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

Asymptotic shape for subadditve processes on groups of polynomial growth

Cristian F. Coletti and Lucas R. de Lima Centro de Matemática, Computação e Cognição, Universidade Federal do ABC
Av. dos Estados, 5001
09210-580 Santo André, São Paulo
Brazil.
[email protected] [email protected]
Abstract.

This study delves into the exploration of the limiting shape theorem for subadditive processes on finitely generated groups with polynomial growth, commonly referred to as virtually nilpotent groups. Investigating the algebraic structures underlying these processes, we present a generalized form of the asymptotic shape theorem within this framework. Extending subadditive ergodic theory in this context, we consider processes which exhibit both at most and at least linear random growth. We conclude with applications and illustrative examples.

Key words and phrases:
Subadditive cocycle, shape theorem, random growth, groups, Cayley graphs
2020 Mathematics Subject Classification:
Primary: 52A22, 60F15; Secondary: 60K35
Funding: Research supported by grants #2017/10555-0 and #2019/19056-2, São Paulo Research Foundation (FAPESP)

1. Introduction

The investigation of the asymptotic shape for subadditive processes on groups with polynomial growth, often synonymous with virtually nilpotent groups, has recently gained significant attention in the mathematical community. This is in part due to the fact that the usage of subadditive ergodic theorems for the limiting shape relies on vertex-transitive properties that are natural for group actions. Typically, these actions involve translations of the underlying space, providing motivation for the investigation of random processes defined on groups. Our study brings to light the algebraic structures inherent in a class of subadditive processes, offering a generalization beyond the fundamental settings of previously studied models.

The findings presented in this paper hold the potential to deepen our comprehension of various mathematical and scientific phenomena. For instance, they could be instrumental in exploring the geometry of random surfaces or modeling the propagation of information or diseases through networks. The techniques used in this paper could also be applied to other types of random processes on graphs or manifolds.

Benjamini and Tessera [4] were the first to establish an asymptotic shape theorem for First-Passage Percolation (FPP) models on finitely generated groups of subexponential growth with i.i.d. random variables having finite exponential moments. Recently, Auffinger and Gorski [2] demonstrated a converse result, revealing that a Carnot-Carathéodory metric on the associated graded nilpotent Lie group serves as the scaling limit for certain FPP models on a Cayley graph under specified conditions. Broadening the investigation, Cantrell and Furman [8] explored the LL^{\infty} limiting shape for subadditive random processes on groups of polynomial growth. From a probabilistic standpoint, there is considerable interest in relaxing the almost-surely bi-Lipschitz condition imposed by LL^{\infty}. Here, we modify this hypothesis by replacing it with L1L^{1} conditions and introducing hypotheses for at least and at most linear growth. The implications and applicability of this new result are illustrated through examples presented at the end of the article. Notably, we enhance our previous result from [9] on a limiting shape theorem obtained for the Frog Model, now extended to a broader class of non-abelian groups.

Addressing this challenge is primarily approached through the utilization of techniques from metric geometry and geometric group theory. The existence of the limiting shape can be viewed as an extension of Pansu’s theorem to random metrics. The primary strategy involves considering the subadditive cocycle determining a pseudo-quasi-random metric, with the standard case on D\operatorname{\mathbb{Z}}^{D} and D\operatorname{\mathbb{R}}^{D} extensively covered in the literature (see, for instance, [5, 6]).

We describe the process and the obtained theorem below, more detailed definitions can be found in the next section.

Basic description and main results

Let (Ω,,)(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}) be a probability space and (Γ,.)(\operatorname{\Gamma},.) a finitely generated group with polynomial growth rate. Set ϑ:Γ(Ω,,)\operatorname{\upvartheta}:\operatorname{\Gamma}\curvearrowright(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}) to be a \operatorname{\mathbbm{P}}-preserving (p.m.p.) ergodic group action. Consider the family {c(x)}xΓ\{c(x)\}_{x\in\operatorname{\Gamma}} of non-negative random variables such that, \operatorname{\mathbbm{P}}-a.s.,

c(xy)c(y)+c(x)ϑy{c}(xy)\leq c(y)+c(x)\circ\operatorname{\upvartheta}_{y} (1.1)

Write c(x,ω)c(x,\omega) for c(x)(ω)c(x)(\omega) and let zω:=ϑz(ω)z\cdot\omega:=\operatorname{\upvartheta}_{z}(\omega). A function c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} satisfying (1.1) is referred to as a subadditive cocycle. Once given a subadditive cocycle cc, there is a correspondent random pseudo-quasi metric dωd_{\omega} defined by

dzω(x,y):=(c(yx1)ϑx)(zω),d_{z\cdot\omega}(x,y):=\big{(}c(yx^{-1})\circ\operatorname{\upvartheta}_{x}\big{)}(z\cdot\omega),

which is Γ\operatorname{\Gamma}-right equivariant, i.e., for all x,y,zΓx,y,z\in\operatorname{\Gamma}, and for every ωΩ\omega\in\operatorname{\Upomega},

dω(x,y)=dzω(xz1,yz1).d_{\omega}(x,y)=d_{z\cdot\omega}(xz^{-1},yz^{-1}).

The correspondence is one-to-one since given a Γ\operatorname{\Gamma}-right equivariant random pseudoquasimetric dωd_{\omega}, one can easily verify that

c(x,ω):=dω(e,x)c(x,\omega):=d_{\omega}(e,x) (1.2)

is a subadditive cocycle.

To avoid dealing with unnecessary technicalities, we initially consider Γ\operatorname{\Gamma} as the group of polynomial growth, which is nilpotent and torsion-free. Later, we address the more general case where Γ\operatorname{\Gamma} is virtually nilpotent. The essential definitions and notation are introduced as we proceed with the text. The group will be associated with a finite symmetric generating set SΓS\subseteq\operatorname{\Gamma}. We write S\|-\|_{S} and dSd_{S} for a word length and a word metric, respectively. The following conditions will be needed throughout the paper. We assume the existence of β>0\upbeta>0 and κ>1\upkappa>1 such that, for all xΓx\in\operatorname{\Gamma},

(c(x)t)g(t)for all t>βxS\operatorname{\mathbbm{P}}\big{(}c(x)\geq t\big{)}\leq{g}(t)\quad\text{for all }t>\upbeta\|x\|_{S} (i)

where g(t)𝒪(1/t2D+κ)g(t)\in\mathcal{O}\left(1/t^{2D+\upkappa}\right) as t+t\uparrow+\infty.

Let [Γ,Γ][\operatorname{\Gamma},\operatorname{\Gamma}] be the commutator subgroup of Γ\operatorname{\Gamma} and set xSab:=infyx[Γ,Γ]yS\|x\|_{S}^{\operatorname{\operatorname{ab}}}:=\inf_{y\in x[\operatorname{\Gamma},\operatorname{\Gamma}]}\|y\|_{S}. Suppose that there exists a>0a>0 such that, for all xΓ[Γ,Γ]x\in\operatorname{\Gamma}\setminus[\operatorname{\Gamma},\operatorname{\Gamma}] there is a sequence {nj}j\{n_{j}\}_{j\in\operatorname{\mathbb{N}}} of positive integers depending on x[Γ,Γ]x[\operatorname{\Gamma},\operatorname{\Gamma}] with limj+nj=+\lim_{j\uparrow+\infty}n_{j}=+\infty and, for all yx[Γ,Γ]y\in x[\operatorname{\Gamma},\operatorname{\Gamma}] and every jj\in\operatorname{\mathbb{N}},

aynjSab𝔼[c(ynj)].a\|y^{n_{j}}\|_{S}^{\operatorname{\operatorname{ab}}}\leq\operatorname{\mathbbm{E}}\left[c\left(y^{n_{j}}\right)\right]. (ii)

We say that the process grows at least linearly when condition (i) is satisfied. Condition (ii) provides a lower bound for the norm of the rescaled process ϕ\phi, which will be defined later.

To obtain the asymptotic result, we will introduce an innerness assumption. Specifically, for each ε>0\upvarepsilon>0, we require the existence of a finite generating set F(ε)Γ[Γ,Γ]F(\upvarepsilon)\subseteq\operatorname{\Gamma}\setminus[\operatorname{\Gamma},\operatorname{\Gamma}] such that, for \operatorname{\mathbbm{P}}-a.s. ωΩ\omega\in\operatorname{\Upomega} and for every xΓx\in\operatorname{\Gamma}, we can write x=znzn1z1x=z_{n}z_{n-1}\dots z_{1} with zn,zn1,,z1F(ε)z_{n},z_{n-1},\dots,z_{1}\in F(\upvarepsilon) satisfying

i=1nc(zi,zi1z1ω)(1+ε)c(x,ω).\sum_{i=1}^{n}c(z_{i},{z_{i-1}\dots z_{1}}\cdot\omega)\leq(1+\upvarepsilon)c(x,\omega). (iii)

When considering First-Passage Percolation models where SΓ[Γ,Γ]S\subseteq\operatorname{\Gamma}\setminus[\operatorname{\Gamma},\operatorname{\Gamma}], condition (iii) is automatically fulfilled (see Section 2.5). Additionally, in the case where Γ\operatorname{\Gamma} is abelian, we can eliminate the need for hypothesis (iii) in the main theorem altogether.

Theorem 1.1 (Limiting Shape for Torsion-Free Nilpotent Groups).

Let (Γ,.)(\operatorname{\Gamma},.) be a torsion-free nilpotent finitely generated group with polynomial growth rate D1D\geq 1 and torsion-free abelianization. Consider c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} to be a subadditive cocycle associated with dωd_{\omega} and a p.m.p. ergodic group action ϑ\operatorname{\upvartheta}.

Suppose that conditions (i), (ii), and (iii) are satisfied for a finite symmetric generating set SΓS\subseteq\operatorname{\Gamma}. Then

(Γ,1ndω,e)GH(G,dϕ,e)-a.s.\quad\quad\left(\operatorname{\Gamma},\frac{1}{n}d_{\omega},e\right)\operatorname{\xrightarrow{\text{GH}}}\left(G_{\infty},d_{\phi},\mathlcal{e}\right)\quad\quad\operatorname{\mathbbm{P}}\text{-a.s.} (1.3)

where GG_{\infty} is a simply connected graded Lie group, and dϕd_{\phi} is a quasimetric homogeneous with respect to a family of homotheties {δt}t>0\{\delta_{t}\}_{t>0}. Moreover, dϕd_{\phi} is bi-Lipschitz equivalent to dd_{\infty} on GG_{\infty}.

In addition, if Γ\operatorname{\Gamma} is abelian, then (1.3) remains true even when condition (iii) is not valid.

The limit space GG_{\infty} is also known as a Carnot group and dd_{\infty} coincides with the Carnot-Carathéodory metric obtained by the asymptotic cone of Γ\operatorname{\Gamma} as the limit of 1ndS\frac{1}{n}d_{S}. More details about its construction and properties will be given in Section 2 along with the definitions of δt\delta_{t} and dϕd_{\phi}. The usage of the pointed Gromov-Hausdorff convergence arises naturally from its correspondence with geometric group theory.

Let now (Γ,.)(\operatorname{\Gamma},.) be a finitely generated group with polynomial growth rate. Gromov’s Theorem [15] establishes the equivalence of polynomial growth and virtual nilpotency in finitely generated groups. Then there exists a normal nilpotent subgroup NΓN\unlhd\operatorname{\Gamma} with finite index κ:=[Γ:N]<+\kappa:=[\operatorname{\Gamma}:N]<+\infty. Set torN\operatorname{\operatorname{tor}}N to be the torsion subgroup of NN and define

Γ:=N/torN.\operatorname{\Gamma}^{\prime}:=N/\operatorname{\operatorname{tor}}N.

Pansu [16] showed that Γ\operatorname{\Gamma} and Γ\operatorname{\Gamma}^{\prime} share the same asymptotic cone. Let us fix z(j)z_{(j)} as a representative of the coset N(j)=z(j)NN_{(j)}=z_{(j)}N such that Γ=j=1κN(j)\operatorname{\Gamma}=\bigcup_{j=1}^{\kappa}N_{(j)}. Consider z(j)=ez_{(j)}=e when N(j)=NN_{(j)}=N. Set πN:ΓN\uppi_{N}:\operatorname{\Gamma}\to N to be given by πN(x)=z(j)1x\uppi_{N}(x)=z_{(j)}^{-1}x for xN(j)x\in N_{(j)}. Define now :ΓΓ\llbracket-\rrbracket:\operatorname{\Gamma}\to\operatorname{\Gamma}^{\prime} to be given by

x:=πN(x).torN.\llbracket x\rrbracket:=\uppi_{N}(x).\operatorname{\operatorname{tor}}N.

