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

On the quasi-isometric rigidity of chambers and walls in cusp-decomposable manifolds

Haydeé Contreras Peruyero Instituto de Matemáticas, Universidad Nacional Autónoma de México, Coyoacán, 04510. CDMX, México [email protected]
Abstract.

A cusp-decomposable manifold is a manifold constructed from a finite number of complete, negatively curved, finite volume manifolds and identifying the boundaries of truncated cusps by diffeomorphisms. Using properties of the electric space of the universal cover of cusp-decomposable manifolds, we show that the inclusion of walls and pieces induces quasi-isometric embeddings. We also show that isomorphisms between fundamental groups of higher graph manifolds preserve the decomposition into pieces.

1. Introduction

In the theory of 3-manifolds, the first definition of a graph manifold is due to Waldhausen in 1967 [26]. Consider the following family of manifolds which extend the ideas of Frigerio, Lafont and Sisto [11].

Definition 1.

[2, Definition 1]

  1. (1)

    For every i=1,,ni=1,...,n with 2nin2\leq n_{i}\leq n, let ViV_{i} a finite volume, complete, non-compact, pinched negatively curved nin_{i}-manifold.

  2. (2)

    Denote by BiB_{i} the compact smooth manifold with boundary obtained by truncating the cusps of ViV_{i}, i.e., by removing from ViV_{i} a non-maximal horospherical open neighbourhood of each cusp.

  3. (3)

    Take fiber bundles ZiBiZ_{i}\longrightarrow B_{i} with fiber NiN_{i} a compact quotient of an aspherical nilpotent simply connected Lie group Ni~\widetilde{N_{i}} by the action of a uniform lattice Γi\Gamma_{i} of dimension nnin-n_{i}, i.e., NiN_{i} is diffeomorphic to Ni~/Γi\widetilde{N_{i}}/\Gamma_{i}.

  4. (4)

    Fix a pair of diffeomorphic boundary components between different ZiZ_{i}, provided one exists, and identify the paired boundary components using diffeomorphism, so that we obtain a connected nn-manifold.

A manifold constructed in this way is called a higher dimension graph manifold. Each ZiZ_{i} is called a piece of MM, when dim(Bi)=n\operatorname{dim}(B_{i})=n we have Zi=BiZ_{i}=B_{i} and we say that the piece is a pure piece. The boundary components of the pieces ZiZ_{i} are called walls and we denote them by WW.

These manifolds were introduced by Bárcenas, Juan-Pineda and Suárez-Serrato and have been shown to satisfy the Borel Conjecture for n6n\geq 6 [2]. This definition also includes the manifolds described by Connell and Suárez-Serrato [7].
Tam Nguyen Phan described a family of graph manifolds called cusp decomposable manifolds [25]. This is a subfamily of the manifolds constructed as in Definition 1 where the pieces are all pure pieces. This is analogous to the JSJ decomposition of 3-manifolds. The family of twisted-doubles described by Aravinda and Farrell is also included in the cusp-decomposable manifolds [1].

We can associate to the universal cover of a higher graph manifold (and a cusp decomposable manifold) a Bass-Serre tree [24]. The preimage of a vertex in the Bass-Serre tree will be called a chamber. The preimage of an edge in the Bass-Serre tree will be called a wall, (see Definition 2).

We say that a piece ZZ of MM is a surface piece if the base of the fiber bundle which defines ZZ is a hyperbolic surface.

Our two main results are:

Theorem 1.

Let M1M_{1}, M2M_{2} be two higher graph manifolds, as in definition 1 and without surface pieces. Let Δ1\Delta_{1} be a subgroup of π1(M1)\pi_{1}(M_{1}) that is conjugate to the fundamental group π1(Z1,j)\pi_{1}(Z_{1,j}) of a piece Z1,jZ_{1,j} in M1M_{1}, and φ:π1(M1)π1(M2)\varphi:\pi_{1}(M_{1})\longrightarrow\pi_{1}(M_{2}) be an isomorphism. Then φ(Δ1)\varphi(\Delta_{1}) is conjugate to the fundamental group π1(Z2,σ(j))\pi_{1}(Z_{2,\sigma(j)}) in π1(M2)\pi_{1}(M_{2}) of a piece Z2,σ(j)Z_{2,\sigma(j)} in M2M_{2}, for some permutation σ\sigma of the boundary components of the pieces.

To prove this Theorem, we begin by explaining the properties of the universal cover M~\widetilde{M} as a tree of spaces and we analyse the Bass-Serre tree TT associated to the decomposition into pieces. With these properties and the fact that the fundamental group of a higher graph manifold has uniformly exponential growth (Lemma 16), we show in Lemma 18 that the stabilizer of a vertex vv in TT can not fix any other vertex. Also, in Lemma 21 we prove the analogous property for walls. In Lemma 24, Corollary 25 and Corollary 26, we describe certain metric properties of walls and chambers that allow us to understand when two walls or two chambers are the same. Then, we prove that the maximal nilpotent subgroups of π1(M)\pi_{1}(M) with Hirsch length n1n-1 are the stabilizers of walls. This characterization of the stabilizers of walls and chambers, together with the Milnor-Svarcˇ\check{\textrm{c}} Lemma enable us to proof Theorem 1, the details are found in Section 3.

Let (X1,d1)(X_{1},\operatorname{d}_{1}) and (X2,d2)(X_{2},\operatorname{d}_{2}) be two metric spaces. A map f:X1X2f:X_{1}\longrightarrow X_{2} is called a (c,d)(c,d)-quasi-isometric embedding if there exist constants c1c\geq 1 and d0d\geq 0 such that for all x,yX1x,y\in X_{1}

1cd1(x,y)dd2(f(x),f(y))cd1(x,y)+d\frac{1}{c}\operatorname{d}_{1}(x,y)-d\leq\operatorname{d}_{2}(f(x),f(y))\leq c\operatorname{d}_{1}(x,y)+d

If, in addition, there exist a constant k0k\geq 0 such that every point in X2X_{2} lies in the kk-neighbourhood of the image of ff, then ff is called a (c,d)(c,d)-quasi-isometry and we say that X1X_{1} and X2X_{2} are quasi-isometric.

Our second main result shows that in a cusp-decomposable manifold MM the fundamental groups of walls and pieces are quasi-isometrically embedded in π1(M)\pi_{1}(M).

Theorem 2.

Let MM be a cusp decomposable manifold. Then, the inclusion of each piece and wall induces quasi-isometric embeddings of their fundamental groups into their image in π1(M)\pi_{1}(M).

This result generalizes Theorem 0.16 of Frigerio, Lafont and Sisto [11]. They mention that their strategy to prove the quasi-isometric embedding can not be generalized to the case of cusp-decomposable manifolds. In order to circumvent these difficulties we use properties of the electric space. The electric space is the space obtained by contracting the boundary of the removed horospheres in the base manifolds BiB_{i} to a point.

As a consequence of Lemma 3.2 of Osin [18], in Proposition 34 we prove that the length of geodesics lying on the horoboundaries of the pieces are bounded above by the lengths of paths that connect different horoboundaries. Relying on previous results by Farb [10], in Lemma 35 we prove that there exists a positive constant such that the length of a good path between any near two points in a wall is bounded by above by the product of this constant and the distance between the points. Also, using the properties electric spaces, in Proposition 36 we construct a non-backtracking good path between two points in the same internal wall, which is the analogous result for cusp-decomposable manifolds to Lemma 7.8 of [11]. This allows us to show that the inclusion of walls and chambers endowed with the path metric are quasi-isometrically embedded in the universal cover M~\widetilde{M} of MM.

The concept of quasi-isometry plays a fundamental role. One of the central questions is to understand what algebraic properties are invariant under quasi-isometries. Stallings shows that the property of splitting over a finite group for a finitely generated group is invariant under quasi-isometries [22, 23]. Papasoglu showed that for a finitely presented group, admitting a splitting over a virtually infinite cyclic group is also invariant under quasi-isometries [19]. Papasoglu also showed that if the vertex groups are fundamental groups of aspherical manifolds and the edge groups are smaller than the vertex groups, then the splitting is preserved under quasi-isometries [20]. As a consequence of this result, if a group GG is an amalgamated product of two aspherical 3-manifolds over a surface group, then any group quasi-isometric to GG also splits over a virtual surface group. Bowditch shows that a one-ended hyperbolic group that is not a triangle group splits over a two-ended group if and only if its Gromov boundary has local cut points. Therefore admitting such splitting is invariant under quasi-isometries for hyperbolic groups [3, 4, 5]. Another fundamental concept is the JSJ decomposition of a manifold MM. This canonical decomposition is unique up to isotopy and its corresponds to the decomposition of the fundamental group of MM as a graph of groups. Kapovich and Leeb proved that this canonical decomposition is invariant under quasi-isometries and as a consequence it is a geometric invariant of the fundamental group [16]. In this paper, we contribute to the understanding of properties that are invariant under quasi-isometries for higher graph manifolds and cusp-decomposable manifolds.

In Section 2, we start with a review fundamental notions of Hadamard manifolds, electric spaces, nilpotent groups, and growth type of groups. In Section 3, we begin with some metric properties of the universal cover of a higher graph manifold MM and we completely characterize the set of wall stabilizers. With this, and the Milnor-Svarcˇ\check{\textrm{c}} Lemma we prove that an isomorphism between the fundamental groups of two higher graph manifolds preserves the decomposition into pieces. In Section 4, we focus on cusp-decomposable manifolds and we prove Theorem 2.

Acknowledgments

This work was supported by a doctoral CONACYT fellowship and the grant IN104819 from PAPIIT-DGAPA-UNAM. The author is very grateful to Pablo Suárez-Serrrato for proposing the problem and helping through the elaboration of this paper. She also wants to thank Noé Bárcenas, Jean-François Lafont and Pierre Py for their valuable comments on this work.

2. Preliminaries

2.1. Hadamard manifolds

For the analysis of complete, finite volume Riemannian manifolds MM with pinched negative sectional curvature, we follow Eberlein [9].
Let V be a Hadamard manifold, i.e., a complete, simply connected Riemannian manifold of non-positive sectional curvature. Denote by d(x,y)\operatorname{d}(x,y) the Riemannian metric on VV and assume that all geodesics in VV have unit speed. We say that two geodesics α,γ\alpha,\gamma of VV are asymptotes if there exists a constant C>0C>0 such that d(α(t),γ(t))C\operatorname{d}(\alpha(t),\gamma(t))\leq C for all t0t\geq 0. A point at infinity of VV is an equivalence class of asymptotic geodesics. The set of all points at infinity will be denoted by V()V(\infty).
Let I(V)\operatorname{I}(V) be the group of isometries of VV. Associate to each isometry φ\varphi a displacement function dφ:pd(p,φp)\operatorname{d}_{\varphi}:p\rightarrow\operatorname{d}(p,\varphi p). An isometry φ\varphi is elliptic if dφ\operatorname{d}_{\varphi} has zero minimum, is hyperbolic if dφ\operatorname{d}_{\varphi} has positive minimum and is parabolic if dφ\operatorname{d}_{\varphi} has no minimum in VV.
Let VV be a finite volume Hadamard manifold with sectional curvature KK, and Λ\Lambda be a subgroup of the isometries of VV. We say that Λ\Lambda is a lattice of VV if Λ\Lambda acts freely and properly discontinuously on VV and the quotient manifold M=V/ΛM=V/\Lambda has finite volume. Moreover MM only has finitely many ends, each end is parabolic and Riemannian collared, and π1(M)\pi_{1}(M) is finitely presented.
Let π:VM\pi:V\rightarrow M be the projection map and denoted by γ~\tilde{\gamma} a geodesic in HH that determines a point at infinity fixed by some parabolic element φ\varphi of Γ\Gamma. We say that an end EE of MM is parabolic if there exists a divergent geodesic ray γ:[0,)M\gamma:\left[0,\infty\right)\rightarrow M that converges to EE and can be expressed as πγ~\pi\circ\tilde{\gamma}.
Suppose that the sectional curvature of VV satisfies the condition bKa<0-b\leq K\leq-a<0 for some positive constants a,ba,b. Then the maximal almost nilpotent subgroups of Λ\Lambda are the nonidentity stability groups Λx\Lambda_{x}, and each group Λx\Lambda_{x} is finitely generated. Here xx is a point at infinity in VV fixed by some parabolic element of Λ\Lambda. Also each nontrivial almost nilpotent subgroup of Λ\Lambda is contained in a unique maximal almost nilpotent subgroup.
We will recall some of the combinatorial properties of the fundamental group of such manifolds, following Farb [10].
Let M~\widetilde{M} be the universal cover of a complete, finite volume, pinched negatively curved Riemannian manifold MM, i.e., there exist constants a,b>0a,b>0 such that b2K(M)a2<0-b^{2}\leq\operatorname{K}(M)\leq-a^{2}<0. In particular, M~\widetilde{M} is a Hadamard manifold.
Let xM~x\in\widetilde{M}, zz is a point at infinity and γ\gamma is the geodesic ray from xx to zz. A horosphere through xx with center zz is the limit as tt\rightarrow\infty of the sphere in M~\widetilde{M} with center in γ(t)\gamma(t) and radius tt. Let dS(x,y)\operatorname{d}_{S}(x,y) denote the path metric in a horosphere SS. That is, if x,ySx,y\in S then dS(x,y)\operatorname{d}_{S}(x,y) is the infimum of the lengths of all paths in SS from xx to yy.

