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

Quasi-redirecting boundaries of groups with linear divergence and 3-manifold groups

Hoang Thanh Nguyen Hoang Thanh Nguyen, Department of Mathematics, FPT University, Hoa Hai ward, Ngu Hanh Son district, Da Nang, Vietnam nthoang.math@gmail.com
Abstract.

The quasi-redirecting (QR) boundary, introduced by Qing and Rafi, generalizes the Gromov boundary for studying the large-scale geometry of finitely generated groups. Although it is not known to exist for all such groups, its existence has been established for several important classes. We prove that if a finitely generated group GG has linear divergence, then its QR-boundary is well-defined and consists of a single point. In addition, we show that all finitely generated 3-manifold groups admit well-defined QR-boundaries.

Key words and phrases:
linear divergence, quasi-redirecting boundary, 3-manifold groups
2010 Mathematics Subject Classification:
20F65, 20F67

1. Introduction

The quasi-redirecting (QR) boundary is a close generalization of the Gromov boundary to all finitely generated groups [QR24]. One of the advantages of the QR-boundary is that it is a new quasi-isometry invariant boundary that is often compact, containing sublinearly Morse boundaries [QRT22], [QRT24] as topological subspaces, capturing a richer spectrum of hyperbolic-like behaviors, making it a promising new tool in geometric group theory.

The QR boundary is defined as follows:

Definition 1.1.

Let α,β:[0,)X\alpha,\beta\colon[0,\infty)\to X be two quasi-geodesic rays in a metric space XX. We say α\alpha can be quasi-redirected to β\beta (and write αβ\alpha\preceq\beta) if there exists a pair of constants (q,Q)(q,Q) such that for every r>0r>0, there exists a (q,Q)(q,Q)–quasi-geodesic ray γ\gamma that is identical to α\alpha inside the ball B(α(0),r)B(\alpha(0),r) and eventually γ\gamma becomes identical to β\beta. We say αβ\alpha\sim\beta if αβ\alpha\preceq\beta and βα\beta\preceq\alpha.

The resulting set of equivalence classes forms a poset, denoted by P(X)P(X) (in this paper, we will call it the QR-poset). This poset P(X)P(X), when equipped with a “cone-like topology” (see [QR24, Section 5]), is called the quasi-redirecting boundary (QR boundary) of XX and denoted by X\partial_{*}X.

Qing and Rafi [QR24] established key properties of the QR-boundary, with further developments in [GQV24]. While QR boundaries are shown to be well-defined for several classes of groups of interest, including relatively hyperbolic groups, Croke-Kleiner admissible groups, non-geometric 3-manifold groups [QR24], [NQ25], its existence for all finitely generated groups remains an open question.

Question 1.2.

[QR24, Question D] Let XX be a Cayley graph of a finitely generated group. Is X\partial_{*}X always defined? Is X\partial_{*}X always compact?

In [QR24, Section 4], the authors show that the QR-boundary of a direct product of two infinite finitely generated groups consists exactly of one point. They also mention the work in [McM] where the author shows the same holds for Baumslag-Solitar group. These groups share a common feature: linear divergence, a quasi-isometry invariant that measures the minimal path length outside a ball connecting two points on its boundary, as a function of the ball’s radius [Gro96], [Ger94a], [DMS10] (see the precise definition in Section 2.3).

This observation naturally raises the following question:

Question 1.3.

For a finitely generated group GG, does linear divergence imply that G\partial_{*}G is a single point?

Our result provides an affirmative answer to Question 1.3.

Theorem 1.4.

Let GG be a finitely generated group. If GG has linear divergence, then the QR-boundary G\partial_{*}G consists of exactly one point.

This result confirms the existence and triviality of the QR-boundary for groups with linear divergence, adding to known examples such as:

  • Lattices in semi-simple Lie groups of \mathbb{Q}–rank 11 and \mathbb{R}–rank 2\geq 2, uniform lattices in higher rank semi-simple Lie groups [DMS10].

  • Thompson groups FF, TT, and VV [GS19] and higher Thompson groups [Kod24].

  • Non-virtually cyclic groups that satisfy a law [DS05], [DMS10].

  • One-ended solvable groups [DS05], [DMS10].

  • Wreath products, permutational wreath products of groups, Houghton groups m\mathcal{H}_{m} with m2m\geq 2, Baumslag-Solitar groups [I23].

Linear divergence is equivalent to wide (i.e., not having cut-points in the asymptotic cones) [DMS10], and wide groups have empty Morse boundary [DMS10]. [QR24, Question 4.4] asks if GG does not have an Morse element, is P(G)P(G) a single point. In [GQV24], the authors answer [QR24, Question 4.4] in the affirmative when GG acts geometrically on a finite-dimensional CAT(0) cube complex. Our result Theorem 1.4 gives the affirmative for [QR24, Question 4.4] for the class of wide groups.

[NQ25, Theorem A, Theorem B] show that the QR poset of graph manifold groups, and more generally of Croke-Kleiner admissible groups [CK02], has QR-poset of height 22. This is connected to the fact that these groups have quadratic divergence. Our result shows that groups with linear divergence have QR poset of height 11. This naturally raises the question of whether there is a systematic relationship between divergence and QR-boundary structure.

Question 1.5.

If a group has divergence that is a polynomial of degree dd, is it true that its QR poset has height dd?

As an application of Theorem 1.4, we establish a comprehensive result for finitely generated 3-manifold groups:

Theorem 1.6.

All finitely generated 3-manifold groups have well-defined QR-boundaries.

This result addresses cases left unresolved in [NQ25]. While [NQ25, Theorem A] showed that QR-boundaries are well-defined for fundamental groups of non-geometric 3-manifolds, the existence of QR-boundaries for geometric 3-manifolds–particularly those modeled on the Sol and Nil geometries and the broader scenario of 3-manifolds with higher genus boundaries was not completely settled.

These cases were excluded precisely because it was unknown whether their fundamental groups Sol and Nil 3-manifolds satisfy the necessary QR-assumptions. Since Sol and Nil 3-manifold groups are known to have linear divergence [Ger94], it follows from Theorem 1.4 that their fundamental groups have well-defined QR-boundary. We use this observation as a first step to conclude that all finitely generated 3-manifold groups admit a well-defined QR-boundary. Theorem 1.6 strengthens the role of QR boundaries as a tool for studying the coarse geometry of finitely generated 3-manifold groups, one of the central topics in geometric group theory.

Overview

This paper is organized as follows. Section 2 reviews preliminary concepts, including the QR-boundary construction and divergence. Section 2 proves Theorem 1.4, demonstrating that groups with linear divergence have a single-point QR-boundary. In Section 4, we give a proof of Theorem 1.6.

Acknowledgments

The author thanks Yulan Qing and Minh Nhat Doan for the helpful conversations.

2. Preliminary

2.1. Coarse geometry

In this section, we recall the construction of quasi-redirecting boundary as presented in [QR24]. Let XX and YY be metric spaces and ff be a map from XX to YY. Let 𝔮=(q,Q)[1,)×[0,){\mathfrak{q}}=(q,Q)\in[1,\infty)\times[0,\infty) be a pair of constants.