To refine the first main theorem, let us introduce some new conditions. Suppose that there exists a>0a>0 such that, for all xΓx\in\operatorname{\Gamma} there is a sequence {nj}j\{n_{j}\}_{j\in\operatorname{\mathbb{N}}} of positive integers depending on x.[Γ,Γ]\llbracket x\rrbracket.[\operatorname{\Gamma}^{\prime},\operatorname{\Gamma}^{\prime}] with nj+n_{j}\uparrow+\infty as j+j\uparrow+\infty,

axnjS𝔼[c(xnj)].a\|x^{n_{j}}\|_{S}\leq\operatorname{\mathbbm{E}}\left[c\left(x^{n_{j}}\right)\right]. (ii)

Let c:Γ×Ω0c^{\prime}:\operatorname{\Gamma}^{\prime}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} by

c(x):=maxyxztorNc(y)ϑz.c^{\prime}\big{(}\llbracket x\rrbracket\big{)}:=\max_{\begin{subarray}{c}y\in\llbracket x\rrbracket\\ z\in\operatorname{\operatorname{tor}}N\end{subarray}}c(y)\circ\operatorname{\upvartheta}_{z}. (1.4)

Fix, for each xΓ\llbracket x\rrbracket\in\operatorname{\Gamma}^{\prime}, a υxx\upupsilon_{x}\in\llbracket x\rrbracket and consider θ:Γ(Ω,,)\theta:\operatorname{\Gamma}^{\prime}\curvearrowright(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}) given by θxϑυx\theta_{\llbracket x\rrbracket}\equiv\operatorname{\upvartheta}_{\upupsilon_{x}} and θz(ω)=zω\theta_{z}(\omega)=z\ast\omega (see Sec. 3.3 and Remark 5 for a detailed discussion). We consider a similar innerness assumption to replace (iii). Suppose that, for each ε>0\upvarepsilon>0, there exists a finite F(ε)N[N,N]F(\upvarepsilon)\subseteq N\setminus[N,N] which is a generating set of Γ\operatorname{\Gamma}^{\prime} such that, \operatorname{\mathbbm{P}}-a.s., for every xΓx\in\operatorname{\Gamma}, one can write x=znzn1z1\llbracket x\rrbracket=z_{n}z_{n-1}\dots z_{1} with zn,zn1,,z1F(ε)z_{n},z_{n-1},\dots,z_{1}\in F(\upvarepsilon) satisfying

i=1nc(zi,zi1z1ω)(1+ε)c(x,ω).\sum_{i=1}^{n}c^{\prime}(z_{i},~{}{z_{i-1}\dots z_{1}}\ast\omega)\leq(1+\upvarepsilon)c^{\prime}\big{(}\llbracket x\rrbracket,~{}\omega\big{)}. (iii)

Similar to (iii), First-Passage Percolation models satisfy (iii) under specific conditions. In the case where Γ=N\operatorname{\Gamma}=N is nilpotent, it suffices to have SN([N,N]torN)S\subseteq N\setminus\big{(}[N,N]\cup\operatorname{\operatorname{tor}}N\big{)} for an FPP model to satisfy (iii). The virtually nilpotent case is treated separately in Section 3.4 with additional conditions imposed on S\llbracket S\rrbracket and ϑ\operatorname{\upvartheta}. Moreover, when Γ\operatorname{\Gamma}^{\prime} is abelian, hypothesis (iii) is not required to verify the theorem below.

Theorem 1.2 (Limiting Shape for Groups with Polynomial Growth).

Let (Γ,.)(\operatorname{\Gamma},.) be a finitely generated group with polynomial growth rate D1D\geq 1 and Γ/[Γ,Γ]\operatorname{\Gamma}^{\prime}/[\operatorname{\Gamma}^{\prime},\operatorname{\Gamma}^{\prime}] torsion-free. Consider c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} to be a subadditive cocycle associated with dωd_{\omega} and a p.m.p. ergodic group action ϑ\operatorname{\upvartheta}.

Suppose that conditions (i), (ii), and (iii) are satisfied for a finite symmetric generating set SΓS\subseteq\operatorname{\Gamma} is so that S\llbracket S\rrbracket generates Γ\operatorname{\Gamma}^{\prime}. Then

(Γ,1ndω,e)GH(G,dϕ,e)-a.s.\quad\quad\left(\operatorname{\Gamma},\frac{1}{n}d_{\omega},e\right)\operatorname{\xrightarrow{\text{GH}}}\left(G_{\infty},d_{\phi},\mathlcal{e}\right)\quad\quad\operatorname{\mathbbm{P}}\text{-a.s.} (1.5)

where GG_{\infty} is a simply connected graded Lie group, and dϕd_{\phi} is a quasimetric homogeneous with respect to a family of homotheties {δt}t>0\{\delta_{t}\}_{t>0}. Moreover, dϕd_{\phi} is bi-Lipschitz equivalent to dd_{\infty} on GG_{\infty}.

Furthermore, if Γ\operatorname{\Gamma}^{\prime} is abelian, then (1.5) remains true even when condition (iii) is not valid.

The primary technique employed in this work involves the approximation of admissible curves through the use of polygonal paths and ergodic theory. In Section 3, we introduce and delve into these tools, presenting their application in proving the theorems and a corollary for FPP models. Section 4 showcases examples dedicated to illustrating the applicability of the theorems.

2. Auxiliary Theory and Methodological Framework

In this section, we delve into the fundamental concepts of geometric group theory, a field that provides tools to comprehend the relationship between algebraic properties and geometric structures. We begin by establishing the basic definitions that serve as the cornerstone for our exploration. Central to our analysis is the construction of the asymptotic cone, a powerful tool that reveals the geometric behavior of groups at infinity. To illustrate the versatility of our framework, we present concrete examples of groups that satisfy the conditions under consideration.

A key focus of our investigation lies in the construction of the norm in GG_{\infty}, providing the foundation for defining the limiting shape. We explore crucial results and properties in the following subsections. For an in-depth discussion on this topic, we refer interested readers to [7, 11, 10, 17]. This construction leverages subadditive ergodic theorems, revealing the asymptotic behavior of sequences in the group. Through this lens, we gain a deeper understanding of the interplay between algebraic properties and geometric structures.

Building on these concepts, we introduce and elaborate on First-Passage Percolation (FPP) models, serving as a suitable example for a comprehensive exploration of limiting shapes and their implications.

2.1. Cayley graphs and volume growth

The interplay among finitely generated groups, Cayley graphs, word metrics, and the convergence of metric spaces establishes a bridge between the algebraic properties of groups and geometric structures.

Let (Γ,.)(\operatorname{\Gamma},.) be a group generated by a finite set SS. The associated Cayley graph 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) represents elements of GG as vertices, with edges connecting xx and yy if and only if y=sxy=sx for some sSs\in S. Formally, the right-invariant Cayley graph 𝒞(Γ,S)=(V,E)\mathcal{C}(\operatorname{\Gamma},S)=(V,E) is defined by

V=ΓandE={{x,sx}:xΓ,sS}.V=\operatorname{\Gamma}\quad\text{and}\quad E=\big{\{}\{x,sx\}:x\in\operatorname{\Gamma},s\in S\big{\}}.

Cayley graphs provide a visual representation of the group structure and are fundamental in the study of geometric group theory.

Denote by uvu\sim v the relation {u,v}E\{u,v\}\in E. Let 𝒫(x,y)\mathscr{P}(x,y) be the set of self-avoiding paths from xx to yy, where each γ𝒫(x,y)\gamma\in\mathscr{P}(x,y) follows γ=(x0,,xm)\gamma=(x_{0},\dots,x_{m}) with mm\in\operatorname{\mathbb{N}}, xixi+1x_{i}\sim x_{i+1}, x0=xx_{0}=x, xm=yx_{m}=y, and xixjx_{i}\neq x_{j} for all iji\neq j. We write 𝚎γ\mathtt{e}\in\gamma for 𝚎={xi,xi+1}E\mathtt{e}=\{x_{i},x_{i+1}\}\in E, and |γ|=m|\gamma|=m represents the length of the path.

The word length on Γ\operatorname{\Gamma} with respect to SS is defined as follows: For any xΓx\in\operatorname{\Gamma}, the length of the shortest word (or self-avoiding path) in SS that represents xx is its word length, denoted by xS=infγ𝒫(e,x)|γ|\|x\|_{S}=\inf_{\gamma\in\mathscr{P}(e,x)}|\gamma|. The word metric dSd_{S} on Γ\operatorname{\Gamma} is given by dS(x,y)=yx1Sd_{S}(x,y)=\|yx^{-1}\|_{S}.

Throughout this article, various distinct metrics will be considered. Therefore, let us consider a (semi-pseudo-quasi) metric dd_{\diamondsuit} on a non-empty set 𝕏\mathbb{X}, where the metric is indexed by \diamondsuit. We define B(x,r):={y𝕏:d(x,y)<r}B_{\diamondsuit}(x,r):=\{y\in\mathbb{X}\colon d_{\diamondsuit}(x,y)<r\} as the open dd_{\diamondsuit}-ball centered at x𝕏x\in\mathbb{X}. To streamline notation, let tt:=max{t,t}t\vee t^{\prime}:=\max\{t,t^{\prime}\} and tt:=min{t,t}t\wedge t^{\prime}:=\min\{t,t^{\prime}\}. Here, the set \operatorname{\mathbb{N}} stands for {1,2,}\{1,2,\dots\} and 0={0}\operatorname{\mathbb{N}}_{0}=\operatorname{\mathbb{N}}\cup\{0\}.

A finitely generated group Γ\operatorname{\Gamma} has polynomial growth with respect to SS when |BS(e,r)|𝒪(nD)|B_{S}(e,r)|\in\mathcal{O}\big{(}n^{D^{\prime}}\big{)} for a D0D^{\prime}\in\operatorname{\mathbb{N}}_{0} as n+n\uparrow+\infty. The growth is associated with the Cayley graph 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S). The polynomial growth rate of Γ\operatorname{\Gamma} is a constant D0D\in\operatorname{\mathbb{N}}_{0} such that there exists 𝚔(1,+)\mathtt{k}\in(1,+\infty) for all r>1r>1 satisfying

𝚔1rD|BS(e,r)|𝚔rD.\mathtt{k}^{-1}r^{D}\leq|B_{S}(e,r)|\leq\mathtt{k}r^{D}.

Thus D=min{D0:|BS(e,r)|𝒪(nD)}D=\min\big{\{}D^{\prime}\in\operatorname{\mathbb{N}}_{0}\colon|B_{S}(e,r)|\in\mathcal{O}\big{(}n^{D^{\prime}}\big{)}\big{\}}. Moreover, one can verify that the polynomial growth rate of 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) does not depend on the choice of SS.

Recall the definition of the commutator element [x,y]=xyx1y1[x,y]=xyx^{-1}y^{-1} and the subgroup [U,V]:=[u,v]:uU,vV[U,V]:=\big{\langle}[u,v]\colon u\in U,v\in V\big{\rangle} for any U,VΓU,V\subseteq\operatorname{\Gamma}. Set Γ0:=Γ\operatorname{\Gamma}_{0}:=\operatorname{\Gamma} and let Γn:=[Γ,Γn1]\operatorname{\Gamma}_{n}:=[\operatorname{\Gamma},\operatorname{\Gamma}_{n-1}] for all nn\in\operatorname{\mathbb{N}}. Thus {Γn}n\{\operatorname{\Gamma}_{n}\}_{n\in\operatorname{\mathbb{N}}} forms a lower central series with ΓnΓn1\operatorname{\Gamma}_{n}\unlhd\operatorname{\Gamma}_{n-1} for all nn\in\operatorname{\mathbb{N}}. The group Γ\operatorname{\Gamma} is called nilpotent when there is an nn\in\operatorname{\mathbb{N}} such that Γn={e}\operatorname{\Gamma}_{n}=\{e\}, i.e., when the sequence stabilizes in the trivial group. More specifically, Γ\operatorname{\Gamma} is nilpotent of class nn when nn is the minimal value such that Γn\operatorname{\Gamma}_{n} is the trivial group. A group is abelian if and only if it is nilpotent of class 11. The abelianization of a group Γ\operatorname{\Gamma} is given by ΓabΓ/[Γ,Γ]\operatorname{\Gamma}^{\operatorname{\operatorname{ab}}}\cong\operatorname{\Gamma}/[\operatorname{\Gamma},\operatorname{\Gamma}].

The group Γ\operatorname{\Gamma} is called virtually nilpotent when there exists a normal subgroup NΓN\unlhd\operatorname{\Gamma} that is nilpotent with finite index κ=[Γ:N]<+\kappa=[\operatorname{\Gamma}\colon N]<+\infty. A noteworthy result obtained by Gromov [15] is that a finitely generated group has polynomial growth exactly when it is virtually nilpotent. Therefore, the growth established by word metrics is strongly related to algebraic properties of the group.