Theorem 3.

(Heintze-Im Hof [15, Theorem 4.9]) Let γ\gamma be a geodesic tangent to a horosphere SS in a pinched Hadamard manifold YY where the sectional curvature satisfies the conditions b2K(M)a2<0-b^{2}\leq\operatorname{K}(M)\leq-a^{2}<0. If p,qp,q are the projections of γ(±)\gamma(\pm\infty) onto SS, then

2bdS(p,q)2a.\frac{2}{b}\leq\operatorname{d}_{S}(p,q)\leq\frac{2}{a}.

The following Lemma is a well known fact from Riemannian geometry. For completeness we include a proof.

Lemma 4.

Let MM be a Riemannian manifold of dimk\operatorname{dim}k with γM\gamma\subseteq M a bi-infinite geodesic and GG acting on MM isometrically. If Gγ=γG\gamma=\gamma then G×HG\leq\mathbb{R}\times H, where HO(k1)×2H\leq O(k-1)\times\mathbb{Z}_{2}.

Proof.

Let γ:GM\gamma:G\rightarrow M be an infinite geodesic in MM, such that Gγ=γG\gamma=\gamma. Consider an orthonormal basis {γ(0),e1,,ek1}\{\gamma^{\prime}(0),e_{1},...,e_{k-1}\} of the tangent space at γ(0)\gamma(0). By parallel transport we have an orthonormal basis at every γ(t)\gamma(t).
Let ρ:G\rho:G\rightarrow\mathbb{R} be a morphism defined by gγ(0)=γ(ρ(g))g\cdot\gamma(0)=\gamma(\rho(g)), for each gGg\in G.
We will use the parallel transport along γ\gamma, with respect to the action of GG to define a morphism ψ\psi as follows:

ψ:GO(k)g[Tγ(0)MdgTγ(ρ(g))M]\begin{array}[]{c c c l}\psi:&G&\rightarrow&O(k)\\ &g&\mapsto&[T_{\gamma(0)}M\xrightarrow{\operatorname{d}_{g}}T_{\gamma(\rho(g))}M]\end{array}

Finally, as the tangent vectors can go to γ(t)\gamma^{\prime}(-t) or γ(t)\gamma^{\prime}(t) and they are all unitary vectors, then the image of ψ\psi is O(k1)×2O(k-1)\times\mathbb{Z}_{2}. Therefore, G×HG\leq\mathbb{R}\times H, where HO(k1)×2H\leq O(k-1)\times\mathbb{Z}_{2}. ∎

2.2. The electric space

Let M~\widetilde{M} be a pinched Hadamard manifold on which π1(M)\pi_{1}(M) acts freely, properly and discontinuous by isometries. Let Γ\Gamma be the Cayley graph associated to π1(M)\pi_{1}(M) and Γ^\widehat{\Gamma} be the coned-off Cayley graph of Γ\Gamma with respect to the cusp subgroup HH. Choose a Γi\Gamma_{i} invariant set of disjoint horoballs centered on the parabolic fixed points. Denote by ZZ the resultant space of deleting the interior of all these horoballs and endow ZZ with the path metric. Each boundary component of ZZ is a totally geodesic horosphere, and Γ\Gamma acts freely and cocompactly by isometries on ZZ. Choosing a base point xZx\in Z on a horosphere, we obtain a quasi-isometry of Γ\Gamma with ZZ given by γγx\gamma\mapsto\gamma\cdot x [10].
The electric space Z^\widehat{Z} is the quotient space obtained from ZZ by identifying points which lie in the same horospherical boundary component of ZZ. The path metric dZ\operatorname{d}_{Z} of ZZ induces a path pseudo-metric dZ^\operatorname{d}_{\widehat{Z}} on Z^\widehat{Z} as follows. Let

dY(x,y)={0, if x,yS for some horoboundary component of ZdZ(x,y), otherwise.\operatorname{d}_{Y}(x,y)=\begin{cases}0,&\textit{ if }x,y\in S\textrm{ for some horoboundary component of }Z\\ \operatorname{d}_{Z}(x,y),&\textrm{ otherwise.}\end{cases}

Then dZ^(x,y)\operatorname{d}_{\widehat{Z}}(x,y) is equal to the infimum of dY(xi,xi+1)\,\sum\operatorname{d}_{Y}(x_{i},x_{i+1})\, over all sequences of points x=x1,x2,,xn=yx=x_{1},x_{2},\dots,x_{n}=y. Observe that dZ^\operatorname{d}_{\widehat{Z}} locally agrees with the path metric dM~\operatorname{d}_{\widetilde{M}} outside the horospheres.
A path γ\gamma in Z^\widehat{Z} is called an electric path. The electric length of an electric path γ\gamma is denoted by lZ^(γ)l_{\widehat{Z}}(\gamma), it is the sum of the lengths in ZZ of the subpaths of γ\gamma lying outside every horosphere. An electric geodesic from xx to yy in Z^\widehat{Z} is a path from xx to yy such that lZ^(γ)l_{\widehat{Z}}(\gamma) is minimal. An electric P-quasi-geodesic is a PP-quasi-geodesic in the pseudo-path-metric space Z^\widehat{Z}.
Let P,CP,C be two positive constants. The (P,C)(P,C)-quasi isometry ff from Γ\Gamma to ZZ induces a quasi-isometry f^:Γ^Z^\widehat{f}:\widehat{\Gamma}\longleftrightarrow\widehat{Z}, defined by f^(v)=f(v)\widehat{f}(v)=f(v) for all vΓv\in\Gamma. Hence, we obtain the following commutative diagram:

Γ{\Gamma}Z{Z}Γ^{\widehat{\Gamma}}Z^{\widehat{Z}}f\scriptstyle{f}f^\scriptstyle{\widehat{f}}

Observe that f^\widehat{f} is a (2RP,C+1)(2RP,C+1)-quasi-isometry, where RR is such that the distance in M~\widetilde{M} between any two horospheres in the collection of removed horospheres is at least RR.
Consider a geodesic γM\gamma\in M that does not intersect a horosphere SS. The visual size of SS with respect to γ\gamma is the diameter of the set T={sS |  t for which γ(t)sS¯={s}}T=\{s\in S\textit{ }|\textit{ }\exists\textit{ }t\textrm{ for which }\overline{\gamma(t)s\cap S}=\{s\}\} in the metric dS\operatorname{d}_{S}. The visual size of SS is the supremum of the visual size of SS with respect to γ\gamma taken over all geodesics that do not intersect SS.

Lemma 5.

(Farb [10, Lemma 4.4]) Horospheres in a pinched Hadamard manifold YY have uniformly bounded visual size.

Let SM~S\subseteq\widetilde{M} be a horosphere and denote by πS\pi_{S} the orthogonal projection onto SS. Let γ\gamma be a geodesic not intersecting SS and suppose another geodesic goes from xγx\in\gamma to some point ySy\in S. Then, if dS(y,πS(x))C\operatorname{d}_{S}(y,\pi_{S}(x))\leq C for some constant CC, the visual size of SS is bounded by 2/a+2C2/a+2C.

Lemma 6.

(Farb [10, Lemma 4.5]) There exist K=K(P)K=K(P), L=L(P)>0L=L(P)>0 with P>0P>0 given with the following property: let β\beta be a electric PP-quasi-geodesic from xx to yy, and let γ\gamma be the geodesic in M~\widetilde{M} from xx to yy. Then any subpath of β\beta which lies outside of NZ^(γ,K)N_{\widehat{Z}}(\gamma,K) must have electric length at most LL. In particular, any electric PP-quasi-geodesic from xx to yy stays completely inside of NZ^(γ,K+L/2)N_{\widehat{Z}}(\gamma,K+L/2).

Another important fact is that the electric space is a δ\delta-hyperbolic pseudometric space for some δ>0\delta>0 (Proposition 4.6 of [10]). The following lemmata compare the length of an electric quasi geodesic β\beta that penetrates a sequence of horospheres with the length of a geodesic α\alpha in M~\widetilde{M} with the same end points. We will say that β\beta electrically tracks α\alpha.

Lemma 7.

(Farb [10, Lemma 4.7]) Let β\beta be an electric PP-quasi-geodesic that does not penetrate SS. Then, there exists a constant D=D(P)D=D(P) such that the projection of β\beta onto SS has a length in SS of at most DlZ^(β)D\cdot l_{\widehat{Z}}(\beta).

We will say that an electric quasi-geodesic βZ^\beta\in\widehat{Z} is a quasi-geodesic without backtracking if for each horosphere SZ^S\in\widehat{Z} which β\beta penetrates, β\beta never returns to SS after leaving SS. The image under f^:Γ^Z^\widehat{f}:\widehat{\Gamma}\rightarrow\widehat{Z} of every geodesic in Γ^\widehat{\Gamma} is a quasi-geodesic without backtracking in Z^\widehat{Z}.

Lemma 8.

(Farb [10, Lemma 4.8]) Let β\beta be an electric PP-quasi-geodesic without backtracking from xx to yy and let γ\gamma be the geodesic in M~\widetilde{M} from xx to yy. Then, there exists a constant D=D(P)D=D(P) such that if any of β\beta or γ\gamma penetrates SS, then the distance in SS from the point of entry of this path into SS to its exit point is at most DD.

Lemma 9.

(Farb [10, Lemma 4.9]) Let β\beta be an electric PP-quasi-geodesic without backtracking from xx to yy and let γ\gamma be the geodesic in M~\widetilde{M} from xx to yy. Then, there exists a constant D=D(P)D=D(P) such that if β\beta and γ\gamma penetrate some horosphere SS, then the entry point of β\beta into SS is at distance DD in SS from the entry point of γ\gamma into SS; similarly for the exit points.

Let GG be a finitely generated group and let Γ\Gamma be its Cayley graph. Let {H1,H2,,Hr}\{H_{1},H_{2},...,H_{r}\} be a finite set of finitely generated subgroups of GG. The coned-off Cayley graph of GG with respect to {H1,H2,,Hr}\{H_{1},H_{2},...,H_{r}\} denoted by Γ^=Γ^({H1,H2,,Hr})\widehat{\Gamma}=\widehat{\Gamma}(\{H_{1},H_{2},...,H_{r}\}) is formed as follows: for each gHigH_{i}, 1ir1\leq i\leq r, add a vertex v(gHi)v(gH_{i}) to Γ\Gamma and add an edge e(gHi)e(gH_{i}) of length 1/21/2 from each ghigHigh_{i}\in gH_{i} to the vertex v(gHi)v(gH_{i}).
Following Farb [10] We say that the group GG is hyperbolic relative to {H1,H2,,Hr}\{H_{1},H_{2},...,H_{r}\} if the coned-off Cayley graph Γ^\widehat{\Gamma} of GG with respect to {H1,H2,,Hr}\{H_{1},H_{2},...,H_{r}\} is a Gromov δ\delta-hyperbolic space for some δ\delta.

Theorem 10.

(Farb [10, Theorem 5.1]) Let MM be a complete, noncompact, finite volume Riemannian nn-manifold with pinched negatively sectional curvature and denote by Γ\Gamma its fundamental group. Let {H1,H2,,Hr}\{H_{1},H_{2},...,H_{r}\} be the cusp subgroups of Γ\Gamma, then Γ\Gamma is hyperbolic relative to the set {H1,H2,,Hr}\{H_{1},H_{2},...,H_{r}\} of cusp subgroups.

2.3. Nilpotent groups

A group GG is solvable if it has an abelian series, i.e., a series 1=G0G1Gn=G1=G_{0}\triangleleft G_{1}\triangleleft...\triangleleft G_{n}=G in which each factor Gi+1/GiG_{i+1}/Gi is abelian. A group GG is called nilpotent if it has a central series, i.e., a normal series 1=G0G1Gn=G1=G_{0}\leq G_{1}\leq...\leq G_{n}=G such that Gi+1/GiG_{i+1}/G_{i} is contained in the center of G/GiG/G_{i} for all ii. A nilpotent group is solvable. A group GG is said to be polycyclic if it has a cyclic series, i.e., if it has a series with cyclic factors. Polycyclic groups are solvable. Observe that if GG is nilpotent then G is polycyclic. In a polycyclic group GG, the number of infinite factors in a cyclic series is independent of the series and hence it is an invariant of GG, known as the Hirsch length [21, 5.4.13]. We will denote by h(G)\operatorname{h}(G) the Hirsch length of GG, i.e., h(G):=#{ i | Gi/Gi+1}\operatorname{h}(G):=\#\{\textrm{ }i\textrm{ }|\textrm{ }G_{i}/G_{i+1}\cong\mathbb{Z}\}.
We will use the following results about the Hirsch length in our proofs below.

