponding part of c¯(y,y)\bar{c}(y,y^{\prime}).

Case 2. zY,wYz\in Y,w^{\prime}\not\in Y. In this case, the modified segment is formed as the concatenation of subsegment of α\alpha joining zz to 𝒬F0{\mathcal{Q}}\cap F_{0} and the fiber geodesic [𝒬F0,𝒬F0]0[{\mathcal{Q}}\cap F_{0},{\mathcal{Q}}^{\prime}\cap F_{0}]_{0}.

w¯\overline{w}zzw¯\overline{w^{\prime}}wwww^{\prime}Fπ(z)F_{\pi(z)}F0F_{0}Fπ(w)F_{\pi(w^{\prime})}𝒬\mathcal{Q}𝒬\mathcal{Q}^{\prime}
Figure 9. Case 2

Case 3. wY,zYw^{\prime}\in Y,z\not\in Y. In this case the modified segment is the concatenation of the segment of α\alpha from 𝒬F0{\mathcal{Q}}\cap F_{0} to ww and the fiber geodesic segment σ\sigma.

z¯\overline{z}w=w¯w=\overline{w}w=w¯w^{\prime}=\overline{w^{\prime}}zzFπ(w)F_{\pi(w^{\prime})}F0F_{0}Fπ(z)F_{\pi(z)}𝒬\mathcal{Q}𝒬\mathcal{Q}^{\prime}
Figure 10. Case 3

Case 4. z,wYz,w^{\prime}\not\in Y. In this case the modified segment is the fiber geodesic [𝒬F0,𝒬F0]0[{\mathcal{Q}}\cap F_{0},{\mathcal{Q}}^{\prime}\cap F_{0}]_{0}.

zzz¯\overline{z}w¯\overline{w^{\prime}}wwww^{\prime}F0F_{0}Fπ(z)F_{\pi(z)}Fπ(w)F_{\pi(w^{\prime})}𝒬\mathcal{Q}𝒬\mathcal{Q}^{\prime}
Figure 11. Case 4

Here, the dashed lines denote the portion of c(y,y)c(y,y^{\prime}), the thick lines denote the portion of c¯(y,y)\bar{c}(y,y^{\prime}) and dotted lines are portions of the qi sections 𝒬,𝒬{\mathcal{Q}},{\mathcal{Q}}^{\prime}.

6. Applications, examples and related results

As the first application of our main theorem, we have the following. Given a short exact sequence of finitely generated groups there is a natural way to associate a metric graph bundle to it as mentioned in Example 1.8 of [MS12]. See also Example 2. Having said that Theorem 5.2 gives the following as an immediate consequence.

Theorem 6.1.

Suppose 1NGπQ11\rightarrow N\rightarrow G\stackrel{{\scriptstyle\pi}}{{\rightarrow}}Q\rightarrow 1 is a short exact sequence of hyperbolic groups where NN is nonelementary hyperbolic. Suppose Q1Q_{1} is a finitely generated, qi embedded subgroup of QQ and G1=π1(Q1)G_{1}=\pi^{-1}(Q_{1}). Then the G1G_{1} is hyperbolic and the inclusion G1GG_{1}\rightarrow G admits the CT map.

The next application is in the context of complexes of hyperbolic groups. Suppose 𝒴\mathcal{Y} is a finite, connected simplicial complex and 𝔾(𝒴){\mathbb{G}}(\mathcal{Y}) is a developable complex of nonelementary hyperbolic groups with qi condition defined over 𝒴\mathcal{Y} (see Section 3.3.2) such that the fundamental group GG of the complex of groups is hyperbolic. Suppose we have a good subcomplex 𝒴1𝒴{\mathcal{Y}}_{1}\subset{\mathcal{Y}} and G1G_{1} is the image of π1(𝒢,𝒴1)\pi_{1}({\mathcal{G}},{\mathcal{Y}}_{1}) in GG under the natural homomorphism π1(𝒢,𝒴1)π1(𝒢,𝒴)\pi_{1}({\mathcal{G}},{\mathcal{Y}}_{1})\rightarrow\pi_{1}({\mathcal{G}},{\mathcal{Y}}). Then we have the following pullback diagram as obtained in Proposition 3.29 satisfying the properties of Theorem 5.2.

X1X_{1}XXB1B_{1}BBffπ1\pi_{1}π\piii
Figure 12.

Thus we have:

Theorem 6.2.

The group G1G_{1} is hyperbolic and the inclusion G1GG_{1}\rightarrow G admits the CT map.

Remark 19.

The rest of the paper is devoted to properties of the boundary of metric (graph) bundles and Cannon-Thurston maps. We recall that qi sections, ladders etc for a metric bundle are defined as transport of the same from the canonical metric graph bundle associated to it. All the results in the rest of the section are meant for metric bundles as well as metric graph bundles. However, using the dictionary provided by Proposition 4.1 it is enough to prove the results only for metric graph bundles. Therefore, we shall state and prove results only for metric graph bundles in what follows starting with the convention below.

Convention 6.3.

(1) For the rest of the paper we shall assume that π:XB\pi:X\rightarrow B is a δ\delta-hyperbolic η\eta-metric graph bundle over BB satisfying the hypothesis H1, H2, H3 and H4 of section 5. (2) By Proposition 2.37 any point of B\partial B can be joined to any point of BBB\cup\partial B and any point of X\partial X can be joined to XXX\cup\partial X by a uniform quasigeodesic ray or line. We shall assume that these are κ0\kappa_{0}-quasigeodesics. (3) We shall assume that any geodesic in BB has a cc-qi lift in XX using the path lifting lemma for metric graph bundles. (4) We recall that through any point of XX there is a K0K_{0}-qi section over BB.

6.1. Some properties of X\partial X

Lemma 6.4.

Suppose α,β:[0,)B\alpha,\beta:[0,\infty)\rightarrow B are two kk-quasigeodesic rays for some k1k\geq 1 with α()=β()=ξ\alpha(\infty)=\beta(\infty)=\xi. Suppose β~\tilde{\beta} is a KK-qi lift of β\beta for some K1K\geq 1. Then there is a KK^{\prime}-qi lift α~\tilde{\alpha} of α\alpha such that α~()=β~()\tilde{\alpha}(\infty)=\tilde{\beta}(\infty) where KK^{\prime} depends on kk, KK, dB(α(0),β(0))d_{B}(\alpha(0),\beta(0)) and the various parameters of the metric graph bundle.

Proof.

Suppose α,β:[0,)B\alpha,\beta:[0,\infty)\rightarrow B are two kk-quasigeodesic rays for some k1k\geq 1 with α()=β()=ξ\alpha(\infty)=\beta(\infty)=\xi. This means Hd(α,β)<Hd(\alpha,\beta)<\infty. Let R=Hd(α,β)R=Hd(\alpha,\beta). Then for all s[0,)s\in[0,\infty) there is t=t(s)[0,)t=t(s)\in[0,\infty) such that dB(α(s),β(t))Rd_{B}(\alpha(s),\beta(t))\leq R. Let ϕts:Fβ(t)Fα(s)\phi_{ts}:F_{\beta(t)}\rightarrow F_{\alpha(s)} be fiber identification maps such that dX(x,ϕts(x))3c+3cRd_{X}(x,\phi_{ts}(x))\leq 3c+3cR for all xFβ(t)x\in F_{\beta(t)}, t[0,)t\in[0,\infty) where c=1c=1 for metric graph bundles. (See Lemma 3.10.) Let β~\tilde{\beta} be a KK-qi lift of β\beta. Now, for all s[0,)s\in[0,\infty) we define α~(s)=ϕts(β~(t))\tilde{\alpha}(s)=\phi_{ts}(\tilde{\beta}(t)). It is easy to verify that α~\tilde{\alpha} thus defined is a uniform qi lift of α\alpha. Also clearly α~N3c+3cR(β~)\tilde{\alpha}\subset N_{3c+3cR}(\tilde{\beta}). It follows that α~()=β~()\tilde{\alpha}(\infty)=\tilde{\beta}(\infty)

Corollary 6.5.

Let ξB\xi\in\partial B and let α\alpha be a quasigeodesic ray in BB joining bb to ξ\xi. Let αξX:={γ():γis a qi lift ofα}\partial^{\xi}_{\alpha}X:=\{\gamma(\infty):\gamma\,\mbox{is a qi lift of}\,\,\alpha\}.

Then αξX\partial^{\xi}_{\alpha}X is independent of α\alpha; it is determined by ξ\xi.

Due to the above corollary, we shall use the notation ξX\partial^{\xi}X for all ξB\xi\in\partial B without further explanation. The following proposition is motivated by a similar result proved by Bowditch ([Bow13, Proposition 2.3.2]).

Proposition 6.6.

Let bBb\in B be an arbitrary point and F=FbF=F_{b}. Then we have

X=Λ(F)(ξBξX).\partial X=\Lambda(F)\cup(\coprod_{\xi\in\partial B}\partial^{\xi}X).
Proof.

We first fix a point xFx\in F. Let γ\gamma be a quasigeodesic ray in XX starting from xx. Let bn=π(γ(n))b_{n}=\pi(\gamma(n)). Let αn\alpha_{n} be a (1,1)(1,1)-quasigeodesic in BB joining bb to bnb_{n}. Let α~n\tilde{\alpha}_{n} be a K0K_{0}-qi lift of αn\alpha_{n} joining γ(n)\gamma(n) to αn~(b)=xnF\tilde{\alpha_{n}}(b)=x_{n}\in F. There are two possibilities.

Suppose {xn}\{x_{n}\} has an unbounded subsequence say {xnk}\{x_{n_{k}}\}. Then d(xnk,x)d(x_{n_{k}},x)\rightarrow\infty. We note that α~nk\tilde{\alpha}_{n_{k}}’s are uniform quasigeodesics in XX whose distance from xx is going to infinity by Lemma 4.13. Hence, by Lemma 2.34 xnkγ()x_{n_{k}}\rightarrow\gamma(\infty) and thus γ()Λ(F)\gamma(\infty)\in\Lambda(F).

Otherwise, suppose {xn}\{x_{n}\} is a bounded sequence.

Claim: In this case πγ\pi\circ\gamma is a quasigeodesic ray.

Proof of claim: We note that by stability of quasigeodesics (Corollary 2.19) and slimness of triangles (Lemma 2.20) Hd(α~n,γ|[0,n])Hd(\tilde{\alpha}_{n},\gamma|_{[0,n]}) is uniformly small for all nn. This implies that Hd(αn,(πγ)|[0,n])Hd(\alpha_{n},(\pi\circ\gamma)|_{[0,n]}) is uniformly small for all nn; in particular dB(bm,αn)d_{B}(b_{m},\alpha_{n}) is uniformly small for all nmn\geq m. Next we note that dB(b,bn)d_{B}(b,b_{n})\rightarrow\infty for otherwise d(γ(n),x)d(\gamma(n),x) will be bounded. Then it follows that limm,n(bm.bn)b=\lim_{m,n\rightarrow\infty}(b_{m}.b_{n})_{b}=\infty. Let ξ=limnbn\xi=\lim_{n\rightarrow\infty}b_{n} and let α\alpha be a κ0\kappa_{0}-quasigeodesic ray in BB joining bb to ξ\xi. Now, to show that πγ\pi\circ\gamma is a quasigeodesic it is enough to show by Lemma 2.5 that πγ\pi\circ\gamma is (1) uniformly close to α\alpha and (2) properly embedded.

(1): Fix an arbitrary mm\in{\mathbb{N}} and consider all nmn\geq m. Since limnbn=α()=ξ\lim_{n\rightarrow\infty}b_{n}=\alpha(\infty)=\xi, by Lemma 2.45(2) for any κ0\kappa_{0}-quasigeodesic ray βn\beta_{n} joining bnb_{n} to ξ\xi we have d(b,βn)d(b,\beta_{n})\rightarrow\infty. Since the triangles with vertices bn,b,ξb_{n},b,\xi are uniformly slim by Lemma 2.38 and dB(bm,αn)d_{B}(b_{m},\alpha_{n}) are uniformly small it follows that bmb_{m} is uniformly close to α\alpha. This shows (1).

(2): Since π\pi is Lipschitz and γ\gamma is a quasigeodesic it follows that πγ\pi\circ\gamma is coarsely Lipschitz. Suppose dB(bn,bm)Dd_{B}(b_{n},b_{m})\leq D for some D0D\geq 0 and m,nm,n\in{\mathbb{N}}, mnm\leq n. We claim that dX(γ(m),γ(n))d_{X}(\gamma(m),\gamma(n)) is uniformly small. Note that this would then imply that nmn-m is uniformly small since γ\gamma is quasigeodesic, and also that γ\gamma is a qi lift of πγ\pi\circ\gamma. We know that Hd(α~n,γ|[0,n])RHd(\tilde{\alpha}_{n},\gamma|_{[0,n]})\leq R for some constant RR independent of nn. Hence, dX(γ(m),α~n)Rd_{X}(\gamma(m),\tilde{\alpha}_{n})\leq R. Let ym,nα~ny_{m,n}\in\tilde{\alpha}_{n} be such that dX(γ(m),ym,n)Rd_{X}(\gamma(m),y_{m,n})\leq R. Since π\pi is 11-Lipschitz we have dX(bm,π(ym,n))Rd_{X}(b_{m},\pi(y_{m,n}))\leq R. Then dB(π(ym,n),bn)dB(π(ym,n),bm)+dB(bm,bn)R+Dd_{B}(\pi(y_{m,n}),b_{n})\leq d_{B}(\pi(y_{m,n}),b_{m})+d_{B}(b_{m},b_{n})\leq R+D. Since α~n\tilde{\alpha}_{n} is K0K_{0}-qi lift of αn\alpha_{n} and πα~(n)=bn\pi\circ\tilde{\alpha}(n)=b_{n} it follows that dX(ym,n,α~(n))=dX(ym,n,γ(n))K0(R+D)+K0d_{X}(y_{m,n},\tilde{\alpha}(n))=d_{X}(y_{m,n},\gamma(n))\leq K_{0}(R+D)+K_{0}. Hence, dX(γ(m),γ(n))dX(γ(m),ym,n)+dX(ym,n,γ(n))R+K0(R+D)+K0d_{X}(\gamma(m),\gamma(n))\leq d_{X}(\gamma(m),y_{m,n})+d_{X}(y_{m,n},\gamma(n))\leq R+K_{0}(R+D)+K_{0}. Since γ\gamma is quasigeodesic it follows that (nm)(n-m) is uniformly small. This proves (2) and along with this the claim. It follows that γ()ξX\gamma(\infty)\in\partial^{\xi}X.