The torsion subgroup of a group HH is denoted by torH\operatorname{\operatorname{tor}}H and it is defined as the set of all elements with finite order. In other words, torH:=hH:n(hn=e)\operatorname{\operatorname{tor}}H:=\big{\langle}h\in H\colon\exists n\in\operatorname{\mathbb{N}}(h^{n}=e)\big{\rangle}. The group HH is called torsion-free when torH\operatorname{\operatorname{tor}}H is the trivial group.

Let [U]ε[U]_{\varepsilon} to be the ε\varepsilon-neighborhood of U𝕏U\subseteq\mathbb{X} of in a metric space (𝕏,d)(\mathbb{X},d_{\diamondsuit}), i.e., the set [U]ε=uUB(e,ε)[U]_{\varepsilon}=\bigcup_{u\in U}B_{\diamondsuit}(e,\varepsilon). The Hausdorff distance dHd_{H} detects the largest variations between sets with respect to the given metric

dH(U,V):=inf{ε>0:U[V]ε and V[U]ε}.d_{H}(U,V):=\inf\{\varepsilon>0\colon U\subseteq[V]_{\varepsilon}\text{ and }V\subseteq[U]_{\varepsilon}\}.

We define the convergence of metric spaces used in the main theorems employing the Hausdorff distance. Let (𝕏n,dn,on)n(\mathbb{X}_{n},d_{\diamondsuit_{n}},o_{n})_{n\in\operatorname{\mathbb{N}}} be a sequence of centered, locally compact metric spaces. Consider {ψn}n\{\psi_{n}\}_{n\in\operatorname{\mathbb{N}}} as a family of isometric embeddings ψn:(𝕏n,dn,on)(𝕏,d,o)\psi_{n}:(\mathbb{X}_{n},d_{\diamondsuit_{n}},o_{n})\to(\mathbb{X},d_{\diamondsuit},o).

The pointed Gromov-Hausdorff convergence of (𝕏n,dn,on)(\mathbb{X}_{n},d_{\diamondsuit_{n}},o_{n}) to (𝕏,d,o)(\mathbb{X},d_{\diamondsuit},o) is denoted by

(𝕏n,dn,on)GH(𝕏,d,o)(\mathbb{X}_{n},d_{\diamondsuit_{n}},o_{n})\operatorname{\xrightarrow{\text{GH}}}(\mathbb{X},d_{\diamondsuit},o)

and it implies, for all r>0r>0,

limn+dH(ψn(Bn(on,r)),B(o,r))=0.\lim_{n\uparrow+\infty}d_{H}\Big{(}\psi_{n}\big{(}B_{\diamondsuit_{n}}(o_{n},r)\big{)},~{}B_{\diamondsuit}(o,r)\Big{)}=0.

The definitions above are immediately extended to random semi-pseudo-quasi metrics, as employed in the main theorems. The assumption of almost sure local compactness is also maintained. We are now prepared to present Pansu’s theorem on the convergence of finitely generated virtually nilpotent groups.

Theorem 2.1 (Pansu [16]).

Let Γ\operatorname{\Gamma} be a virtually nilpotent group generated by a symmetric and finite SΓS\subseteq\operatorname{\Gamma}. Then

(Γ,1ndS,e)GH(G,d,e),\left(\operatorname{\Gamma},\frac{1}{n}d_{S},e\right)\operatorname{\xrightarrow{\text{GH}}}(G_{\infty},d_{\infty},\mathlcal{e}),

where GG_{\infty} is a simply connected real graded Lie group (Carnot group). The metric dd_{\infty} is a right-invariant sub-Riemannian (Carnot-Caratheodory) metric which is homogeneous with respect to a family of homotheties {δt}t>0\{\delta_{t}\}_{t>0}, i.e., d(δt(g),δt(h))=td(g,h)d_{\infty}\big{(}\delta_{t}(\mathlcal{g}),\delta_{t}(\mathlcal{h})\big{)}=t~{}d_{\infty}(\mathlcal{g},\mathlcal{h}) for all t>0t>0 and g,hG\mathlcal{g},\mathlcal{h}\in G_{\infty}.

Note that Theorems 1.1 and 1.2 are generalizations of the theorem above. Therefore, the shape theorems under investigation can be interpreted as the convergence of random metric spaces in large-scale geometry. The next subsection is dedicated to the construction of the asymptotic cone GG_{\infty} and related results.

2.2. Rescaled distance and asymptotic cone

Consider for now Γ\operatorname{\Gamma} as a nilpotent and torsion-free group, unless stated otherwise. We also assume that its abelianization is torsion-free. In this subsection, we use Γ\operatorname{\Gamma} instead of Γ\operatorname{\Gamma}^{\prime} to simplify notation, but we will subsequently extend the results to the more general case.

Let GG denote the real Mal’cev completion of Γ\operatorname{\Gamma}. The group GG can be defined as the smallest simply connected real Lie group such that ΓG\operatorname{\Gamma}\leq G and, for all zΓz\in\operatorname{\Gamma} and nn\in\operatorname{\mathbb{N}}, there exists zG\mathlcal{z}\in G with zn=z\mathlcal{z}^{n}=z. In this case, GG is nilpotent of the same order of Γ\operatorname{\Gamma} and it is uniquely defined. Furthermore, GG is simply connected it is associated with the Lie algebra (𝔤,[,]1)(\operatorname{\mathfrak{g}},[\cdot,\cdot]_{1}) where Γ\operatorname{\Gamma} is cocompact in GG. We write log:G𝔤\log:G\to\operatorname{\mathfrak{g}} for the Lie logarithm map.

Define 𝔤1:=𝔤\operatorname{\mathfrak{g}}^{1}:=\operatorname{\mathfrak{g}} and 𝔤i+1:=[𝔤,𝔤i]1\operatorname{\mathfrak{g}}^{i+1}:=[\operatorname{\mathfrak{g}},\operatorname{\mathfrak{g}}^{i}]_{1}. It follows from the nilpotency of Γ\operatorname{\Gamma} that threre exists ll\in\operatorname{\mathbb{N}} such that Γl={e}\operatorname{\Gamma}_{l}=\{e\}. Thus 𝔤l+1=(0)\operatorname{\mathfrak{g}}^{l+1}=(0). Since [𝔤i,𝔤j]1𝔤i+j[\operatorname{\mathfrak{g}}^{i},\operatorname{\mathfrak{g}}^{j}]_{1}\subseteq\operatorname{\mathfrak{g}}^{i+j} and, in particular, [𝔤i+1,𝔤j]1,[𝔤i,𝔤j+1]1𝔤i+j+1[\operatorname{\mathfrak{g}}^{i+1},\operatorname{\mathfrak{g}}^{j}]_{1},[\operatorname{\mathfrak{g}}^{i},\operatorname{\mathfrak{g}}^{j+1}]_{1}\subseteq\operatorname{\mathfrak{g}}^{i+j+1}, the Lie bracket on 𝔤\operatorname{\mathfrak{g}} determines a bilinear map

(𝔤i/𝔤i+1)(𝔤j/𝔤j+1)𝔤i+j/𝔤i+j+1(\operatorname{\mathfrak{g}}^{i}/\operatorname{\mathfrak{g}}^{i+1})\otimes(\operatorname{\mathfrak{g}}^{j}/\operatorname{\mathfrak{g}}^{j+1})\longrightarrow\operatorname{\mathfrak{g}}^{i+j}/\operatorname{\mathfrak{g}}^{i+j+1}

which in turn defines a Lie bracket [,][\cdot,\cdot]_{\infty} on

𝔤:=i=1l𝔳iwith𝔳i:=𝔤i/𝔤i+1.\operatorname{\mathfrak{g}}_{\infty}:=\bigoplus\limits_{i=1}^{l}\operatorname{\mathfrak{v}}_{i}~{}~{}~{}\mbox{with}~{}~{}\operatorname{\mathfrak{v}}_{i}:=\operatorname{\mathfrak{g}}^{i}/\operatorname{\mathfrak{g}}^{i+1}.

Consider the decomposition 𝔤=V1Vl\operatorname{\mathfrak{g}}=V_{1}\oplus\cdots\oplus V_{l} given by 𝔤i:=ViVl\operatorname{\mathfrak{g}}^{i}:=V_{i}\oplus\cdots\oplus V_{l}. Thus, (𝔤,[,])(\operatorname{\mathfrak{g}}_{\infty},[\cdot,\cdot]_{\infty}) is a graded Lie algebra. Let us define a family of linear maps δt:𝔤𝔤\delta_{t}:\operatorname{\mathfrak{g}}_{\infty}\to\operatorname{\mathfrak{g}}_{\infty} given by

δt(v1+v2++vl)=tv1+t2v2++tlvl\delta_{t}(v_{1}+v_{2}+\dots+v_{l})=tv_{1}+t^{2}v_{2}+\dots+t^{l}v_{l}

for each t>0t>0 and vi𝔳iv_{i}\in\mathfrak{v}_{i} with i{1,,l}i\in\{1,\dots,l\}. It follows from the definition of δt\delta_{t} that δt([u,v])=[δt(u),δt(v)]\delta_{t}([u,v]_{\infty})=[\delta_{t}(u),\delta_{t}(v)]_{\infty} and δtt=δtδt\delta_{tt^{\prime}}=\delta_{t}\circ\delta_{t^{\prime}} for all u,v𝔤u,v\in\operatorname{\mathfrak{g}}_{\infty} and t,t>0t,t^{\prime}>0. Hence, {δt}t>0\{\delta_{t}\}_{t>0} defines a family of automorphisms in the graded Lie group G:=exp[𝔤]G_{\infty}:=\exp_{\infty}\left[\operatorname{\mathfrak{g}}_{\infty}\right]. Here we write exp:𝔤G\exp_{\infty}:\operatorname{\mathfrak{g}}_{\infty}\to G_{\infty} and exp:𝔤G\exp:\operatorname{\mathfrak{g}}\to G to differentiate the distinct exponential maps of 𝔤\operatorname{\mathfrak{g}}_{\infty} and 𝔤\operatorname{\mathfrak{g}}. Similarly, log\log_{\infty} and log\log stand for their correspondent Lie logarithm maps.

Let 𝔤=V1Vl\operatorname{\mathfrak{g}}=V_{1}\oplus\cdots\oplus V_{l} be the decomposition given by 𝔤i=ViVl\operatorname{\mathfrak{g}}^{i}=V_{i}\oplus\cdots\oplus V_{l}. Set L:𝔤𝔤L:\operatorname{\mathfrak{g}}\to\operatorname{\mathfrak{g}}_{\infty} to be an linear map such that L(Vi)=𝔳iL(V_{i})=\operatorname{\mathfrak{v}}_{i}. Consider now σt\sigma_{t} to be the linear automorphism on 𝔤\operatorname{\mathfrak{g}} so that σt(vi)=tivi\sigma_{t}(v_{i})=t^{i}v_{i} for each viViv_{i}\in V_{i} and i{1,,l}i\in\{1,\dots,l\}. Define the Lie brackets [,]t[\cdot,\cdot]_{t} on 𝔤\operatorname{\mathfrak{g}} by

[v,w]t=σ1/t([σt(v),σt(w)]), for all t>0,[v,w]_{t}=\sigma_{1/t}\big{(}[\sigma_{t}(v),\sigma_{t}(w)]\big{)},~{}\mbox{ for all }t>0,

thus (𝔤,[,]t)(\operatorname{\mathfrak{g}},[\cdot,\cdot]_{t}) is isomorphic to (𝔤,[,]1)(\operatorname{\mathfrak{g}},[\cdot,\cdot]_{1}) via σt\sigma_{t}. Furthermore,

[L(v),L(w)]=limt+[v,w]t[L(v),L(w)]_{\infty}=\lim\limits_{t\uparrow+\infty}[v,w]_{t}

since, given vViv\in V_{i} and wVjw\in V_{j}, one has that the main term belongs to Vi+jV_{i+j}, the other terms of superior order belong to Vi+j+1VlV_{i+j+1}\oplus\cdots\oplus V_{l} and it makes them insignificant in the rescaled limit (see [7, 16] for a detailed discussion). Set

1tnxn:=(expLσ1/tnlog)(xn).\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{n}:=(\exp_{\infty}\circ L\circ\sigma_{1/t_{n}}\circ\log)(x_{n}).

The convergence established by Theorem 2.1 determines the metric dd_{\infty} such that

(Γ,1ndS,e)GH(G,d,e).\left(\Gamma,\frac{1}{n}d_{S},e\right)\operatorname{\xrightarrow{\text{GH}}}\left(G_{\infty},d_{\infty},\mathlcal{e}\right).