Theorem 11.

(Robinson [21]) For a short exact sequence of polycyclic groups

1HGG/H11\longrightarrow H\longrightarrow G\longrightarrow G/H\longrightarrow 1

the following is true

h(G)=h(H)+h(G/H).\operatorname{h}(G)=\operatorname{h}(H)+\operatorname{h}(G/H).
Lemma 12.

(Robinson [21]) If HGH\leq G are polycyclic groups, then h(H)h(G)\operatorname{h}(H)\leq\operatorname{h}(G).

Let GG be a group with torsion free, finite index subgroups, then all such subgroups have the same cohomological dimension. This common dimension is called the virtual cohomological dimension of GG and it is denoted by vcd(G)\operatorname{vcd}(G).

Theorem 13.

(Brown [6]) If GG is a polycyclic group, then h(G)=vcd(G)\operatorname{h}(G)=\operatorname{vcd}(G), where vcd\operatorname{vcd} denotes the cohomological virtual dimension. Let MM be closed aspherical manifold of dimension nn and G=π1(M)G=\pi_{1}(M). Then vcd(G)=n\operatorname{vcd}(G)=n.

As a consequence of this Theorem, we have that h(M)=n\operatorname{h}(M)=n when MM is a infranilmanifold.

Proposition 14.

(Serre [24, Proposition 27]) Let GG be a finitely generated nilpotent group acting on a tree XX. Then only the following mutually exclusive cases are possible:

  1. (1)

    GG has a fixed point.

  2. (2)

    There is a straight path TT stable under GG on which GG acts by translations by means of a non-trivial homomorphism GG\longrightarrow\mathbb{Z}.

2.4. Injections of fundamental groups

Denote by cic_{i} the number of boundary components of each piece ZiZ_{i}, i.e., ci=#π0(Zi)c_{i}=\#\pi_{0}(\partial Z_{i}). We will denote the set of disjoint boundary components of Zi,ZjZ_{i},Z_{j} as follows:

Zi=α=1ciWαi\displaystyle\partial Z_{i}=\coprod\limits_{\alpha=1}^{c_{i}}W_{\alpha}^{i}
Zj=α=1cjWαj.\displaystyle\partial Z_{j}=\coprod\limits_{\alpha=1}^{c_{j}}W_{\alpha}^{j}.

We denote the gluing diffeomorphism between the pieces Zi,ZjZ_{i},Z_{j} by dij,k\operatorname{d}_{ij,k}, where the index kk denotes the kk-th wall of the piece ZiZ_{i}, that is, dij,k:WkiWd(k)j\operatorname{d}_{ij,k}:W_{k}^{i}\rightarrow W_{\operatorname{d}(k)}^{j}. We sometimes omit the index kk when it is clear from context. This diffeomorphism induces an isomorphism between the fundamental groups,

dij:π1(Wi)π1(Wj).\operatorname{d}_{ij_{\star}}:\pi_{1}(W^{i})\rightarrow\pi_{1}(W^{j}).

An end of ViV_{i} will be denoted by Ei,jE_{i,j}.
A well known result from Eberlein [9] tell us the following. Each ViV_{i} can be retracted to Bi=Vii=1kSiB_{i}=V_{i}-\cup_{i=1}^{k}S_{i}. Here SiS_{i} are non-maximal horospheres removed from ViV_{i}. Then, for each end EiE_{i} of ViV_{i} the map π1(Ei)π1(Bi)\pi_{1}(E_{i})\rightarrow\pi_{1}(B_{i}) is injective. We have the following commutative diagram:

0{0}π1(Ni){\pi_{1}(N_{i})}π1(Wij){\pi_{1}(W^{ij})}π1(Ei){\pi_{1}(E_{i})}0{0}0{0}π1(Ni){\pi_{1}(N_{i})}π1(Zi){\pi_{1}(Z_{i})}π1(Bi){\pi_{1}(B_{i})}0{0}

Therefore the map π1(Wij)π1(Zi)\pi_{1}(W^{ij})\hookrightarrow\pi_{1}(Z_{i}) is injective. As a consequence of this, the fundamental group of a higher graph manifold MM constructed as in Definition 1 is isomorphic to the fundamental group of a graph of groups 𝒢M\mathcal{G}_{M}. The vertex groups are the fundamental groups of the pieces π1(Zi)\pi_{1}(Z_{i}) and the edge groups are π1(Wij)\pi_{1}(W^{ij}) (see [8, 11]).

2.5. Uniformly exponential growth of the fundamental group

Let Γ\Gamma be a group finitely generated by the finite set of generators SS. The word length lS(γ)l_{S}(\gamma) of an element γΓ\gamma\in\Gamma is defined as the smallest integer nn for which there exist s1,s2,,snSS1s_{1},s_{2},...,s_{n}\in S\cup S^{-1} such that γ=s1s2sn\gamma=s_{1}s_{2}\cdots s_{n}. The word metric dS(γ1,γ2)\operatorname{d}_{S}(\gamma_{1},\gamma_{2}) is defined as the length lS(γ11γ2)l_{S}(\gamma_{1}^{-1}\gamma_{2}). With this metric the group Γ\Gamma is a metric space.

Proposition 15.

(Milnor-Svarcˇ\check{\textrm{c}} Lemma, [13]) Let XX be a proper, geodesic metric space. Let Γ\Gamma be a group acting by isometries from the left on XX. Assume that the action is proper and the quotient Γ\X\Gamma\backslash X is compact. Then the group Γ\Gamma is finitely generated and quasi-isometric to XX. Moreover, for every point x0Xx_{0}\in X, the map ΓX\Gamma\rightarrow X given by γγx0\gamma\mapsto\gamma x_{0}, is a quasi-isometry.

A useful consequence of this Lemma is that if MM is a compact Riemannian manifold with Riemannian universal covering M~\widetilde{M}, then the fundamental group of MM is quasi-isometric to M~\widetilde{M}.
The growth function β(Γ,S;k)\beta(\Gamma,S;k) is the number of elements γΓ\gamma\in\Gamma such that lS(γ)kl_{S}(\gamma)\leq k. The growth type of the pair (Γ,S)(\Gamma,S) is classified as follows:

  1. (1)

    The group Γ\Gamma is of exponential growth, if there exist constants A,B>0A,B>0 such that for n0n\geq 0 the growth function satisfies β(Γ,S;n)AexpBn\beta(\Gamma,S;n)\geq A\exp^{Bn}.

  2. (2)

    The group Γ\Gamma has polynomial growth, if there exist constants d,cd,c such that for n0n\geq 0 the growth function satisfies β(Γ,S;n)cnd\beta(\Gamma,S;n)\leq cn^{d}.

  3. (3)

    The group Γ\Gamma is of intermediate growth, if it is neither of exponential nor of polynomial growth.

The growth type of a finitely generated group is a quasi-isometry invariant, i.e., quasi-isometric finitely generated groups have the same growth type [17, Corollary 6.2.6]. If Γ\Gamma is a finitely generated nilpotent group, then Γ\Gamma has polynomial growth [17, Theorem 6.3.6].
The type of exponential growth of the pair (Γ,S)(\Gamma,S) is ω(Γ,S)=limkβ(Γ,S;k)k\omega(\Gamma,S)=\underset{k\rightarrow\infty}{\lim}\sqrt[k]{\beta(\Gamma,S;k)}. Denote by ω(Γ)=inf{ω(Γ,S):S is a finite generating set of Γ}\omega(\Gamma)=\inf\{\omega(\Gamma,S):S\ \text{ is a finite generating set of }\Gamma\}. We have the following classification:

  1. (1)

    The group Γ\Gamma has exponential growth if ω(Γ,S)>1\omega(\Gamma,S)>1.

  2. (2)

    The group Γ\Gamma has subexponential growth if ω(Γ,S)=1\omega(\Gamma,S)=1.

  3. (3)

    The group Γ\Gamma has uniformly exponential growth if ω(Γ)>1\omega(\Gamma)>1.

A well known result of de la Harpe and Bucher [14] states that if CC is a subgroup of two finitely generated groups A,BA,B and they satisfy the condition ([A:C]1)([B:C]1)2([A:C]-1)([B:C]-1)\geq 2 then the free product with amalgamation ACBA\ast_{C}B has uniformly exponential growth. Using this we obtain the following result (see [8]).

Lemma 16.

The fundamental group of a higher graph manifold MM as in Definition 1 has uniformly exponential growth.

Proof.

The proof is by induction on the pieces. Let Zi,ZjZ_{i},Z_{j} be two adjacent pieces and Wi,jW_{i,j} be their common wall. We know that the map π1(Wi,j)π1(Zi)\pi_{1}(W_{i,j})\hookrightarrow\pi_{1}(Z_{i}) is injective and the same for the map π1(Wi,j)π1(Zj)\pi_{1}(W_{i,j})\hookrightarrow\pi_{1}(Z_{j}). Then by the main Theorem of [14] the free product with amalgamation π1(Zi)π1(Wi,j)π1(Zj)\pi_{1}(Z_{i})\ast_{\pi_{1}(W_{i,j})}\pi_{1}(Z_{j}) has uniformly exponential growth.
Let ZkZ_{k} be another piece adjacent to ZjZ_{j} and Wj,kW_{j,k} their common wall. Observe that π1(Wk)\pi_{1}(W_{k}) can be seen as a subgroup of π1(Zi)π1(Wi,j)π1(Zj)\pi_{1}(Z_{i})\ast_{\pi_{1}(W_{i,j})}\pi_{1}(Z_{j}) so we can again consider the free product with amalgamation, apply the same result of [14] and conclude the proof. ∎

3. Isomorphisms preserve pieces

The objective of this section is to prove Theorem 1. Throughout this section we will assume that MM is a higher graph manifold as in definition 1, and M~\widetilde{M} is its universal covering.
Let MM be a manifold constructed as in definition 1. We will now describe the universal cover of a higher graph manifold MM. Let XX be a graph with vertices V(X)V(X) and oriented edges E(X)E(X). Consider a group GG which acts on XX. An inversion is a pair consisting of an element gGg\in G and an edge xE(X)x\in E(X) such that gx=x¯gx=\bar{x}, here x¯\bar{x} denotes the reverse orientation on xx. If there is no such pair we say that GG acts without inversion. Remember that if 𝒢\mathcal{G} is a graph of groups, the fundamental group π1(𝒢)\pi_{1}(\mathcal{G}) acts without inversion on the Bass-Serre tree TT associated to 𝒢\mathcal{G} [24, Section 5.4].
Suppose that (M~,ϕ,T)(\widetilde{M},\phi,T) is a tree of spaces, where M~\widetilde{M} is the universal cover of MM and TT is the Bass-Serre tree associated to the decomposition of MM.

Definition 2.
  1. (1)

    A wall of M~\widetilde{M} is the closure of the pre-image under ϕ\phi of the interior of an edge of TT. We will denote by dW\operatorname{d}_{W} the path metric induced on WW by the restriction to WW of the Riemannian structure of M~\widetilde{M}.

  2. (2)

    A chamber CM~C\subseteq\widetilde{M} is the pre-image under ϕ\phi of a vertex of TT. We will denote by dC\operatorname{d}_{C} the path metric induced on CC by the restriction of the Riemannian structure of M~\widetilde{M}.

We say that two chambers are adjacent if their corresponding vertices in TT are joined by an edge. A wall WW is adjacent to a chamber CC if WCW\cap C\neq\emptyset. If WW is a wall, then WW is adjacent to a chamber CC if and only if the vertex corresponding to CC is the end point of the edge corresponding to WW.
Every boundary component of ZiZ_{i} is π1\pi_{1}-injective in ZiZ_{i}. Hence every piece and every boundary component of a piece is π1\pi_{1}-injective in MM. In combination with the construction of a Bass-Serre tree of spaces we obtain the following result.

Corollary 17.

Let MM be a higher graph manifolds as in definition 1. If M~\widetilde{M} is it universal cover and (M~,ϕ,T)(\widetilde{M},\phi,T) is a tree of spaces, we have the following:

  1. (1)

    If CC is a chamber of M~\widetilde{M}, then CC is homeomorphic to Zi~Bi~×Ni~\widetilde{Z_{i}}\cong\widetilde{B_{i}}\times\widetilde{N_{i}}. Here Zi~,Bi~,Ni~\widetilde{Z_{i}},\widetilde{B_{i}},\widetilde{N_{i}} are the universal covers of ZiZ_{i}, BiB_{i} and NiN_{i}, in that order.

  2. (2)

    If WW is a wall of M~\widetilde{M}, then WW is homeomorphic to Bi~×Ni~\partial\widetilde{B_{i}}\times\widetilde{N_{i}} or Bj~×Nj~\partial\widetilde{B_{j}}\times\widetilde{N_{j}}.