Definition 2.1.
  1. (1)

    We say that ff is a (q,Q)(q,Q)–quasi-isometric embedding if for all x,yXx,y\in X,

    1qd(x,x)Qd(f(x),f(x))qd(x,x)+Q.\frac{1}{q}d(x,x^{\prime})-Q\leq d(f(x),f(x^{\prime}))\leq qd(x,x^{\prime})+Q.
  2. (2)

    We say that ff is a (q,Q)(q,Q)–quasi-isometry if it is a (q,Q)(q,Q)–quasi-isometric embedding such that Y=NQ(f(X))Y=N_{Q}(f(X)).

Definition 2.2.

A quasi-geodesic in a metric space XX is a quasi-isometric embedding α:IX\alpha:I\to X where II\subset\mathbb{R} is a (possibly infinite) interval. That is α:IX\alpha:I\to X is a (q,Q)(q,Q)–quasi-geodesic if for all s,tIs,t\in I, we have

|ts|qQdX(α(s),α(t))q|st|+Q\frac{|t-s|}{q}-Q\leq d_{X}\big{(}\alpha(s),\alpha(t)\big{)}\leq q|s-t|+Q
Remark 2.3.

We can always assume α\alpha is (2q+2Q)(2q+2Q)–Lipschitz, and hence, α\alpha is continuous. By [QR24, Lemma 2.3] the Lipschitz assumption can be made without loss of generality.

Notation:

Let oo be a fixed base-point in XX. We use 𝔮=(q,Q)[1,)×[0,){\mathfrak{q}}=(q,Q)\in[1,\infty)\times[0,\infty) to indicate a pair of constants. For instance, one can say Φ:XY\Phi\colon X\to Y is a 𝔮{\mathfrak{q}}–quasi-isometry and α\alpha is a 𝔮{\mathfrak{q}}–quasi-geodesic ray or segment.

  • By a 𝔮{\mathfrak{q}}–ray we mean a 𝔮{\mathfrak{q}}–quasi-geodesic ray α:[0,)X\alpha:[0,\infty)\to X such that α(0)=o\alpha(0)=o.

  • If points x,yXx,y\in X on the image of α\alpha are given, we denote the sub-segment of α\alpha connecting xx to yy by [x,y]α[x,y]_{\alpha}.

  • For r>0r>0, let BrXB_{r}^{\circ}\subset X be the open ball of radius rr centered at oo, let BrB_{r} be the closed ball centered at oo and let Brc=XBrB_{r}^{c}=X-B_{r}^{\circ}. For a 𝔮\mathfrak{q}–ray α\alpha and r>0r>0, we let tr0t_{r}\geq 0 denote the first time when α\alpha first intersects BrcB_{r}^{c}.

    Lastly, if pp is a point on a 𝔮{\mathfrak{q}}–ray α\alpha, we use α[p,)\alpha_{[p,\infty)} to denote the tail of α\alpha starting from the point pp.

2.2. QR-Assumptions

In this section, we briefly review the notion QR-poset and QR-redirecting boundary from [QR24].

Definition 2.4.

Let XX be a geodesic metric space. Let α,β\alpha,\beta and γ\gamma be quasi-geodesic rays in XX. We say

  1. (1)

    γ\gamma eventually coincides with β\beta if there are times tβ,tγ>0t_{\beta},t_{\gamma}>0 such that, for ttγt\geq t_{\gamma}, we have γ(t)=β(t+tβ)\gamma(t)=\beta(t+t_{\beta}).

  2. (2)

    For r>0r>0, we say γ\gamma quasi-redirects α\alpha to β\beta at radius rr if γ|r=α|r\gamma|_{r}=\alpha|_{r} and β\beta eventually coincides with γ\gamma. If γ\gamma is a 𝔮{\mathfrak{q}}–ray, we say α\alpha can be 𝔮{\mathfrak{q}}–quasi-redirected to β\beta at radius rr or α\alpha can be 𝔮{\mathfrak{q}}–quasi-redirected to β\beta by γ\gamma at radius rr. We refer to tγt_{\gamma} as the landing time.

  3. (3)

    We say α\alpha is quasi-redirected to β\beta, denoted by αβ\alpha\preceq\beta, if there is 𝔮[1,)×[0,){\mathfrak{q}}\in[1,\infty)\times[0,\infty) such that for every r>0r>0, α\alpha can be 𝔮{\mathfrak{q}}–quasi-redirected to β\beta at radius rr.

Definition 2.5.

Define αβ\alpha\simeq\beta if and only if αβ\alpha\preceq\beta and βα\beta\preceq\alpha. Then \simeq is an equivalence relation on the space of all quasi-geodesic rays in XX.

Let P(X)P(X) denote the set of all equivalence classes of quasi-geodesic rays under \simeq. For a quasi-geodesic ray α\alpha, let [α]P(X)[\alpha]\in P(X) denote the equivalence class containing α\alpha. We extend \preceq to P(X)P(X) by defining [α][β][\alpha]\preceq[\beta] if αβ\alpha\preceq\beta. Note that this does not depend on the chosen representative in the given class. The relation \preceq is a partial order on elements of P(X)P(X). We call P(X)P(X) the QR-poset of XX.

QR-Assumption 0: (No dead ends)

The metric space XX is proper and geodesic. Furthermore, there exists a pair of constants 𝔮0{\mathfrak{q}}_{0} such that every point xXx\in X lies on an infinite 𝔮0{\mathfrak{q}}_{0}–quasi-geodesic ray.

QR-Assumption 1: (Quasi-geodesic representative)

For 𝔮0{\mathfrak{q}}_{0} as in QR-Assumption 0, every equivalence class of quasi-geodesics 𝐚P(X)\mathbf{a}\in P(X) contains a 𝔮0{\mathfrak{q}}_{0}–ray. We fix such a 𝔮0{\mathfrak{q}}_{0}–ray, denote it by a¯𝐚\underline{a}\in\mathbf{a}, and call it a central element of 𝐚\mathbf{a}.

QR-Assumption 2: (Uniform redirecting function)

For every 𝐚P(X)\mathbf{a}\in P(X), there is a function

f𝐚:[1,)×[0,)[1,)×[0,),f_{\mathbf{a}}:\,[1,\infty)\times[0,\infty)\to[1,\infty)\times[0,\infty),

called the redirecting function of the class 𝐚\mathbf{a}, such that if 𝐛𝐚\mathbf{b}\prec\mathbf{a} then any 𝔮{\mathfrak{q}}–ray β𝐛\beta\in\mathbf{b} can be f𝐚(𝔮)f_{\mathbf{a}}({\mathfrak{q}})–quasi-redirected to a¯\underline{a}.

Quasi-redirecting boundary (QR-boundary):

Once a proper geodesic metric space XX satisfies all three QR-Assumptions, there is a “cone-line” topology on the poset P(X)P(X) described on [QR24]. This poset P(X)P(X), when equipped with this topology, is called the quasi-redirecting boundary (QR boundary) of XX and denoted by X\partial_{*}X. Since we don’t use this topology on P(X)P(X) in an essential way in this paper, we refer the reader to [QR24] for the detailed discussion.

A remarkable fact about QR-boundary is the following result.

Theorem 2.6 ([QR24, Theorem B, Theorem C]).

Let X,YX,Y be proper geodesic metric spaces satisfying all three QR-Assumptions.

  1. (1)

    A quasi-isometry f:XYf\colon X\to Y induces a homeomorphism between X\partial_{*}X and Y\partial_{*}Y.

  2. (2)

    Sublinearly Morse boundaries are topological subspaces of X\partial_{*}X.

2.3. Divergence of groups

In this section, we briefly review the definition of divergence from [DMS10].