Hence, limn+1tnxn=g\lim_{n\uparrow+\infty}\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{n}=\mathlcal{g} exactly when 1tnxn\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{n} converges to g\mathlcal{g} in (G,d)(G_{\infty},d_{\infty}). The corresponding metric statement shows that, given sequences {xn}n\{x_{n}\}_{n\in\operatorname{\mathbb{N}}}, {xn}n\{x_{n}^{\prime}\}_{n\in\operatorname{\mathbb{N}}} in Γ\Gamma, and tn+t_{n}\uparrow+\infty as n+n\uparrow+\infty with limn+1tnxn=g\lim_{n\uparrow+\infty}\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{n}=\mathlcal{g} and limn+1tnxn=g\lim_{n\uparrow+\infty}\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{n}^{\prime}=\mathlcal{g}^{\prime},

d(g,g)=limn+1tndS(xn,xn).d_{\infty}(\mathlcal{g},\mathlcal{g}^{\prime})=\lim_{n\uparrow+\infty}\frac{1}{t_{n}}d_{S}(x_{n},x_{n}^{\prime}).

The abelianized Lie algebras are defined by 𝔤ab:=𝔤/[𝔤,𝔤]𝔳1\operatorname{\mathfrak{g}}_{\infty}^{\operatorname{\operatorname{ab}}}:=\operatorname{\mathfrak{g}}_{\infty}/[\operatorname{\mathfrak{g}}_{\infty},\operatorname{\mathfrak{g}}_{\infty}]_{\infty}\cong\mathfrak{v}_{1} and 𝔤ab:=𝔤/[𝔤,𝔤]1\operatorname{\mathfrak{g}}^{\operatorname{\operatorname{ab}}}:=\operatorname{\mathfrak{g}}/[\operatorname{\mathfrak{g}},\operatorname{\mathfrak{g}}]_{1}. In particular, 𝔤ab𝔤ab\operatorname{\mathfrak{g}}^{\operatorname{\operatorname{ab}}}\cong\operatorname{\mathfrak{g}}_{\infty}^{\operatorname{\operatorname{ab}}}. By the Frobenius integrability criterion, the integrable curves in GG_{\infty}, the tangent vectors must belong to 𝔳1\mathfrak{v}_{1} at each point of the curve. An admissible (or curve) in GG_{\infty} is a Lipschitz curve γ:[t0,t1]G\upgamma:[t_{0},t_{1}]\to G_{\infty} such that the tangent vector γ(t)𝔳1\upgamma^{\prime}(t)\in\mathfrak{v}_{1} for all t[t0,t1]t\in[t_{0},t_{1}]. Let ϕ:𝔤ab[0,+)\phi:\operatorname{\mathfrak{g}}_{\infty}^{\operatorname{\operatorname{ab}}}\to[0,+\infty) be a norm in the abelianized algebra. Then the ϕ\ell_{\phi}-length of the admissible γ\upgamma is

ϕ(γ):=t0t1ϕ(γ(t))𝑑t.\ell_{\phi}(\upgamma):=\int_{t_{0}}^{t_{1}}\phi\big{(}\upgamma^{\prime}(t)\big{)}dt.

Set dϕd_{\phi} to be the inner metric of the length space (G,ϕ)(G,\ell_{\phi}) given by

dϕ(g,g):=inf{ϕ(γ):γ is an admissible curve from g to g in G}.d_{\phi}(\mathlcal{g},\mathlcal{g}^{\prime}):=\inf\big{\{}\ell_{\phi}(\upgamma)\colon\upgamma\text{ is an admissible curve from }\mathlcal{g}\text{ to }\mathlcal{g}^{\prime}\text{ in }G_{\infty}\big{\}}. (2.1)

In fact, the construction of dϕd_{\phi} can be employed to define dd_{\infty}. The bi-Lipschitz property is a consequence of the results in Section 2.4. One can also verify that the metric dd_{\infty} is right-invariant and homogeneous with respect to δt\delta_{t}. Let us define the projections

π:𝔤𝔤abandπ:𝔤𝔳1𝔤ab\pi:\operatorname{\mathfrak{g}}\to\operatorname{\mathfrak{g}}^{\operatorname{\operatorname{ab}}}\quad\text{and}\quad\pi_{\infty}:\operatorname{\mathfrak{g}}_{\infty}\to\mathfrak{v}_{1}\cong\operatorname{\mathfrak{g}}_{\infty}^{\operatorname{\operatorname{ab}}}

so that, if v=i=1lvi𝔤v=\sum_{i=1}^{l}v_{i}\in\operatorname{\mathfrak{g}}_{\infty} with vi𝔳𝔦v_{i}\in\mathfrak{v_{i}}, then π(v)=v1\pi_{\infty}(v)=v_{1} and π=L1πL\pi=L^{-1}\circ\pi_{\infty}\circ L. The next lemma compiles several well-known results that will be employed throughout the text. We state the results and their proofs can be found in [8, 10, 16].

Lemma 2.2.

Consider Γ\operatorname{\Gamma} a finitely generated torsion-free nilpotent group, then all of the following hold true:

  • (i)

    Let gG\mathlcal{g}\in G_{\infty}. Then there exists a sequence {xn}nΓ\{x_{n}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\Gamma} such that

    limn+1nxn=g.\lim_{n\uparrow+\infty}\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}x_{n}=\mathlcal{g}.
  • (ii)

    Let {xn}n,{yn}nΓ\{x_{n}\}_{n\in\operatorname{\mathbb{N}}},\{y_{n}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\Gamma}, g,hG\mathlcal{g},\mathlcal{h}\in G_{\infty}, and tn+t_{n}\uparrow+\infty as n+n\uparrow+\infty be such that limn+1tnxn=g\lim_{n\uparrow+\infty}\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{n}=\mathlcal{g} and limn+1tnyn=h\lim_{n\uparrow+\infty}\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}y_{n}=\mathlcal{h}. Then

    limn+1tnxnyn=gh.\lim_{n\uparrow+\infty}\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{n}y_{n}=\mathlcal{gh}.
  • (iii)

    Let xΓx\in\operatorname{\Gamma}, then

    limn+1nxn\displaystyle\lim_{n\uparrow+\infty}\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}x^{n} =(expLπlog)(x)\displaystyle=\left(\exp_{\infty}\circ L\circ\pi\circ\log\right)(x)
    =(expπLlog)(x).\displaystyle=\left(\exp_{\infty}\circ~{}\pi_{\infty}\circ L\circ\log\right)(x).
Remark 1.

The conditions imposed on Γ\operatorname{\Gamma} might appear somewhat restrictive. However, we will subsequently regain many properties by making necessary adjustments for virtually nilpotent Γ\operatorname{\Gamma} through the quotient Γ=N/torN\operatorname{\Gamma}^{\prime}=N/\operatorname{\operatorname{tor}}N (see Section 3.3).

The item (iii) in Lemma 2.2 has direct implications for the application of subadditive ergodic theorems. To address this constraint, we overcome it by approximating the lengths using polygonal curves. We present, without proof, Lemma 3.7 from [8], which will be employed in the approximation.

Lemma 2.3.

Consider Γ\operatorname{\Gamma} nilpotent. Let {yi}i=1mΓ\{y_{i}\}_{i=1}^{m}\subseteq\operatorname{\Gamma} and ε>0\varepsilon>0 be given. Then there exist ξ>0\xi>0 and n \accentset{\rule{2.79996pt}{0.5pt}}{n}\in\operatorname{\mathbb{N}} so that, for all n>n n>\accentset{\rule{2.79996pt}{0.5pt}}{n}, for all nj{0,1,,ξn }n_{j}\in\{0,1,\dots,\lfloor\xi\accentset{\rule{2.79996pt}{0.5pt}}{n}\rfloor\},

1ndS(ymnnmym1nnm1y1nn1,ymnym1ny1n)<ε.\frac{1}{n}d_{S}\left(y_{m}^{n-n_{m}}y_{m-1}^{n-n_{m-1}}\dots y_{1}^{n-n_{1}},~{}y_{m}^{n}y_{m-1}^{n}\dots y_{1}^{n}\right)<\varepsilon.

One standout example that exemplifies several properties presented above is the discrete Heisenberg group. As a prime example of a nilpotent group, it offers valuable insights into the fusion of algebraic structures with geometric phenomena in both geometric group theory and metric geometry.

Example 2.1 (The discrete Heisenberg group).

The discrete Heisenberg group can be visualized as a collection of integer lattice points in a three-dimensional space, with a unique group structure derived from matrix multiplication. The nilpotent nature is the key to understand its intricate geometric properties. Let RR be a commutative ring with identity and set H3(R):={(𝚡,𝚢,𝚣):𝚡,𝚢,𝚣R}H_{3}(R):=\{(\mathtt{x},\mathtt{y},\mathtt{z}):\mathtt{x},\mathtt{y},\mathtt{z}\in R\} to be the set of upper triangular matrices with

(𝚡,𝚢,𝚣):=(𝟷𝚡𝚣𝟶𝟷𝚢𝟶𝟶𝟷).(\mathtt{x},\mathtt{y},\mathtt{z}):=\begin{pmatrix}\mathtt{1}&\mathtt{x}&\mathtt{z}\\ \mathtt{0}&\mathtt{1}&\mathtt{y}\\ \mathtt{0}&\mathtt{0}&\mathtt{1}\end{pmatrix}.
Refer to caption
Figure 1. A section of the Heisenberg discrete Cayley graph 𝒞(H3(),S)\mathcal{C}\big{(}H_{3}(\mathbb{Z}),S\big{)} embedded in 3\operatorname{\mathbb{R}}^{3}.

The Heisenberg group on RR is H3(R)H_{3}(R) with the matrix multiplication. In particular, H3()H_{3}(\mathbb{Z}) is known as discrete Heisenberg group. Let Γ=H3()\operatorname{\Gamma}=H_{3}(\mathbb{Z}), 𝚇=(1,0,0)\mathtt{X}=(1,0,0), 𝚈=(0,1,0)\mathtt{Y}=(0,1,0), 𝚉=(0,0,1)\mathtt{Z}=(0,0,1), and S={𝚇±1,𝚈±1}S=\{\mathtt{X}^{\pm 1},\mathtt{Y}^{\pm 1}\}. Observe that

(𝚡,𝚢,𝚣).(𝚡,𝚢,𝚣)\displaystyle(\mathtt{x},\mathtt{y},\mathtt{z}).(\mathtt{x}^{\prime},\mathtt{y}^{\prime},\mathtt{z}^{\prime}) =(𝚡+𝚡,𝚢+𝚢,𝚣+𝚣+𝚡𝚢),\displaystyle=(\mathtt{x}+\mathtt{x}^{\prime},~{}\mathtt{y}+\mathtt{y}^{\prime},~{}\mathtt{z}+\mathtt{z}^{\prime}+\mathtt{x}\mathtt{y}^{\prime}),
(𝚡,𝚢,𝚣)1\displaystyle(\mathtt{x},\mathtt{y},\mathtt{z})^{-1} =(𝚡,𝚢,𝚡𝚢𝚣), and\displaystyle=(-\mathtt{x},-\mathtt{y},\mathtt{xy}-\mathtt{z}),\text{ and}
[(𝚡,𝚢,𝚣),(𝚡,𝚢,𝚣)]\displaystyle\big{[}(\mathtt{x},\mathtt{y},\mathtt{z}),~{}(\mathtt{x}^{\prime},\mathtt{y}^{\prime},\mathtt{z}^{\prime})\big{]} =(0,0,𝚡𝚢𝚡𝚢).\displaystyle=(0,0,\mathtt{xy}^{\prime}-\mathtt{x}^{\prime}\mathtt{y}).

Therefore, for all m,nm,n\in\mathbb{Z},

𝚇m=(m,0,0),𝚈n=(0,n,0),and[𝚇m,𝚈n]=𝚉mn=(0,0,mn).\mathtt{X}^{m}=(m,0,0),\quad\mathtt{Y}^{n}=(0,n,0),~{}~{}\text{and}\quad[\mathtt{X}^{m},\mathtt{Y}^{n}]=\mathtt{Z}^{m\cdot n}=(0,0,m\cdot n). (2.2)

One can easily see that SS is a finite generating set of Γ\operatorname{\Gamma}. Furthermore, Γ1=[Γ,Γ]=𝚉\operatorname{\Gamma}_{1}=[\operatorname{\Gamma},\operatorname{\Gamma}]=\langle\mathtt{Z}\rangle and Γ2=[Γ,Γ1]={e}\operatorname{\Gamma}_{2}=[\operatorname{\Gamma},\operatorname{\Gamma}_{1}]=\{e\}. Hence, Γ\operatorname{\Gamma} is nilpotent of class 22 and SΓ[Γ,Γ]S\subseteq\operatorname{\Gamma}\setminus[\operatorname{\Gamma},\operatorname{\Gamma}]. Consider S\|-\|_{S} the word norm of 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S). It follows from (2.2) that

𝚉mS𝒪(m)as m+.\|\mathtt{Z}^{m}\|_{S}\in\mathcal{O}(\sqrt{m})\quad\text{as }m\uparrow+\infty.