Let (M~,ϕ,T)(\widetilde{M},\phi,T) be the tree of spaces given by Corollary 17. The tree TT is called the Bass-Serre tree of π1(M)\pi_{1}(M) with respect to the isomorphism π1(M)π1(𝒢M)\pi_{1}(M)\cong\pi_{1}(\mathcal{G}_{M}). Let V(T)V(T) denote the vertices of TT and E(T)E(T) denote its edges. The action of π1(M)\pi_{1}(M) on M~\widetilde{M} induces an action of π1(M)\pi_{1}(M) on TT. The fundamental group of a piece of MM coincides with the stabilizer of a vertex of TT, and the fundamental group of a wall WijW_{ij} corresponds to the stabilizer of an edge of TT.

Lemma 18.

Let MM be a higher graph manifolds as in definition 1 and TT the Bass-Serre tree associated to the decomposition of MM. For every vertex vV(T)v\in V(T) and every edge eE(T)e\in E(T) we denote by Gv,GeG_{v},G_{e} their stabilizers in π1(M)\pi_{1}(M), in that order. If vv is a vertex in TT, then vv is the unique vertex which is fixed by GvG_{v}.

Proof.

Suppose that GvG_{v} fixes another vertex vvv^{\prime}\neq v, then GvG_{v} fixes an edge ee that contains vv. This implies that GvG_{v} is contained in the stabilizer of the edge ee. By Corollary 3.3 of [9], the stabilizers of edges are virtually nilpotent groups. By a well known result from Gromov we know that virtually nilpotent groups have polynomial growth [12].
As a consequence of Lemma 16, the stabilizer of a vertex has uniformly exponential growth. Therefore, the stabilizer of an edge can not be contained in the stabilizer of a vertex. We conclude that GvG_{v} only fixes vv. ∎

Lemma 19.

Let MM be a higher graph manifolds as in definition 1. Let Z1,Z2Z_{1},Z_{2} be two pieces of MM. Let Gi<π1(M)G_{i}<\pi_{1}(M) be groups conjugate to the fundamental groups of ZiZ_{i} for i=1,2i=1,2. Then:

  1. (1)

    The normalizer of G1G_{1} in π1(M)\pi_{1}(M) is equal to G1G_{1}.

  2. (2)

    If G1G_{1} is conjugate to G2G_{2} in π1(M)\pi_{1}(M) then Z1Z_{1} is isometric to Z2Z_{2}.

Proof.

We will consider the action of π1(M)\pi_{1}(M) on the Bass-Serre tree TT.

  1. (1)

    Let v1V(T)v_{1}\in V(T), by Lemma 18, v1v_{1} is the only vertex fixed by Gv1G_{v_{1}} and therefore fixed by G1G_{1}. Let gπ1(M)g\in\pi_{1}(M) and suppose that gg normalizes G1G_{1}. So, for all xG1x\in G_{1} there exists yG1y\in G_{1} such that gx=yggx=yg. This implies gxv1=ygv1gxv_{1}=ygv_{1}, because xv1=v1xv_{1}=v_{1}, thus gv1=ygv1gv_{1}=ygv_{1}. Therefore, G1G_{1} fixes gv1gv_{1} and as v1v_{1} is the only vertex fixed by G1G_{1}, then gv1=v1gv_{1}=v_{1}. Hence, gG1g\in G_{1}.

  2. (2)

    Suppose there exists gg such that gG1g1=G2gG_{1}g^{-1}=G_{2} and let v1,v2v_{1},v_{2} be the vertices which are fixed by G1G_{1} and G2G_{2} in that order. As G1G_{1} is conjugate to G2G_{2}, G1G_{1} fixes v2v_{2} and g(v1)g(v_{1}), hence v2=g(v1)v_{2}=g(v_{1}). Therefore the covering automorphism g:M~M~g:\tilde{M}\longrightarrow\tilde{M} sends the chamber covering Z1Z_{1} to the chamber covering Z2Z_{2}, from which Z1Z_{1} is isometric to Z2Z_{2}. ∎

Remember that the graph manifold MM is formed by a finite union of pieces ZiZ_{i} and that each one is defined by a fiber bundle with base BiB_{i} and fiber NiN_{i}. We will say that an element gπ1(M)g\in\pi_{1}(M) corresponds to the fiber direction of π1(Zi)\pi_{1}(Z_{i}) if gg for ρ:π1(M)π1(Bi)\rho:\pi_{1}(M)\longrightarrow\pi_{1}(B_{i}) the morphism induced by the projection MBiM\rightarrow B_{i}, we have that ρ(g)=eπ1(Bi)\rho(g)=e\in\pi_{1}(B_{i}). That is to say, it fixes the direction associated to the base of a piece of ZiZ_{i}.

Lemma 20.

Let MM be a higher graph manifolds as in definition 1 and TT the Bass-Serre tree associated to the decomposition of MM. Let W1,W2W_{1},W_{2} be two distinct walls of M~\widetilde{M} and let vV(T)v\in V(T) be a vertex so that every path which connects W1W_{1} and W2W_{2} intersects the chamber corresponding to vv. If gπ1(M)g\in\pi_{1}(M) is such that g(Wi)=Wig(W_{i})=W_{i} for i=1,2i=1,2, then gg is an element that corresponds to the fiber direction in π1(Zi)\pi_{1}(Z_{i}), here ZiZ_{i} is the piece of MM that corresponds to the vertex vv in TT.

Proof.

Let Zi~M~\widetilde{Z_{i}}\subset\widetilde{M} be the chamber associated to vv and denote the piece of MM corresponding to Zi~\widetilde{Z_{i}} by ZiZ_{i}. Let Zi1,Zi2Z_{i}^{1},Z_{i}^{2} be the boundary components of (Zi~)\partial(\widetilde{Z_{i}}) such that g(Zi~)=Zi~g(\widetilde{Z_{i}})=\widetilde{Z_{i}}, g(Zi1)=Zi1g(Z_{i}^{1})=Z_{i}^{1} and g(Zi2)=Zi2g(Z_{i}^{2})=Z_{i}^{2}. Thus we have that gGvg\in G_{v}.
Recall that GvG_{v} is identified with the fundamental group of the piece corresponding to vv, i.e., with π1(Zi)\pi_{1}(Z_{i}), and that Zi~Bi~×Ni~\widetilde{Z_{i}}\cong\widetilde{B_{i}}\times\widetilde{N_{i}}. Let ρ:π1(M)π1(Bi)\rho:\pi_{1}(M)\longrightarrow\pi_{1}(B_{i}) be the morphism induced by the projection MBiM\rightarrow B_{i}. The boundary components Zi~\widetilde{Z_{i}} are in bijection with the boundary components of Bi~\widetilde{B_{i}}.
Let γ\gamma be a bi-infinite geodesic that is completely contained in BiB_{i}. By definition the stabilizer of γ\gamma satisfies Stab(γ)=γ\operatorname{Stab}(\gamma)=\gamma, then by Lemma 4, Stab(γ)×H\operatorname{Stab}(\gamma)\leq\mathbb{R}\times H. If ρ(g)γ=γ\rho(g)\gamma=\gamma then there exists pγp\in\gamma such that ρ(g)p=p\rho(g)p=p. This implies |ρ(g)|<|\rho(g)|<\infty and therefore ρ(g)=Id\rho(g)=\textrm{Id}. We conclude that gKer(ρ)g\in\operatorname{Ker}(\rho). ∎

Lemma 21.

Let MM be a higher graph manifolds as in definition 1 and TT the Bass-Serre tree associated to the decomposition of MM. Let WW be a wall in M~\widetilde{M} and HH its stabilizer in π1(M)\pi_{1}(M). Then WW is the unique wall which is stabilized by HH.

Proof.

First, we need to note that the Hirsch length of HH is h(H)=n1\operatorname{h}(H)=n-1, because HH is aspherical, here n=dimMn=\operatorname{dim}M. Suppose that HH stabilizes another wall WWW^{\prime}\neq W. By Lemma 20, HH is contained in π1(Nv)\pi_{1}(N_{v}), for vv a vertex of TT and NvN_{v} the corresponding fiber of the piece associated to vv. As NvN_{v} is an aspherical manifold, using Theorem 13 we have that the Hirsch length of π1(Nv)\pi_{1}(N_{v}) is equal to its virtual cohomological dimension. Now, by Theorem 11 we have that h(π1(Nv))n2\operatorname{h}(\pi_{1}(N_{v}))\leq n-2. This gives us a contradiction. We conclude that WW is the unique wall which is stabilized by HH. ∎

Lemma 22.

Let MM be a higher graph manifolds as in definition 1. Let Z1,Z2Z_{1},Z_{2} be two pieces of MM and WiW_{i} be a component of Zi\partial Z_{i}, for i=1,2i=1,2. Let Hi<π1(M)H_{i}<\pi_{1}(M) be groups conjugate to π1(Wi)\pi_{1}(W_{i}). Then:

  1. (1)

    The normalizer of H1H_{1} in π1(M)\pi_{1}(M) is equal to H1H_{1}.

  2. (2)

    If H1H_{1} is conjugate to H2H_{2} in π1(M)\pi_{1}(M) then W1W_{1} is isometric to W2W_{2} in MM.

Proof.

Both proofs are analogous to those of Lemma 19. ∎

We will now present some metric properties of M~\widetilde{M}. We will denote by d\operatorname{d} the distance associated to the Riemannian structure of M~\widetilde{M}. For every r>0r>0 and XM~X\subseteq\widetilde{M}, we will denote by Nr(X)M~N_{r}(X)\subseteq\widetilde{M} the rr-neighborhood of XX, with respect to the Riemannian metric d\operatorname{d} of M~\widetilde{M}. Recall that dC\operatorname{d}_{C} denotes the path metric on a chamber CM~C\in\widetilde{M} and it is the induced metric on CC defined by the restriction of the Riemannian structure of M~\widetilde{M}.

Lemma 23.

Let MM be a higher graph manifolds as in definition 1. If CC is a chamber of M~\widetilde{M}, then there exists a function g:++g:\mathbb{R}^{+}\longrightarrow\mathbb{R}^{+} such that g(t)g(t)\rightarrow\infty as t+t\rightarrow+\infty and d(x,y)g(dC(x,y))\operatorname{d}(x,y)\geq g(\operatorname{d}_{C}(x,y)) for each x,yCx,y\in C.

Proof.

As d\operatorname{d} and dC\operatorname{d}_{C} induce the same topology in CC, it is enough to prove the result for a fixed point xx.
Let {yi}\{y_{i}\} be a sequence of points in CC such that dC(x,yi)\operatorname{d}_{C}(x,y_{i})\longrightarrow\infty. Now, as M~\widetilde{M} is proper, if we suppose that d(x,yi)\operatorname{d}(x,y_{i}) is bounded, then passing to a sub sequence if necessary, there exists yM~y\in\widetilde{M} such that limiyi=y\displaystyle\lim_{i\longrightarrow\infty}y_{i}=y. We know that CC is closed in M~\widetilde{M}, and yy is necessarily in CC. This contradicts dC(x,yi)d_{C}(x,y_{i})\longrightarrow\infty. ∎

The proof of the following Lemma follows the same line of argument of Lemma 2.19 [11].

Lemma 24.

Let MM be a higher graph manifolds as in definition 1. Let W1,W2W_{1},W_{2} be two walls of M~\widetilde{M}, and suppose that there exists r>0r>0 such that W1Nr(W2)W_{1}\subset N_{r}(W_{2}), then W1=W2W_{1}=W_{2}.

Proof.

Consider the realization of M~\widetilde{M} as tree of spaces, we can then reduce to the case when W1W_{1} and W2W_{2} are walls adjacent to a given chamber CC. By the previous lemma, we can assume that W1W_{1} is contained in an rr-neighborhood of W2W_{2} with respect to the metric of CC.
Let ZiZ_{i} be the corresponding piece to the chamber CC and let Bi,NiB_{i},N_{i} be its base and fiber, in that order. We known that CBi~×Ni~C\cong\widetilde{B_{i}}\times\widetilde{N_{i}}. As MiM_{i} is a manifold that admits a negative curvature metric from which a finite number of non-maximal horospheres were removed, we have that by construction W1,W2W_{1},W_{2} are projected onto two horospheres O1,O2O_{1},O_{2} and O1O_{1} is contained in the rr-neighborhood of O2O_{2} with respect to the metric in Bi~\widetilde{B_{i}}. However, the metric B~i\widetilde{B}_{i} is bounded above by the locally symmetric metric, this forces O1=O2O_{1}=O_{2}. Therefore, W1=W2W_{1}=W_{2}. ∎

As a consequence of this lemma we obtain the following two results.

Corollary 25.

Let MM be a higher graph manifolds as in definition 1. Let WW be a wall of M~\widetilde{M} and let CC be a chamber of M~\widetilde{M}. If WNr(C)W\subset N_{r}(C) for some r0r\geq 0, then WW is a wall adjacent to CC.

Proof.

Note that WW is contained in the rr-neighbourhood of an adjacent wall to a chamber CC. Using Lemma 24, we obtain that WW is adjacent to CC. ∎

Corollary 26.

Let MM be a higher graph manifolds as in definition 1. Let C1,C2C_{1},C_{2} be two chambers of M~\widetilde{M} and suppose that there exists an r0r\geq 0 such that C1Nr(C2)C_{1}\subset N_{r}(C_{2}), then C1=C2C_{1}=C_{2}.