Definition 2.7.

Let \mathcal{F} be the collection of all functions from positive reals to positive reals. Let ff and gg be arbitrary elements of \mathcal{F}. The function ff is dominated by a function gg, denoted by fgf\preceq g, if there are positive constants AA, BB, CC, DD and EE such that

f(x)Ag(Bx+C)+Dx+Efor all x.f(x)\leq A\,g(Bx+C)+Dx+E\quad\text{for all $x$.}

Two functions ff and gg are equivalent, denoted by fgf\sim g, if fgf\preceq g and gfg\preceq f.

Remark 2.8.

The relation \sim is an equivalence relation on the set \mathcal{F}. Let ff and gg be two polynomial functions with degree at least 11 in \mathcal{F}, then it is not hard to show that they are equivalent if and only if they have the same degree. Moreover, all exponential functions of the form abx+ca^{bx+c}, where a>1a>1, b>0b>0 are equivalent.

Since we mainly work on Cayley graphs of finitely generated groups, we assume in this section that our metric spaces are geodesic, proper and periodic spaces.

For such a space XX, given three points a,b,cXa,b,c\in X and parameters δ(0,1)\delta\in(0,1) and γ0\gamma\geq 0 we define divergence divγ(a,b,c;δ)\operatorname{div}_{\gamma}(a,b,c;\delta) to be the infimum of lengths of paths that connect aa to bb outside Bo(c,δrγ)B^{o}(c,\delta r-\gamma), the open ball around cc of radius δrγ\delta r-\gamma, if this exists. We define it to be infinite otherwise. Here r=min{d(c,a),d(c,b)}r=\operatorname{min}\{d(c,a),d(c,b)\}. We then define

Div(n,δ)=supa,b,cX,d(a,b)ndiv(a,b,c;δ)\operatorname{Div}(n,\delta)=\operatorname{sup}_{a,b,c\in X,d(a,b)\leq n}\operatorname{div}(a,b,c;\delta)

Since a space has more than one end then its divergence is infinite, we thus restrict to one-ended spaces. Furthermore, we can fix a third point c=x0c=x_{0} in the definition and assume that a,ba,b are in the sphere Sr:=S(x0,r)S_{r}:=S(x_{0},r) and Divγ(n,δ)\operatorname{Div}_{\gamma}(n,\delta) can be modified to:

Divγ(n,δ)=supa,bS(x0,r)divγ(a,b,x0;δ)\operatorname{Div}_{\gamma}(n,\delta)=\operatorname{sup}_{a,b\in S(x_{0},r)}\operatorname{div}_{\gamma}(a,b,x_{0};\delta)

It is shown by [DMS10] that Divγ(n,δ)\operatorname{Div}_{\gamma}(n,\delta) is independent of γ\gamma and δ\delta up to \sim for any δ1/2\delta\leq 1/2 and γ2\gamma\geq 2 and is invariant under quasi-isometry up to \sim. Thus in this paper, we think of Div(X)\operatorname{Div}(X) as a function of nn, defining it to be equal to Div2(n,1/2)\operatorname{Div}_{2}(n,1/2). We say that the divergence is linear if Div2(n,1/2)n\operatorname{Div}_{2}(n,1/2)\sim n, quadratic if Div2(n,1/2)n2\operatorname{Div}_{2}(n,1/2)\sim n^{2}, and so on.

3. Quasi-redirecting boundaries of groups with linear divergence

In this section, we are going to prove Theorem 1.4. To prove Theorem 1.4, we show that the cayley graph XX of GG (with respect to a finite generating set), any two quasi-geodesic rays α\alpha and β\beta are equivalent under the QR-relation. We achieve this by constructing quasi-geodesic rays that direct α\alpha to β\beta by using the linear divergence property.

We need the following lemmas.

Lemma 3.1 ([QR24, Lemma 2.6]).

Let XX be a metric space that satisfies QR-Assumption 0. (Nearest-point projection surgery) Consider a point xXx\in X and a (q,Q)(q,Q)–quasi-geodesic segment β\beta connecting a point zXz\in X to a point wXw\in X. Let yy be a closest point in β\beta to xx. Then

γ=[x,y][y,z]β\gamma=[x,y]\cup[y,z]_{\beta}

is a (3q,Q)(3q,Q)–quasi-geodesic.

Lemma 3.2.

[NQ25, Lemma 2.9] Let α,β\alpha,\beta be quasi-geodesic rays. Suppose there exists constants 𝔮\mathfrak{q} and a sequence of points {xn}\{x_{n}\} on α\alpha such that normxn\operatorname{norm}{x_{n}}\to\infty and the following holds. For every nn, there exists a 𝔮\mathfrak{q}-ray γn\gamma_{n} such that γn\gamma_{n} eventually coincides with β\beta, and γn\gamma_{n} and α\alpha are identical on the subsegment [o,xn]α[o,x_{n}]_{\alpha}. Then α\alpha can be 𝔮\mathfrak{q}-quasi-redirected to β\beta.

The following lemma follows from the proof of [QR24, Lemma 3.5].

Lemma 3.3.

[QR24, Lemma 3.5] Let XX be a proper, geodesic, metric space and let α\alpha be a 𝔮\mathfrak{q}–ray. Then there exists a geodesic ray α¯\bar{\alpha} such that α¯\bar{\alpha} is (3q,Q)(3q,Q)–quasi-redirected to α\alpha.

Lemma 3.4.

[Tra19, Lemma 3.3]] For each C>1C>1 and ρ(0,1]\rho\in(0,1] there is a constant L=L(C,ρ)1L=L(C,\rho)\geq 1 such that the following holds. Let rr be an arbitrary positive number and γ\gamma a path with the length (γ)<Cr\ell(\gamma)<Cr and d(x,y)>rd(x,y)>r. Then there is an (L,0)(L,0)–quasi-geodesic α\alpha connecting two points xx, yy such that the image of α\alpha lies in the ρr\rho r–neighborhood of γ\gamma and (α)<(γ)\ell(\alpha)<\ell(\gamma).

Lemma 3.5 (Annulus Surgery).

Let XX be a proper geodesic metric space. Given δ>ϵ>0\delta>\epsilon>0 and a constant C>0C>0. Given two pairs of constants (q1,Q1),(q2,Q2)[1,)×[0,)(q_{1},Q1),(q_{2},Q_{2})\in[1,\infty)\times[0,\infty). Then there exists a constant M=M(δ,ϵ,q1,Q1,q2,Q2,C)M=M(\delta,\epsilon,q_{1},Q_{1},q_{2},Q_{2},C) such that the following holds.

Let α\alpha be a (q1,Q1)(q_{1},Q_{1})–quasi-geodesic based at oo with α+\alpha_{+} in the sphere Sϵr:=S(o,ϵr)S_{\epsilon r}:=S(o,\epsilon r) and lies entirely in the ball BϵrB_{\epsilon r} with r>1r>1. Let ζ\zeta be a geodesic ray based at oo and passes through α+\alpha_{+}. Let pp denote the intersection point of ζ\zeta with the sphere SδrS_{\delta r}. Let γ\gamma be a (q2,Q2)(q_{2},Q_{2})–quasi-geodesic based at pp such that β\beta lies entirely outside the open ball BδroB^{o}_{\delta r} and (β)Cr\ell(\beta)\leq Cr. Then the concatenation σ:=α[α+,p]β\sigma:=\alpha\cup[\alpha_{+},p]\cup\beta is a (M,M)(M,M)–quasi-geodesic.