It highlights how the rescaled norm 1nxnS\frac{1}{n}\|x^{n}\|_{S} vanishes as n+n\uparrow+\infty when x[Γ,Γ]x\in[\operatorname{\Gamma},\operatorname{\Gamma}].

Due to the properties above, one can write (𝚡,𝚢)=(𝚡,𝚢,𝚣)[Γ,Γ](\mathtt{x},\mathtt{y})=(\mathtt{x},\mathtt{y},\mathtt{z})[\operatorname{\Gamma},\operatorname{\Gamma}]. Note that Sab={(±1,0),(0,±1)}S^{\operatorname{\operatorname{ab}}}=\left\{(\pm 1,0),(0,\pm 1)\right\} is a finite generating set of the abelianized group Γab=Γ/[Γ,Γ]\operatorname{\Gamma}^{\operatorname{\operatorname{ab}}}=\operatorname{\Gamma}/[\operatorname{\Gamma},\operatorname{\Gamma}] which yields an isomorphism of 𝒞(Γab,Sab)\mathcal{C}(\operatorname{\Gamma}^{\operatorname{\operatorname{ab}}},S^{\operatorname{\operatorname{ab}}}) and the square 2\mathbb{Z}^{2} lattice.

By construction of the asymptotic cone, the Mal’cev completion GG of ΓΓ\operatorname{\Gamma}\simeq\operatorname{\Gamma}^{\prime} is the continuous Heisenberg group H3()H_{3}(\operatorname{\mathbb{R}}) with its associated Lie algebra 𝔥=𝔤\mathfrak{h}=\operatorname{\mathfrak{g}}, in this case, 𝔤𝔤\operatorname{\mathfrak{g}}\simeq\operatorname{\mathfrak{g}}_{\infty} and GGG\simeq G_{\infty}. The Heisenberg algebra 𝔥\mathfrak{h} is given by 𝔥=span{e12,e13,e23}\mathfrak{h}=\operatorname{span}_{\operatorname{\mathbb{R}}}\{e_{12},e_{13},e_{23}\} with {eij:i,j{1,2,3}}\big{\{}e_{ij}:i,j\in\{1,2,3\}\big{\}} the canonical basis of M3×3()M_{3\times 3}(\operatorname{\mathbb{R}}).

Since for all 𝙰,𝙱𝔥\mathtt{A},\mathtt{B}\in\mathfrak{h} one has [𝙰,𝙱]=𝙰𝙱𝙱𝙰span{e13}[\mathtt{A},\mathtt{B}]_{\infty}=\mathtt{A}\mathtt{B}-\mathtt{B}\mathtt{A}\in\operatorname{span}_{\operatorname{\mathbb{R}}}\{e_{13}\} by matrix multiplication, it then follows that 𝔥=𝔳1𝔳2\mathfrak{h}=\mathfrak{v}_{1}\oplus\mathfrak{v}_{2} with 𝔳1span{e12,e23}\mathfrak{v}_{1}\simeq\operatorname{span}_{\operatorname{\mathbb{R}}}\{e_{12},e_{23}\} and 𝔥ab𝔤ab𝔳1\mathfrak{h}^{\operatorname{\operatorname{ab}}}\simeq\operatorname{\mathfrak{g}}^{\operatorname{\operatorname{ab}}}_{\infty}\simeq\mathfrak{v}_{1}.

Let 𝙰=𝚞e12+𝚟e23+𝚠e13\mathtt{A}=\mathtt{u}\cdot e_{12}+\mathtt{v}\cdot e_{23}+\mathtt{w}\cdot e_{13}, then exp(𝙰)=(𝚞,𝚟,𝚠+12𝚞𝚟)\exp_{\infty}(\mathtt{A})=\left(\mathtt{u},\mathtt{v},\mathtt{w}+\frac{1}{2}\mathtt{uv}\right). Since (𝚡,𝚢,𝚣)n=(n𝚡,n𝚢,n𝚣+n(n1)2𝚡𝚢)(\mathtt{x},\mathtt{y},\mathtt{z})^{n}=\left(n\mathtt{x},n\mathtt{y},n\mathtt{z}+\frac{n(n-1)}{2}\mathtt{xy}\right) one can verify by the procedure defined in this section that 1n(𝚡,𝚢,𝚣)n=(𝚡,𝚢,1n𝚣1n𝚡𝚢+12𝚡𝚢)G\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}(\mathtt{x},\mathtt{y},\mathtt{z})^{n}=\left(\mathtt{x},\mathtt{y},\frac{1}{n}\mathtt{z}-\frac{1}{n}\mathtt{xy}+\frac{1}{2}\mathtt{xy}\right)\in G_{\infty}. It implies that, for all 𝚡,𝚢,𝚣\mathtt{x},\mathtt{y},\mathtt{z}\in\mathbb{Z},

limn+1n(𝚡,𝚢,𝚣)n=(𝚡,𝚢,12𝚡𝚢)=exp(π(log(𝚡,𝚢,𝚣))).\lim_{n\uparrow+\infty}\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}(\mathtt{x},\mathtt{y},\mathtt{z})^{n}=\left(\mathtt{x},\mathtt{y},\frac{1}{2}\mathtt{xy}\right)=\exp_{\infty}\left(\pi_{\infty}\Big{(}\log(\mathtt{x},\mathtt{y},\mathtt{z})\Big{)}\right).

2.3. Some examples of virtually nilpotent groups

In this subsection, our focus shifts to examples of virtually nilpotent groups that can be constructed through direct and outer semidirect products. The discussion of the virtually nilpotent case will be explored more extensively later in the text.

Let L\mathrm{L} be a nilpotent group and consider MM a finite group. Then the direct product

K=L×MK=\mathrm{L}\times M

is a group with the binary operation given by (x,m).(y,m)=(xy,mm)(x,m).(y,m^{\prime})=(xy,mm^{\prime}). Note that the commutator is [(x,m),(y,m)]=([x,y],[m,m])\big{[}(x,m),(y,m^{\prime})\big{]}=\big{(}[x,y],[m,m^{\prime}]\big{)}. It follows that, for all A,ALA,A^{\prime}\subseteq\mathrm{L} and B,BMB,B^{\prime}\subseteq M,

[A×B,A×B]=[A,B]×[A,B].\big{[}A\times B,A^{\prime}\times B^{\prime}\big{]}=[A,B]\times[A^{\prime},B^{\prime}].

Hence, KK is a nilpotent group if, and only if, MM is nilpotent. On the other hand, for all finite group MM, KK is virtually nilpotent.

Set SLS_{\mathrm{L}} and SMS_{M} to be finite symmetric generating sets of L\mathrm{L} and MM, respectively.

(SL×{e})({e}×SM)\big{(}S_{\mathrm{L}}\times\{e\}\big{)}\cup\big{(}\{e\}\times S_{M})

is a finite generating set of KK. We will consider another useful example of generating set of KK. Let SeS_{\square}^{e} stand for S{e}S_{\square}\cup\{e\}. Then

S=SL×SMeS=S_{\mathrm{L}}\times S_{M}^{e}

is also a symmetric generating set of KK. Furthermore, if L\mathrm{L} is torsion-free, then S\llbracket S\rrbracket is analog to SLS_{\mathrm{L}} while ΓL\operatorname{\Gamma}^{\prime}\simeq\mathrm{L}.

Example 2.2.

Let SL(2,3)\mathrm{SL}(2,3) be the of degree two over a field of three elements determined by

SL(2,3)=ρ1,ρ2,ρ3:ρ13=ρ23=ρ33=ρ1ρ2ρ3\mathrm{SL}(2,3)=\left\langle\rho_{1},\rho_{2},\rho_{3}\colon\rho_{1}^{3}=\rho_{2}^{3}=\rho_{3}^{3}=\rho_{1}\rho_{2}\rho_{3}\right\rangle

A remarkable property of SL(2,3)\mathrm{SL}(2,3) is that it is the smallest group that is not nilpotent. Let m=ρ0\operatorname{\mathbb{Z}}_{m}=\langle\rho_{0}\rangle the cyclic group with ρ0m=e\rho_{0}^{m}=e and consider H3()H_{3}(\operatorname{\mathbb{Z}}) to be the discrete Heisenberg group, as defined in Example 2.1. Set

Γ=(H3()×m)×SL(2,3).\operatorname{\Gamma}=\big{(}H_{3}(\operatorname{\mathbb{Z}})\times\operatorname{\mathbb{Z}}_{m}\big{)}\times\mathrm{SL}(2,3).

Then Γ\operatorname{\Gamma} is virtually nilpotent with N=H3()×m×{e}ΓN=H_{3}(\operatorname{\mathbb{Z}})\times\operatorname{\mathbb{Z}}_{m}\times\{e\}\unlhd\operatorname{\Gamma} such that κ=[Γ:N]=|SL(2,3)|=24\kappa=[\operatorname{\Gamma}:N]=|\mathrm{SL}(2,3)|=24. Hence, considering this notation:

NH3()×m,torN={e}×3×{e}3Γ=N/torNH3().N\simeq H_{3}(\operatorname{\mathbb{Z}})\times\operatorname{\mathbb{Z}}_{m},\quad\operatorname{\operatorname{tor}}N=\{e\}\times\operatorname{\mathbb{Z}}_{3}\times\{e\}\simeq\operatorname{\mathbb{Z}}_{3}\quad\operatorname{\Gamma}^{\prime}=N/\operatorname{\operatorname{tor}}N\simeq H_{3}(\operatorname{\mathbb{Z}}).

Let us write SL(2,3)={zj}j=124\mathrm{SL}(2,3)=\{z_{j}\}_{j=1}^{24} and fix z(j)=(e,e,zj)z_{(j)}=(e,e,z_{j}) as representatives for each coset in Γ/N\operatorname{\Gamma}/N. Thus,

πN(x,y,z)=(x,y,e), and(x,y,z)={x}×3×{e}xH3().\uppi_{N}(x,y,z)=(x,y,e),\text{ and}\quad\big{\llbracket}(x,y,z)\big{\rrbracket}=\{x\}\times\operatorname{\mathbb{Z}}_{3}\times\{e\}\cong x\in H_{3}(\operatorname{\mathbb{Z}}).

Now, set

SH3()={𝚇±1,𝚈±1},Sm={ρ0±1}, andSSL(2,3)={ρ1±1,ρ2±1,ρ3±1}.S_{H_{3}(\operatorname{\mathbb{Z}})}=\big{\{}\mathtt{X}^{\pm 1},\mathtt{Y}^{\pm 1}\big{\}},\quad S_{\operatorname{\mathbb{Z}}_{m}}=\left\{\rho_{0}^{\pm 1}\right\},\text{ and}\quad S_{\mathrm{SL}(2,3)}=\left\{\rho_{1}^{\pm 1},\rho_{2}^{\pm 1},\rho_{3}^{\pm 1}\right\}.

Then

S=SH3()×Sme×SSL(2,3)eS=S_{H_{3}(\operatorname{\mathbb{Z}})}\times S_{\operatorname{\mathbb{Z}}_{m}}^{e}\times S_{\mathrm{SL}(2,3)}^{e}

is a finite symmetric generating set of Γ\operatorname{\Gamma}. Moreover, the Cayley graph 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) is homomorphic equivalent to 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma}^{\prime},\llbracket S\rrbracket), which is isomorphic to 𝒞(H3(),SH3())\mathcal{C}\left(H_{3}(\operatorname{\mathbb{Z}}),S_{H_{3}(\operatorname{\mathbb{Z}})}\right).

More generally, one can also obtain a virtually nilpotent group by the outer semidirect product. Consider NN a nilpotent and HH a finite group. Let φ\varphi be a group homomorphism φ:HAut(N)\varphi:H\to\operatorname{Aut}(N), where Aut(N)\operatorname{Aut}(N) is the automorphism group of NN. Then the semidirect group is

Γ=NφH\operatorname{\Gamma}=N\rtimes_{\varphi}H

whose elements are the same of N×HN\times H but the binary operation is characterized by

(x,h).(y,h)\displaystyle(x,h).(y,h^{\prime}) =(xφh(y),hh),\displaystyle=\big{(}x\varphi_{h}(y),hh^{\prime}\big{)},
(x,h)1\displaystyle(x,h)^{-1} =(φh1(x1),h1), and\displaystyle=\big{(}\varphi_{h^{-1}}(x^{-1}),h^{-1}\big{)},\text{ and}
[(x,h),(y,h)]\displaystyle\Big{[}(x,h),(y,h^{\prime})\Big{]} =(xφh(y)φhhh1(x1)φ[h,h](y1),[h,h]).\displaystyle=\Big{(}x\varphi_{h}(y)\varphi_{hh^{\prime}h^{-1}}(x^{-1})\varphi_{[h,h^{\prime}]}(y^{-1}),~{}[h,h^{\prime}]\Big{)}.