Proof.

Let W1,W2W_{1},W_{2} be distinct walls, both of them adjacent to C1C_{1}. By Corollary 25, they are adjacent to C2C_{2}. Thus C1=C2C_{1}=C_{2}. ∎

We now want to completely characterize the fundamental group of the pieces of a higher graph manifold MM. We will focus on the action of π1(M)\pi_{1}(M) on TT, and describe the set of stabilizers of walls of M~\widetilde{M}.
We say that a piece ZZ of MM is a surface piece if the base of the fiber bundle which defines ZZ is a hyperbolic surface.
In order to understand the stabilizers of walls, we define the following set:

𝒩(π1(M))={H<π1(M) | H a maximal nilpotent subgroup of π1(M) and h(H)=n1}\mathcal{N}(\pi_{1}(M))=\{H<\pi_{1}(M)\textrm{ }|\textrm{ }H\textrm{ a maximal nilpotent subgroup of }\pi_{1}(M)\textrm{ and }\operatorname{h}(H)=n-1\}

Here h(H)\operatorname{h}(H) is the Hirsch length of HH.

Proposition 27.

Let MM be a higher graph manifolds as in definition 1 without surface pieces. Let HH be a subgroup of π1(M)\pi_{1}(M). Then H𝒩(π1(M))H\in\mathcal{N}(\pi_{1}(M)) if and only if, HH is a maximal nilpotent subgroup of the stabilizer of a vertex vv in the Bass-Serre tree TT associated to the decomposition of MM in pieces and has h(H)=n1\operatorname{h}(H)=n-1.

Proof.

Let ZvZ_{v} be the piece corresponding to the vertex vv and let NvN_{v} be the fiber as in definition 1.
Suppose that H𝒩(π1(M))H\in\mathcal{N}(\pi_{1}(M)), then by Proposition 14 either HH fixes a unique vertex, or there exists a path γ\gamma, fixed as a set under the action HH, on which HH acts by translations. We will prove that HH can not act by translations on γ\gamma.
By contradiction, assume HH acts by translations on the path γ\gamma. Denote by φ:H\varphi:H\longrightarrow\mathbb{Z} the homomorphism through which HH acts by translations in γ\gamma and K=Ker(φ)K=\operatorname{Ker}(\varphi). As h(H)=n1\operatorname{h}(H)=n-1 and h()=1\operatorname{h}(\mathbb{Z})=1, Lemma 11 implies h(K)=n2\operatorname{h}(K)=n-2. Moreover, KK is a subgroup of HH which fixes the path γ\gamma.
For every point on γ\gamma, in particular for vv, KK acts on the chamber CvC_{v} which corresponds to vv and KK stabilizes two walls W1,W2W_{1},W_{2} of CvC_{v} for which γ\gamma enters and leaves. By Lemma 20, Kπ1(Nv)K\leq\pi_{1}(N_{v}). Hence the Hirsch length of KK is less than the Hirsch length of NvN_{v}, that is h(K)<h(Nv)\operatorname{h}(K)<\operatorname{h}(N_{v}). By hypothesis we do not have surface pieces, then h(Nv)<n3\operatorname{h}(N_{v})<n-3, which is a contradiction. Thus KK only fixes a unique vertex of γ\gamma. Therefore HH is contained in the stabilizer of a vertex.
Suppose HH is a maximal nilpotent subgroup of the stabilizer of a vertex vV(T)v\in V(T) and is such that h(H)=n1\operatorname{h}(H)=n-1. Let HH^{\prime} be a maximal nilpotent subgroup of π1(M)\pi_{1}(M), with h(H)=n1\operatorname{h}(H^{\prime})=n-1 and containing HH. There are two cases to consider.
If vv is the unique vertex fixed by HH, then as H<HH<H^{\prime}, HH^{\prime} also fixes vv. Therefore HH^{\prime} is contained in the stabilizer of vv and by maximality H=HH=H^{\prime}.
If HH fixes another vertex wvw\neq v, then HH must also fix an edge ee of TT exiting from vv. As HH is a maximal subgroup of the stabilizer of vv, HH coincides with the stabilizer of ee.
In both cases we conclude that H𝒩(π1(M))H\in\mathcal{N}(\pi_{1}(M)). ∎

Lemma 28.

Let MM be a higher graph manifold as in definition 1. If H<π1(M)H<\pi_{1}(M) is a wall stabilizer then H𝒩(π1(M))H\in\mathcal{N}(\pi_{1}(M)). On the other hand, if H𝒩(π1(M))H\in\mathcal{N}(\pi_{1}(M)), then:

  1. (1)

    either HH is a wall stabilizer, or

  2. (2)

    there exists a unique vertex vV(T)v\in V(T) fixed by HH, and this vertex corresponds to a surface piece of MM.

Proof.

If HH is a wall stabilizer, then HH is a nilpotent maximal subgroup of the stabilizer of a vertex of TT, hence Proposition 27 implies H𝒩(π1(M))H\in\mathcal{N}(\pi_{1}(M)).
Now, suppose H𝒩(π1(M))H\in\mathcal{N}(\pi_{1}(M)) is not a wall stabilizer. By Proposition 27, HH is contained in the stabilizer of a vertex vV(T)v\in V(T). Moreover, vv is the unique vertex which is fixed by HH, because otherwise HH would fix an edge and by maximality, HH will be a wall stabilizer.
Let ZvZ_{v} be the piece corresponding to vv and Nvk,BvnkN_{v}^{k},B_{v}^{n-k} the fiber and base of ZvZ_{v}, in that order. Suppose by contradiction that ZvZ_{v} is not a surface piece, hence kn3k\leq n-3. Now, the Hirsch length of the projection of HH on π1(Bv)\pi_{1}(B_{v}) is at least nk12n-k-1\geq 2 and it is therefore contained in a cusp subgroup. By maximality, this implies that HH is a wall stabilizer, a contradiction. ∎

Lemma 29.

Let MM be a higher graph manifold as in definition 1. Let CC be a chamber in M~\widetilde{M}. Let W1,W2W_{1},W_{2} be two different walls adjacent to CC and let H1,H2H_{1},H_{2} be their stabilizers. Let H¯𝒩(π1(M)){H1,H2}\overline{H}\in\mathcal{N}(\pi_{1}(M))\setminus\{H_{1},H_{2}\}. Then, for each k0k\geq 0, there exist points w1W1Cw_{1}\in W_{1}\cap C, w2W2Cw_{2}\in W_{2}\cap C, which can be joined by a path γ:[0,l]C\gamma:[0,l]\longrightarrow C that does not cross the neighborhood of radius kk, Nk(W¯)N_{k}(\overline{W}), of W¯\overline{W}. Here W¯\overline{W} is the wall which is stabilized by H¯\overline{H}.

The proof is analogous to the one of Lemma 7.8 in [11], here we only used the wall stabilizer H𝒩(π1(M))H\in\mathcal{N}(\pi_{1}(M)). We include the proof for completeness.

Proof.

As H¯{H1,H2}\overline{H}\notin\{H_{1},H_{2}\} we can then assume that W¯\overline{W} is disjoint from CW1W2C\cup W_{1}\cup W_{2}. We thus need to consider the following two cases:

  1. (1)

    W¯\overline{W} lies in the connected component of M~\{W1,W2}\widetilde{M}\backslash\{W_{1},W_{2}\} that contains CC.

  2. (2)

    W¯\overline{W} and CC lie in different connected components of M~\{W1,W2}\widetilde{M}\backslash\{W_{1},W_{2}\}.

Case 1: If W¯\overline{W} lies in the connected component of M~\{W1,W2}\widetilde{M}\backslash\{W_{1},W_{2}\} that contains CC, then there exists a wall W3W1,W2W_{3}\neq W_{1},W_{2} adjacent to CC such that every path that connects W¯\overline{W} with W1W2W_{1}\cup W_{2} must pass through W3W_{3}.
Every path that connects W1W_{1} with W2W_{2} but that does not pass through Nk(W3)N_{k}(W_{3}) also avoids W¯\overline{W}, because we have assumed that W¯\overline{W} is disjoint from CW1W2C\cup W_{1}\cup W_{2} and W3W_{3} is adjacent to CC. Let dC\operatorname{d}_{C} be the path metric on CC. By Lemma 23, it is enough to construct a path γ\gamma that joins w1w_{1} with w2w_{2} and such that for all t[0,l]t\in[0,l] and any given constant k>0k^{\prime}>0 we have dC(γ(t),W3)k\operatorname{d}_{C}(\gamma(t),W_{3})\geq k^{\prime}.
Let p:CBp:C\rightarrow B be the projection of chamber CC onto its base BB. For i=1,2,3i=1,2,3, let ξi\xi_{i} be points at infinity on BB and OiO_{i} the horospheres centered at ξi\xi_{i} defined as follows:

  1. (1)

    O1=p(W1C)O_{1}=p(W_{1}\cap C)

  2. (2)

    O2=p(W2C)O_{2}=p(W_{2}\cap C)

  3. (3)

    O3=p(W3C)O_{3}=p(W_{3}\cap C)

Refer to caption
Figure 1. Construction of the projection of the path γ\gamma between w1w_{1} and w2w_{2}, from Lemma 29.

Consider points xO1,yO2x\in O_{1},y\in O_{2} and the geodesic rays c1:[0,]Bc_{1}:[0,\infty]\rightarrow B from xx to ξ3\xi_{3} and c2:[0,]Bc_{2}:[0,\infty]\rightarrow B from yy to ξ3\xi_{3}. For every t[0,]t\in[0,\infty], we can consider a sequence of horospheres defined by the level sets of the Busemann function fci(t)f_{c_{i}}(t). For every sufficiently small ϵ>0\epsilon>0, we want to construct a path from any point on O1O_{1} to any other point on O2O_{2} and such that this path does not intersect O3O_{3}. Consider the concatenation of the following paths:

  1. (1)

    The subpath of c1c_{1} from xx to the intersection point, c1(ϵ)c_{1}(\epsilon), of one horospheres defined by fc1(ϵ)f_{c_{1}}(\epsilon) and c1c_{1}.

  2. (2)

    Consider the horosphere defined by the Bussemann function fc3(t)f_{c_{3}}(t) that goes through the point c1(ϵ)c_{1}(\epsilon) and call it Oc3O_{c_{3}}.

  3. (3)

    The subpath of c2c_{2} that goes from yy to the intersection point, c2(ϵ)c_{2}(\epsilon^{\prime}), of c2c_{2} and the horosphere of previous point.

  4. (4)

    The path that goes from c1(ϵ)c_{1}(\epsilon) to c2(ϵ)c_{2}(\epsilon^{\prime}) trough the horosphere Oc3O_{c_{3}}.

By construction, the concatenation of these three paths does not intersect O3O_{3}. Let γ:[0,l]C\gamma:[0,l]\rightarrow C the lift of this path to CC. Therefore there exists a constant K(ϵ)K(\epsilon), which depends on ϵ\epsilon and tends to ++\infty as ϵ\epsilon goes to 0, such that for every t[0,l]t\in[0,l], dC(γ(t),W¯)K(ϵ)\operatorname{d}_{C}(\gamma(t),\overline{W})\geq K(\epsilon). With this, we conclude the proof of the first case.
Case 2: Now, suppose that W¯\overline{W} and CC lie in different connected components of M~\{W1,W2}\widetilde{M}\backslash\{W_{1},W_{2}\}. Then we have that every path joining W1W_{1} to W¯\overline{W} must pass through W2W_{2} and that every path joining W2W_{2} to W¯\overline{W} must pass through W1W_{1}. Suppose that the first of these possibilities happens, by symmetry, the second case is analogous.
Then, there exists a chamber C¯\overline{C} different to CC and adjacent to W2W_{2}. Choose a fiber F¯\overline{F} of C¯W2\overline{C}\cap W_{2} and let F2F_{2} be the corresponding subspace of W2CW_{2}\cap C. By Lemma 23, if kk^{\prime} is a given constant, then we want to construct a path γ\gamma joining W2W_{2} to W¯\overline{W} and such that dC(γ(t),F2¯)k\operatorname{d}_{C}(\gamma(t),\overline{F_{2}})\geq k^{\prime} for every t[0,l]t\in[0,l].
Let B,FB,F be the base and fiber of CC, in that order. Then, we can choose pFp\in F such that B×{p}CB\times\{p\}\subseteq C intersects FF in a proper subspace of W(B×{p})W\cap(B\times\{p\}). We know that BB can be identified with a negatively curved space without a finite number of non maximal horospheres removed, therefore we can identify W2(B×{p})W_{2}\cap(B\times\{p\}) with a horosphere O2O_{2} centered at infinity and W1(B×{p})W_{1}\cap(B\times\{p\}) with a horosphere O1O_{1}, also centered at infinity. Using the same strategy of case 1, we can construct a path γ\gamma as a concatenation of certain special subpaths in such a way that γ\gamma and the neighbourhood of finite radius O¯\overline{O} of F(B×{p})F\cap(B\times\{p\}) do not intersect each other. Finally, if the constant kk^{\prime} is large enough, then any path in B×{p}B\times\{p\} that connects W2W_{2} and W¯\overline{W} and that avoids the neighbourhood of radius kk^{\prime} of F(B×{p})F\cap(B\times\{p\}) also avoids the neighbourhood of radius kk^{\prime} of FF. ∎

Corollary 30.