γ\gamma
Refer to caption
ζ\zeta
α\alpha
SδrS_{\delta r}
SϵrS_{\epsilon r}
pp
α+\alpha_{+}
Figure 1. The cacatenation α[α+,p]γ\alpha\cup[\alpha_{+},p]\cup\gamma is a quasi-geodesic.
Proof.

To see this, for every xyσx\neq y\in\sigma, we are going to show that the ratio

([x,y]σ)d(x,y)\frac{\ell([x,y]_{\sigma})}{d(x,y)}

is bounded above by a uniform constant.

According to Lemma 3.1, the concatenations

α[α+,p],[α+,p]γ\alpha\cup[\alpha_{+},p],[\alpha_{+},p]\cup\gamma

are (3q1,Q1)(3q_{1},Q_{1})–quasi-geodesic and (3q2,Q2)(3q_{2},Q_{2})–quasi-geodesic respectively.

We thus only need to consider the case xαx\in\alpha and yγy\in\gamma. On a one hand, we have

([x,y]σ)\displaystyle\ell([x,y]_{\sigma}) (α)+d(α+,p)+(γ)\displaystyle\leq\ell(\alpha)+d(\alpha_{+},p)+\ell(\gamma)
q1d(o,α+)+Q1+(δϵ)r+Cr\displaystyle\leq q_{1}d(o,\alpha_{+})+Q_{1}+(\delta-\epsilon)r+Cr
(q1+δϵ+C)r+Q1<(q1+δϵ+C+Q1)r\displaystyle\leq(q_{1}+\delta-\epsilon+C)r+Q_{1}<(q_{1}+\delta-\epsilon+C+Q_{1})r

On the other hand, since x,yx,y lie outside the annulus Bδr\BϵroB_{\delta r}\backslash B^{o}_{\epsilon r}, we have

d(x,y)(δϵ)rd(x,y)\geq(\delta-\epsilon)r

Thus we have

([x,y]σ)d(x,y)(q1+Q1+C+δϵ)r(δϵ)r=q1+Q1+C+δϵδϵ\frac{\ell([x,y]_{\sigma})}{d(x,y)}\leq\frac{(q_{1}+Q_{1}+C+\delta-\epsilon)r}{(\delta-\epsilon)r}=\frac{q_{1}+Q_{1}+C+\delta-\epsilon}{\delta-\epsilon}

Combining with cases x,yα[α+,p]x,y\in\alpha\cup[\alpha_{+},p] , x,y[α+,p]γ]x,y\in[\alpha_{+},p]\cup\gamma] (which are (3q1,Q1)(3q_{1},Q_{1}) and (3q2,Q2)(3q_{2},Q_{2})–quasi-geodesics respectively), there is a constant MM depending only on constants q1,Q1,q2,Q2,δ,ϵ,Cq_{1},Q_{1},q_{2},Q_{2},\delta,\epsilon,C so that σ\sigma is a (M,M)(M,M)–quasi-geodesic. ∎

The following lemma is a slight modification of [QRT22, Lemma 4.3]. We include it here for completeness.

Lemma 3.6.

(Quasi-geodesic ray to geodesic ray surgery) Let ϵ(0,1)\epsilon\in(0,1). Let β\beta be a geodesic ray and γ\gamma be a (q,Q)(q,Q)–ray. Suppose that there exists an increasing sequence {rn}\{r_{n}\} such that for every nn, d(β(rn),γ)ϵrnd(\beta(r_{n}),\gamma)\leq\epsilon r_{n}. Then γ\gamma can be (9q,Q)(9q,Q)–quasi-redirected to the geodesic ray β\beta.

Proof.

Let qnq_{n} be a point in γ\gamma that is closest to β(rn)\beta(r_{n}) and let Rn>0R_{n}>0 be such that the ball of radius RnR_{n} centered at oo contains [o,qn]γ[o,q_{n}]_{\gamma}. Now let qnq^{\prime}_{n} be the point in [o,qn]γ[o,q_{n}]_{\gamma} closest to β(Rn)\beta(R_{n}). Then

d(o,qn)\displaystyle d(o,q^{\prime}_{n}) d(o,β(Rn))d(β(Rn),qn)=Rnd(β(Rn),qn)\displaystyle\geq d(o,\beta(R_{n}))-d(\beta(R_{n}),q^{\prime}_{n})=R_{n}-d(\beta(R_{n}),q_{n})
Rn(d(β(Rn),β(rn))+d(β(rn),qn))\displaystyle\geq R_{n}-(d(\beta(R_{n}),\beta(r_{n}))+d(\beta(r_{n}),q_{n}))
Rn(Rnrn)ϵrn=(1ϵ)rn\displaystyle\geq R_{n}-(R_{n}-r_{n})-\epsilon r_{n}=(1-\epsilon)r_{n}

By Lemma 3.1, the concatenation ζ:=[o,qn]γ[qn,β(Rn)]\zeta:=[o,q^{\prime}_{n}]_{\gamma}\cup[q^{\prime}_{n},\beta(R_{n})] is a (3q,Q)(3q,Q)–quasi-geodesic. Furthermore, d(o,qn)Rd(o,q^{\prime}_{n})\leq R, it follows that the projection of any point on the geodesic β([Rn,))\beta([R_{n},\infty)) to ζ\zeta is the point β(Rn)\beta(R_{n}). By Lemma 3.1 again, the concatenation ξ:=ζβ([Rn,)\xi:=\zeta\cup\beta([R_{n},\infty) is a (9q,Q)(9q,Q)–quasi-geodesic. Since d(o,qn)(1ϵ)rnd(o,q^{\prime}_{n})\geq(1-\epsilon)r_{n}, it follows that ξ\xi is identical with γ\gamma in the open ball B(1ϵ)rnoB^{o}_{(1-\epsilon)r_{n}}. In other words, γ\gamma can be (9q,Q)(9q,Q)–quasi-redirected to the geodesic ray β\beta at the radius rn:=(1ϵ)rnr^{\prime}_{n}:=(1-\epsilon)r_{n}. As rnr^{\prime}_{n}\to\infty, it follows from Lemma 3.2 that γ\gamma can be (9q,Q)(9q,Q)–quasi-redirected to β\beta. The lemma is proved. ∎

We are now ready for the proof of Theorem 1.4.

Proof of Theorem 1.4.

Fix a finite generating set for GG, and let XX the Cayley graph of GG with respect to this generating set. We aim to prove that the poset P(X)P(X) consists of exactly one point and satisfies all three QR-Assumptions.

Consider two 𝔮\mathfrak{q}-rays α\alpha and β\beta in XX, both based at a vertex oo. By Lemma 3.3, there exist geodesic rays α¯\bar{\alpha} and β¯\bar{\beta} in XX such that α¯α\bar{\alpha}\preceq\alpha and β¯β\bar{\beta}\preceq\beta. We will show that αβ¯\alpha\preceq\bar{\beta}, which implies αβ\alpha\preceq\beta. By a symmetric argument, we can also establish βα\beta\preceq\alpha, thus proving αβ\alpha\sim\beta (and hence the poset P(X)P(X) consists of exactly one point).