Let SNS_{N} and SHS_{H} be finite symmetric generating sets of NN and HH, respectively. Hence, similarly to the direct product,

(SN×{e})({e}×SH)\big{(}S_{N}\times\{e\}\big{)}\cup\big{(}\{e\}\times S_{H}\big{)}

is a finite symmetric generating set of Γ\operatorname{\Gamma}. Moreover, SN×SHeS_{N}\times S_{H}^{e} is also a finite generating set, but not necessarily symmetric. However,

(hHφh(SN))×H\left(\bigcup_{h\in H}\varphi_{h}(S_{N})\right)\times H

is finite, symmetric, and generates Γ\operatorname{\Gamma}. The next example illustrates how some properties of the outer semidirect product groups change in comparison to the direct product.

Example 2.3 (Generalized dihedral group).

Let (N,+)(N,+) be a finitely generated abelian group with polynomial growth rate D1D\geq 1 and (2,+)(\operatorname{\mathbb{Z}}_{2},+) with 2={0,1}\operatorname{\mathbb{Z}}_{2}=\{0,1\}. Fix φ:2Aut(N)\varphi:\operatorname{\mathbb{Z}}_{2}\to\operatorname{Aut}(N) such that φ0=id\varphi_{0}=id and φ1=id\varphi_{1}=-id. The generalized virtually nilpotent diheral group is

Dih(N):=Nφ2.\operatorname{Dih}(N):=N\rtimes_{\varphi}\operatorname{\mathbb{Z}}_{2}.

Consider Γ=Dih(N)\operatorname{\Gamma}=\operatorname{Dih}(N), then for all (x,r),(y,r)Γ(x,r),(y,r^{\prime})\in\operatorname{\Gamma},

(x,r).(y,r)\displaystyle(x,r).(y,r^{\prime}) =(x+φr(y),r+r),\displaystyle=\big{(}x+\varphi_{r}(y),r+r^{\prime}\big{)},
(x,r)1\displaystyle(x,r)^{-1} =((1)r+1x,r),\displaystyle=\big{(}(-1)^{r+1}x,r\big{)},
[(x,r),(y,r)]\displaystyle\Big{[}(x,r),(y,r^{\prime})\Big{]} =((1(1)r)x(1(1)r)y, 0).\displaystyle=\Big{(}\big{(}1-(-1)^{r^{\prime}}\big{)}x-\big{(}1-(-1)^{r}\big{)}y,\ 0\Big{)}.

Therefore, Γ\operatorname{\Gamma} is non-abelian and Γ1=[Γ,Γ]=2N×{0}\operatorname{\Gamma}_{1}=[\operatorname{\Gamma},\operatorname{\Gamma}]=2N\times\{0\}. One can easily verify that all elements of Γ2=[Γ,Γ1]\operatorname{\Gamma}_{2}=[\operatorname{\Gamma},\operatorname{\Gamma}_{1}] are

[(x,r),(2y,0)]=(2((1)r1)y, 0).\Big{[}(x,r),(2y,0)\Big{]}=\Big{(}2\big{(}(-1)^{r}-1\big{)}y,\ 0\Big{)}.

Hence, for all nn\in\operatorname{\mathbb{N}}, one has Γn2nN\operatorname{\Gamma}_{n}\simeq 2^{n}N. We can conclude that Γ\operatorname{\Gamma} is not nilpotent while it is virtually nilpotent since NΓN\unlhd\operatorname{\Gamma}.

2.4. Establishing a Candidate for the Limiting Shape

From this point until Section 3.2, let us once again regard Γ\operatorname{\Gamma} as a torsion-free nilpotent group. Our objective is to characterize the limiting shape using a norm that defines a metric in GG_{\infty}. In this section, we achieve the desired norm through the application of a subadditive ergodic theorem. The convergence is not directly established as uniform convergence in D\operatorname{\mathbb{R}}^{D} because of the constraints imposed by admissible curves.

Set c (x):=𝔼[c(x)]\accentset{\rule{2.79996pt}{0.5pt}}{c}(x):=\operatorname{\mathbbm{E}}[c(x)], due to the subadditivity of the cocycle

c (xy)c (y)+c (x),\accentset{\rule{2.79996pt}{0.5pt}}{c}(xy)\leq\accentset{\rule{2.79996pt}{0.5pt}}{c}(y)+\accentset{\rule{2.79996pt}{0.5pt}}{c}(x),

for all x,yΓx,y\in\operatorname{\Gamma}. Thus c (x)bxS\accentset{\rule{2.79996pt}{0.5pt}}{c}(x)\leq b\|x\|_{S} with b=maxsS{c (s)}b=\max_{s\in S}\big{\{}\accentset{\rule{2.79996pt}{0.5pt}}{c}(s)\big{\}}. It follows from (ii) that there exists a subsequence of c(xn)/nc(x^{n})/n such that c(xnj)/njaxSabc(x^{n_{j}})/n_{j}\geq a\|x\|_{S}^{\operatorname{\operatorname{ab}}} \operatorname{\mathbbm{P}}-a.s. for sufficiently large jj.

Recall that Γab=Γ/[Γ,Γ]\operatorname{\Gamma}^{\operatorname{\operatorname{ab}}}=\operatorname{\Gamma}/[\operatorname{\Gamma},\operatorname{\Gamma}] and consider xab=x[Γ,Γ]x^{\operatorname{\operatorname{ab}}}=x[\operatorname{\Gamma},\operatorname{\Gamma}], To simplify notation, we also use xabx^{\operatorname{\operatorname{ab}}} interchangeably with (πLlog)(x)(\pi_{\infty}\circ L\circ\log)(x) when it is clear from the context. Let

xSab:=infyx[Γ,Γ]yS.\|x\|_{S}^{\operatorname{\operatorname{ab}}}:=\inf_{y\in x[\operatorname{\Gamma},\operatorname{\Gamma}]}\|y\|_{S}.

Since Sab\|-\|_{S}^{\operatorname{\operatorname{ab}}} is discrete, there exists yx[Γ,Γ]y\in x[\operatorname{\Gamma},\operatorname{\Gamma}] such that xSab=yS\|x\|_{S}^{\operatorname{\operatorname{ab}}}=\|y\|_{S}. Hence, for all x,yΓx,y\in\operatorname{\Gamma}, there exist x,x′′x[Γ,Γ]x^{\prime},x^{\prime\prime}\in x[\operatorname{\Gamma},\operatorname{\Gamma}] and y,y′′y[Γ,Γ]y^{\prime},y^{\prime\prime}\in y[\operatorname{\Gamma},\operatorname{\Gamma}] such that

xySab=xyS,xSab=x′′S, andySab=y′′S;\|xy\|_{S}^{\operatorname{\operatorname{ab}}}=\|x^{\prime}y^{\prime}\|_{S},\quad\|x\|_{S}^{\operatorname{\operatorname{ab}}}=\|x^{\prime\prime}\|_{S},\text{ and}\quad\|y\|_{S}^{\operatorname{\operatorname{ab}}}=\|y^{\prime\prime}\|_{S};

which implies the subadditivity

xySab=xySx′′y′′Sx′′S+y′′S=xSab+ySab.\|xy\|_{S}^{\operatorname{\operatorname{ab}}}=\|x^{\prime}y^{\prime}\|_{S}\leq\|x^{\prime\prime}y^{\prime\prime}\|_{S}\leq\|x^{\prime\prime}\|_{S}+\|y^{\prime\prime}\|_{S}=\|x\|_{S}^{\operatorname{\operatorname{ab}}}+\|y\|_{S}^{\operatorname{\operatorname{ab}}}.

Now, regarding xSab=0\|x\|_{S}^{\operatorname{\operatorname{ab}}}=0 whenever x[Γ,Γ]x\in[\operatorname{\Gamma},\operatorname{\Gamma}], one has for all x[Γ,Γ]x\in[\operatorname{\Gamma},\operatorname{\Gamma}] and yΓy\in\operatorname{\Gamma}, xySab=ySab\|xy\|_{S}^{\operatorname{\operatorname{ab}}}=\|y\|_{S}^{\operatorname{\operatorname{ab}}}. Let y=smsj+1sjsj1s1y=s_{m}\dots s_{j+1}s_{j}s_{j-1}\dots s_{1} with sj[Γ,Γ]s_{j}\in[\operatorname{\Gamma},\operatorname{\Gamma}]. Since [Γ,Γ][\operatorname{\Gamma},\operatorname{\Gamma}] is a normal subgroup of Γ\operatorname{\Gamma}, s¯j=(sj1s1)1sj(sj1s1)[Γ,Γ]\bar{s}_{j}=(s_{j-1}\dots s_{1})^{-1}s_{j}(s_{j-1}\dots s_{1})\in[\operatorname{\Gamma},\operatorname{\Gamma}] is such that y=smsj+1sj1s1s¯jy=s_{m}\dots s_{j+1}s_{j-1}\dots s_{1}\bar{s}_{j}. Hence

ySab=smsj+1sj1s1Sab.\|y\|_{S}^{\operatorname{\operatorname{ab}}}=\|s_{m}\dots s_{j+1}s_{j-1}\dots s_{1}\|_{S}^{\operatorname{\operatorname{ab}}}.

Therefore, ySab=yS=m\|y\|_{S}^{\operatorname{\operatorname{ab}}}=\|y\|_{S}=m if, and only if, there exists {si}i=1mS[Γ,Γ]\{s_{i}\}_{i=1}^{m}\subseteq S\setminus[\operatorname{\Gamma},\operatorname{\Gamma}] such that y=sms1y=s_{m}\dots s_{1}. Observe that Γab\operatorname{\Gamma}^{\operatorname{\operatorname{ab}}} is a topological lattice of GabG^{\operatorname{\operatorname{ab}}} and GabΓabdim𝔳1𝔤abG^{\operatorname{\operatorname{ab}}}\simeq\operatorname{\Gamma}^{\operatorname{\operatorname{ab}}}\otimes\operatorname{\mathbb{R}}\simeq\operatorname{\mathbb{R}}^{\operatorname{dim}\mathfrak{v}_{1}}\simeq\operatorname{\mathfrak{g}}^{\operatorname{\operatorname{ab}}}. Let \|-\| be an Euclidean norm on GabG^{\operatorname{\operatorname{ab}}} and fix a ,b >0\accentset{\rule{2.79996pt}{0.5pt}}{a},\accentset{\rule{2.79996pt}{0.5pt}}{b}>0 such that

a :=min{s[Γ,Γ]:sS[Γ,Γ]}, andb :=max{s[Γ,Γ]:sS[Γ,Γ]},\accentset{\rule{2.79996pt}{0.5pt}}{a}:=\min\big{\{}\|s[\operatorname{\Gamma},\operatorname{\Gamma}]\|\colon s\in S\setminus[\operatorname{\Gamma},\operatorname{\Gamma}]\big{\}},\text{ and}\quad\accentset{\rule{2.79996pt}{0.5pt}}{b}:=\max\big{\{}\|s[\operatorname{\Gamma},\operatorname{\Gamma}]\|\colon s\in S\setminus[\operatorname{\Gamma},\operatorname{\Gamma}]\big{\}},

Due to the properties of a normed vector space, one has, for all xΓx\in\operatorname{\Gamma},

a xSabx[Γ,Γ]b xSab.\accentset{\rule{2.79996pt}{0.5pt}}{a}\|x\|_{S}^{\operatorname{\operatorname{ab}}}\leq\big{\|}x[\operatorname{\Gamma},\operatorname{\Gamma}]\big{\|}\leq\accentset{\rule{2.79996pt}{0.5pt}}{b}\|x\|_{S}^{\operatorname{\operatorname{ab}}}. (2.3)

Set f:Γab0f:\operatorname{\Gamma}^{\operatorname{\operatorname{ab}}}\to\operatorname{\mathbb{R}}_{\geq 0} to be given by

f(xab)=inf{c (y):yx[Γ,Γ]}.f(x^{\operatorname{\operatorname{ab}}})=\inf\{\accentset{\rule{2.79996pt}{0.5pt}}{c}(y):y\in x[\Gamma,\Gamma]\}.

It follows immediately from the subadditivity of c \accentset{\rule{2.79996pt}{0.5pt}}{c} and the definition of ff that

f(xabyab)f(xab)+f(yab).f(x^{\operatorname{\operatorname{ab}}}y^{\operatorname{\operatorname{ab}}})\leq f(x^{\operatorname{\operatorname{ab}}})+f(y^{\operatorname{\operatorname{ab}}}).

We are now able to state the following a subadditive ergodic theorem obtained by Austin [3] and improved by Cantrell and Furman [8].

Proposition 2.4 (Subadditive Ergodic Theorem).