It remains to check that for all ξ1,ξ2B\xi_{1},\xi_{2}\in\partial B, ξ1Xξ2X\partial^{\xi_{1}}X\cap\partial^{\xi_{2}}X\neq\emptyset implies ξ1=ξ2\xi_{1}=\xi_{2}. Suppose γi\gamma_{i} is a κ0\kappa_{0}-quasigeodesic ray in BB joining bb to ξi\xi_{i}, i=1,2i=1,2. Suppose γ~i\tilde{\gamma}_{i} is a qi lift of γi\gamma_{i}, i=1,2i=1,2 such that γ~1()=γ~2()\tilde{\gamma}_{1}(\infty)=\tilde{\gamma}_{2}(\infty), i.e. Hd(γ~1,γ~2)<Hd(\tilde{\gamma}_{1},\tilde{\gamma}_{2})<\infty. Then Hd(γ1,γ2)<Hd(\gamma_{1},\gamma_{2})<\infty because π:XB\pi:X\rightarrow B is 11-Lipschitz. Thus ξ1=ξ2\xi_{1}=\xi_{2}. This finishes the proof. ∎

Corollary 6.7.

Suppose FF is a bounded metric space. Then X=ξBξX\partial X=\coprod_{\xi\in\partial B}\partial^{\xi}X.

For instance suppose Σ1,Σ2{\Sigma}_{1},{\Sigma}_{2} are two qi sections and 𝕃=𝕃(Σ1,Σ2){\mathbb{L}}={\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}) then by Corollary 4.7 there is a metric graph subbundle πZ:ZB\pi_{Z}:Z\rightarrow B of XX where the bundle map ZXZ\rightarrow X is a qi embedding onto a finite neighborhood of 𝕃{\mathbb{L}}. It follows that ZZ is hyperbolic and fibers are uniformly quasiisometric to intervals. Therefore, the conclusion of Corollary 6.7 applies to the metric bundle ZZ too. Hence, informally speaking we have the following.

Corollary 6.8.

For any ladder 𝕃=𝕃(Σ1,Σ2){\mathbb{L}}={\mathbb{L}}({\Sigma}_{1},{\Sigma}_{2}) we have

𝕃=ξBξ𝕃.\partial{\mathbb{L}}=\coprod_{\xi\in\partial B}\,\partial^{\xi}{\mathbb{L}}.
Lemma 6.9.

Suppose bBb\in B and αn:[0,)B\alpha_{n}:[0,\infty)\rightarrow B is a sequence of uniform quasigeodesic rays starting from bb. Suppose α~n\tilde{\alpha}_{n} is a uniform qi lift of αn\alpha_{n} for all nn such that the set {α~n(0)}\{\tilde{\alpha}_{n}(0)\} has finite diameter. If α~n()zX\tilde{\alpha}_{n}(\infty)\rightarrow z\in\partial X then limnαn()\lim_{n\rightarrow\infty}\alpha_{n}(\infty) exists. If ξ=limnαn()\xi=\lim_{n\rightarrow\infty}\alpha_{n}(\infty) and α:[0,)B\alpha:[0,\infty)\rightarrow B is a κ0\kappa_{0}-quasigeodesic ray joining bb to ξ\xi then there is a uniform qi lift α~\tilde{\alpha} of α\alpha such that α~()=z\tilde{\alpha}(\infty)=z.

Proof.

Since α~n()ξ\tilde{\alpha}_{n}(\infty)\rightarrow\xi there is a constant DD such that for all M>0M>0 there is N=N(M)>0N=N(M)>0 with Hd(α~m|[0,M],α~n|[0,M])DHd(\tilde{\alpha}_{m}|_{[0,M]},\tilde{\alpha}_{n}|_{[0,M]})\leq D for all m,nNm,n\geq N by Lemma 2.45(1). It follows that for all M>0M>0, Hd(αm|[0,M],αn|[0,M])DHd(\alpha_{m}|_{[0,M]},\alpha_{n}|_{[0,M]})\leq D for all m,nNm,n\geq N. Hence, again by Lemma 2.45(1) αn()\alpha_{n}(\infty) converges to a point of ξB\xi\in\partial B. Let α\alpha be a κ0\kappa_{0}-quasigeodesic ray in BB joining bb to ξ\xi. We claim zξXz\in\partial^{\xi}X. Given any t[0,)t\in[0,\infty) by Lemma 2.45(2) there is N=N(t)N^{\prime}=N^{\prime}(t)\in{\mathbb{N}} such that d(α(t),αn)Dd(\alpha(t),\alpha_{n})\leq D^{\prime} for all nNn\geq N^{\prime} where DD^{\prime} depends only on κ0\kappa_{0} and δ\delta. Let N0=max{N(t),N(t)}N_{0}=\max\{N(t),N^{\prime}(t)\}. Let tt^{\prime} be such that dX(α(t),αN0(t))Dd_{X}(\alpha(t),\alpha_{N_{0}}(t^{\prime}))\leq D^{\prime}. Define α~(t)=ϕuv(α~N0(t))\tilde{\alpha}(t)=\phi_{uv}(\tilde{\alpha}_{N_{0}}(t^{\prime})) where u=αN0(t),v=α(t)u=\alpha_{N_{0}}(t^{\prime}),v=\alpha(t) and ϕuv\phi_{uv} is a fiber identification map. It is now easy to check that this defines a qi section over α\alpha and z=α~()z=\tilde{\alpha}(\infty).

Corollary 6.10.

If fibers of the metric (graph) bundle are of finite diameter then the map X=ξBξXB\partial X=\cup_{\xi\in\partial B}\partial^{\xi}X\rightarrow\partial B defined by sending ξX\partial^{\xi}X to ξ\xi for all ξB\xi\in\partial B is continuous.

6.2. Cannon-Thurston lamination

Suppose b0Bb_{0}\in B is an arbitrary point and F=Fb0F=F_{b_{0}}. Then we know that the inclusion i=iF,X:FXi=i_{F,X}:F\hookrightarrow X admits the CT map i:FX\partial i:\partial F\rightarrow\partial X. For any set SS we define

S(2)={(a,b)S×S:ab}.S^{(2)}=\{(a,b)\in S\times S:a\neq b\}.

Now, following Mitra([Mit97]) we define the following.

Definition 6.11.

(1)(1) (Cannon-Thurston lamination) Let X(2)(F)={(α,β)(2)F:i(α)=i(β)}\partial^{(2)}_{X}(F)=\{(\alpha,\beta)\in\partial^{(2)}F:\partial i(\alpha)=\partial i(\beta)\}.

(2)(2) Suppose ξB\xi\in\partial B. Let ξ,X(2)(F)={(α,β)(2)F:i(α)=i(β)ξX}\partial^{(2)}_{\xi,X}(F)=\{(\alpha,\beta)\in\partial^{(2)}F:\partial i(\alpha)=\partial i(\beta)\in\partial^{\xi}X\}. We shall denote ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F) simply by ξ(2)(F)\partial^{(2)}_{\xi}(F) when XX is understood.

In this subsection we are going to discuss the various properties of the CT lamination. First we need some definitions. We recall that for all b,sBb,s\in B we have the fiber identification map ϕbs:FbFs\phi_{bs}:F_{b}\rightarrow F_{s} which is a uniform quasiisometry depending on dB(b,s)d_{B}(b,s). This induces a bijection ϕbs:FbFs\partial\phi_{bs}:\partial F_{b}\rightarrow\partial F_{s}. Suppose zFbz\in\partial F_{b}. Let zs=ϕbs(z)z_{s}=\partial\phi_{bs}(z) for all sBs\in B.

Convention 6.12.

For the rest of the subsection by ‘quasigeodesic rays’ or ‘lines’, we shall always mean κ0\kappa_{0}-quasigeodesic rays and lines in the fibers of a metric (graph) bundle unless otherwise specified,

Definition 6.13.

(1) (Semi-infinite ladders) Suppose Σ1{\Sigma}_{1} is a qi section over BB in XX. For all sBs\in B let γsFs\gamma_{s}\subset F_{s} be a (uniform) quasigeodesic ray joining Σ1Fs{\Sigma}_{1}\cap F_{s} to zs=ϕbs(z)z_{s}=\partial\phi_{bs}(z). The union of all the rays will be denoted by 𝕃(Σ1;z){\mathbb{L}}({\Sigma}_{1};z).

This set is coarsely well-defined by Lemma 2.38. We shall refer to this as the semi-infinite ladder defined by Σ1{\Sigma}_{1} and zz.

(2) (Bi-infinite ladders) Suppose bBb\in B and z,zFbz,z^{\prime}\in\partial F_{b}, zzz\neq z^{\prime}. Now for all sBs\in B join zs=ϕbs(z)z_{s}=\partial\phi_{bs}(z) to zs=ϕbs(z)z^{\prime}_{s}=\partial\phi_{bs}(z^{\prime}) by a (uniform) quasigeodesic line in FsF_{s}. The union of all these lines will be denoted by 𝕃(z;z){\mathbb{L}}(z;z^{\prime}).

As before, this set is coarsely well-defined by Lemma 2.38. We shall refer to this as the bi-infinite ladder defined by zz and zz^{\prime}.

We shall refer to either of these ladders as an ‘infinite girth ladder’.

Lemma 6.14.

(Properties of infinite girth ladders) Suppose 𝕃{\mathbb{L}} is an infinite girth ladder.

  1. (1)

    (Coarse retract) There is a uniformly coarsely Lipschitz retraction π𝕃:X𝕃\pi_{{\mathbb{L}}}:X\rightarrow{\mathbb{L}} such that for all bBb\in B and xFbx\in F_{b}, π𝕃(x)\pi_{{\mathbb{L}}}(x) is a (uniform approximate) nearest point projection of xx in FbF_{b} on 𝕃Fb{\mathbb{L}}\cap F_{b}.

    Consequently, infinite girth ladders are uniformly quasiconvex and their uniformly small neighborhoods are qi embedded in XX.

  2. (2)

    (QI sections in ladders) Through any point of 𝕃{\mathbb{L}}, there exists a uniform qi section contained in 𝕃{\mathbb{L}}.

  3. (3)

    (QI sections coarsely bisect ladders) Any qi section in 𝕃{\mathbb{L}} coarsely bisects it into two subladders.

Proof.

We shall briefly indicate the proofs comparing with the proof of the analogous results for finite girth ladders. (3) follows exactly as Lemma 5.4. (2) is immediate from (1). In fact given x𝕃x\in{\mathbb{L}} one takes a K0K_{0}-qi section Σ{\Sigma} in XX containing xx and then π𝕃(Σ)\pi_{{\mathbb{L}}}({\Sigma}) is the required qi section. Therefore, we are left with proving (1). This is an exact analog of Proposition 4.6(1). The reader is referred to [Mit97, Theorem 4.6] for supporting arguments. ∎

Convention 6.15.

All semi-infinite ladders 𝕃(Σ;z){\mathbb{L}}({\Sigma};z) are formed by K0K_{0}-qi section Σ{\Sigma}. We shall assume that through any point of an infinite girth ladder there is a K¯0\bar{K}_{0}-qi section contained in the ladder. Also, all infinite girth ladders are assumed to be λ¯0\bar{\lambda}_{0}-quasiconvex.

6.2.1. Properties of the CT lamination X(2)(F)\partial^{(2)}_{X}(F)

In this subsection, we prove many properties of the CT lamination using coarse bisection of ladders by qi sections. These are motivated by analogous results proved in [Mit97] and [Bow13]. For the rest of the subsection, we will use the following set up. Let b0Bb_{0}\in B and F=Fb0F=F_{b_{0}}. Suppose (z1,z2)(2)=X(2)(F)(z_{1},z_{2})\in\partial^{(2)}=\partial^{(2)}_{X}(F) and 𝕃=𝕃(z1;z2){\mathbb{L}}={\mathbb{L}}(z_{1};z_{2}). Let γ:F\gamma:{\mathbb{R}}\rightarrow F be a κ0\kappa_{0}-quasigeodesic line in FF joining z1z_{1} to z2z_{2} such that Im(γ)=𝕃FIm(\gamma)={\mathbb{L}}\cap F. Let iF,X:FXi_{F,X}:F\rightarrow X denote the inclusion map and iF,X:FX\partial i_{F,X}:\partial F\rightarrow\partial X denote the CT map.

Lemma 6.16.