Refer to caption
oo
ζ\zeta
β¯\bar{\beta}
α\alpha
pp^{\prime}
aa
pp
bb
ee
dd
β¯(s)\bar{\beta}(s)
Refer to caption
S35s72S_{\frac{35s}{72}}
S71s1442S_{\frac{71s}{144}-2}
SsS_{s}
S(1+C)sS_{(1+C)s}
σ\sigma
Refer to caption
Figure 2. The figure demonstrates the concatenation γ:=[o,p]αp,a][a,b][b,d]σ[d,e]β¯|[e,)\gamma:=[o,p^{\prime}]_{\alpha}\cup p^{\prime},a]\cup[a,b]\cup[b,d]_{\sigma}\cup[d,e]\cup\bar{\beta}|_{[e,\infty)} is the desired quasi-geodesic which quasi-redirects α\alpha to β¯\bar{\beta} at radius 35s72\frac{35s}{72}

Since the divergence of XX is linear, it follows that there exists a constant C>0C>0 such that for every r>1r>1 then

Div2(r,1/2)Cr\operatorname{Div}_{2}(r,1/2)\leq Cr

That is for every xyx\neq y in the sphere S(o,r)S(o,r), there exists a path α\alpha connecting xx to yy lying outside the open ball Br22oB^{o}_{\frac{r}{2}-2} such that (α)Cr\ell(\alpha)\leq Cr.

Recall that trt_{r} denote the first time the path α\alpha first intersects Brc:=X\BroB^{c}_{r}:=X\backslash B^{o}_{r}. Since β¯\bar{\beta} is a geodesic ray, we thus have β¯(tr)=β¯(r)\bar{\beta}(t_{r})=\bar{\beta}(r).

We consider the following cases.

Case 1: There exists a strictly increasing sequence (rn)n(r_{n})_{n\in\mathbb{N}} such that for each nn\in\mathbb{N}, the inequality

d(β¯(rn),α(t35rn72))39rn72d(\bar{\beta}(r_{n}),\alpha(t_{\frac{35r_{n}}{72}}))\leq\frac{39r_{n}}{72}

holds. It follows that

d(β(rn),α)39rn72d(\beta(r_{n}),\alpha)\leq\frac{39r_{n}}{72}

By Lemma 3.6, α\alpha can be (9q,Q)(9q,Q)–quasi-redirected to β¯\bar{\beta}.

Case 2: Case 1 does not hold, that is, for every increasing sequence r1<r2<r_{1}<r_{2}<\ldots, there exists n0n_{0}\in\mathbb{N} such that

d(β¯(rn0),α(t35rn072))>39rn072d(\bar{\beta}(r_{n_{0}}),\alpha(t_{\frac{35r_{n_{0}}}{72}}))>\frac{39r_{n_{0}}}{72}

It follows that there exists an increasing subsequence 4<s1<s2<4<s_{1}<s_{2}<\ldots of {rn}\{r_{n}\} such that

d(β¯(s),α(t35s72))>39s72d(\bar{\beta}(s),\alpha(t_{\frac{35s}{72}}))>\frac{39s}{72}

for all s=sis=s_{i}.

Let ζ\zeta be a geodesic ray based at oo and passes through p:=α(t35s72)p^{\prime}:=\alpha(t_{\frac{35s}{72}}). Let pp denote the intersection point of ζ\zeta with S(o,s)S(o,s). We have

d(β¯(s),p)d(β¯(s),p)d(p,p)39s72(s35s72)=s36d(\bar{\beta}(s),p)\geq d(\bar{\beta}(s),p^{\prime})-d(p^{\prime},p)\geq\frac{39s}{72}-(s-\frac{35s}{72})=\frac{s}{36}

Claim 1: There exists a constant L=L(C)L=L(C), and an (L,0)(L,0)–quasi-geodesic σ\sigma connecting pp to β¯(s)\bar{\beta}(s) such that (σ)<Cs\ell(\sigma)<Cs and σ\sigma lies in the annulus B(1+C)s\B71s1442oB_{(1+C)s}\backslash B^{o}_{\frac{71s}{144}-2}.

Indeed, since both points β¯(s)\bar{\beta}(s) and pp lie in the sphere S(o,s)S(o,s) and the divergence of XX is linear, it follows that there is a path τ\tau connecting pp to β¯(s)\bar{\beta}(s) lying outside the open ball Bo(o,s/22)B^{o}(o,s/2-2) and the length of τ\tau satisfies

(τ)Cs\ell(\tau)\leq Cs

By Lemma 3.4, there exists a constant LL depending only on CC, an (L,0)(L,0)–quasi-geodesic σ\sigma from pp to β¯(s)\bar{\beta}(s) so that (σ)<(τ)<Cs\ell(\sigma)<\ell(\tau)<Cs and σ𝒩s144(τ)\sigma\subset\mathcal{N}_{\frac{s}{144}}(\tau) where 𝒩s144(τ)\mathcal{N}_{\frac{s}{144}}(\tau) we mean the s144\frac{s}{144}–neighborhood of τ\tau. Since τ\tau lies outside the ball B(o,s/22)B(o,s/2-2), it follows that σ\sigma lies outside the ball B71s1442B_{\frac{71s}{144}-2}.

Let R:=max{d(x,o):xσ}R:=\operatorname{max}\{d(x,o):x\in\sigma\}.Then we have σBR\sigma\subset B_{R}. Note that R(σ)+d(β¯(s),o)Cs+s=(1+C)sR\leq\ell(\sigma)+d(\bar{\beta}(s),o)\leq Cs+s=(1+C)s. Thus σ\sigma lies inside the ball B(1+C)sB_{(1+C)s}.

In the rest of the proof, we will construct the desired quasi-geodesic.

Let aa be the intersection point of the geodesic ray ζ\zeta with the sphere S71s1442S_{\frac{71s}{144}-2}. Let bσb\in\sigma be the nearest point to aa. Since σ\sigma is an (L,0)(L,0)–quasi-geodesic, it follows that the concatenation γ1:=[a,b][b,β¯(s)]σ\gamma_{1}:=[a,b]\cup[b,\bar{\beta}(s)]_{\sigma} is a (3L,0)(3L,0)–quasi-geodesic by Lemma 3.1.

Let e:=β¯((1+C)s)e:=\bar{\beta}((1+C)s), and let dγ1=[a,b][b,β¯(s)]σd\in\gamma_{1}=[a,b]\cup[b,\bar{\beta}(s)]_{\sigma} be the nearest point to ee (we refer the reader to Figure 2). Again by Lemma 3.1, we have that the concatenation γ2:=[e,d][d,a]γ1\gamma_{2}:=[e,d]\cup[d,a]_{\gamma_{1}} is a (9L,0)(9L,0)–quasi-geodsic.

We have

(γ2)\displaystyle\ell(\gamma_{2}) d(e,d)+d(a,b)+(σ)\displaystyle\leq d(e,d)+d(a,b)+\ell(\sigma)
d(e,β¯(s))+d(a,p)+Cs\displaystyle\leq d(e,\bar{\beta}(s))+d(a,p)+Cs
Cs+2+74s144+Cs\displaystyle\leq Cs+2+\frac{74s}{144}+Cs
(75144+1+2C)s\displaystyle\leq(\frac{75}{144}+1+2C)s

Applying Lemma 3.5, we have that the concatenation

ξ:=[o,p]α[p,a]γ2\xi:=[o,p^{\prime}]_{\alpha}\cup[p^{\prime},a]\cup\gamma_{2}

is a (M,M)(M,M)–quasi-geodesic where MM is the constant given by Lemma 3.5.