Let the subadditive cocycle c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} associated with a p.m.p. ergodic group action ϑ:ΓΩ\operatorname{\upvartheta}:\operatorname{\Gamma}\curvearrowright\operatorname{\Upomega} be such that c(x)L1(Ω,,)c(x)\in L^{1}(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}) for all xΓx\in\operatorname{\Gamma}. Then there exists a unique homogeneous subadditive function ϕ:𝔤ab0\phi:\mathfrak{g}_{\infty}^{\operatorname{\operatorname{ab}}}\to\operatorname{\mathbb{R}}_{\geq 0} such that, for every xΓx\in\operatorname{\Gamma},

limn+1nc(xn)=ϕ(xab)-a.s. and in L1.\lim_{n\uparrow+\infty}\frac{1}{n}c(x^{n})=\phi(x^{\operatorname{\operatorname{ab}}})\quad\operatorname{\mathbbm{P}}\text{-a.s. and in }L^{1}.

Moreover, ϕ\phi is given by

ϕ(xab)=limn+1nf(nxab)=infn11nf(nxab).\phi(x^{\operatorname{\operatorname{ab}}})=\lim_{n\uparrow+\infty}\frac{1}{n}f\left(n\cdot x^{\operatorname{\operatorname{ab}}}\right)=\inf_{n\geq 1}\frac{1}{n}f\left(n\cdot x^{\operatorname{\operatorname{ab}}}\right). (2.4)
Remark 2.

The function ϕ\phi obtained above is naturally associated with the abelianized space considering the well-known fact of the convergence of 1nxn\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}x^{n} to the projection of xx onto a subspace isomorphic to 𝔤ab\operatorname{\mathfrak{g}}_{\infty}^{\operatorname{\operatorname{ab}}}. It will allow us to measure distances in GG_{\infty} with ϕ\ell_{\phi} by considering the rescaling of the subadditive cocyle.

The bi-Lipschitz property established in the following lemma is crucial for the main results.

Lemma 2.5.

Let c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} be a subadditive cocycle under the assumptions of Proposition 2.4. Set ϕ\phi as in (2.4). Consider cc satisfying (i) and (ii). Then there exist a,b>0a^{\prime},b^{\prime}>0 such that, for all xΓx\in\operatorname{\Gamma},

axabϕ(xab)bxab.a^{\prime}\|x^{\operatorname{\operatorname{ab}}}\|\leq\phi(x^{\operatorname{\operatorname{ab}}})\leq b^{\prime}\|x^{\operatorname{\operatorname{ab}}}\|.
Proof.

Observe that condition (i) implies cL1(Ω,,)c\in L^{1}(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}). Consider a ,b >0\accentset{\rule{2.79996pt}{0.5pt}}{a},\accentset{\rule{2.79996pt}{0.5pt}}{b}>0 as in (2.3) and fix a:=a/b a^{\prime}:=a/\accentset{\rule{2.79996pt}{0.5pt}}{b}. By (ii), one has

f(njxab)=infyx[Γ,Γ]c (y)axnjSabanjxab.f(n_{j}\cdot x^{\operatorname{\operatorname{ab}}})=\inf_{y\in x[\operatorname{\Gamma},\operatorname{\Gamma}]}\accentset{\rule{2.79996pt}{0.5pt}}{c}(y)\geq a\|x^{n_{j}}\|_{S}^{\operatorname{\operatorname{ab}}}\geq a^{\prime}n_{j}\|x^{\operatorname{\operatorname{ab}}}\|.

We know by Proposition 2.4 that ϕ\phi exists and

ϕ(xab)=infn1nc (xab)=limj+1njf(njxab)axab.\phi(x^{\operatorname{\operatorname{ab}}})=\inf_{n\in\operatorname{\mathbb{N}}}\frac{1}{n}\accentset{\rule{2.79996pt}{0.5pt}}{c}(x^{\operatorname{\operatorname{ab}}})=\lim_{j\uparrow+\infty}\frac{1}{n_{j}}f(n_{j}\cdot x^{\operatorname{\operatorname{ab}}})\geq a^{\prime}\|x^{\operatorname{\operatorname{ab}}}\|.

It follows from (i) and subaditivity that there exists b>0b>0 such that, for all xΓx\in\operatorname{\Gamma},

c (x)bxS.\accentset{\rule{2.79996pt}{0.5pt}}{c}(x)\leq b\|x\|_{S}.

Let us fix b:=b/a b^{\prime}:=b/\accentset{\rule{2.79996pt}{0.5pt}}{a}, then

ϕ(xab)=infn1ninfyxn[Γ,Γ]c (y)binfn1nxnSabbxab,\phi(x^{\operatorname{\operatorname{ab}}})=\inf_{n\in\operatorname{\mathbb{N}}}\frac{1}{n}\inf_{y\in x^{n}[\operatorname{\Gamma},\operatorname{\Gamma}]}\accentset{\rule{2.79996pt}{0.5pt}}{c}(y)\leq b\inf_{n\in\operatorname{\mathbb{N}}}\frac{1}{n}\|x^{n}\|_{S}^{\operatorname{\operatorname{ab}}}\leq b^{\prime}\|x^{\operatorname{\operatorname{ab}}}\|,

which is our assertion. ∎

Remark 3.

The Subadditive Ergodic Theorem guarantees the \operatorname{\mathbbm{P}}-a.s. existence of the limn+c(xn)/n\lim_{n\uparrow+\infty}c(x^{n})/n. By combining this fact with previous assertions and the L1L^{1} convergence, we obtain the existence of 0<ab<+0<a\leq b<+\infty such that

axSabc (x)bxS.a\|x\|_{S}^{\operatorname{\operatorname{ab}}}\leq\accentset{\rule{2.79996pt}{0.5pt}}{c}(x)\leq b\|x\|_{S}. (2.5)

Furthermore, one has from (2.5) that axSabϕ(xab)bxSa\|x\|^{\operatorname{\operatorname{ab}}}_{S}\leq\phi(x^{\operatorname{\operatorname{ab}}})\leq b\|x\|_{S} for all xΓx\in\operatorname{\Gamma}. Since there exists yx[Γ,Γ]y\in x[\operatorname{\Gamma},\operatorname{\Gamma}] with xSab=yS\|x\|_{S}^{\operatorname{\operatorname{ab}}}=\|y\|_{S} and xab=yabx^{\operatorname{\operatorname{ab}}}=y^{\operatorname{\operatorname{ab}}}, one has by (2.3)

ab xabaxSabϕ(xab)bxSabba xab.\frac{~{}a~{}}{\accentset{\rule{2.79996pt}{0.5pt}}{b}}\|x^{\operatorname{\operatorname{ab}}}\|\leq a\|x\|^{\operatorname{\operatorname{ab}}}_{S}\leq\phi(x^{\operatorname{\operatorname{ab}}})\leq b\|x\|_{S}^{\operatorname{\operatorname{ab}}}\leq\frac{~{}b~{}}{\accentset{\rule{2.79996pt}{0.5pt}}{a}}\|x^{\operatorname{\operatorname{ab}}}\|.

Recall the definition of dϕd_{\phi} in (2.1). Therefore, there is a bi-Lipschitz relation between dd_{\infty} and dϕd_{\phi}. We now define Φ:G[0,+)\Phi:G_{\infty}\to[0,+\infty) by

Φ(g):=dϕ(e,g).\Phi(\mathlcal{g}):=d_{\phi}(\mathlcal{e},\mathlcal{g}).

The next subsection deals with one of the most relevant models to be considered in our context.

2.5. First-Passage Percolation models

Hammersley and Welsh introduced the First-Passage Percolation (FPP) as a mathematical model in 1965 to study the spread of fluid through a porous medium. In FPP models, a graph with random edge weights is considered, where these weights represent the time taken for the fluid to pass through the corresponding edge. These concepts will be revisited in Section 3.4 and illustrated with examples in Section 4.

Let 𝒢=(V,E)\mathcal{G}=(V,E) be a graph and set τ={τ(𝚎)}𝚎E\tau=\{\tau(\mathtt{e})\}_{\mathtt{e}\in E} to be a collection of non-negative random variables. We may regard each τ(u,v)\tau(u,v) as random length (also passage time or weight) of an edge {u,v}E\{u,v\}\in E. It turns (𝒢,τ)(\mathcal{G},\tau) into a random length space and it motivates the following construction.

The random passage time of a path γ𝒫(x,y)\gamma\in\mathscr{P}(x,y) is given by T(γ)=𝚎γτ(𝚎)T(\gamma)=\sum_{\mathtt{e}\in\gamma}\tau(\mathtt{e}). Let us now define the first-passage time of yy with the process starting at xx by

T(x,y):=infγ𝒫(x,y)T(γ).T(x,y):=\inf_{\gamma\in\mathscr{P}(x,y)}T(\gamma).

The random variable T(x,y)T(x,y) is also known as first-hitting time. Observe that T(x,y)T(x,y) is a random intrinsic pseudometric, i.e., xyx\neq y does not imply in T(x,y)>0T(x,y)>0. We can now consider the group action ϑ:Γ(Ω,,)\operatorname{\upvartheta}:\operatorname{\Gamma}\curvearrowright(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}) as a translation such that c(x):=T(e,x)c(x):=T(e,x) is a subadditive cocycle (see 1.2) with τ(x,sx)ϑy=τ(xy1,sxy1)\tau(x,sx)\circ\operatorname{\upvartheta}_{y}=\tau(xy^{-1},sxy^{-1}) for all x,yΓx,y\in\operatorname{\Gamma} and sSs\in S when 𝒢=𝒞(Γ,S)\mathcal{G}=\mathcal{C}(\operatorname{\Gamma},S).

By requiring ϑ\operatorname{\upvartheta} to be ergodic, we obtain for the FPP model that, for all xΓx\in\operatorname{\Gamma} and sSs\in S,

c(s)ϑx=τ(x,sx)τ(e,s)=c(s).c(s)\circ\operatorname{\upvartheta}_{x}=\tau(x,sx)\sim\tau(e,s)=c(s).

It also follows that τ(e,s)τ(e,s1)\tau(e,s)\sim\tau(e,s^{-1}). Therefore, each direction of 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) determines a common distribution for its random lengths in a FPP model. Example 4.1 portraits an FPP model with dependend and identically distributed random lengths. While the FPP the random variables of Example 4.2 are independent but not identically distributed.

As passage times are preserved under translation, condition (iii) is immediately satisfied, as it suffices to consider a geodesic path when S=F(ε)S=F(\varepsilon). However, other examples of subadditive interacting particle systems do not exhibit these properties. For instance, the Frog Model (see Example 4.3) can be described by a subadditive cocycle satisfying (i), (ii), and (ii). If we denote τ(x,sx)=|T(x)T(sx)|\tau(x,sx)=|T(x)-T(sx)|, then τ\tau describes the growth of the process, and

τ(x,sx)θxτ(e,s)whileτ(x,sx)≁τ(e,s).\tau(x,sx)\circ\theta_{x}\sim\tau(e,s)\quad\text{while}\quad\tau(x,sx)\not\sim\tau(e,s).

The results and properties highlighted above will be crucial in the study of the asymptotic shape and its applications in the subsequent discussions.

3. The Limiting Shape

This section begins by introducing auxiliary results, offering essential tools to be employed in our subsequent analysis. Subsequently, our focus shifts to the proof of the two main theorems. The concluding subsection is dedicated to exploring a corollary specifically tailored for FPP models.

3.1. Approximation of admissible curves along polygonal paths

The proof strategy for the main theorem involves approximating geodesic curves with polygonal paths. Throughout the following discussion, we assume that cc is a subadditive cocycle, and Γ\operatorname{\Gamma} is finitely generated by the symmetric set SS with polynomial growth rate D1D\geq 1. To set the stage, we begin by stating Proposition 3.1 from [8].

Proposition 3.1.

Let γ:[0,1]G\upgamma:[0,1]\to G_{\infty} be a Lipschitz curve and let ε^(0,1)\hat{\varepsilon}\in(0,1). Then there exists k0=k0(γ,ε^)>0k_{0}=k_{0}(\upgamma,\hat{\varepsilon})>0 so that one can find, for all k>k0k>k_{0} {yj}j=1kΓ\{y_{j}\}_{j=1}^{k}\subseteq\operatorname{\Gamma}, p>0p>0 and n0>0n_{0}>0 such that, for all n>n0n>n_{0},

j=1kd(1npyjnyj1ny1n,γ(jk))<ε^\sum_{j=1}^{k}d_{\infty}\left(\frac{1}{np}\operatorname{{\scriptscriptstyle\bullet}}y_{j}^{n}y_{j-1}^{n}\dots y_{1}^{n},\upgamma\left(\frac{j}{k}\right)\right)<\hat{\varepsilon}

Moreover, for ϕ:𝔤ab0\upphi:\operatorname{\mathfrak{g}}_{\infty}^{\operatorname{\operatorname{ab}}}\to\operatorname{\mathbb{R}}_{\geq 0} a subadditive homogeneous function bi-Lipschitz with respect to \|-\|, one has that

|1p(ϕ(ykab)++ϕ(y1ab))ϕ(γ)|<ε^.\left|\frac{1}{p}\big{(}\upphi(y_{k}^{\text{ab}})+\cdots+\upphi(y_{1}^{\text{ab}})\big{)}-\ell_{\upphi}(\upgamma)\right|<\hat{\varepsilon}.