Suppose Σ\Sigma is any qi section contained in 𝕃{\mathbb{L}}. Then iF,X(zi)Λ(Σ)\partial i_{F,X}(z_{i})\in\Lambda({\Sigma}), i=1,2i=1,2.

Proof.

Let Σ{\Sigma} be a qi section contained in 𝕃{\mathbb{L}}. Then Σ{\Sigma} coarsely separates 𝕃{\mathbb{L}} in XX into 𝕃1=𝕃(Σ;z1){\mathbb{L}}_{1}={\mathbb{L}}({\Sigma};z_{1}) and 𝕃2=𝕃(Σ;z2){\mathbb{L}}_{2}={\mathbb{L}}({\Sigma};z_{2}). We note that iF,X(z1)=iF,X(z2)Λ(𝕃1)Λ(𝕃2)\partial i_{F,X}(z_{1})=\partial i_{F,X}(z_{2})\in\Lambda({\mathbb{L}}_{1})\cap\Lambda({\mathbb{L}}_{2}). Hence we are done by Lemma 2.53. ∎

Lemma 6.17.

Suppose (z1,z2)X(2)(F)(z_{1},z_{2})\in\partial^{(2)}_{X}(F) and 𝕃=𝕃(z1;z2){\mathbb{L}}={\mathbb{L}}(z_{1};z_{2}). There is a unique ξB\xi\in\partial B such that (z1,z2)ξ,X(2)(F)(z_{1},z_{2})\in\partial^{(2)}_{\xi,X}(F). Moreover, for any κ0\kappa_{0}-quasigeodesic β:[0,)B\beta:[0,\infty)\rightarrow B joining b0b_{0} to ξ\xi and any qi section Σ\Sigma contained in 𝕃{\mathbb{L}}, if β~\tilde{\beta} is the lift of β\beta in Σ\Sigma then β~()=iF,X(z1)=iF,X(z2)\tilde{\beta}(\infty)=\partial i_{F,X}(z_{1})=\partial i_{F,X}(z_{2}).

In particular X(2)(F)=ξBξ(2)(F)\partial^{(2)}_{X}(F)=\coprod_{\xi\in\partial B}\partial^{(2)}_{\xi}(F).

Proof.

Let σ:BX\sigma:B\rightarrow X be a qi section with image Σ\Sigma contained in 𝕃{\mathbb{L}}. By Lemma 6.16 iF,X(z1)Λ(Σ)\partial i_{F,X}(z_{1})\in\Lambda({\Sigma}). But Λ(Σ)=σ(B)\Lambda({\Sigma})=\partial\sigma(\partial B) by Lemma 2.55. Hence, there is a κ0\kappa_{0}-quasigeodesic ray β:[0,)B\beta:[0,\infty)\rightarrow B such that σ(β())=iF,X(z1)\partial\sigma(\beta(\infty))=\partial i_{F,X}(z_{1}). Let ξ=β()\xi=\beta(\infty). If β~=σβ\tilde{\beta}=\sigma\circ\beta then β~\tilde{\beta} is a qi lift of β\beta and iF,X(z1)=β~()ξX\partial i_{F,X}(z_{1})=\tilde{\beta}(\infty)\in\partial^{\xi}X. Thus (z1,z2)ξ,X(2)(F)(z_{1},z_{2})\in\partial^{(2)}_{\xi,X}(F). This shows the existence of ξ\xi. Thus we have X(2)(F)=ξBξ(2)(F)\partial^{(2)}_{X}(F)=\bigcup_{\xi\in\partial B}\partial^{(2)}_{\xi}(F). Also for ξ,ξB\xi,\xi^{\prime}\in\partial B, ξξ\xi\neq\xi^{\prime} we have ξ1Xξ2X=\partial^{\xi_{1}}X\cap\partial^{\xi_{2}}X=\emptyset by Proposition 6.6 which immediately implies ξ,X(2)(F)ξ,X(2)(F)=\partial^{(2)}_{\xi,X}(F)\cap\partial^{(2)}_{\xi^{\prime},X}(F)=\emptyset. This shows that the point ξ\xi is independent of the chosen section Σ\Sigma in 𝕃{\mathbb{L}}. The last part of the lemma is immediate from these observations. ∎

We next aim to show that the sets ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F) are closed subsets of X(2)(F)\partial^{(2)}_{X}(F). Let β:[0,)B\beta:[0,\infty)\rightarrow B be a continuous, arc length parameterized κ0\kappa_{0}-quasigeodesic in BB with β(0)=b0\beta(0)=b_{0} and β()=ξ\beta(\infty)=\xi as in the proof of Lemma 6.17. Let A=β([0,))A=\beta([0,\infty)). Let Y=π1(A)Y=\pi^{-1}(A) be the restriction of the bundle XX over AA. Let iY,X:YXi_{Y,X}:Y\rightarrow X, iF,Y:FYi_{F,Y}:F\rightarrow Y be inclusion maps.

Lemma 6.18.

If (z1,z2)ξ,X(2)(F)(z_{1},z_{2})\in\partial^{(2)}_{\xi,X}(F) then iF,Y(z1)=iF,Y(z2)\partial i_{F,Y}(z_{1})=\partial i_{F,Y}(z_{2}), i.e. (z1,z2)ξ,Y(2)(F)(z_{1},z_{2})\in\partial^{(2)}_{\xi,Y}(F).

Proof.

Let Σn{\Sigma}_{n} be any qi section in 𝕃{\mathbb{L}} over BB passing through γ(n)\gamma(n), nn\in{\mathbb{Z}}. Then by Lemma 6.17, ΣmY{\Sigma}_{m}\cap Y and ΣnY{\Sigma}_{n}\cap Y are asymptotic for all m,nm,n\in{\mathbb{Z}} in XX. Since YY is properly embedded in XX by Lemma 5.20 they are still asymptotic in YY. Clearly dY(γ(0),ΣnY)d_{Y}(\gamma(0),{\Sigma}_{n}\cap Y)\rightarrow\infty as n±n\rightarrow\pm\infty. Thus by Lemma 2.45(1) limn±γ(n)=β~0()\lim_{n\rightarrow\pm\infty}\gamma(n)=\tilde{\beta}_{0}(\infty) in YY where β~0\tilde{\beta}_{0} is the lift of β\beta in Σ0{\Sigma}_{0}. This completes the proof. ∎

Corollary 6.19.

Let β~\tilde{\beta} be any qi lift of β\beta in 𝕃{\mathbb{L}}. Then β~()=iF,X(z1)\tilde{\beta}(\infty)=\partial i_{F,X}(z_{1}). In particular any two qi lifts of β\beta in 𝕃{\mathbb{L}} are asymptotic.

Proof.

We know that β~\tilde{\beta} coarsely separates 𝕃Y{\mathbb{L}}\cap Y into two semi-infinite ladders, 𝕃+{\mathbb{L}}^{+} and 𝕃{\mathbb{L}}^{-} in YY. It follows that Λ(𝕃+)Λ(𝕃)=Λ(β~)=β~()\Lambda({\mathbb{L}}^{+})\cap\Lambda({\mathbb{L}}^{-})=\Lambda(\tilde{\beta})=\tilde{\beta}(\infty). It then follows that the limit of γ(n)\gamma(n) in 𝕃\partial{\mathbb{L}} is β~()\tilde{\beta}(\infty). ∎

Corollary 6.20.

(1) (𝕃Y)\partial({\mathbb{L}}\cap Y) is a point. (2) ΛY(𝕃Y)\Lambda_{Y}({\mathbb{L}}\cap Y) is a point. (3) ΛX(𝕃Y)\Lambda_{X}({\mathbb{L}}\cap Y) is a point.

Proof.

We know by Proposition 6.14(1) (see also Proposition 4.6(4)) that a small neighborhood, say 𝕃Y=NR(𝕃Y){\mathbb{L}}^{\prime}_{Y}=N_{R}({\mathbb{L}}\cap Y), of 𝕃Y{\mathbb{L}}\cap Y in YY is qi embedded in YY and hence it is a hyperbolic metric space by its own right. Also, this is a subbundle of YY by Corollary 4.7.

(1) The first part is an informal way of saying that (𝕃Y)\partial({\mathbb{L}}^{\prime}_{Y}) is a point. However, this is immediate from Proposition 6.6 and Corollary 6.19.

(2) By Lemma 2.55 ΛY(𝕃Y)\Lambda_{Y}({\mathbb{L}}^{\prime}_{Y}) is the image of the CT map for the inclusion 𝕃YY{\mathbb{L}}^{\prime}_{Y}\rightarrow Y since 𝕃Y{\mathbb{L}}^{\prime}_{Y} is qi embedded in YY. But 𝕃Y\partial{\mathbb{L}}^{\prime}_{Y} is a point by the first part. Thus ΛY(𝕃Y)\Lambda_{Y}({\mathbb{L}}^{\prime}_{Y}) is a singleton. Finally, ΛY(𝕃Y)=ΛY(𝕃Y)\Lambda_{Y}({\mathbb{L}}^{\prime}_{Y})=\Lambda_{Y}({\mathbb{L}}\cap Y) by Lemma 2.52. Hence we are done.

(3) Lastly, it follows that 𝕃Y{\mathbb{L}}\cap Y is quasiconvex in XX too since by Corollary 6.19 𝕃Y{\mathbb{L}}\cap Y is the union of qi lifts of β\beta contained in 𝕃Y{\mathbb{L}}\cap Y all of which converge to the same point of X\partial X. Hence 𝕃Y{\mathbb{L}}^{\prime}_{Y} is also quasiconvex in XX. Since YY is properly embedded in XX by Lemma 5.20 and 𝕃Y{\mathbb{L}}^{\prime}_{Y} is qi embedded in YY it follows that 𝕃Y{\mathbb{L}}^{\prime}_{Y} is properly embedded in XX. Thus 𝕃Y{\mathbb{L}}^{\prime}_{Y} is qi embedded in XX by Lemma 2.24(2). As in (2) we are done by Lemma 2.55. ∎

Corollary 6.21.

We have Y(2)(F)=ξ,Y(2)(F)=ξ,X(2)(F)\partial^{(2)}_{Y}(F)=\partial^{(2)}_{\xi,Y}(F)=\partial^{(2)}_{\xi,X}(F).

In particular, each ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F) is a closed subset of (2)F\partial^{(2)}F.

Proof.

The first equality follows from Lemma 6.17 applied to the metric bundle YY over AA. We will now prove the second one. Since iF,X=iY,XiF,Y\partial i_{F,X}=\partial i_{Y,X}\circ\partial i_{F,Y}, clearly ξ,Y(2)(F)ξ,X(2)(F)\partial^{(2)}_{\xi,Y}(F)\subset\partial^{(2)}_{\xi,X}(F). The opposite inclusion is an immediate consequence of Lemma 6.18.

Since iF,Y\partial i_{F,Y} is continuous it follows that ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F) is a closed subset of (2)F\partial^{(2)}F. One has to use the standard fact that the Gromov boundaries are Hausdorff spaces. ∎

The following three results are motivated by similar results proved in [Mit97]. The proof ideas are very similar. However, we get rid of the group actions that were there and in our setting properness is never needed.

Definition 6.22.

Suppose Z1,Z2Z_{1},Z_{2} are hyperbolic metric spaces. Suppose f:Z1Z2f:Z_{1}\rightarrow Z_{2} is a metrically proper map that admits the CT map. If γZ1\gamma\subset Z_{1} is a quasigeodesic line such that f(γ())=f(γ())\partial f(\gamma(\infty))=\partial f(\gamma(-\infty)) then we refer to γ\gamma as a leaf of the CT lamination Z2(2)(Z1)\partial^{(2)}_{Z_{2}}(Z_{1}).

We recall that in our context the quasigeodesic lines are assumed to be κ0\kappa_{0}-quasigeodesic lines.

Lemma 6.23.

Suppose ξ1ξ2B\xi_{1}\neq\xi_{2}\in\partial B. Given D>0D>0 there exists R=R6.23(D)>0R=R_{\ref{transversality}}(D)>0 such that the following holds:

Suppose γ1\gamma_{1} is a leaf of ξ1,X(2)(F)\partial^{(2)}_{\xi_{1},X}(F) and γ2\gamma_{2} is a leaf of ξ2,X(2)(F)\partial^{(2)}_{\xi_{2},X}(F). Then γ1ND(γ2)\gamma_{1}\cap N_{D}(\gamma_{2}) has diameter less than RR.

Proof.

Let α\alpha be a κ0\kappa_{0}-quasigeodesic line in BB joining ξ1,ξ2\xi_{1},\xi_{2}. Let b0αb^{\prime}_{0}\in\alpha be a nearest point projection of b0b_{0} on α\alpha. Let cc be a geodesic in BB joining b0b_{0} to b0b^{\prime}_{0}. Let αi\alpha_{i} be the concatenation of cc with the portion of α\alpha joining b0b^{\prime}_{0} to ξi\xi_{i}, i=1,2i=1,2. We note that κ0\kappa_{0}-quasigeodesics in BB are D2.17(δ0,κ0,κ0)D_{\ref{stab-qg}}(\delta_{0},\kappa_{0},\kappa_{0})-quasiconvex by stability of quasigeodesics. Let K=D2.17(δ0,κ0,κ0)K=D_{\ref{stab-qg}}(\delta_{0},\kappa_{0},\kappa_{0}). Hence, αi\alpha_{i}’s are K2.25(δ0,K,κ0,1)K_{\ref{subqc-elem}}(\delta_{0},K,\kappa_{0},1)-quasigeodesics by Lemma 2.25(2). Let k=K2.25(δ0,K,κ0,1)k=K_{\ref{subqc-elem}}(\delta_{0},K,\kappa_{0},1).