Now, every point uβ¯|(1+C)su\in\bar{\beta}|_{(1+C)s}, its nearest point projection on ξ\xi is ee since ξ\xi lies entirely in the ball B(1+C)sB_{(1+C)s}. According to Lemma 3.1, the concatenation ξβ¯|(1+C)s\xi\cup\bar{\beta}|_{\geq(1+C)s} is a (3M,M)(3M,M)–quasi-geodesic.

Thus, we can redirect α\alpha to the geodesic ray β¯\bar{\beta} at radius ss. Since this can be done for an increasing sequence of radius ss (i.e, the sequence s1<s2<s_{1}<s_{2}<\ldots), it follows from Lemma 3.2 that α\alpha can be (3M,M)(3M,M)–quasi-redirected to β¯\bar{\beta}.

In conclusion, enlarging constants MM if necessary, we have that α\alpha can be (3M,M)(3M,M)–quasi-redirected to β¯\bar{\beta} in both Case 1 and Case 2.

Since β¯\bar{\beta} can be (3q,Q)(3q,Q)–quasi-redirected to β\beta by Lemma 3.3, it follows from [QR24, Lemma 3.2] that α\alpha can be (3q+3M+1,Q+M)(3q+3M+1,Q+M)–quasi-redirected to β\beta.

Using symmetric argument, it can be shown that β\beta is (3q+3M+1,Q+M)(3q+3M+1,Q+M)–quasi-redirected to α\alpha. Therefore

αβ\alpha\sim\beta

In particular, the poset P(X)P(X) consists of exactly one point and satisfies all three QR-Assumptions. Therefore X\partial_{*}X is well-defined and consists of only one point.

4. Quasi-redirecting boundary of finitely generated 3-manifold groups

In this section, we are going to prove Theorem 1.6 which says all finitely generated 3-manifold groups have well-defined QR-boundaries.

A compact orientable irreducible 33–manifold MM with empty or tori boundary is called geometric if its interior admits a geometric structure in the sense of Thurston which are 33–sphere, Euclidean 33–space, hyperbolic 33–space, S2×S^{2}\times\mathbb{R}, 2×\mathbb{H}^{2}\times\mathbb{R}, SL~(2,)\widetilde{SL}(2,\mathbb{R}), Nil and Sol. Otherwise, it is called non-geometric. The QR-boundaries have been studied in [NQ25].

Theorem 4.1.

[NQ25, Theorem A] Let MM be an non-geometric 3-manifold. Then the fundamental group G=π1(M)G=\pi_{1}(M) satisfies the QR-Assumptions and hence G\partial_{*}G is well-defined.

We now address the geometric case.

Lemma 4.2 ( Geometric case).

Let MM be an geometric 3-manifold. Then π1(M)\partial_{*}\pi_{1}(M) is well-defined.

Proof.

MM has a geometric structure modeled on eight geometries in the sense of Thurston: S3S^{3}, 3\mathbb{R}^{3}, S2×S^{2}\times\mathbb{R}, Nil, SL(2,)~\widetilde{SL(2,\mathbb{R})}, 2×\mathbb{H}^{2}\times\mathbb{R}, 3\mathbb{H}^{3}, and Sol.

  1. (1)

    If the geometry of MM is spherical, then its fundamental group is finite, and hence (π1(M))\partial_{*}(\pi_{1}(M)) is empty.

  2. (2)

    If the geometry of MM is 𝔼3\mathbb{E}^{3}, then there exists a finite index subgroup Kπ1(M)K\leq\pi_{1}(M) such that KK is isomorphic to 3\mathbb{Z}^{3}. It follows that π1(M)\pi_{1}(M) is quasi-isometric to 𝔼3\mathbb{E}^{3}, and hence (π1(M))\partial_{*}(\pi_{1}(M)) consists of only one point as 𝔼3\mathbb{E}^{3} has linear divergence.

  3. (3)

    If the geometry of MM is S2×S^{2}\times\mathbb{R}, then there exists a finite index subgroup Kπ1(M)K\leq\pi_{1}(M) such that KK is isomorphic to \mathbb{Z}. It follows that π1(M)\pi_{1}(M) is quasi-isometric to \mathbb{Z}, and hence (π1(M))\partial_{*}(\pi_{1}(M)) is well-defined and consists of only two points.

  4. (4)

    If the geometry of MM is 2×\mathbb{H}^{2}\times\mathbb{R}, then MM is finitely covered by M=Σ×S1M^{\prime}=\Sigma\times S^{1} where Σ\Sigma is a compact surface with negative Euler characteristic. It follows that π1(M)\pi_{1}(M) is quasi-isometric to the direct product π1(Σ)×\pi_{1}(\Sigma)\times\mathbb{Z}, and hence (π1(M))\partial_{*}(\pi_{1}(M)) is well-defined and consists of only one point.

  5. (5)

    If MM has a geometry modeled on SL(2,)~\widetilde{SL(2,\mathbb{R})}, then (π1(M))\partial_{*}(\pi_{1}(M)) is well-defined and consists of only one point since two geometries 2×\mathbb{H}^{2}\times\mathbb{R} and SL(2,)~\widetilde{SL(2,\mathbb{R})} are quasi-isometric and QR-boundary is a quasi-isometric invariant (Theorem 2.6).

  6. (6)

    If MM has a geometry modeled on 3\mathbb{H}^{3} then MM is a hyperbolic 3-manifold with finite volume. (π1(M))\partial_{*}(\pi_{1}(M)) is well-defined as shown in [QR24].

  7. (7)

    Finally, If MM has a geometry modeled on Sol or Nil then π1(M)\pi_{1}(M) has linear divergence (see the first and second paragraphs in the proof of [Ger94, Theorem 4], and thus Theorem 1.4 implies that (π1(M))\partial_{*}(\pi_{1}(M)) is well-defined and consists of only one point.

Below, we explain how one might reduce the study of all finitely generated 3–manifold groups to the case of compact, orientable, irreducible, \partial–irreducible 3–manifold groups.

Let MM be 3-manifold with finitely generated fundamental group. By Scott’s Core Theorem, MM contains a compact codimension zero submanifold whose inclusion map is a homotopy equivalence [Sco73], and thus also an isomorphism on fundamental groups. We thus can assume MM is compact. Since QR-boundary is a quasi-isometric invariant (Theorem 2.6), we can assume that M is orientable by passing to a double cover if necessary.

We can also assume that MM is irreducible and \partial–irreducible by the following reason: Since MM is a compact, orientable 33–manifold, it decomposes into irreducible, \partial–irreducible pieces M1,,MkM_{1},\dots,M_{k} (by the sphere-disc decomposition). In particular, π1(M)\pi_{1}(M) is the free product

π1(M)=π1(M1)π1(M2)π1(Mk)\pi_{1}(M)=\pi_{1}(M_{1})*\pi_{1}(M_{2})*\cdots*\pi_{1}(M_{k})

Let Gi=π1(Mi)G_{i}=\pi_{1}(M_{i}). We remark here that π1(M)\pi_{1}(M) is hyperbolic relative to the collection ={G1,,Gk}\mathbb{P}=\{G_{1},\cdots,G_{k}\}. According to [NQ25, Theorem D], if the QR-boundary exist for each peripheral subgroup GiG_{i} then the QR-boundary of π1(M)\pi_{1}(M) exists. In other words, for the purpose of showing QR-boundary exists for π1(M)\pi_{1}(M), we only need to focus on the case where the manifold MM is compact, connected, orientable, irreducible, and \partial–irreducible.