Let M1,M2M_{1},M_{2} be two higher graph manifolds as in definition 1. Let φ:π1(M1)π1(M2)\varphi:\pi_{1}(M_{1})\longrightarrow\pi_{1}(M_{2}) be an isomorphism. If H1H_{1} is a subgroup of π1(M1)\pi_{1}(M_{1}), then H1H_{1} is the stabilizer of a wall in M~1\widetilde{M}_{1} if and only if φ(H1)\varphi(H_{1}) is the stabilizer of a wall in M~2\widetilde{M}_{2}.

Proof.

Suppose H1H_{1} is the stabilizer of a wall in M~1\widetilde{M}_{1}, then H1H_{1} is contained in the stabilizer of an edge ee of the Bass-Serre tree T1T_{1} associated to M1~\widetilde{M_{1}}. As φ\varphi is an isomorphism, φ(H)\varphi(H) is contained in the stabilizer of an edge e2e_{2} of the Bass-Serre tree T2T_{2} associated to M2~\widetilde{M_{2}}, i.e., it is contained in gH2g1gH_{2}g^{-1}. We only need to show that φ(H1)\varphi(H_{1}) coincides with gH2g1gH_{2}g^{-1}. The subgroup φ1(gH2g1)\varphi^{-1}(gH_{2}g^{-1}) of π1(M1)\pi_{1}(M_{1}) contains H1H_{1} and again it is contained in the stabilizer of an edge of the Bass-Serre tree T1T_{1}. We conclude, φ1(gH2g1)=H1\varphi^{-1}(gH_{2}g^{-1})=H_{1}. ∎

Proposition 31.

([11, Proposition 4.13]) Let M1,M2M_{1},M_{2} be two higher graph manifolds as in definition 1. Let f:M~1M~2f:\widetilde{M}_{1}\longrightarrow\widetilde{M}_{2} be a (k,c)(k,c)-quasi-isometry and let gg be its quasi-inverse. Suppose that there exists a λ\lambda with the property that for each wall W1W_{1} of M~1\widetilde{M}_{1}, there exists a wall W2W_{2} of M~2\widetilde{M}_{2} with Hausdorff distance between f(W1)f(W_{1}) and W2W_{2} bounded by λ\lambda, and that if we change W1W_{1} with W2W_{2} we have the same for gg.
Then, there exists a universal constant HH, with the property that for each chamber C1C_{1} of M~1\widetilde{M}_{1}, there exists a unique chamber C2C_{2} in M~2\widetilde{M}_{2} such that the Hausdorff distance between f(C1)f(C_{1}) and C2C_{2} is bounded by HH.

We include the proof for completeness, following the one presented by [11], which also applies to MM.

Proof.

Let W1,W1W_{1},W_{1}^{\prime} be walls adjacent to a fixed chamber C1M~1C_{1}\subseteq\widetilde{M}_{1}. By hypothesis, there exist two walls W2,W2W_{2},W_{2}^{\prime} in M2~\widetilde{M_{2}} such that f(W1)f(W_{1}) and f(W1)f(W_{1}^{\prime}) are at finite Hausdorff distance of W2W_{2} and W2W_{2}^{\prime} respectively, and by Lemma 24 these walls are unique.
We want to prove that there exists a chamber C2M2~C_{2}\subseteq\widetilde{M_{2}} such that W2W_{2} and W2W_{2}^{\prime} are adjacent walls of C2C_{2}.
Suppose that there exists a wall P2M2~P_{2}\subseteq\widetilde{M_{2}} different to W2W_{2} and W2W_{2}^{\prime} such that every path which connects W2W_{2} with W2W_{2}^{\prime} crosses P2P_{2}. Then, there exists a wall P1M1~P_{1}\subseteq\widetilde{M_{1}} such that f(P1)f(P_{1}) is at most at finite distance λ\lambda of P2P_{2}. Since ff and gg are quasi-isometries, if P2P_{2} separates W2W_{2} and W2W_{2}^{\prime}, then there exists a constant D>0D>0 such that every path which connects W1W_{1} with W1W_{1}^{\prime} intersects ND(P1)N_{D}(P_{1}). However, by Lemma 29 this is a contradiction. Therefore W2W_{2} and W2W_{2}^{\prime} are adjacent walls to C2C_{2}.
Let did_{i} be the diameter of a chamber ZiZ_{i} and let hh be the maximum of di/2d_{i}/2. Then for each p1C1p_{1}\in C_{1}, there exists p1W1p_{1}^{\prime}\in W_{1} with d(p1,p1)hd(p_{1},p_{1}^{\prime})\leq h, here W1W_{1} is a adjacent wall to C1C_{1}. Thus

d(f(p1),C2)d(f(p1),f(p1))+d(f(p1),C2)kh+c+λ.\begin{array}[]{rcl}d(f(p_{1}),C_{2})&\leq&d(f(p_{1}),f(p_{1}^{\prime}))+d(f(p_{1}),C_{2})\\ &\leq&kh+c+\lambda.\end{array}

From which we obtain that f(C1)f(C_{1}) is in the neighbourhood of radius kh+c+λkh+c+\lambda of C2C_{2}. In a similar way, we can see that g(C2)g(C_{2}) is in the neighbourhood of radius kh+c+λkh+c+\lambda of some chamber C1C_{1}^{\prime}, but by Corollary 26 C1=C1C_{1}=C_{1}^{\prime}.
Since gg is the quasi-inverse of ff, if q2C2q_{2}\in C_{2}, then d(q2,f(g(q2)))cd(q_{2},f(g(q_{2})))\geq c and also, there exists q1C1q_{1}\in C_{1} with d(g(q2),q1)kh+c+λd(g(q_{2}),q_{1})\geq kh+c+\lambda. We now want to estimate the distance between each element q2C2q_{2}\in C_{2} and f(q1)f(q_{1}), where q1C1q_{1}\in C_{1}:

d(q2,f(q1))d(q2,f(g(q2)))+d(f(g(q2)),f(q1))c+kd(g(q2),q1)+c2c+k(kh+c+λ)\begin{array}[]{rcl}d(q_{2},f(q_{1}))&\leq&d(q_{2},f(g(q_{2})))+d(f(g(q_{2})),f(q_{1}))\\ &\leq&c+kd(g(q_{2}),q_{1})+c\\ &\leq&2c+k(kh+c+\lambda)\end{array}

With L=2c+k(kh+c+λ)L=2c+k(kh+c+\lambda) we obtain the result. Finally, the uniqueness of C2C_{2} is a consequence of Lemma 26. ∎

We are now ready to prove Theorem 1.

Proof of Theorem 1.

Let Λ1<π1(M1)\Lambda_{1}<\pi_{1}(M_{1}) be the fundamental group of a piece Z1Z_{1} of M1M_{1}. As a consequence of Proposition 31 and the Milnor-Svarcˇ\check{\textrm{c}} Lemma (Proposition 15), the Hausdorff distance dH(φ(Λ1),gΛ2g1)\operatorname{d}_{H}(\varphi(\Lambda_{1}),g\Lambda_{2}g^{-1}) is bounded above by LL for some Λ2<π1(M2)\Lambda_{2}<\pi_{1}(M_{2}) which is the fundamental group of a piece in M2M_{2} and some gπ1(M2)g\in\pi_{1}(M_{2}).
Without loss of generality, we may assume that g=idg=id.
If hΛ1h\in\Lambda_{1}, we obtain that:

φ(h)φ(Λ1)=φ(hΛ1)=φ(Λ1)\varphi(h)\varphi(\Lambda_{1})=\varphi(h\Lambda_{1})=\varphi(\Lambda_{1})

Since the Hausdorff distance dH(φ(Λ1),Λ2)\operatorname{d}_{H}(\varphi(\Lambda_{1}),\Lambda_{2}) is bounded above, then φ(h)Λ2\varphi(h)\Lambda_{2} is at bounded Hausdorff distance of Λ2\Lambda_{2}. By Milnor-Svarcˇ\check{\textrm{c}} Lemma, if C2C_{2} is a chamber fixed by Λ2\Lambda_{2}, then dH(φ(h)C2,C2)\operatorname{d}_{H}(\varphi(h)C_{2},C_{2}) is finite. This yields φ(h)C2=C2\varphi(h)C_{2}=C_{2} and φ(h)Λ2\varphi(h)\in\Lambda_{2}, and therefore φ(Λ1)Λ2\varphi(\Lambda_{1})\subseteq\Lambda_{2}.
Using the quasi-inverse φ1\varphi^{-1} of φ\varphi we can prove that φ1(Λ2)Λ1\varphi^{-1}(\Lambda_{2})\subseteq\Lambda_{1}.
By Proposition 31, for hΛ1h\in\Lambda_{1} the Hausdorff distance dH(φ1(Λ2),hΛ1h1)<H\operatorname{d}_{H}(\varphi^{-1}(\Lambda_{2}),h\Lambda_{1}h^{-1})<H.
Again, we can assume that h=idh=id, so for gΛ2g\in\Lambda_{2} we have that

φ1(g)φ1(Λ2)=φ1(gΛ2)=φ1(Λ2).\varphi^{-1}(g)\varphi^{-1}(\Lambda_{2})=\varphi^{-1}(g\Lambda_{2})=\varphi^{-1}(\Lambda_{2}).

Then, as dH(φ1(Λ2),hΛ1h1)<H\operatorname{d}_{H}(\varphi^{-1}(\Lambda_{2}),h\Lambda_{1}h^{-1})<H, we find that dH(φ1(g)Λ1,Λ1)\operatorname{d}_{H}(\varphi^{-1}(g)\cdot\Lambda_{1},\Lambda_{1}) is bounded.
If Λ1\Lambda_{1} fixes a chamber C1M~1C_{1}\in\widetilde{M}_{1}, then dH(φ1(g)(C1),C1)<\operatorname{d}_{H}(\varphi^{-1}(g)(C_{1}),C_{1})<\infty. Therefore, by Corollary 26 φ1(g)(C1)=C1\varphi^{-1}(g)(C_{1})=C_{1}, so φ1(g)Λ1\varphi^{-1}(g)\in\Lambda_{1} and φ1(Λ2)Λ1\varphi^{-1}(\Lambda_{2})\subseteq\Lambda_{1}.
We conclude that φ(Λ1)=Λ2\varphi(\Lambda_{1})=\Lambda_{2}. ∎

4. Cusp-Decomposable manifolds

In this section we will prove Theorem 2 for cusp-decomposable manifolds.
Frigerio, Lafont and Sisto mention that except for Proposition 7.8 all the results of their section VII [11] can be modified to hold for cusp-decomposable manifolds. They mention that a proof of Proposition 7.8 for cusp-decomposable manifolds can not be similar to the one that they presented, (Remark 7.9 [11]). Armed with knowledge of the electric space associated to a pinched Hadamard manifold, we will prove the corresponding result to Proposition 7.8 for the case of cusp decomposable manifolds in Proposition 36.
Tam Nguye^~\tilde{\hat{\textrm{e}}}n-Phan described the following family of graph manifolds that are a subfamily of the manifolds in Definition 1 where the pieces are all pure pieces [25].
Let VV be a locally symmetric, complete, finite volume, noncompact, connected, negatively curved manifold of dimension n3n\geq 3. It is know that VV has a finite number of cusps and each cusp is diffeomorphic to S×[0,]S\times[0,\infty], where SS is a compact (n1)(n-1)-dimension manifold [9]. Let bb be large enough so that the boundary components S×{b}S\times\{b\} of different cusps do not intersect each other. Now delete S×(b,)S\times(b,\infty) from each cusp, then the resulting space ZZ is a compact manifold with boundary. The lifts of the boundaries components of ZZ are horospheres in Z~\widetilde{Z}. This manifold is called a bounded cusp manifold with horoboundary.

Definition 3.

(Tam Nguyen Phan,[25]) A cusp decomposable manifold MM is a manifold which is obtained by taking a finite number of bounded cusp manifolds with horoboundary ZiZ_{i} and identifying their horoboundaries using affine diffeomorphisms.