Next suppose xi,xiγix_{i},x^{\prime}_{i}\in\gamma_{i}, i=1,2i=1,2 are such that dF(x1,x2)Dd_{F}(x_{1},x_{2})\leq D and dF(x1,x2)Dd_{F}(x^{\prime}_{1},x^{\prime}_{2})\leq D. Let Σi,Σi{\Sigma}_{i},{\Sigma}^{\prime}_{i} be two qi sections in each 𝕃i=𝕃(γi(),γi()){\mathbb{L}}_{i}={\mathbb{L}}(\gamma_{i}(\infty),\gamma_{i}(-\infty)) passing through xix_{i} and xix^{\prime}_{i} respectively, i=1,2i=1,2. Let α~i\tilde{\alpha}_{i} and α~i\tilde{\alpha}^{\prime}_{i} be lifts of αi\alpha_{i} in 𝕃i{\mathbb{L}}_{i} through xix_{i} and xix^{\prime}_{i} respectively for i=1,2i=1,2. We now look at the quasigeodesic hexagon in XX with vertices xi,xi,ξix_{i},x^{\prime}_{i},\xi_{i}, i=1,2i=1,2 where α~i\tilde{\alpha}_{i}’s and α~i\tilde{\alpha}^{\prime}_{i}’s form four sides and the other two sides are formed by geodesics joining x1x_{1} to x2x_{2} and x1x^{\prime}_{1} to x2x^{\prime}_{2} respectively. We note that the infinite sides of this polygon are all (kK¯0+k+K¯0)(k\bar{K}_{0}+k+\bar{K}_{0})-quasigeodesics. Let k~=kK¯0+k+K¯0\tilde{k}=k\bar{K}_{0}+k+\bar{K}_{0}. Hence, such a hexagon is R2.39(δ,k~,6)R_{\ref{ideal polygons are slim}}(\delta,\tilde{k},6)-slim by Corollary 2.39. Let R1=R2.39(δ,k~,6)R_{1}=R_{\ref{ideal polygons are slim}}(\delta,\tilde{k},6). Let b2b_{2} be a point on α2\alpha_{2} such that dB(b2,α1)=D+R1+1=Rd_{B}(b_{2},\alpha_{1})=D+R_{1}+1=R, say and let y2=α~2(b2)y_{2}=\tilde{\alpha}_{2}(b_{2}). Then y2NR1(α~2)y_{2}\in N_{R_{1}}(\tilde{\alpha}^{\prime}_{2}). In particular, y2NR(Σ2)y_{2}\in N_{R}({\Sigma}^{\prime}_{2}). Hence, by Lemma 4.13 db2(Σ2Fb2,Σ2Fb2)R4.13(K¯0,R)d_{b_{2}}({\Sigma}_{2}\cap F_{b_{2}},{\Sigma}^{\prime}_{2}\cap F_{b_{2}})\leq R_{\ref{distance from qi section}}(\bar{K}_{0},R). It follows by bounded flaring that db0(x2,x2)μk~(R4.13(K¯0,R))d_{b_{0}}(x_{2},x^{\prime}_{2})\leq\mu_{\tilde{k}}(R_{\ref{distance from qi section}}(\bar{K}_{0},R)). ∎

Lemma 6.24.

If ξnξ\xi_{n}\rightarrow\xi in B\partial B, (zn,wn)ξn,X(2)(F)(z_{n},w_{n})\in\partial^{(2)}_{\xi_{n},X}(F) and (zn,wn)(z,w)(2)F(z_{n},w_{n})\rightarrow(z,w)\in\partial^{(2)}F. Then (z,w)ξ,X(2)(F)(z,w)\in\partial^{(2)}_{\xi,X}(F).

Proof.

Since iF,X(zn)=iF,X(wn)\partial i_{F,X}(z_{n})=\partial i_{F,X}(w_{n}) for all nn and iF,X\partial i_{F,X} is continuous it follows that iF,X(z)=iF,X(w)\partial i_{F,X}(z)=\partial i_{F,X}(w) whence (z,w)X(2)F(z,w)\in\partial^{(2)}_{X}F. Let [zn,wn],[zn,z],[wn,w][z_{n},w_{n}],[z_{n},z],[w_{n},w] and [z,w][z,w] denote κ0\kappa_{0}-quasigeodesic lines in FF joining these pairs of points. Let x[z,w]Fx\in[z,w]\cap F and let α\alpha be a κ0\kappa_{0}-quasigeodesic ray in BB joining bb to ξ\xi.

Claim: There is a uniform qi lift α~\tilde{\alpha} of α\alpha through xx such that α~()=iF,X(z)=iF,X(w)\tilde{\alpha}(\infty)=\partial i_{F,X}(z)=\partial i_{F,X}(w).

Since znzz_{n}\rightarrow z and wnww_{n}\rightarrow w by Lemma 2.45(1), we have db0(x,[zn,z])d_{b_{0}}(x,[z_{n},z])\rightarrow\infty and db0(x,[wn,w])d_{b_{0}}(x,[w_{n},w])\rightarrow\infty. Hence, by Corollary 2.39 there is NN\in{\mathbb{N}} such that db0(x,[zn,wn])R=R2.39(δ0,κ0,4)d_{b_{0}}(x,[z_{n},w_{n}])\leq R=R_{\ref{ideal polygons are slim}}(\delta_{0},\kappa_{0},4) for all nNn\geq N. Now, let xn[zn,wn]x_{n}\in[z_{n},w_{n}] such that db0(x,xn)Rd_{b_{0}}(x,x_{n})\leq R. Let αn\alpha_{n} be a κ0\kappa_{0}-quasigeodesic ray in BB joining bb to ξn\xi_{n}. Then by Corollary 6.19 we know that there is a uniform qi lift α~n\tilde{\alpha}_{n} of each αn\alpha_{n}, nNn\geq N such that α~n(0)=xn\tilde{\alpha}_{n}(0)=x_{n} and α~n()=iF,X(zn)\tilde{\alpha}_{n}(\infty)=\partial i_{F,X}(z_{n}). Hence, by Lemma 6.9 and Lemma 6.4 there is a qi lift α~\tilde{\alpha} starting from xx such that α~()=iF,X(z)=iF,X(w)\tilde{\alpha}(\infty)=\partial i_{F,X}(z)=\partial i_{F,X}(w). This proves the claim.

However, this means that iF,X(z)=iF,X(w)ξX\partial i_{F,X}(z)=\partial i_{F,X}(w)\in\partial^{\xi}X. Therefore, (z,w)ξ,X(2)(F)(z,w)\in\partial^{(2)}_{\xi,X}(F). ∎

6.2.2. Leaves of CT laminations for pullback bundles

The following result is motivated by a similar result proved in [KS] for trees of hyperbolic spaces which in turn was suggested by Mahan Mj. We gratefully acknowledge the same.

Suppose we have the hypotheses of Theorem 5.2. We identify YY as a subspace of XX and AA as a subspace of BB. Similarly, A\partial A is identified as a subset of B\partial B. With that in mind, we have the following:

Theorem 6.25.

Suppose we have a metric graph bundle satisfying the hypotheses of Theorem 5.2 such that the fibers of the bundle are all proper metric spaces. Suppose γ\gamma is a quasigeodesic line in YY such that (γ(),γ())X(2)(Y)(\gamma(\infty),\gamma(-\infty))\in\partial^{(2)}_{X}(Y). Let F=FbF=F_{b} be any fiber of YY.

Then (1) γ(±)iF,Y(F)\gamma(\pm\infty)\in\partial i_{F,Y}(\partial F).

(2) There is a point ξBA\xi\in\partial B\setminus\partial A determined by γ(±)\gamma(\pm\infty) such that if z±Fz_{\pm}\in\partial F with F,Y(z±)=γ(±)\partial_{F,Y}(z_{\pm})=\gamma(\pm) then (z+,z)ξ,X(2)(F)(z_{+},z_{-})\in\partial^{(2)}_{\xi,X}(F).

(3) π(γ)\pi(\gamma) is bounded. Moreover, γ\gamma is within a finite Hausdorff distance from a κ0\kappa_{0}-quasigeodesic line σ\sigma of FF so that iF,Y(σ(±))=γ(±)\partial i_{F,Y}(\sigma(\pm\infty))=\gamma(\pm\infty). Also, (σ(),σ())ξ,X(2)(F)(\sigma(\infty),\sigma(-\infty))\in\partial^{(2)}_{\xi,X}(F) for some ξBA\xi\in\partial B\setminus\partial A.

(4) If bb is a nearest point projection of ξ\xi on AA. Then σ\sigma (as defined in (3)) is a uniform quasigeodesic line in YY.

Proof.

We have Y=ΛY(F)(ξAξY)\partial Y=\Lambda_{Y}(F)\cup(\cup_{\xi\in\partial A}\partial^{\xi}Y) by Proposition 6.6. Also since FF is a proper metric space, by Lemma 2.55 ΛY(F)=iF,Y(F)\Lambda_{Y}(F)=\partial i_{F,Y}(\partial F). Thus Y=iF,Y(F)(ξAξY)\partial Y=\partial i_{F,Y}(\partial F)\cup(\cup_{\xi\in\partial A}\partial^{\xi}Y). We shall use the following observation a few times in the proof which are immediate from the fact that AA is qi embedded in BB.

Suppose α\alpha is a quasigeodesic ray in AA and α~\tilde{\alpha} is a qi lift of α\alpha in YY. Then α~\tilde{\alpha} is a quasigeodesic ray in YY as well as in XX. Also any pair of such rays are asymptotic in YY if and only if they are asymptotic in XX since YY is properly embedded in XX.

(1) The proof of this assertion is by elimination of the possibilities coming from the decomposition iF,Y(F)(ξAξY)\partial i_{F,Y}(\partial F)\cup(\cup_{\xi\in\partial A}\partial^{\xi}Y) of Y\partial Y.

Suppose γ()ξ1Y\gamma(\infty)\in\partial^{\xi_{1}}Y and γ()ξ2Y\gamma(-\infty)\in\partial^{\xi_{2}}Y for some ξ1,ξ2A\xi_{1},\xi_{2}\in\partial A. However, this case is not possible due to the above observation.

Suppose γ()ξY\gamma(\infty)\in\partial^{\xi}Y for some ξA\xi\in\partial A and γ()iF,Y(F)ξAξY\gamma(-\infty)\in\partial i_{F,Y}(\partial F)\setminus\cup_{\xi\in\partial A}\partial^{\xi}Y or vice versa. We show below that this case is also not possible.

Let α\alpha be a κ0\kappa_{0}-quasigeodesic ray in AA joining bb to ξ\xi and let α~\tilde{\alpha} be a K0K_{0}-qi lift of α\alpha in YY such that α~()=γ()\tilde{\alpha}(\infty)=\gamma(\infty). Also let β\beta be a κ0\kappa_{0}-quasigeodesic ray in FF such that iF,Y(β())=γ()\partial i_{F,Y}(\beta(\infty))=\gamma(-\infty). Now, for all nn\in{\mathbb{N}} let Σn{\Sigma}_{n} be a K0K_{0}-qi section in XX passing through β(n)\beta(n) and let 𝕃n=𝕃(Σn,β()){\mathbb{L}}_{n}={\mathbb{L}}({\Sigma}_{n},\beta(\infty)). Then 𝕃n{\mathbb{L}}_{n} is λ¯0\bar{\lambda}_{0}-quasiconvex in XX. Clearly γ()=α~()ΛX(𝕃n)\gamma(\infty)=\tilde{\alpha}(\infty)\in\Lambda_{X}({\mathbb{L}}_{n}). Hence, by Lemma 2.54 α~\tilde{\alpha} is asymptotic to 𝕃n{\mathbb{L}}_{n}. It follows by Proposition 4.6 and Lemma 4.15 that π𝕃n(α~)\pi_{{\mathbb{L}}_{n}}(\tilde{\alpha}) is a uniform qi lift of α\alpha and it is asymptotic to α~\tilde{\alpha}. Since YY properly embedded in XX by Lemma 5.20, it follows that these qi lifts are asymptotic in YY too. In particular, π𝕃n(α~)()=γ()\pi_{{\mathbb{L}}_{n}}(\tilde{\alpha})(\infty)=\gamma(\infty). Now, since dF(β(0),β(n))d_{F}(\beta(0),\beta(n))\rightarrow\infty, by Lemma 4.15 dY(β(0),π𝕃n(α~))d_{Y}(\beta(0),\pi_{{\mathbb{L}}_{n}}(\tilde{\alpha}))\rightarrow\infty. It follows from Lemma 2.45 that limnβ(n)=γ()\lim_{n\rightarrow\infty}\beta(n)=\gamma(\infty) in Y\partial Y. This gives a contradiction since limnβ(n)=γ()γ()\lim_{n\rightarrow\infty}\beta(n)=\gamma(-\infty)\neq\gamma(\infty).

Therefore, the only possibility is that

γ(±)iF,Y(F)ξAξY\gamma(\pm\infty)\in\partial i_{F,Y}(\partial F)\setminus\cup_{\xi\in\partial A}\partial^{\xi}Y

proving part (1) of the theorem.

Let z,zFz,z^{\prime}\in\partial F be such that iF,Y(z)=γ()\partial i_{F,Y}(z)=\gamma(\infty) and iF,Y(z)=γ()\partial i_{F,Y}(z^{\prime})=\gamma(-\infty).