If MM has empty or toroidal boundary, the existence of QR-boundary of π1(M)\pi_{1}(M) would follow from Theorem 4.1 and Lemma 4.2.

We note that the compact, connected, orientable, irreducible and \partial–irreducible manifold MM could have boundary components that are higher genus surfaces. The following lemma addresses certain manifolds with higher genus boundary.

Lemma 4.3.

Let MM be a compact, orientable, irreducible, \partial–irreducible 33–manifold which has at least one boundary component of genus at least 22. Then the fundamental group G=π1(M)G=\pi_{1}(M) satisfies the QR-Assumptions and hence G\partial_{*}G is well-defined.

Proof.

As in [Sun20, Section 6.3], we can paste compact hyperbolic 3-manifolds with totally geodesic boundaries to the higher genus boundary components of MM to obtain a finite volume hyperbolic manifold NN (in case MM has trivial torus decomposition) or a mixed 3-manifold (in case MM has non-trivial torus decomposition). By [NQ25, Theorem A], π1(N)\pi_{1}(N) has well-defined QR-boundary. The new manifold NN satisfies the following properties.

  1. (1)

    MM is a submanifold of NN with incompressible tori boundary.

  2. (2)

    The torus decomposition of MM also gives the torus decomposition of NN.

  3. (3)

    Each piece of MM with a boundary component of genus at least 22 is contained in a hyperbolic piece of NN.

Since NN contains at least one hyperbolic piece, we equip NN with a nonpositively curved metric as in [Leeb95]. This metric induces a metric on the universal cover N~\tilde{N}.

Let M1,,MkM^{\prime}_{1},\dots,M^{\prime}_{k} be the pieces of MM that satisfies (3). Let NiN^{\prime}_{i} be the hyperbolic piece of NN such that MiM^{\prime}_{i} is contained in NiN^{\prime}_{i}. We remark here that it has been proved in [NS19] that the inclusion of the subgroup π1(M)\pi_{1}(M) in π1(N)\pi_{1}(N) is a quasi-isometric embedding (see the proof of Case 1.2 in the proof of Theorem 1.3 in [NS19]), and hence the inclusion M~N~\widetilde{M}\to\widetilde{N} is a (λ1,c1)(\lambda_{1},c_{1})–quasi-isometric embedding for some uniform constants λ1,c1\lambda_{1},c_{1}.

Since QR-boundary is a quasi-isometric invariant (see Theorem 2.6), it suffices to show that the QR-boundary exists for the universal cover M~\widetilde{M}.

Fix a base point oM~o\in\widetilde{M}. Let α\alpha and β\beta be two 𝔮\mathfrak{q}–rays in M~\widetilde{M} based at oo. Since the inclusion M~N~\widetilde{M}\to\widetilde{N} is a (λ1,c1)(\lambda_{1},c_{1})–quasi-isometric embedding, α\alpha and β\beta are 𝔮\mathfrak{q}^{\prime}–rays in N~\widetilde{N} for some q=q(q,Q,λ1,c1)q^{\prime}=q^{\prime}(q,Q,\lambda_{1},c_{1}) and Q=Q(q,Q,λ1,c1)Q^{\prime}=Q^{\prime}(q,Q,\lambda_{1},c_{1}).

Claim 1: There exists a function :[1,)×[0,)[1,)×[0,)\mathcal{M}\colon[1,\infty)\times[0,\infty)\to[1,\infty)\times[0,\infty) such that the following holds. At every radius rr, if α\alpha can be quasi-redirected to β\beta at radius rr via a (A,B)(A,B)–quasi-geodesic γ\gamma in N~\widetilde{N} then α\alpha can be quasi-redirected to β\beta at radius rr via a (A,B)\mathcal{M}(A,B)–quasi-geodesic γ\gamma^{\prime} which lies entirely in M~N~\widetilde{M}\subset\widetilde{N}.

Proof of the claim.

Indeed, let ss be the landing time of γ\gamma on β\beta. By the construction of the manifold NN, the restriction of γ\gamma on [0,s][0,s] can be decomposed into a concatenation

γ|[0,s]=α1β1α2β2αβα+1\gamma|_{[0,s]}=\alpha_{1}\cdot\beta_{1}\cdot\alpha_{2}\cdot\beta_{2}\cdots\alpha_{\ell}\cdot\beta_{\ell}\cdot\alpha_{\ell+1}

such that:

  • For each jj, the subpath αj\alpha_{j} is a subset of M~\tilde{M}, and βj\beta_{j} intersects M~\tilde{M} only at its endpoints. Here α1\alpha_{1} and α+1\alpha_{\ell+1} might degenerate to points.

  • Moreover, there are pieces M~j\tilde{M}^{\prime}_{j} and N~j\tilde{N}^{\prime}_{j} of M~\tilde{M} and N~\tilde{N} respectively such that M~jN~j\tilde{M}^{\prime}_{j}\subset\tilde{N}^{\prime}_{j}, βjN~j\beta_{j}\subset\tilde{N}^{\prime}_{j}, and the endpoints of βj\beta_{j} lies in Σ~jM~j\widetilde{\Sigma}_{j}\subset\tilde{M}^{\prime}_{j} where Σ~j\widetilde{\Sigma}_{j} is the universal cover of some boundary component of genus at least 2 of MjM_{j}.

Since each inclusion π1(Mj)π1(Nj)\pi_{1}(M_{j})\to\pi_{1}(N_{j}) is a quasi-isometric embedding, and π1(Nj)\pi_{1}(N_{j}) is a hyperbolic group, the subgroup π1(Mj)\pi_{1}(M_{j}) is a quasi-convex in π1(Nj)\pi_{1}(N_{j}) by Morse Lemma.

Since there are finitely many pieces MjM_{j}’s, it follows that there exists a function :[1,)×[0,)[1,)×[0,)\mathcal{M}^{\prime}\colon[1,\infty)\times[0,\infty)\to[1,\infty)\times[0,\infty) such that for every (λ,c)(\lambda,c)–quasi-geodesic β\beta in N~j\widetilde{N}_{j} with endpoints in M~j\widetilde{M}_{j} will eventually lies in the (λ,c)\mathcal{M}^{\prime}(\lambda,c)–distance from a geodesic in M~j\widetilde{M}_{j} connecting the two endpoints of the quasi-geodesic β\beta. Thus each (A,B)(A,B)–quasi-geodesic βj\beta_{j} in N~j\widetilde{N}_{j} lies within a RR-distance from a geodesic in M~j\widetilde{M}_{j} connecting two endpoints of βj\beta_{j}, denoted βj\beta^{\prime}_{j} in M~j\widetilde{M}_{j}. Define

γ¯:=α1β1α2β2αβα+1\bar{\gamma}:=\alpha_{1}\cdot\beta^{\prime}_{1}\cdot\alpha_{2}\cdot\beta^{\prime}_{2}\cdots\alpha_{\ell}\cdot\beta^{\prime}_{\ell}\cdot\alpha_{\ell+1}

and

γ:=γ¯γ|[s,)\gamma^{\prime}:=\bar{\gamma}\cup\gamma|_{[s,\infty)}

that lies entirely in M~\widetilde{M}. Since each βj\beta_{j} lies within a RR–distance from βj\beta^{\prime}_{j} and αjγ\alpha_{j}\subset\gamma and γ\gamma is a quasi-geodesic ray, it follows that γ\gamma^{\prime} is a (A,B))\mathcal{M}(A,B))–quasi-geodesic ray. ∎

Claim 2: M~\widetilde{M} satisfies all three QR-Assumptions, and hence it has well-defined QR-boundary.