Throughout this section MM will denote a cusp-decomposable manifold.
Let MM be a cusp-decomposable manifold of dimension nn. The universal cover M~\widetilde{M} has a natural structure as a tree of spaces as follows. A chamber CM~C\in\widetilde{M} will be the preimage of a connected component in M~\widetilde{M} of ZZ without the deleted cusp S×(b,)S\times(b,\infty). A wall WM~W\in\widetilde{M} will be the preimage of a connected component of S×[0,b]S\times[0,b].
For each wall WW of M~\widetilde{M} its boundary can be decomposed into two connected components W+W_{+} and WW_{-} that are the intersection of a chamber with one adjacent wall. We will call each of the two components W+,WW_{+},W_{-} the thin walls associated to WW. Denote by d±\operatorname{d}_{\pm} the path metric of W±W_{\pm}, induced by the restriction of the Riemannian structure of M~\widetilde{M}. For each wall WW, let sW:W+Ws_{W}:W_{+}\longrightarrow W_{-} be a map that sends pW+p\in W_{+} to the point sW(p)Ws_{W}(p)\in W_{-} which is joined to pp. We think of sWs_{W} as the map that glues the two pieces, W+,WW_{+},W_{-}, together.

Lemma 32.

Let MM be a cusp-decomposable manifold as in definition 3 and let WW be a wall in M~\widetilde{M}. If CC is a chamber that contains W+W_{+}, then the inclusion of (W+,dW+)(C,dC)(W_{+},\operatorname{d}_{W_{+}})\hookrightarrow(C,\operatorname{d}_{C}) is an isometry.

Proof.

Recall that if ZZ is the piece of the manifold MM associated to the chamber CC, then every component of C\partial C is a convex horosphere in the metric sense (see [10]).
As a consequence, if W+W_{+} is a thin wall contained in CC, then W+W_{+} is a convex hypersurface of CC. Therefore, the path metric induced in W+W_{+} by the Riemannian structure on M~\widetilde{M} is isometric to the restriction of dC\operatorname{d}_{C}. ∎

Proposition 33.

Let MM be a cusp-decomposable manifold as in definition 3. Let WW be a wall in M~\widetilde{M} and W±W_{\pm} the thin walls associated to WW. Then, the inclusion (W±,dW±)(W,dW)(W_{\pm},\operatorname{d}_{{W}_{\pm}})\hookrightarrow(W,\operatorname{d}_{W}) is a bi-Lipschitz embedding and a quasi-isometry.

Proof.

The inclusion i:W±Wi:W_{\pm}\hookrightarrow W is 1-Lipschitz, by definition of the induced path metric. This map induces an isomorphism on fundamental groups, so by the Milnor-Svarcˇ\check{\textrm{c}} Lemma ii is a quasi-isometry. Therefore ii is a bi-Lipschitz embedding at large scale, i.e. there exist constants b0b\geq 0, c1c\geq 1 such that if dW±(x,y)b\operatorname{d}_{W_{\pm}}(x,y)\geq b then dW±(x,y)cdW(x,y)\operatorname{d}_{W_{\pm}}(x,y)\leq c\cdot\operatorname{d}_{W}(x,y).
We now only need to analyze the case where 0dW±(x,y)b0\leq\operatorname{d}_{W_{\pm}}(x,y)\leq b. Let TT be the subset of W±×W±W_{\pm}\times W_{\pm} defined by the inequality 0dW±(x,y)b0\leq\operatorname{d}_{W_{\pm}}(x,y)\leq b. Let ZZ be a piece of MM and let SS be the infranilmanifold of its boundary. Then, S~\widetilde{S} is homeomorphic to W+W_{+} and WW_{-}. Consider the following action of π1(S)\pi_{1}(S) on TT:

ρ:π1(S)×TT(g,(p,sW(p)))(gp,sW(gp))\begin{array}[]{rcl}\rho:\pi_{1}(S)\times T&\rightarrow&T\\ (g,(p,s_{W}(p)))&\mapsto&(gp,s_{W}(gp))\end{array}

Therefore, TT is invariant under this action and the quotient space T/π1(S)T/\pi_{1}(S) is compact.
Define a function f:Tf:T\rightarrow\mathbb{R} as follows:

f(x,y)={1for every pair (p,sW(p))(W+,W),dW±(x,y)/dW(x,y)if (x,y)T\{(p,sW(p))|pW+ and sW(p)W}f(x,y)=\left\{\begin{array}[]{ll}1&\textit{for every pair }(p,s_{W}(p))\in(W_{+},W_{-}),\\ \operatorname{d}_{{W}_{\pm}}(x,y)/\operatorname{d}_{W}(x,y)&\textit{if }(x,y)\in T\backslash\{(p,s_{W}(p))|p\in W_{+}\textit{ and }s_{W}(p)\in W_{-}\}\end{array}\right.

This is a positive continuous function and the compactness of T/π1(S)T/\pi_{1}(S) implies that it is bounded above by some constant cc^{\prime}. If C=max{c,c}C=\max\{c,c^{\prime}\}, then the inclusion map ii is C-bi-Lipschitz and we conclude the proof. ∎

Let W±W_{\pm} be a thin wall and x,yW±x,y\in W_{\pm}. Let γ\gamma be a path which connects xx with yy in M~\widetilde{M}. We will say that γ\gamma is a non-backtracking path in W±W_{\pm} if γ\gamma only intersects the wall W±W_{\pm} in its endpoints.
The following proposition is a consequence of Lemma 3.2 of [18] and the proof is the same as that of the Proposition 7.4 of [11].

Proposition 34.

Let ZZ be a bounded cusp manifold with horoboundary. Then there exists a constant λ\lambda that depends only on ZZ such that the following is true. Let γZ\gamma\subseteq Z be a loop obtained by concatenating a finite number of paths α1,γ1,,αn,γn\alpha_{1},\gamma_{1},...,\alpha_{n},\gamma_{n}, where

  • Each αi\alpha_{i} is a geodesic in the horosphere SiZS_{i}\subseteq\partial Z.

  • Each γi\gamma_{i} is a path in ZZ that connects the final point of αi\alpha_{i} with the initial point of αi+1\alpha_{i+1}.

  • The final points of each γi\gamma_{i} are in different walls.

Let Λ{1,,n}\Lambda\subseteq\{1,...,n\} be a subset of indices such that SkSiS_{k}\neq S_{i} for each kΛk\in\Lambda, i{1,n}i\in\{1,...n\}, iki\neq k. Then,

kΛL(αk)λi=1nL(γi).\sum_{k\in\Lambda}L(\alpha_{k})\leq\lambda\cdot\sum_{i=1}^{n}L(\gamma_{i}).

We say that γ\gamma is minimal, if for each chamber CC the set γC̊\gamma\cap\mathring{C} is only a finite collection of paths and each of these paths connects different walls of CC. Moreover, we say that γ\gamma is good if it is minimal and for each thin wall W±W_{\pm} contained in CC, there are at most two final points of paths in γC̊\gamma\cap\mathring{C} that belong to W±W_{\pm}. Another characterization of good paths is as follows. Let W±W_{\pm} be a thin wall contained in a wall WW. A path γ:[t0,t1]M~\gamma:[t_{0},t_{1}]\rightarrow\widetilde{M} is external to W±W_{\pm} if γ(t0),γ(t1)W±\gamma(t_{0}),\gamma(t_{1})\in W_{\pm} and γ|(t0,t1)\gamma|_{(t_{0},t_{1})} is supported on M~\W\widetilde{M}\backslash W. Let γ\gamma be a minimal path and nn be the number of external subpaths of γ\gamma to W±W_{\pm}, the exceeding number of γ\gamma on W±W_{\pm} is defined as max{0,n1}\max\{0,n-1\}. The exceeding number e(γ)e(\gamma) of γ\gamma is defined as the sum over all the thin walls of the exceeding number of γ\gamma. Denote by j(γ)j(\gamma) the sum over all the chambers of M~\widetilde{M} of the number of connected components of γC̊\gamma\cap\mathring{C}. A path γ\gamma is good if its minimal and e(γ)=0e(\gamma)=0.

Lemma 35.

Let MM be a cusp-decomposable manifold as in definition 3. Let x,yWM~x,y\in W\subset\widetilde{M} be two points in the same wall and β1\beta\geq 1. Let DD be the electric constant of Lemma 8. If d(x,y)D\operatorname{d}(x,y)\leq D then there exists a good path γM~\gamma\in\widetilde{M} from xx to yy such that L(γ)βd(x,y)L(\gamma)\leq\beta\operatorname{d}(x,y).

Proof.

Consider the set of pairs (x,y)W(x,y)\in W such that d(x,y)D\operatorname{d}(x,y)\leq D. This set is a compact set. Now, pick good paths γxy\gamma_{xy} for each such pair. We have that the map (x,y)L(γxy)(x,y)\mapsto L(\gamma_{xy}) is continuous and compact, therefore it is bounded above by a constant.
So for each pair x,yx,y there exists a constant βxy\beta_{xy} such that L(γxy)βxyd(x,y)L(\gamma_{xy})\leq\beta_{xy}\operatorname{d}(x,y). Set β\beta equal to the maximum over all βxy\beta_{xy} to obtain the result. ∎

Proposition 36.

Let MM be a cusp-decomposable manifold as in definition 3. Let x,yx,y be points in the same wall of M~\widetilde{M}. Then there exists a constant β1\beta\geq 1 that depends only on the geometry of M~\widetilde{M}, such that the following is true. There exists a good path γ\gamma on M~\widetilde{M} that connects xx with yy and such that L(γ)βd(x,y)L(\gamma)\leq\beta\cdot\operatorname{d}(x,y).

Proof.
Refer to caption
Figure 2. Construction of a good path interpolating between x1x_{1} and y2y_{2}, that goes through the points z1z_{1} and z2z_{2}, from Proposition 36.

Let M^\widehat{M} be the electric space associated to MM. Remember that RR is the distance between any two of the deleted horospheres of VV. By Lemma 5, we know that the horospheres are visually bounded. Let Δ>0\Delta>0 be a constant such that every horospherical boundary component of each ZiZ_{i} is visually bounded by Δ/2\Delta/2.
Let γ0\gamma_{0} be a geodesic joining xx and yy in M~\widetilde{M}. As every thin wall is convex, then every geodesic in M~\widetilde{M} is minimal, therefore γ0\gamma_{0} is minimal.
We want to modify a minimal path γi\gamma_{i} with e(γi)>0e(\gamma_{i})>0 by a new minimal path γi+1\gamma_{i+1} such that j(γi+1)<j(γi)j(\gamma_{i+1})<j(\gamma_{i}) and L(γi+1)L(γi)+KL(\gamma_{i+1})\leq L(\gamma_{i})+K, where KK is a positive constant. As j(γ0)L(γ0)/R=d(x,y)/Rj(\gamma_{0})\leq L(\gamma_{0})/R=\operatorname{d}(x,y)/R then after at most d(x,y)/R\operatorname{d}(x,y)/R steps, we will obtain a new minimal path γ\gamma such that e(γ)=0e(\gamma)=0.
The reader might find it useful to see Figure 2 while following the next construction. Suppose that we have some external subpaths α1=[x1,y1]\alpha_{1}=[x_{1},y_{1}] and α2=[x2,y2]\alpha_{2}=[x_{2},y_{2}] of γi\gamma_{i}, contained in the interior of CC, and such that x1,y1,x2,y2W±x_{1},y_{1},x_{2},y_{2}\in W_{\pm} where W±W_{\pm} is a thin wall of a chamber CC. Choose two points z1,z2γiC̊z_{1},z_{2}\in\gamma_{i}\cap\mathring{C} in such a way that L([z1,y1])<1/2L([z_{1},y_{1}])<1/2 and L([y2,z2])<1/2L([y_{2},z_{2}])<1/2. The new path γi+1\gamma_{i+1}^{\prime} is constructed as follows. Start by taking the subpath α1=[x1,z1]\alpha_{1}^{\prime}=[x_{1},z_{1}], then consider a path α3\alpha_{3} in C̊\mathring{C} from z1z_{1} to z2z_{2} but such that L([z1,z2])<Δ+1L([z_{1},z_{2}])<\Delta+1 and the subpath α2=[z2,y2]\alpha_{2}^{\prime}=[z_{2},y_{2}].
The path γi+1\gamma_{i+1}^{\prime} satisfies j(γi+1)=j(γi)1j(\gamma_{i+1}^{\prime})=j(\gamma_{i})-1 and its external number equals zero. So, we have a good path between x1x_{1} and y2y_{2}. By Lemma 35, there exists a constant β(Δ/2)\beta(\Delta/2), such that L(γi+1)β(Δ/2)d(x1,y2)L(\gamma_{i+1}^{\prime})\leq\beta(\Delta/2)\cdot\operatorname{d}(x_{1},y_{2}).
We need to carry out these replacements for all the external subpaths of γi\gamma_{i}. After performing all these replacements, we end with a good path that satisfies the length condition. ∎

Lemma 37.

Let MM be a cusp-decomposable manifold as in definition 3. Let WM~W\subseteq\widetilde{M} be a fixed wall, x,yx,y points in W±W_{\pm}, and CC the chamber that contains x,yx,y. Suppose that there exists α1\alpha\geq 1 such that if x,yx,y can be joined by a path γ\gamma in M~\widetilde{M} which does not backtrack in W±W_{\pm}, then

dC(x,y)αL(γ).\operatorname{d}_{C}(x,y)\leq\alpha\cdot L(\gamma).

Therefore the inclusion of WW into M~\widetilde{M} is a bi-Lipschitz embedding.

Proof.