(2) Since iF,X=iY,XiF,Y\partial i_{F,X}=\partial i_{Y,X}\circ\partial i_{F,Y} by Lemma 2.50(1), we have (z,z)X(2)(F)(z,z^{\prime})\in\partial^{(2)}_{X}(F) and hence (z,z)ξ,X(2)(F)(z,z^{\prime})\in\partial^{(2)}_{\xi,X}(F) for some ξB\xi\in\partial B by Lemma 6.17. From Corollary 6.21 it follows that ξBA\xi\in\partial B\setminus\partial A. This proves part (2) of the theorem.

(3) Let 𝕃=𝕃(z;z){\mathbb{L}}={\mathbb{L}}(z;z^{\prime}) be the bi-infinite ladder in XX formed by z,zz,z^{\prime}. Let σ=𝕃F\sigma={\mathbb{L}}\cap F which is an arc length parameterized κ0\kappa_{0}-quasigeodesic line in FF joining z,zz,z^{\prime}. Let α\alpha be a κ0\kappa_{0}-quasigeodesic ray in BB joining bb to ξ\xi.

Let Σn\Sigma_{n} be a K¯0\bar{K}_{0}-qi section in 𝕃{\mathbb{L}} passing through σ(n)\sigma(n), nn\in{\mathbb{N}}. By Corollary 6.19 qi lifts of α\alpha contained in these qi sections are asymptotic. Denote the qi section of α\alpha contained in Σn{\Sigma}_{n} by α~n\tilde{\alpha}_{n}. We note that these are k=(K¯0κ0+K¯0+κ0)k=(\bar{K}_{0}\kappa_{0}+\bar{K}_{0}+\kappa_{0})-quasigeodesics by Lemma 2.3(2). Hence, by Lemma 2.38 given m,nm,n\in{\mathbb{N}} we have α~n(i)NR(α~m)\tilde{\alpha}_{n}(i)\in N_{R}(\tilde{\alpha}_{-m}) (and α~m(i)NR(α~n)\tilde{\alpha}_{-m}(i)\in N_{R}(\tilde{\alpha}_{n})) where R=D2.38(δ,k)R=D_{\ref{ideal triangles are slim}}(\delta,k) as long as α~n(i)\tilde{\alpha}_{n}(i) (resp. α~m(i)\tilde{\alpha}_{-m}(i)) is not contained in the RR-neighborhood of any 11-quasigeodesic joining σ(m),σ(n)\sigma(-m),\sigma(n). In particular for such ii we have α~n(i)NR(Σm)\tilde{\alpha}_{n}(i)\in N_{R}({\Sigma}_{-m}), α~m(i)NR(Σn)\tilde{\alpha}_{-m}(i)\in N_{R}({\Sigma}_{n}). Hence, by Lemma 4.13 we have

dα(i)(α~n(i),α~m(i))R1=R4.13(R,K¯0)d_{\alpha(i)}(\tilde{\alpha}_{n}(i),\tilde{\alpha}_{-m}(i))\leq R_{1}=R_{\ref{distance from qi section}}(R,\bar{K}_{0})

for all such ii. Let R2=max{R1,MK¯0}R_{2}=\max\{R_{1},M_{\bar{K}_{0}}\}. Thus for all nn\in{\mathbb{N}}, Un=UR2(Σn,Σn)U_{n}=U_{R_{2}}({\Sigma}_{n},{\Sigma}_{-n})\neq\emptyset. Let bnUnb_{n}\in U_{n} be a nearest point projection of bb on UnU_{n} and let bnb^{\prime}_{n} be a nearest point projection of bnb_{n} on AA. Then it follows from Lemma 5.18 that the concatenation of the segments of α~n,α~n\tilde{\alpha}_{n},\tilde{\alpha}_{-n} over the portion of α\alpha joining b,bnb,b^{\prime}_{n} and the fiber geodesic segment 𝕃Fbn{\mathbb{L}}\cap F_{b^{\prime}_{n}} is a uniform quasigeodesic in YY joining σ(±n)\sigma(\pm n). Call it γn\gamma^{\prime}_{n}. Since limnσ(n)limnσ(n)\lim_{n\rightarrow\infty}\sigma(n)\neq\lim_{n\rightarrow\infty}\sigma(-n) in YY there is a constant D0D\geq 0 such that dY(σ(0),γn)Dd_{Y}(\sigma(0),\gamma^{\prime}_{n})\leq D by Lemma 2.34. We claim that this means dB(b,bn)d_{B}(b,b^{\prime}_{n}) is bounded. In fact dY(σ(0),α~±n)d_{Y}(\sigma(0),\tilde{\alpha}_{\pm n})\rightarrow\infty by Lemma 4.13. Thus for all large nn we have dY(σ(0),𝕃Fbn)Dd_{Y}(\sigma(0),{\mathbb{L}}\cap F_{b^{\prime}_{n}})\leq D whence dB(b,bn)Dd_{B}(b,b^{\prime}_{n})\leq D. It follows from Proposition 4.6(3) that the Hausdorff distance of 𝕃Fbn{\mathbb{L}}\cap F_{b^{\prime}_{n}} and the segment of σ\sigma between σ(n)\sigma(n) and σ(n)\sigma(-n) is at most (1+2K0)C4.6(K¯0)(1+2K_{0})C_{\ref{ladders are qi embedded}}(\bar{K}_{0}). Since σ\sigma is a proper embedding in YY it follows by Lemma 2.5 that σ\sigma is a uniform quasigeodesic in YY depending on DD. Let K1K\geq 1 be such that both σ\sigma and γ\gamma are KK-quasigeodesics in YY. Then, since YY is δ\delta^{\prime}-hyperbolic, Hd(σ,γ)R2.39(δ,K,2)Hd(\sigma,\gamma)\leq R_{\ref{ideal polygons are slim}}(\delta^{\prime},K,2). Thus diam(π(γ))R2.39(δ,K,2)diam(\pi(\gamma))\leq R_{\ref{ideal polygons are slim}}(\delta^{\prime},K,2).

We note here that diam(π(γ))diam(\pi(\gamma)) as well as the quasigeodesic constant of σ\sigma depends only on max{dB(b,bn)}\max\{d_{B}(b,b^{\prime}_{n})\}.

(4) Use shall use the notation of the proof of (3). Thus we know that there is Dn0D_{n}\geq 0 such that for all iDni\geq D_{n} we have dα(i)(α~n(i),α~n(i))R1d_{\alpha(i)}(\tilde{\alpha}_{n}(i),\tilde{\alpha}_{-n}(i))\leq R_{1} whence α(i)Un\alpha(i)\in U_{n} for all iDni\geq D_{n}. Also we know that the sets UnU_{n} are K4.11(K¯0)K_{\ref{qc-level-set-new}}(\bar{K}_{0})-quasiconvex in BB by Lemma 4.11. Let tnmax{Dn,dB(b,bn)}t_{n}\geq\max\{D_{n},d_{B}(b,b_{n})\}. Then α(tn)Un\alpha(t_{n})\in U_{n}. Thus [b,bn]B[bn.α(tn)]B[b,b_{n}]_{B}*[b_{n}.\alpha(t_{n})]_{B} is a K2.25(δ0,K4.11(K¯0),κ0,ϵ)K_{\ref{subqc-elem}}(\delta_{0},K_{\ref{qc-level-set-new}}(\bar{K}_{0}),\kappa_{0},\epsilon)-quasigeodesic segment. Let K=max{κ0,K2.25(δ0,K4.11(K¯0),κ0,ϵ)}K^{\prime}=\max\{\kappa_{0},K_{\ref{subqc-elem}}(\delta_{0},K_{\ref{qc-level-set-new}}(\bar{K}_{0}),\kappa_{0},\epsilon)\}. Hence, by stability of quasigeodesics (Lemma 2.17) we get that bnNR(α)b_{n}\in N_{R^{\prime}}(\alpha) where R=D2.17(δ0,K,K)R^{\prime}=D_{\ref{stab-qg}}(\delta_{0},K^{\prime},K^{\prime}). We also note that dB(b,bn)d_{B}(b,b_{n})\rightarrow\infty by the bounded flaring condition (Lemma 3.12) since db(α~n(0),α~n(0))d_{b}(\tilde{\alpha}_{n}(0),\tilde{\alpha}_{-n}(0))\rightarrow\infty. This implies that bnξb_{n}\rightarrow\xi. Hence by Lemma 2.56 there exists N>0N>0 such that d(bn,b)R2.56(δ0,k0)d(b^{\prime}_{n},b)\leq R_{\ref{qc last lemma}}(\delta_{0},k_{0}) for all nNn\geq N since BB is δ0\delta_{0}-hyperbolic and AA is k0k_{0}-quasiconvex. Hence, we are done by the note left at the end of the proof of (3). ∎

Surjectivity of the CT maps

Theorem 6.26.

Suppose we have the hypotheses of Theorem 5.2 such that the fibers of the bundle are proper metric spaces. Let FF be the fiber over a point bAb\in A. Suppose the CT map iF,X:FX\partial i_{F,X}:\partial F\rightarrow\partial X is surjective. Then the CT map iF,Y:FY\partial i_{F,Y}:\partial F\rightarrow\partial Y is also surjective.

Conversely for any geodesic ray α:[0,)B\alpha:[0,\infty)\rightarrow B with α(0)=b\alpha(0)=b let Yα=π1(α)Y_{\alpha}=\pi^{-1}(\alpha). If for all zBz\in\partial B and for some (any) geodesic ray α\alpha joining bb to zz the CT map F,Yα:FYα\partial_{F,Y_{\alpha}}:\partial F\rightarrow\partial Y_{\alpha} is surjective then the CT map F,X:FX\partial_{F,X}:\partial F\rightarrow\partial X is also surjective.

Proof.

Let ξY\xi\in\partial Y. We want to show that ξIm(iF,Y)\xi\in Im(\partial i_{F,Y}). Since iF,X:FX\partial i_{F,X}:\partial F\rightarrow\partial X is surjective there exists zFz\in\partial F such that iF,X(z)=iY,X(ξ)\partial i_{F,X}(z)=\partial i_{Y,X}(\xi). If iF,Y(z)=ξ\partial i_{F,Y}(z)=\xi we are done. Suppose not. However, iF,X=iY,XiF,Y\partial i_{F,X}=\partial i_{Y,X}\circ\partial i_{F,Y}. Hence, iY,X(iF,Y(z))=iY,X(ξ)\partial i_{Y,X}(\partial i_{F,Y}(z))=\partial i_{Y,X}(\xi). Then by Theorem 6.25(3) we are done.

The converse part is a direct consequence of Corollary 6.5 and Proposition 6.6. ∎

Corollary 6.27.

Suppose π:XB\pi:X\rightarrow B is a metric (graph) bundle such that X,BX,B are hyperbolic and the fibers are all proper, uniformly quasiisometric to the hyperbolic plane 2\mathbb{H}^{2}. Then for all bBb\in B, the CT map Fb,X:FbX\partial_{F_{b},X}:\partial F_{b}\rightarrow\partial X is surjective.

Proof.

This is an immediate consequence of the second part of Theorem 6.26 and the following proposition of Bowditch. ∎

Proposition 6.28.

([Bow13, Proposition 2.6.1]) Suppose π:XB\pi:X\rightarrow B is a metric (graph) bundle where B=[0,)B=[0,\infty), XX is hyperbolic and the fibers are all uniformly quasiisometric to the hyperbolic plane 2\mathbb{H}^{2}. Then for all bBb\in B, the CT map Fb,X:FbX\partial_{F_{b},X}:\partial F_{b}\rightarrow\partial X is surjective.

We would like to remark that Bowditch stated the above proposition in case the fibers are all isometric to the hyperbolic plane, but the same proof goes through for fibers uniformly quasiisometric to the hyperbolic plane.

A special case of the following result was proved by E. Field ([Fie20, Theorem B]).

Theorem 6.29.

Suppose 1NGπQ11\rightarrow N\rightarrow G\stackrel{{\scriptstyle\pi}}{{\rightarrow}}Q\rightarrow 1 is a short exact sequence of infinite hyperbolic groups. Suppose AQA\subset Q is qi embedded and Y=π1(A)Y=\pi^{-1}(A). Then the CT map NY\partial N\rightarrow\partial Y is surjective.

Proof.

Since NN is a normal subgroup of the hyperbolic group GG it is a standard fact that Λ(N)=G\Lambda(N)=\partial G. Thus by Lemma 2.55 the CT map NG\partial N\rightarrow\partial G is surjective. Now we are done by Corollary 6.26. ∎

Fibers of the CT maps

Theorem 6.30.

Suppose XX is a metric (graph) bundle over BB satisfying the hypotheses of Theorem 5.2 such that XX is a proper metric space. Let F=FbF=F_{b} where bBb\in B. Suppose F\partial F is not homeomorphic to a dendrite and also the CT map FX\partial F\rightarrow\partial X is surjective.

Then for all ξB\xi\in\partial B we have ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F)\neq\emptyset.

Proof.