For every quasi-geodesic ray α\alpha in M~\widetilde{M}, it is also a quasi-geodesic ray in N~\widetilde{N} since the inclusion M~N~\widetilde{M}\to\widetilde{N} is a quasi-isometric embedding. In [NQ25], the class [α]P(N~)[\alpha]\in P(\widetilde{N}) contains a geodesic representative α¯\underline{\alpha} and this geodesic lies in M~\widetilde{M} by our construction of NN (i.e, the torus decomposition of MM also gives the torus decomposition of NN) . We thus consider α¯\underline{\alpha} is a representative of [a]P(M~)[a]\in P(\widetilde{M}) as well.

Given 𝐚P(M~)\mathbf{a}\in P(\widetilde{M}), we consider i(𝐚)P(N~)i(\mathbf{a})\in P(\widetilde{N}) where i:P(M~)P(N~)i\colon P(\widetilde{M})\to P(\widetilde{N}) is the inclusion. By [NQ25, Theorem A] there is a function

fi(𝐚):[1,)×[0,)[1,)×[0,),f_{i(\mathbf{a})}:\,[1,\infty)\times[0,\infty)\to[1,\infty)\times[0,\infty),

called the redirecting function of the class i(𝐚)i(\mathbf{a}). Now let 𝐛P(M~)\mathbf{b}\in P(\widetilde{M}) such that 𝐛𝐚\mathbf{b}\prec\mathbf{a}. In particular, 𝐛i(𝐚)\mathbf{b}\prec i(\mathbf{a}) such that if 𝐛i(𝐚)\mathbf{b}\prec i(\mathbf{a}) where 𝐛P(N~)\mathbf{b}\in P(\widetilde{N}) then any 𝔮{\mathfrak{q}}–ray β𝐛\beta\in\mathbf{b} can be fi(𝐚)(𝔮)f_{i(\mathbf{a})}({\mathfrak{q}})–quasi-redirected to a¯\underline{a} via fi(𝐚)(𝔮)f_{i(\mathbf{a})}(\mathfrak{q})–rays in N~\widetilde{N}. According to Claim 1, such fi(𝐚)(𝔮)f_{i(\mathbf{a})}(\mathfrak{q})–rays in N~\widetilde{N} can be modified to be (fi(𝐚)(𝔮))\mathcal{M}(f_{i(\mathbf{a})}(\mathfrak{q}))–rays in M~\widetilde{M} that quasi-redirect β\beta to a¯\underline{a}. This shows that M~\widetilde{M} would satisfy all three QR-Assumptions and thus π1(M)\pi_{1}(M) has well-defined QR-boundary. ∎

Proof of Theorem 1.6.

By applying Scott’s Core Theorem and, if needed, passing to a double cover, combined with the quasi-isometric invariance of the QR-boundary, we may assume MM is compact and orientable. We then employ sphere-disc decomposition to decompose MM into irreducible, \partial-irreducible manifolds M1,M2,,MkM_{1},M_{2},\ldots,M_{k}. Consequently, the fundamental group decomposes as a free product: π1(M)=π1(M1)π1(M2)π1(Mk)\pi_{1}(M)=\pi_{1}(M_{1})*\pi_{1}(M_{2})*\cdots*\pi_{1}(M_{k}). By [NQ25, Theorem D], the existence of a QR-boundary for each π1(Mi)\pi_{1}(M_{i}) implies its existence for π1(M)\pi_{1}(M). This follows from Theorem 4.1, Lemma 4.2, and Lemma 4.3, completing the proof. ∎

References

  • [CK02] C. Croke, B. Kleiner. The geodesic flow of a nonpositively curved graph manifold. Geom. Funct. Anal., 12(3):479–545, 2002.
  • [DMS10] C. Drutu, S. Mozes, M. Sapir. Divergence in lattices in semisimple Lie groups and graphs of groups. Trans. Amer. Math. Soc. 362 (2010), no. 5, 2451–2505.
  • [DS05] C. Drutu, M. Sapir. Tree-graded spaces and asymptotic cones of groups. Topology 44 (2005), no. 5, 959–1058. With an appendix by Denis Osin and Mark Sapir.
  • [Ger94] S. Gersten. Divergence in 3-manifold groups. Geometric and Functional Analysis (4), 633–647 (1994).
  • [Ger94a] S. Gersten. Quadratic divergence of geodesics in CAT(0)-spaces. Geometric and Functional Analysis (4) (1994) no.1, 37-51.
  • [GQV24] J. Garcia, Y. Qing, E. Vest. Topological and Dynamic Properties of the Sublinearly Morse boundary and the Quasi-Redirecting Boundary. https://arxiv.org/abs/2408.10105.
  • [Gro96] M. Gromov. Geometric group theory, Vol. 2: Asymptotic invariants of infinite groups. Bull. Amer. Math. Soc 33 (1996), no. 0273, 0979.
  • [GS19] G. Golan, M. Sapir. Divergence functions of Thompson groups. Geom. Dedicata 201 (2019), 227–242.
  • [I23] L. Issini. On linear divergence in finitely generated groups. https://arxiv.org/abs/2311.12938.
  • [Kod24] Y. Kodama. Divergence functions of higher-dimensional Thompson’s groups. Pacific J. Math. 333 (2024) 309-329.
  • [Leeb95] B. Leeb. 3-manifolds with(out) metrics of nonpositive curvature. Invent.Math., 122(2):277-289, 1995.
  • [McM] A. McMullin. The Redirecting Boundary of Baumslag-Solitar Groups. In preparation.
  • [NQ25] H. Nguyen, Y. Qing. Quasi-redirecting boundaries of non-positively curved groups. https://arxiv.org/abs/2503.22994.
  • [NS19] H. Nguyen, H. Sun. Subgroup distortion of 3-manifold groups. Trans. Amer. Math. Soc, 373 (2020), no. 9, 6683-6711.
  • [NT23] H. Nguyen, H. Tran. Strongly quasiconvex subgroups in amalgams and HNN extension. Int. J. Algebra Comput., Vol. 33, No. 03, pp. 435-444 (2023).
  • [QR24] Y. Qing, K Rafi. The quasi-redirecting Boundary. https://arxiv.org/abs/2406.16794.
  • [QRT22] Y. Qing, K. Rafi, G. Tiozzo. Sublinearly Morse boundary I: CAT (0) spaces. Adv.Math., 404(part B):Paper No. 108442, 51, 2022.
  • [QRT24] Y. Qing, K. Rafi, G. Tiozzo. Sublinearly Morse boundary, II: Proper geodesic spaces. Geom. Topol., 28(4):1829–1889, 2024.
  • [Sco73] P. Scott. Compact submanifolds of 3-manifolds. J. London Math. Soc. (2) 7 (1973), no.3, 246-250.
  • [Sun20] H. Sun. A characterization on separable subgroups of 3-manifold groups. Journal of Topology 13 (2020), no.1, 187–236.
  • [Tra19] H. Tran. On strongly quasiconvex subgroups. Geom. Topol., 23(3):1173–1235, 2019.