As the inclusion (W,dW)M~(W,\operatorname{d}_{W})\hookrightarrow\widetilde{M} is 1-Lipschitz, we only need to prove that there exists a constant k1k\geq 1 such that for all p,qWp,q\in W we have that dW(p,q)kd(p,q)\operatorname{d}_{W}(p,q)\leq k\cdot\operatorname{d}(p,q). By Proposition 36, there exists a good path γ\gamma joining pp and qq such that for β1\beta\geq 1, L(γ)βd(p,q)L(\gamma)\leq\beta\operatorname{d}(p,q).
Let mm be the number of chambers adjacent to WW whose interior intersects γ\gamma.
Let us relabel the following subpaths of γ\gamma as follows:

  1. (1)

    For i=1,,mi=1,...,m, let γiW\gamma_{i}^{W} be a path contained in WW.

  2. (2)

    For i=1,,mi=1,...,m, let γiW±\gamma_{i}^{W_{\pm}} be a good path with endpoints on the thin walls W±W_{\pm} and which does not backtrack in W±W_{\pm}.

With this notation, we can decompose γ\gamma in the following form:

γ=γ1Wγ1W±γmWγmW±γm+1W.\gamma=\gamma_{1}^{W}\gamma_{1}^{W_{\pm}}...\gamma_{m}^{W}\gamma_{m}^{W_{\pm}}\gamma_{m+1}^{W}.

Let CC be a chamber adjacent to WW. We want to replace the paths γiW±\gamma_{i}^{W_{\pm}} by paths ηi\eta_{i} in CC and such that the total length of the curve obtained does not exceed αL(γ)\alpha\cdot L(\gamma).
Let W1,W2W_{1},W_{2} be two walls different from WW and adjacent to CC. As γiW±\gamma_{i}^{W_{\pm}} is a good path, then there exist subpaths of γiW±\gamma_{i}^{W_{\pm}} that connect different walls and these subpaths only cross the walls in two points.
Figure 3 illustrates the following construction. Let p,qp^{\prime},q^{\prime} be the endpoints of the good path γ1W±\gamma_{1}^{W_{\pm}}. Suppose that γ1,1W±\gamma_{1,1}^{W_{\pm}} is the subpath of γ1W±\gamma_{1}^{W_{\pm}} that connects WW with W1W_{1} and that γ1,2W±\gamma_{1,2}^{W_{\pm}} is the subpath of γ2W±\gamma_{2}^{W_{\pm}} that connects W1W_{1} with W2W_{2}. Let p1p_{1} be the point of W1,±W_{1,\pm} where γ1,1W±\gamma_{1,1}^{W_{\pm}} crosses W1,±W_{1,\pm} and let q1q_{1} be the point of W1,±W_{1,\pm} where γ1,2W±\gamma_{1,2}^{W_{\pm}} crosses W1,±W_{1,\pm}. Let ηp1,q1C\eta_{p_{1},q_{1}}^{C} be the geodesic path inside CC between p1p_{1} and q1q_{1}. With this process, we are cutting the path inside W1W_{1} and replacing it by a path contained in CC and such that its length is less than kk times the length of the path inside W1W_{1}, for some positive constant kk. Now, repeat this process for all the other subpaths γ1,iW±\gamma_{1,i}^{W_{\pm}} of γ1W±\gamma_{1}^{W_{\pm}} between the rest of the walls adjacent to CC. Let η1\eta_{1} be the path formed as follows. Start in pp^{\prime} and follow the subpath of γ1,1W±\gamma_{1,1}^{W_{\pm}} between pp^{\prime} and p1p_{1}, then on p1p_{1} switch to following the path ηp1,q1C\eta_{p_{1},q_{1}}^{C} and repeat this procedure until returning to qq^{\prime}. The path η1\eta_{1} between pp^{\prime} and qq^{\prime} is a path completely contained in CC and by hypothesis its length is bounded by α1L(γ1W±)\alpha_{1}\cdot L(\gamma_{1}^{W_{\pm}}).

Refer to caption
Figure 3. Here we illustrate how the paths ηpi,qiC\eta_{p_{i},q_{i}}^{C} replace γiW\gamma_{i}^{W} by paths inside CC. With the same notation as in Lemma 37, the points p,qp^{\prime},q^{\prime} lie on W±W_{\pm} and the points pi,qip_{i},q_{i} lie on the thin walls Wi,±W_{i,\pm}. The paths γi,j±\gamma_{i,j}^{\pm} are good paths joining the thin walls Wi,±,Wj,±W_{i,\pm},W_{j,\pm} and passing through pi+1,qip_{i+1},q_{i}.

Repeat this for all the paths γiW±\gamma_{i}^{W_{\pm}} in γ\gamma. After replacing all the curves, the resultant curve η=γ1Wη1γmWηmγm+1W\eta=\gamma_{1}^{W}\eta_{1}...\gamma_{m}^{W}\eta_{m}\gamma_{m+1}^{W} has length less that αL(γ)\alpha\cdot L(\gamma). Moreover, we have that

dW(p,q)L(η)αL(γ)\operatorname{d}_{W}(p,q)\leq L(\eta)\leq\alpha\cdot L(\gamma)

If k=αβk=\alpha\cdot\beta then we have shown that dW(p,q)kd(p,q)\operatorname{d}_{W}(p,q)\leq k\cdot\operatorname{d}(p,q). ∎

As a consequence of the previous results, we are now able to prove that the walls and chambers embeddings in M~\widetilde{M} are bi-Lipschitz.

Proposition 38.

Let MM be a cusp-decomposable manifold as in definition 3 and let WM~W\subseteq\widetilde{M} be a wall. Then the inclusion of (W,dW)M~(W,\operatorname{d}_{W})\hookrightarrow\widetilde{M} is a bi-Lipschitz embedding. In particular, it is a quasi-isometric embedding. Moreover, the bi-Lipschitz constant of the embedding only depends on the geometry of M~\widetilde{M}.

Proof.

Let W+W_{+} be a thin wall of WW and CC be the chamber which contains W+W_{+}. Let p,qp,q be two points on W+W_{+}. Proposition 36 guarantees that there exists a good path γ\gamma in M~\widetilde{M} that joins p,qp,q and which does not backtrack on W+W_{+}. By Lemma 37, we only need to prove that there exists a constant α1\alpha\geq 1 only depending on the geometry of M~\widetilde{M} such that the following condition it is true:

dC(p,q)αL(γ).\operatorname{d}_{C}(p,q)\leq\alpha\cdot L(\gamma).

Observe that as dC(p,q)\operatorname{d}_{C}(p,q) is the path metric over CC, by definition this is the infimum over all the paths that joins p,qp,q therefore with α=1\alpha=1 we obtain the result. ∎

Proposition 39.

Let MM be cusp decomposable manifold, then the inclusion of a chamber in M~\widetilde{M} is a bi-Lipschitz embedding.

Proof.

Let p,qp,q be points in a chamber CM~C\in\widetilde{M} and γ\gamma a geodesic from pp to qq. We can decompose γ\gamma as follows. Let γi\gamma_{i} be geodesics in CC and let pi,qiWi,+p_{i},q_{i}\in W_{i,+} be the endpoints of a path ηi\eta_{i} where Wi,+W_{i,+} is a thin wall adjacent to CC. So we may write γ=γ1η1γnηnγn+1\gamma=\gamma_{1}\eta_{1}...\gamma_{n}\eta_{n}\gamma_{n+1}.
By Lemma 37, there exists a constant a1a\geq 1 such that dWi(pi,qi)ad(pi,qi)\operatorname{d}_{W_{i}}(p_{i},q_{i})\leq a\cdot\operatorname{d}(p_{i},q_{i}). Then dWi,+(pi,qi)ad(pi,qi)\operatorname{d}_{W_{i,+}}(p_{i},q_{i})\leq a\cdot\operatorname{d}(p_{i},q_{i}). We can replace each ηi\eta_{i} with a path ηiWi,+\eta_{i}^{\prime}\in W_{i,+} that has the same endpoints as ηi\eta_{i} but such that its length is less than ad(pi,qi)a\operatorname{d}(p_{i},q_{i}). So, the new path γ=γ1η1γnηnγn+1\gamma^{\prime}=\gamma_{1}\eta_{1}^{\prime}...\gamma_{n}\eta_{n}^{\prime}\gamma_{n+1} is contained in CC and has length at most ad(p,q)a^{\prime}\operatorname{d}(p,q). Therefore dC(p,q)ad(p,q)\operatorname{d}_{C}(p,q)\leq a^{\prime}\operatorname{d}(p,q). ∎

We can now present the main result of this section.

Theorem 40.

Let MM be a cusp-decomposable manifold. Then, the inclusion of chambers and walls (with their path metric) in M~\widetilde{M} are quasi-isometric embeddings.

Proof.

Lemma 32 implies that C,WM~C,W\hookrightarrow\widetilde{M} are isometric embeddings. Lemma 38 and Proposition 39 imply that C,WM~C,W\hookrightarrow\widetilde{M} are bi-Lipschitz embeddings. Therefore chambers and walls are quasi-isometrically embedded in M~\widetilde{M}. ∎

References

  • [1] Aravinda, C. S., & Farrell, F. T., Twisted doubles and nonpositive curvature. Bulletin of the London Mathematical Society, 41(6), 1053-1059 (2009).
  • [2] Bárcenas, N., Juan-Pineda, D., & Suárez-Serrato, P., Topological rigidity of higher graph manifolds. Boletín de la Sociedad Matemática Mexicana, 23(1), 119-127 (2017).
  • [3] Bowditch, B. H. Cut points and canonical splittings of hyperbolic groups, Acta Math. 180, 145–186 (1998).
  • [4] Bowditch, B. H. Group actions on trees and dendrons. Topology 37, 1275–1298 (1998).
  • [5] Bowditch, B. H. Treelike Structures Arising from Continua and Convergence Groups. Mem. Amer. Math. Soc. 139, A. M. S., Providence, RI (1999).
  • [6] Brown, K. S., Cohomology of groups (Vol. 87) Springer Science & Business Media (2012).
  • [7] Connell, C., & Suárez-Serrato, P., On higher graph manifolds. International Mathematics Research Notices, 2019(5), 1281–1311 (2019).
  • [8] Contreras-Peruyero, H., Sobre la geometría y topología de las variedades de gráficas., Master Thesis, Universidad Nacional Autónoma de México (2016).
    http://132.248.9.195/ptd2016/agosto/0749031/Index.html
  • [9] Eberlein, P., Lattices in spaces of nonpositive curvature. Annals of Mathematics, 111(3), 435-476 (1980).
  • [10] Farb, B., Relatively hyperbolic groups. Geometric and functional analysis, 8(5), 810-840 (1998).
  • [11] Frigerio, R., Lafont, J.-F., & Sisto, A., Rigidity of high dimensional graph manifolds. Astérisque, 372 (2015).
  • [12] Gromov, M. Groups of polynomial growth and expanding maps. Publications Mathématiques de l’Institut des Hautes Études Scientifiques, 53(1), 53-78 (1981).
  • [13] de La Harpe, P., Topics in geometric group theory. University of Chicago Press (2000).
  • [14] de la Harpe, P., & Bucher, M., Free products with amalgamation and HNN-extensions of uniformly exponential growth. Mathematical Notes, 67(6), 686-689 (2000).
  • [15] Heintze, E., & Im Hof, H. C., Geometry of horospheres. Journal of Differential Geometry, 12(4), 481-491 (1977).
  • [16] Kapovich, M., & Leeb, B. Quasi-isometries preserve the geometric decomposition of Haken manifolds. Inventiones mathematicae, 128(2), 393-416 (1997).
  • [17] Löh, C. Geometric Group Theory: An Introduction. Springer International Publishing AG (2017).
  • [18] Osin, D. V., Relatively hyperbolic groups: intrinsic geometry, algebraic properties, and algorithmic problems. Memoirs American Mathematical Society, 843 (2006).
  • [19] Papasoglu, P., Quasi-isometry invariance of group splittings. Annals of mathematics, 759-830 (2005).
  • [20] Papasoglu, P., Group splittings and asymptotic topology. Journal f"u"{u}r die reine und angewandte Mathematik, 2007(602), 1-16 (2007).
  • [21] Robinson, D. J., A Course in the Theory of Groups, Vol. 80 (2012), Springer Science & Business Media.
  • [22] Stallings, J. R. On torsion-free groups with infinitely many ends. Annals of Mathematics, 312-334 (1968).
  • [23] Stallings, J. Group theory and three-dimensional manifolds. Yale Math. Monographs, Yale Univ. Press, New Haven, 241(4) (1971).
  • [24] Serre, J. P., Trees. Springer (1980).
  • [25] Tam Nguyen Phan, T., Smooth (non) rigidity of cusp-decomposable manifolds. Commentarii Mathematici Helvetici, 87, 789-804 (2012).
  • [26] Waldhausen F., Eine Klasse von 3-dimensionalen Mannigfaltigkeiten. I. Inventiones Mathematicae, 3(4), 308-333 (1967).