Suppose α\alpha is an arc length parameterized κ0\kappa_{0}-quasigeodesic ray in BB joining bb to ξ\xi. Let Y=π1(α)Y=\pi^{-1}(\alpha). Since the CT map FX\partial F\rightarrow\partial X is surjective, the map iF,Y:FY\partial i_{F,Y}:\partial F\rightarrow\partial Y is also surjective by Theorem 6.26. Now, ξ,X(2)(F)=ξ,Y(2)(F)\partial^{(2)}_{\xi,X}(F)=\partial^{(2)}_{\xi,Y}(F) by Corollary 6.21. Hence, it is enough to show that ξ,Y(2)(F)\partial^{(2)}_{\xi,Y}(F)\neq\emptyset. However, ξ,Y(2)(F)=\partial^{(2)}_{\xi,Y}(F)=\emptyset if and only if iF,Y\partial i_{F,Y} is injective. It follows that ξ,Y(2)(F)=\partial^{(2)}_{\xi,Y}(F)=\emptyset if and only if iF,Y\partial i_{F,Y} is bijective. Since XX is proper, so are FF and YY. Hence, F\partial F and Y\partial Y are compact metrizable spaces. (See [BH99, Theorem Proposition 3.7, Proposition 3.21, Chapter III.H] for instance.) Hence, iF,Y\partial i_{F,Y} is bijective implies iF,Y\partial i_{F,Y} is a homeomorphism between F\partial F and Y\partial Y. Since F\partial F is not a dendrite this is impossible due to the following result of Bowditch. Hence, ξ,Y(2)(F)\partial^{(2)}_{\xi,Y}(F)\neq\emptyset. ∎

Theorem 6.31.

([Bow13, Proposition 2.5.2]) Suppose XX is hyperbolic metric (graph) bundle over B=[0,)B=[0,\infty) satisfying the hypotheses H1-H4 of section 5. Suppose moreover that XX is a proper metric space. Then X\partial X is a dendrite.

We note that a special case of interest of Theorem 6.30 is when the fibers are uniformly quasiisometric to the hyperbolic plane. For instance, we have the following.

Corollary 6.32.

Suppose we have an exact sequence of infinite hyperbolic groups 1NGQ11\rightarrow N\rightarrow G\rightarrow Q\rightarrow 1 where NN is either the fundamental group of an orientable closed surface of genus g2g\geq 2 or a free group FnF_{n} on n3n\geq 3 generators. Then for all ξQ\xi\in\partial Q, ξ,G(2)(N)\partial^{(2)}_{\xi,G}(N)\neq\emptyset.

Remark 20.

We remark that much stronger results than the above corollary were already proved by Mj and Rafi in [MR18]. For instance, see Theorem 3.12, Theorem 5.7 and Proposition 5.8 there.

Another context is that of complexes of groups where Theorem 6.30 can be applied.

Corollary 6.33.

Suppose GG is the fundamental group of a finite developable complexes of nonelementary hyperbolic groups (𝒢,𝒴)({\mathcal{G}},{\mathcal{Y}}) with qi condition. Suppose XX is the metric bundle over BB obtained from this data as constructed in Example 3.3.2. Suppose GG is hyperbolic.

Then for all ξB\xi\in\partial B and any vertex group GvG_{v}, vV(𝒴)v\in V({\mathcal{Y}}) we have ξ,G(2)(Gv)\partial^{(2)}_{\xi,G}(G_{v})\neq\emptyset.

Proof.

We need to check the hypotheses of Theorem 6.30. It is a standard fact that the boundary of a hyperbolic group is not a dendrite. Since the fibers of the metric bundle under consideration are quasiisometric to nonelementary hyperbolic groups F\partial F is not a dendrite for any fiber FF. We also note that the metric bundle satisfies H1-H4 of section 5. Finally, GG acts on XX and BB so that the map π:XB\pi:X\rightarrow B is equivariant, the action of GG on XX is proper and cocompact and on BB is cocompact. Thus any orbit map GXG\rightarrow X is a qi by Milnor-Svarc lemma and therefore induces a homeomorphism XG\partial X\rightarrow\partial G.

Now, given any fiber FF and gGg\in G, gFgF is another fiber of the metric bundle. By Lemma 3.10(1) Hd(F,gF)<Hd(F,gF)<\infty. Hence, by Lemma 2.52 Λ(F)=Λ(gF)=gΛ(F)\Lambda(F)=\Lambda(gF)=g\Lambda(F). It is a standard fact that the action of a nonelementary hyperbolic group on its boundary is minimal, i.e. the only invariant closed subsets are the empty set and the whole set. Hence, it follows that Λ(F)=X\Lambda(F)=\partial X. By Lemma 2.55 we have Λ(F)=iF,X(F)\Lambda(F)=\partial i_{F,X}(\partial F). Thus the CT map iF,X:FX\partial i_{F,X}:\partial F\rightarrow\partial X is surjective. Finally, clearly XX is a proper metric space. Hence, we have ξ,X(2)(F)\partial^{(2)}_{\xi,X}(F)\neq\emptyset by Theorem 6.30. Finally since GvG_{v} acts properly and cocompactly on XvX_{v}, any orbit map GvXvG_{v}\rightarrow X_{v} is a quasiisometry. Hence, this induces a homeomorphism GvXv\partial G_{v}\rightarrow\partial X_{v}. Therefore, taking F=XvF=X_{v} we are done.∎

Definition 6.34.

Suppose ZZ is any hyperbolic metric space and SZS\subset Z. Then a point zΛ(S)Zz\in\Lambda(S)\subset\partial Z will be called a conical limit point of SS if for some (any) quasigeodesic γ\gamma converging to zz in ZZ there is a constant D>0D>0 such that ND(γ)SN_{D}(\gamma)\cap S is a subset of infinite diameter in ZZ.

Proposition 6.35.

Suppose we have the hypotheses of Theorem 5.2. Let iY,X:YX\partial i_{Y,X}:\partial Y\to\partial X be the CT map. If ξX\xi\in\partial X is a conical limit point of YY, then |iY,X1(ξ)|=1|\partial i_{Y,X}^{-1}(\xi)|=1.

Proof.

Suppose zzYz\neq z^{\prime}\in\partial Y such that iY,X(z)=iY,X(z)=ξ\partial i_{Y,X}(z)=\partial i_{Y,X}(z^{\prime})=\xi. Then by Theorem 6.25 there is ξBBA\xi_{B}\in\partial B\setminus\partial A and a qi lift of γ\gamma of a quasigeodesic ray joining bb to ξB\xi_{B} such that ξ=γ()\xi=\gamma(\infty). Since ξBBA\xi_{B}\in\partial B\setminus\partial A and AA is quasiconvex ξB\xi_{B} is not a limit point of AA in B\partial B. Thus it is clear that ξ\xi is not a conical limit point of YY. This gives a contradiction and proves the proposition.∎

6.3. QI embedding fibers in a product of bundles

The lemma below is the product of answering a question due to Misha Kapovich.

Lemma 6.36.

Suppose π:X\pi:X\rightarrow{\mathbb{R}} is a metric (graph) bundle satisfying the hypotheses of section 5 and X±X^{\pm} are the restrictions of it to [0,)[0,\infty) and (,0](-\infty,0] respectively. Then the diagonal embedding f:F0X+×Xf:F_{0}\rightarrow X^{+}\times X^{-} is a qi embedding where the latter is given the l2l_{2} metric.

Proof.

Without loss of generality, we assume (X,d)(X,d) is a metric graph bundle. Let d±d_{\pm} be the induced length metric on X±X^{\pm} respectively. Then the l2l_{2} metric dYd_{Y} on Y:=X+×XY:=X^{+}\times X^{-} is given by dY((x1,x2),(y1,y2))2=d+(x1,y1)2+d(x2,y2)2d_{Y}((x_{1},x_{2}),(y_{1},y_{2}))^{2}=d_{+}(x_{1},y_{1})^{2}+d_{-}(x_{2},y_{2})^{2} for all x1,y1X+x_{1},y_{1}\in X^{+} and x2,y2Xx_{2},y_{2}\in X^{-}. We note that the inclusion maps F0X±F_{0}\rightarrow X^{\pm} are 11-Lipschitz.

Let x,yF0x,y\in F_{0}. Then, dY(f(x),f(y))2=dY((x,x),(y,y))2=d+(x,y)2+d(x,y)2d0(x,y)2+d0(x,y)2=2d0(x,y)2d_{Y}(f(x),f(y))^{2}=d_{Y}((x,x),(y,y))^{2}=d_{+}(x,y)^{2}+d_{-}(x,y)^{2}\leq d_{0}(x,y)^{2}+d_{0}(x,y)^{2}=2d_{0}(x,y)^{2}, which implies that dY(f(x),f(y))2d0(x,y)d_{Y}(f(x),f(y))\leq\sqrt{2}d_{0}(x,y). A reverse inequality is obtained as follows.

Let Σ,Σ{\Sigma},{\Sigma}^{\prime} be a pair of K0K_{0}-qi sections in XX through x,yx,y respectively. Let 𝕃=𝕃(Σ,Σ){\mathbb{L}}={\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) be the ladder formed by them. Let λ=𝕃F0\lambda={\mathbb{L}}\cap F_{0}. This is a geodesic in F0F_{0} joining x,yx,y. Now, suppose c(x,y)c(x,y) is a uniform quasigeodesic in XX joining x,yx,y constructed as in section 5 by decomposing 𝕃{\mathbb{L}} into subladders using the the qi sections Σi{\Sigma}_{i}’s and Σj{\Sigma}^{\prime}_{j}’s. Let c¯+:=c¯+(x,y),c¯:=c¯(x,y)\bar{c}_{+}:=\bar{c}_{+}(x,y),\bar{c}_{-}:=\bar{c}_{-}(x,y) be the modified paths joining x,yx,y in X+,XX^{+},X^{-} respectively. By our main theorem in section 5, c¯+,c¯\bar{c}_{+},\bar{c}_{-} are uniform quasigeodesics in X+,XX^{+},X^{-} respectively. Suppose these are KK-quasigeodesics. As in the discussion at the end of section 5, suppose 𝒬,𝒬{\mathcal{Q}},{\mathcal{Q}}^{\prime} are consecutive qi sections in the decomposition of 𝕃=𝕃(Σ,Σ){\mathbb{L}}={\mathbb{L}}({\Sigma},{\Sigma}^{\prime}) and z,w𝒬,w𝒬z,w\in{\mathcal{Q}},w^{\prime}\in{\mathcal{Q}}^{\prime} with b=π(w)=π(w)b^{\prime}=\pi(w)=\pi(w^{\prime}) are such that 𝕃(𝒬,𝒬)c(y,y){\mathbb{L}}({\mathcal{Q}},{\mathcal{Q}}^{\prime})\cap c(y,y^{\prime}) is made of the fiber geodesic [w,w]b[w,w^{\prime}]_{b^{\prime}} and the lift of [π(z),π(w)]B[\pi(z),\pi(w)]_{B} in 𝒬{\mathcal{Q}}. However, if b[0,)b^{\prime}\in[0,\infty) then λ𝕃(𝒬,𝒬)c¯\lambda\cap{\mathbb{L}}({\mathcal{Q}},{\mathcal{Q}}^{\prime})\subset\bar{c}_{-} and similarly if b(,0]b^{\prime}\in(-\infty,0] then λ𝕃(𝒬,𝒬)c¯+\lambda\cap{\mathbb{L}}({\mathcal{Q}},{\mathcal{Q}}^{\prime})\subset\bar{c}_{+}. Thus λc¯+c¯\lambda\subset\bar{c}_{+}\cup\bar{c}_{-}. Therefore we have,

d0(x,y)\displaystyle d_{0}(x,y) l+(c~+)+l(c~)Kd+(x,y)+K+Kd(x,y)+K\displaystyle\leq l_{+}(\tilde{c}_{+})+l_{-}(\tilde{c}_{-})\leq Kd_{+}(x,y)+K+Kd_{-}(x,y)+K
=K(d+(x,y)+d(x,y))+2K\displaystyle=K(d_{+}(x,y)+d_{-}(x,y))+2K
=2KdY((x,x),(y,y))+2K=2KdY(f(x),f(y))+2K.\displaystyle=2Kd_{Y}((x,x),(y,y))+2K=2Kd_{Y}(f(x),f(y))+2K.

Thus, 1+12Kd0(x,y)dY(f(x),f(y))2d0(x,y)-1+\frac{1}{2K}d_{0}(x,y)\leq d_{Y}(f(x),f(y))\leq\sqrt{2}d_{0}(x,y). Hence, ff is (2K,1)(2K,1)-qi embedding. ∎

In the same way, we obtain the following.

Lemma 6.37.

If v0v_{0} is a cut point of BB and removing it produces two quasiconvex subsets A1,A2A_{1},A_{2} and Y1,Y2Y_{1},Y_{2} are the restrictions of the bundle to A1,A2A_{1},A_{2} respectively then the diagonal map Fv0Y1×Y2F_{v_{0}}\rightarrow Y_{1}\times Y_{2} is a qi embedding.

Corollary 6.38.

If v0v_{0} is a cut point of BB and removing it produces finitely many quasiconvex subsets AiA_{i}, 1in1\leq i\leq n and YiY_{i}’s are the restrictions of the bundle to AiA_{i}’s respectively then the diagonal map Fv0ΠiYiF_{v_{0}}\rightarrow\Pi_{i}Y_{i} is a qi embedding.

Remark 21.

In [Mit97] Mitra defined an ending lamination for an exact sequence of groups. Given any point ξQ\xi\in\partial Q he defined a lamination Λξ\Lambda_{\xi} and then showed that Λξ=ξ,X(2)(F)\Lambda_{\xi}=\partial^{(2)}_{\xi,X}(F). However, for formulating and proving these sorts of results one needs additional structure on the bundle, e.g. action of a group on the bundle through morphisms which has uniformly bounded quotients when restricted to the fibers. Results of this type are proved in [MR18, Section 3]; see also [Bow13, Section 4.4].

Acknowledgements: The authors gratefully acknowledge all the helpful comments, inputs, and suggestions received from Mahan Mj and Michael Kapovich. We are very thankful to the anonymous referee also for suggesting many changes that helped to improve the exposition of the paper and for pointing out a number of gaps and inaccuracies in an earlier version of the paper. The second author was partially supported by DST INSPIRE grant DST/INSPIRE/04/2014/002236 and DST MATRICS grant MTR/2017/000485 of the Govt of India. Finally, we thank Sushil Bhunia for a careful reading of an earlier draft of the paper and for making numerous helpful suggestions.

Appendix A Flaring in metric bundle and its canonical metric graph bundle

Suppose π:XB\pi^{\prime}:X^{\prime}\rightarrow B^{\prime} is an (η,c)(\eta,c)-metric bundle and π:XB\pi:X\rightarrow B is the canonical metric graph bundle associated to it. We shall assume that BB^{\prime} and BB are both δ\delta-hyperbolic. However, there will be no assumption about the fibers of the bundles. We shall freely use the notation from section 4 of the paper. The purpose of this appendix is to show that a metric bundle satisfies a sort of ’generalized flaring property’ (see property ()(\dagger) below) iff the associated canonical metric graph bundle satisfies a flaring condition.

Note: If b0,b1,,bnb_{0},b_{1},\cdots,b_{n} are consecutive vertices on a geodesic in BB then α:ibi\alpha^{\prime}:i\mapsto b_{i} is a dotted (1,3)(1,3)-quasigeodesic of BB^{\prime} by Lemma 2.8. Thus there is a constant D0D_{0} such that if β\beta^{\prime} is any (1,1)(1,1)-quasigeodesic in BB^{\prime} joining b0,bnb_{0},b_{n} then Hd(α,β)D0Hd(\alpha^{\prime},\beta^{\prime})\leq D_{0}. We will preserve D0D_{0} to denote this constant for the rest of this section.

Suppose bV(B)b\in V(B) and pBp\in B^{\prime} are such that dB(p,b)D0d_{B^{\prime}}(p,b)\leq D_{0}. Then for any xπ1(b)x\in\pi^{-1}(b) we can lift a (1,1)(1,1)-quasigeodesic of BB^{\prime} joining bb to pp to XX^{\prime} which starts from xx and ends at xx^{\prime}, say. This way we get a ‘fiber identification map’ V(π1(b))π1(p)V(\pi^{-1}(b))\rightarrow\pi^{\prime-1}(p). If we denote this map by fbpf_{bp} then we have the following lemma. Since the proof is evident we skip it.

Lemma A.1.

We have C0+1C0db(x,y)dp(fbp(x),fbp(y))C0+C0db(x,y)-C_{0}+\frac{1}{C_{0}}d_{b}(x,y)\leq d^{\prime}_{p}(f_{bp}(x),f_{bp}(y))\leq C_{0}+C_{0}d_{b}(x,y) for all x,yπ1(b)x,y\in\pi^{-1}(b) and for some uniform constant C0C_{0} where dbd_{b} is the fiber distance in π1(b)\pi^{-1}(b) for the metric graph bundle XX and dpd^{\prime}_{p} is the fiber distance in π1(p)\pi^{\prime-1}(p) for the metric bundle XX^{\prime}.

Suppose α\alpha is a geodesic in BB and α~\tilde{\alpha} is a CC-qi lift of α\alpha in XX. Let α\alpha^{\prime} be a (1,1)(1,1)-quasigeodesic in BB^{\prime} joining the end points of α\alpha. Let σ:αα\sigma:\alpha\rightarrow\alpha^{\prime} be any map such that dB(b,σ(b))D0d_{B^{\prime}}(b,\sigma(b))\leq D_{0} for all bαb\in\alpha. Let p~=fbp(α~(b))π1(p)\tilde{p}=f_{bp}(\tilde{\alpha}(b))\in\pi^{\prime-1}(p) for all bαb\in\alpha where p=σ(b)p=\sigma(b). Now it is easy to find a uniform qi lift α~\tilde{\alpha}^{\prime} of α\alpha^{\prime} such that α~(σ(b))=p~\tilde{\alpha}^{\prime}(\sigma(b))=\tilde{p} where p=σ(b)p=\sigma(b) for all bαb\in\alpha. We record this as a lemma.

Lemma A.2.

There is a constant CC^{\prime} depending on CC and a CC^{\prime}-qi lift α~\tilde{\alpha}^{\prime} of α\alpha^{\prime} such that α~(σ(b))=p~\tilde{\alpha}^{\prime}(\sigma(b))=\tilde{p} where p=σ(b)p=\sigma(b) for all bαb\in\alpha.

The following lemma roughly says that if two qi leaves start flaring in one direction then they keep on flaring in the same direction. The proof follows immediately from the definition of flaring. One may also look up the proof of [MS12, Lemma 2.17(1)].

Lemma A.3.

(Persistence of flaring in graph bundles) Suppose the metric graph bundle satisfies (νk,Mk,nk)(\nu_{k},M_{k},n_{k})-flaring condition for all k1k\geq 1. Suppose α:[m,n]B\alpha:[-m,n]\rightarrow B is a geodesic where mnk,nnkm\geq n_{k},n\geq n_{k} and α~1\tilde{\alpha}_{1} and α~2\tilde{\alpha}_{2} are two kk-qi lifts of α\alpha in XX with dα(0)(α~1(0),α~2(0))Mkd_{\alpha(0)}(\tilde{\alpha}_{1}(0),\tilde{\alpha}_{2}(0))\geq M_{k}. Suppose

dα(snk)(α~1(snk),α~2(snk))νkdα(0)(α~1(0),α~2(0))d_{\alpha(sn_{k})}(\tilde{\alpha}_{1}(sn_{k}),\tilde{\alpha}_{2}(sn_{k}))\geq\nu_{k}d_{\alpha(0)}(\tilde{\alpha}_{1}(0),\tilde{\alpha}_{2}(0))

where ss is either 11 or 1-1. Let tt be the largest integer smaller than n/nkn/n_{k} or m/nkm/n_{k} according as s=1s=1 or 1-1. Then for all integer 1lt1\leq l\leq t we have

dα(lsnk)(α~1(lsnk),α~2(lsnk))νkldα(0)(α~1(0),α~2(0)).d_{\alpha(lsn_{k})}(\tilde{\alpha}_{1}(lsn_{k}),\tilde{\alpha}_{2}(lsn_{k}))\geq\nu^{l}_{k}d_{\alpha(0)}(\tilde{\alpha}_{1}(0),\tilde{\alpha}_{2}(0)).

The same idea of proof gives the next lemma also. We will need a definition.

Property ()(\dagger): We shall say that the metric bundle XX^{\prime} has the property ()(\dagger) if for any k1k\geq 1, there exist νk>1\nu_{k}>1 and nk,Mkn_{k},M_{k}\in\mathbb{N} such that the following holds:

Suppose α:[nk,nk]B\alpha^{\prime}:[-n_{k},n_{k}]\rightarrow B^{\prime} is a 11-quasigeodesic and α~1\tilde{\alpha}^{\prime}_{1} and α~2\tilde{\alpha}^{\prime}_{2} are two kk-qi lifts of α\alpha^{\prime} in XX^{\prime}. If dγ(0)(γ1~(0),γ2~(0))Mkd_{\gamma(0)}(\tilde{\gamma_{1}}(0),\tilde{\gamma_{2}}(0))\geq M_{k} then we have

νkdα(0)(α~1(0),α~2(0))max{dα(nk)(α~1(nk),α~2(nk)),dα(nk)(α~1(nk),α~2(nk))}.\nu_{k}\cdot d_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}_{1}(0),\tilde{\alpha}^{\prime}_{2}(0))\leq\max\{d_{\alpha^{\prime}(n_{k})}(\tilde{\alpha}^{\prime}_{1}(n_{k}),\tilde{\alpha}^{\prime}_{2}(n_{k})),d_{\alpha^{\prime}(-n_{k})}(\tilde{\alpha}^{\prime}_{1}(-n_{k}),\tilde{\alpha}^{\prime}_{2}(-n_{k}))\}.

Note that one could define flaring condition for a length metric bundle using the property ()(\dagger).

Lemma A.4.

(Persistence of flaring in metric bundles) Suppose the metric bundle satisfies ()(\dagger). Let k1k\geq 1. Suppose α:[m,n]B\alpha^{\prime}:[-m,n]\rightarrow B is a geodesic where mnk,nnkm\geq n_{k},n\geq n_{k} and α~1\tilde{\alpha}^{\prime}_{1} and α~2\tilde{\alpha}^{\prime}_{2} are two kk-qi lifts of α\alpha^{\prime} in XX with dα(0)(α~1(0),α~2(0))Mkd_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}_{1}(0),\tilde{\alpha}^{\prime}_{2}(0))\geq M_{k}. Suppose

dα(snk)(α~1(snk),α~2(snk))νkdα(0)(α~1(0),α~2(0))d_{\alpha^{\prime}(sn_{k})}(\tilde{\alpha}^{\prime}_{1}(sn_{k}),\tilde{\alpha}^{\prime}_{2}(sn_{k}))\geq\nu_{k}d_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}_{1}(0),\tilde{\alpha}^{\prime}_{2}(0))

where ss is either 11 or 1-1. Let tt be the largest integer smaller than or equal to n/nkn/n_{k} or m/nkm/n_{k} according as s=1s=1 or 1-1. Then for all integer ltl\leq t we have

dα(lsnk)(α~1(lsnk),α~2(lsnk))νkldα(0)(α~1(0),α~2(0)).d_{\alpha^{\prime}(lsn_{k})}(\tilde{\alpha}^{\prime}_{1}(lsn_{k}),\tilde{\alpha}^{\prime}_{2}(lsn_{k}))\geq\nu^{l}_{k}d_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}_{1}(0),\tilde{\alpha}^{\prime}_{2}(0)).

Following is one of the main results of this appendix.

Lemma A.5.

Suppose the metric bundles XX^{\prime} has the property ()(\dagger). Then the canonical metric graph bundle π:XB\pi:X\rightarrow B associated to XX^{\prime} satisfies a (ν^k,M^k,λ^k)(\hat{\nu}_{k},\hat{M}_{k},\hat{\lambda}_{k})-flaring condition.

In particular if a geodesic metric bundle satisfies flaring condition (see [MS12, Definition 1.12]) then its canonical metric graph bundle satisfies flaring condition.

Proof.

Suppose α:[n,n]B\alpha:[-n,n]\rightarrow B is a geodesic and α~,α~~\tilde{\alpha},\tilde{\tilde{\alpha}} are two kk-qi lifts of α\alpha in XX where nn\in{\mathbb{N}} and k1k\geq 1. Let α\alpha^{\prime} be a (1,1)(1,1)-quasigeodesic in BB^{\prime} joining α(n),α(n)\alpha(n),\alpha(-n). Then there are kk^{\prime}-qi lifts α~,α~~\tilde{\alpha}^{\prime},\tilde{\tilde{\alpha}}^{\prime} of α\alpha^{\prime} respectively as in Lemma A.2. We shall choose a parametrization α:[m,n]B\alpha^{\prime}:[-m^{\prime},n^{\prime}]\rightarrow B^{\prime} so that α(n)=α(n),α(m)=α(n)\alpha^{\prime}(n^{\prime})=\alpha(n),\alpha^{\prime}(-m^{\prime})=\alpha(-n) and dB(α(0),α(0))D0d_{B^{\prime}}(\alpha(0),\alpha^{\prime}(0))\leq D_{0}. Note that dα(0)(α~(0),α~~(0))C0+1C0dα(0)(α~(0),α~~(0))d^{\prime}_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}(0),\tilde{\tilde{\alpha}}^{\prime}(0))\geq-C_{0}+\frac{1}{C_{0}}d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0)) by Lemma A.1. Hence, if we assume dα(0)(α~(0),α~~(0))C0(C0+Mk)d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0))\geq C_{0}(C_{0}+M_{k^{\prime}}) then dα(0)(α~(0),α~~(0))Mkd^{\prime}_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}(0),\tilde{\tilde{\alpha}}^{\prime}(0))\geq M_{k^{\prime}}. Clearly, if we choose nn large enough then we have nk<min{m,n}n_{k^{\prime}}<\min\{m^{\prime},n^{\prime}\}. (In the course of the proof we will be more precise.) Without loss of generality we shall assume νkdα(0)(α~(0),α~~(0))dα(nk)(α~(nk),α~~(nk))\nu_{k}\cdot d^{\prime}_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}(0),\tilde{\tilde{\alpha}}^{\prime}(0))\leq d^{\prime}_{\alpha^{\prime}(n_{k^{\prime}})}(\tilde{\alpha}^{\prime}(n_{k^{\prime}}),\tilde{\tilde{\alpha}}^{\prime}(n_{k^{\prime}})). Let ll be the greatest integer less than or equal to n/nkn^{\prime}/n_{k^{\prime}}. Then by Lemma A.4

(1) νkldα(0)(α~(0),α~~(0))dα(lnk)(α~(lnk),α~~(lnk)).\nu^{l}_{k}\cdot d^{\prime}_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}(0),\tilde{\tilde{\alpha}}^{\prime}(0))\leq d^{\prime}_{\alpha^{\prime}(ln_{k^{\prime}})}(\tilde{\alpha}^{\prime}(ln_{k^{\prime}}),\tilde{\tilde{\alpha}}^{\prime}(ln_{k^{\prime}})).

Note that dB(α(lnk),α(n))2nkd_{B^{\prime}}(\alpha^{\prime}(ln_{k^{\prime}}),\alpha^{\prime}(n^{\prime}))\leq 2n_{k^{\prime}}. Let b=α(lnk)b^{\prime}=\alpha^{\prime}(ln_{k^{\prime}}) and b′′=α(n)b^{\prime\prime}=\alpha^{\prime}(n^{\prime}). Then by Corollary 3.11 the fiber identification map ϕbb′′\phi_{b^{\prime}b^{\prime\prime}} referred to in that corollary is a K3.11(2nk)K_{\ref{fibers unif qi}}(2n_{k^{\prime}})-quasiisometry. Let K=K3.11(2nk)K=K_{\ref{fibers unif qi}}(2n_{k^{\prime}}). Note that

(2) dX(α~(lnk),α~(n))k+2nkkd_{X^{\prime}}(\tilde{\alpha}^{\prime}(ln_{k^{\prime}}),\tilde{\alpha}^{\prime}(n^{\prime}))\leq k^{\prime}+2n_{k^{\prime}}k^{\prime}

since α~\tilde{\alpha}^{\prime} is a kk^{\prime}-qi section and dB(α(lnk),α(n))2nkd_{B^{\prime}}(\alpha^{\prime}(ln_{k^{\prime}}),\alpha^{\prime}(n^{\prime}))\leq 2n_{k^{\prime}}. Also, by Corollary 3.9 we have

(3) dX(α~(lnk),ϕbb′′(α~(lnk)))3c+6cnk.d_{X^{\prime}}(\tilde{\alpha}^{\prime}(ln_{k^{\prime}}),\phi_{b^{\prime}b^{\prime\prime}}(\tilde{\alpha}^{\prime}(ln_{k^{\prime}})))\leq 3c+6cn_{k^{\prime}}.

Using the inequalities (2) and (3) we have

dX(α~(n),ϕbb′′(α~(lnk)))3c+6cnk+k+2nkk.d_{X^{\prime}}(\tilde{\alpha}^{\prime}(n^{\prime}),\phi_{b^{\prime}b^{\prime\prime}}(\tilde{\alpha}^{\prime}(ln_{k^{\prime}})))\leq 3c+6cn_{k^{\prime}}+k^{\prime}+2n_{k^{\prime}}k^{\prime}.

Since XX^{\prime} is an (η,c)(\eta,c)-metric bundle we have

dα(n)(α~(n),ϕbb′′(α~(lnk)))η(3c+6cnk+k+2nkk).d^{\prime}_{\alpha^{\prime}(n^{\prime})}(\tilde{\alpha}^{\prime}(n^{\prime}),\phi_{b^{\prime}b^{\prime\prime}}(\tilde{\alpha}^{\prime}(ln_{k^{\prime}})))\leq\eta(3c+6cn_{k^{\prime}}+k^{\prime}+2n_{k^{\prime}}k^{\prime}).

In the same way we have

dα(n)(α~~(n),ϕbb′′(α~~(lnk)))η(3c+6cnk+k+2nkk).d^{\prime}_{\alpha^{\prime}(n^{\prime})}(\tilde{\tilde{\alpha}}^{\prime}(n^{\prime}),\phi_{b^{\prime}b^{\prime\prime}}(\tilde{\tilde{\alpha}}^{\prime}(ln_{k^{\prime}})))\leq\eta(3c+6cn_{k^{\prime}}+k^{\prime}+2n_{k^{\prime}}k^{\prime}).

Now using the fact that ϕbb′′\phi_{b^{\prime}b^{\prime\prime}} is a KK-quasiisometry and letting R1=2η(3c+6cnk+k+2nkk)R_{1}=2\eta(3c+6cn_{k^{\prime}}+k^{\prime}+2n_{k^{\prime}}k^{\prime}) we have by triangle inequality

(4) dα(lnk)(α~(lnk),α~~(lnk))(K2+2R1K)+Kdα(n)(α~(n),α~~(n)).d^{\prime}_{\alpha^{\prime}(ln_{k^{\prime}})}(\tilde{\alpha}^{\prime}(ln_{k^{\prime}}),\tilde{\tilde{\alpha}}^{\prime}(ln_{k^{\prime}}))\leq(K^{2}+2R_{1}K)+Kd^{\prime}_{\alpha^{\prime}(n^{\prime})}(\tilde{\alpha}^{\prime}(n^{\prime}),\tilde{\tilde{\alpha}}^{\prime}(n^{\prime})).

However, by Lemma 2.8 and Proposition 4.1(2) we have

(5) dα(n)(α~(n),α~~(n))3+dα(n)(α~(n),α~~(n))d^{\prime}_{\alpha^{\prime}(n^{\prime})}(\tilde{\alpha}^{\prime}(n^{\prime}),\tilde{\tilde{\alpha}}^{\prime}(n^{\prime}))\leq 3+d_{\alpha(n)}(\tilde{\alpha}(n),\tilde{\tilde{\alpha}}(n))

Then it follows from the inequalities (1), (4) and (5) that

(6) νkldα(0)(α~(0),α~~(0))RK+Kdα(n)(α~(n),α~~(n))\nu^{l}_{k}\cdot d^{\prime}_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}(0),\tilde{\tilde{\alpha}}^{\prime}(0))\leq RK+Kd_{\alpha(n)}(\tilde{\alpha}(n),\tilde{\tilde{\alpha}}(n))

where R=3+K+2R1R=3+K+2R_{1}. Finally since dα(0)(α~(0),α~~(0))C0+1C0dα(0)(α~(0),α~~(0))d^{\prime}_{\alpha^{\prime}(0)}(\tilde{\alpha}^{\prime}(0),\tilde{\tilde{\alpha}}^{\prime}(0))\geq-C_{0}+\frac{1}{C_{0}}d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0)) using (6) we have

(7) νkl(C0+1C0dα(0)(α~(0),α~~(0)))RK+Kdα(n)(α~(n),α~~(n)).\nu^{l}_{k^{\prime}}(-C_{0}+\frac{1}{C_{0}}d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0)))\leq RK+Kd_{\alpha(n)}(\tilde{\alpha}(n),\tilde{\tilde{\alpha}}(n)).

Recall that we assumed dα(0)(α~(0),α~~(0)))C0(C0+Mk)d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0)))\geq C_{0}(C_{0}+M_{k^{\prime}}). Hence,

(8) C0+1C0dα(0)(α~(0),α~~(0))(1C01C0+Mk)dα(0)(α~(0),α~~(0)).-C_{0}+\frac{1}{C_{0}}d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0))\geq(\frac{1}{C_{0}}-\frac{1}{C_{0}+M_{k^{\prime}}})d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0)).

Let

λ=1K(1C01C0+Mk).\lambda=\frac{1}{K}(\frac{1}{C_{0}}-\frac{1}{C_{0}+M_{k^{\prime}}}).

Then we have, using (7) and (8),

(9) νklλdα(0)(α~(0),α~~(0))R+dα(n)(α~(n),α~~(n)).\nu^{l}_{k^{\prime}}\cdot\lambda\cdot d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0))\leq R+d_{\alpha(n)}(\tilde{\alpha}(n),\tilde{\tilde{\alpha}}(n)).

It is clear that

R+νklλdα(0)(α~(0),α~~(0))12λνkldα(0)(α~(0),α~~(0))-R+\nu^{l}_{k^{\prime}}\cdot\lambda\cdot d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0))\geq\frac{1}{2}\lambda\nu^{l}_{k^{\prime}}d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0))

if dα(0)(α~(0),α~~(0))2Rλνkl.d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0))\geq\frac{2R}{\lambda\nu^{l}_{k^{\prime}}}. In particular, since νkl>1\nu^{l}_{k^{\prime}}>1, we have

(10) 12λνkldα(0)(α~(0),α~~(0))dα(n)(α~(n),α~~(n))\frac{1}{2}\lambda\nu^{l}_{k^{\prime}}d_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0))\leq d_{\alpha(n)}(\tilde{\alpha}(n),\tilde{\tilde{\alpha}}(n))

using (9) if dα(0)(α~(0),α~~(0))2Rλd_{\alpha(0)}(\tilde{\alpha}(0),\tilde{\tilde{\alpha}}(0))\geq\frac{2R}{\lambda}. Thus it is enough to choose

M^k=max{2Rλ,C0(C0+Mk)},λ^k=2\hat{M}_{k}=\max\{\frac{2R}{\lambda},C_{0}(C_{0}+M_{k^{\prime}})\},\,\hat{\lambda}_{k}=2

and to show that if nn is sufficiently large then ll is so large that 12λνkl2\frac{1}{2}\lambda\nu^{l}_{k^{\prime}}\geq 2 which will give a choice for n^k\hat{n}_{k}. This is easy to verify and hence left to the reader. ∎

Converse of Lemma A.5 is also true and has an exactly similar proof. However, in this case one uses Lemma A.3 instead of Lemma A.2. We state it without proof to avoid repetition.

Lemma A.6.

Suppose the metric graph bundle π:XB\pi:X\rightarrow B satisfies a (νk,Mk,nk)(\nu_{k},M_{k},n_{k})-flaring condition for all k1k\geq 1. Then the metric bundle π:XB\pi^{\prime}:X^{\prime}\rightarrow B^{\prime} satisfies the condition ()(\dagger) for three functions νk,Mk,nk\nu^{\prime}_{k},M^{\prime}_{k},n^{\prime}_{k} of kk.

References

  • [ABC+91] J. Alonso, T. Brady, D. Cooper, V. Ferlini, M. Lustig, M. Mihalik, M. Shapiro, and H. Short, Notes on word hyperbolic groups, Group Theory from a Geometrical Viewpoint (E. Ghys, A. Haefliger, A. Verjovsky eds.) (1991), 3–63.
  • [Bes04] M. Bestvina, Geometric group theory problem list, M. Bestvina’s home page: http:math.utah.edu/\simbestvina/eprints/questions-updated (2004).
  • [BF92] M. Bestvina and M. Feighn, A Combination theorem for Negatively Curved Groups, J. Differential Geom., vol 35 (1992), 85–101.
  • [BH99] M. Bridson and A Haefliger, Metric spaces of nonpositive curvature, Grundlehren der mathematischen Wissenchaften, Vol 319, Springer-Verlag (1999).
  • [Bow13] B. H. Bowditch, Stacks of hyperbolic spaces and ends of 3 manifolds, Geometry and Topology Down Under, Contemp. Math. 597 (2013), 65–138.
  • [BR13] O. Baker and T. R. Riley, Cannon-Thurston maps do not always exist, Forum Math., vol 1, e3 (2013).
  • [Cor92] J. M. Corson, Complexes of groups, Proc. London Math. Soc. (3) 65 (1992), 199–224.
  • [CT85] J. Cannon and W. P. Thurston, Group Invariant Peano Curves, preprint, Princeton (1985).
  • [DK18] C. Drutu and M. Kapovich, Geometric group theory, AMS Colloquium Publications, vol 63 (2018).
  • [Far98] B. Farb, Relatively hyperbolic groups, Geom. Funct. Anal. 8 (1998), 810–840.
  • [Fie20] E. Field, Trees, dendrites, and the Cannon-Thurston map, Algbr. Geom. Topol., vol 20(6) (2020), 3083–3126.
  • [Gd90] E. Ghys and P. de la Harpe(eds.), Sur les groupes hyperboliques d’apres Mikhael Gromov, Progress in Math. vol 83, Birkhauser, Boston Ma. (1990).
  • [Gro87] M. Gromov, Hyperbolic Groups, in Essays in Group Theory, ed. Gersten, MSRI Publ.,vol.8, Springer Verlag (1987), 75–263.
  • [Hae92] A. Haefliger, Extension of complexes of groups, Ann. Inst. Fourier, Grenoble, 42, 1-2 (1992), 275–311.
  • [Ham05] U. Hamenstadt, Word hyperbolic extensions of surface groups, preprint, arXiv:math/0505244 (2005).
  • [KS] Michael Kapovich and Pranab Sardar, Hyperbolic trees of spaces (in preparation), https://www.math.ucdavis.edu/ kapovich/EPR/trees.pdf.
  • [MG20] M. Mj and P. Ghosh, Regluing graphs of free groups, https://arxiv.org/abs/2011.01256 (2020).
  • [Min11] Honglin Min, Hyperbolic graphs of surface groups, Algebr. Geom. Topol. 11(1) (2011), 449–476.
  • [Mit97] M. Mitra, Ending Laminations for Hyperbolic Group Extensions, Geom. Funct. Anal. 7 (1997), 379–402.
  • [Mit98a] by same author, Cannon-Thurston Maps for Hyperbolic Group Extensions, Topology 37 (1998), 527–538.
  • [Mit98b] by same author, Cannon-Thurston Maps for Trees of Hyperbolic Metric Spaces, J. Differential Geom. 48 (1998), 135–164.
  • [Mj19] Mahan Mj, Cannon-thurston maps, Proceedings of the International Congress of Mathematicians (ICM 2018), ISBN 978-981-3272-87-3 (2019), pp 885–917.
  • [MR18] Mahan Mj and Kasra Rafi, Algebraic ending laminations and quasiconvexity, Algbr Geom Topol 18 (2018), 1883–1916.
  • [MS12] Mahan Mj and Pranab Sardar, A combination theorem for metric bundles, Geom. Funct. Anal. Vol. 22 (2012), 1636–1707.
  • [V0̈5] Jussi Väisälä, Gromov hyperbolic spaces, Expo. Math. 23 (2005), no. 3, 187–231. MR 2164775
  • [Why10] K. Whyte, Coarse bundles, https://arxiv.org/pdf/1006.3347.pdf (2010).