The approximation technique outlined in the upcoming proposition will be utilized in the subsequent subsections. It extends the guarantees of the subadditive ergodic theorem for the decomposition of polygonal paths under certain properties.

Proposition 3.2.

Let Γ\operatorname{\Gamma} be a torsion-free nilpotent finitely generated group with torsion-free abelianization. Consider c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} a subadditive cocycle associated with an ergodic group action ϑ\operatorname{\upvartheta} satisfying (i). Then for all integer j>1j>1 and {yi}i=1jΓ\{y_{i}\}_{i=1}^{j}\subseteq\operatorname{\Gamma},

limn1nc(yjn)ϑyj1ny1n=ϕ(yjab)a.s.\lim_{n\uparrow\infty}\frac{1}{n}c(y_{j}^{n})\circ\operatorname{\upvartheta}_{{y_{j-1}^{n}\dots y_{1}^{n}}}=\phi(y_{j}^{\text{ab}})\quad\operatorname{\mathbbm{P}}-a.s.

In particular, if we let εˇ(0,1)\check{\varepsilon}\in(0,1), then there exists, \operatorname{\mathbbm{P}}-a.s., a random M0>0M_{0}>0 depending on εˇ\check{\varepsilon} and on i=1jyiab\sum_{i=1}^{j}\|y_{i}^{\operatorname{\operatorname{ab}}}\| such that, for all n>M0n>M_{0},

|1nc(yjn,yj1ny1nω)ϕ(yjab)|<εˇ.\left|\frac{1}{n}c(y_{j}^{n},{y_{j-1}^{n}\dots y_{1}^{n}}\cdot\omega)~{}-~{}\phi(y_{j}^{\text{ab}})\right|<\check{\varepsilon}.

Before proving Proposition 3.2 we show the following lemma.

Lemma 3.3.

Let ε(0,1)\varepsilon\in(0,1) and consider a subadditive cocycle cc that satisfies condition (i). There exists, \operatorname{\mathbbm{P}}-a.s., M1>0M_{1}>0 such that if {xn}n\{x_{n}\}_{n\in\operatorname{\mathbb{N}}}, {yn}n\{y_{n}\}_{n\in\operatorname{\mathbb{N}}}, {un}n\{u_{n}\}_{n\in\operatorname{\mathbb{N}}}, and {vn}n\{v_{n}\}_{n\in\operatorname{\mathbb{N}}} are sequences in Γ\Gamma satisfying, for a n0=n0(ε)n_{0}=n_{0}(\varepsilon)\in\operatorname{\mathbb{N}} and all n>n0n>n_{0}:

  1. (i)

    There exist elements x,uG\mathlcal{x},\mathlcal{u}\in G_{\infty} and 𝚌x,u>0\mathtt{c}_{\mathlcal{x},\mathlcal{u}}>0 such that

    d(1nxn,x)<ε,d(1nun,u)<ε,d_{\infty}\left(\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}x_{n},\mathlcal{x}\right)<\varepsilon,\quad d_{\infty}\left(\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}u_{n},\mathlcal{u}\right)<\varepsilon,

    and xnS,unS<𝚌x,un\ \|x_{n}\|_{S},\|u_{n}\|_{S}<\mathtt{c}_{\mathlcal{x},\mathlcal{u}}\cdot n;

  2. (ii)

    dS(un,vn)nεanddS(xnun,ynvn)nεd_{S}(u_{n},v_{n})\leq n\varepsilon\quad\text{and}\quad d_{S}(x_{n}u_{n},y_{n}v_{n})\leq n\varepsilon.

Then

|c(xn)ϑunc(yn)ϑvn|<2βnε\big{|}c(x_{n})\circ\operatorname{\upvartheta}_{u_{n}}-c(y_{n})\circ\operatorname{\upvartheta}_{v_{n}}\big{|}<2\upbeta n\varepsilon

for all n>max{n0,M1,exp((2𝚌x,u+3)D)}n>\max\big{\{}n_{0},M_{1},\exp\big{(}(2\mathtt{c}_{\mathlcal{x},\mathlcal{u}}+3)^{D}\big{)}\big{\}}.

Proof.

Fix an:=vnun1a_{n}:=v_{n}u_{n}^{-1} and bn:=ynvn(xnun)1b_{n}:=y_{n}v_{n}(x_{n}u_{n})^{-1}. Then anS=an1S=dS(un,vn)\|a_{n}\|_{S}=\|a_{n}^{-1}\|_{S}=d_{S}(u_{n},v_{n}) and bnS=bn1S=dS(xnun,ynvn)\|b_{n}\|_{S}=\|b_{n}^{-1}\|_{S}=d_{S}(x_{n}u_{n},y_{n}v_{n}). Observe that yn=bnxnan1y_{n}=b_{n}x_{n}a_{n}^{-1} and xn=bn1ynanx_{n}=b_{n}^{-1}y_{n}a_{n}. We thus obtain the \operatorname{\mathbbm{P}}-almost surely inequalities below:

c(yn)ϑvn\displaystyle c(y_{n})\circ\operatorname{\upvartheta}_{v_{n}} c(bn)ϑxnun+c(xn)ϑun+c(an1)ϑvn\displaystyle\leq c(b_{n})\circ\operatorname{\upvartheta}_{x_{n}u_{n}}+c(x_{n})\circ\operatorname{\upvartheta}_{u_{n}}+c(a_{n}^{-1})\circ\operatorname{\upvartheta}_{v_{n}} (3.1)
c(xn)ϑun\displaystyle c(x_{n})\circ\operatorname{\upvartheta}_{u_{n}} c(bn1)ϑynvn+c(yn)ϑvn+c(an)ϑun.\displaystyle\leq c(b_{n}^{-1})\circ\operatorname{\upvartheta}_{y_{n}v_{n}}+c(y_{n})\circ\operatorname{\upvartheta}_{v_{n}}+c(a_{n})\circ\operatorname{\upvartheta}_{u_{n}}. (3.2)

Observe now that, by items (i) and (ii), for n>n0(ε)n>n_{0}(\varepsilon),

xnun,ynvn,un,vnBS(e,2𝚌x,un+3nε).x_{n}u_{n},y_{n}v_{n},u_{n},v_{n}\in B_{S}\left(e,2\mathtt{c}_{\mathlcal{x},\mathlcal{u}}\cdot n+3n\varepsilon\right).

Hence, by combining (3.1) and (3.2),

|c(xn)ϑuvc(yn)ϑvn|2supySnεzSlog(n)Dn{c(y)ϑz}|c(x_{n})\circ\operatorname{\upvartheta}_{u_{v}}-c(y_{n})\circ\operatorname{\upvartheta}_{v_{n}}|\leq\hskip 10.0pt2\sup_{\begin{subarray}{c}\|y\|_{S}\leq n\varepsilon\\ \|z\|_{S}\leq\sqrt[D]{\log(n)}n\end{subarray}}\{c(y)\circ\operatorname{\upvartheta}_{z}\} (3.3)

for all n>max{n0,exp((2𝚌x,u+3)D)}n>\max\big{\{}n_{0},\exp\big{(}(2\mathtt{c}_{\mathlcal{x},\mathlcal{u}}+3)^{D}\big{)}\big{\}}. It follows from (i) that there exists 𝙲>0\mathtt{C}>0 such that

(supySnεzSlog(n)Dn{c(y)ϑz}βnε)𝙲n2Dlog(n)g(βεn)𝒪(log(n)/nκ),\operatorname{\mathbbm{P}}\left(\sup_{\begin{subarray}{c}\|y\|_{S}\leq n\varepsilon\\ \|z\|_{S}\leq\sqrt[D]{\log(n)}n\end{subarray}}\{c(y)\circ\operatorname{\upvartheta}_{z}\}\geq\upbeta n\varepsilon\right)\leq\mathtt{C}{n^{2D}}\log(n)g(\upbeta\varepsilon n)\in\mathcal{O}(\log(n)/n^{\upkappa}), (3.4)

for n>max{n0,exp((x+u+3)D)}n>\max\big{\{}n_{0},\exp\big{(}(\|\mathlcal{x}\|_{\infty}+\|\mathlcal{u}\|_{\infty}+3)^{D}\big{)}\big{\}}. Since n=1+log(n)nκ=ζ(κ)<+\sum_{n=1}^{+\infty}\frac{\log(n)}{n^{\upkappa}}=-\upzeta^{\prime}(\upkappa)<+\infty for κ>1\upkappa>1 where ζ\upzeta^{\prime} is the derivative of the Riemann zeta function, the proof is completed by applying Borel-Cantelli Lemma to (3.3) and (3.4). ∎

Remark 4.

If {xn}n,{un}nΓ\{x_{n}\}_{n\in\operatorname{\mathbb{N}}},\{u_{n}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\Gamma} are such that limn+1nxn=x\lim_{n\uparrow+\infty}\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}x_{n}=\mathlcal{x} and limn+1nun=u\lim_{n\uparrow+\infty}\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}u_{n}=\mathlcal{u} in (G,d)(G_{\infty},d_{\infty}), then item (i) of Lemma 3.3 is immediately satisfied (see Section 2.2).

Using the lemma above, the Proposition 3.2 becomes a straightforward extension of Theorem 3.3 of [8]. The result can be verified by replacing the Parallelogram inequality with Lemma 3.3. To be self-contained, let us first define, for each EE\in\operatorname{\mathscr{F}}, ωΩ\omega\in\operatorname{\Upomega}, xΓx\in\operatorname{\Gamma}, ξ>0\xi>0, and nn\in\operatorname{\mathbb{N}},

𝖭x,nξ(E,ω):=#{n{0,1,,ξn1}:ϑxnn(ω)E}.\mathsf{N}_{x,n}^{\xi}(E,\omega):=\#\big{\{}n^{\prime}\in\{0,1,\dots,\lceil\xi n\rceil-1\}\colon\operatorname{\upvartheta}_{x^{n-n^{\prime}}}(\omega)\in E\big{\}}.

Set

Ξ(E,x,ξ):={ωΩ:lim infn+𝖭x,nξ(E,ω)ξn>0},and\displaystyle\Xi^{\star}(E,x,\xi):=\left\{\omega\in\operatorname{\Upomega}\colon\liminf_{n\uparrow+\infty}\frac{\mathsf{N}_{x,n}^{\xi}(E,\omega)}{\xi n}>0\right\},\quad\text{and}
Ξm(E,x,ξ):={ωΩ:nm(𝖭x,nξ(E,ω)ξn>0)}.\displaystyle\Xi_{m}^{\star}(E,x,\xi):=\left\{\omega\in\operatorname{\Upomega}\colon\forall n\geq m\left(\frac{\mathsf{N}_{x,n}^{\xi}(E,\omega)}{\xi n}>0\right)\right\}.

We now state Lemma 3.6 of [8] without proof before the proving Proposition 3.2.

Lemma 3.4.

Let xΓx\in\operatorname{\Gamma}, ξ>0\xi>0, and EE\in\operatorname{\mathscr{F}}. Then, for all ε(0,1)\varepsilon\in(0,1), there is m0>0m_{0}>0 such that, for m>mom>m_{o},

(Ξ(E,x,ξ))(E)and(Ξm(E,x,ξ))>(Ξ(E,x,ξ))ε.\operatorname{\mathbbm{P}}\big{(}\Xi^{\star}(E,x,\xi)\big{)}\geq\operatorname{\mathbbm{P}}(E)\quad\text{and}\quad\operatorname{\mathbbm{P}}\big{(}\Xi_{m}^{\star}(E,x,\xi)\big{)}>\operatorname{\mathbbm{P}}\big{(}\Xi^{\star}(E,x,\xi)\big{)}-\varepsilon.

We proceed below with the proof of ergodic subadditive approximation via polygonal paths.

Proof of Proposition 3.2.

Consider ε(0,1)\varepsilon^{\prime}\in(0,1) and {yi}i=1jΓ\{y_{i}\}_{i=1}^{j}\subseteq\operatorname{\Gamma} fixed. Let ξ>0\xi>0 and n \accentset{\rule{2.79996pt}{0.5pt}}{n}\in\operatorname{\mathbb{N}} be given by Lemma 2.3 for ε=ε\varepsilon=\varepsilon^{\prime}. Set η(0,12j)\eta\in(0,\frac{1}{2j}) and mm\in\operatorname{\mathbb{N}} sufficiently large so that, for each i{1,,j}i\in\{1,\dots,j\}, one has by Proposition 2.4,

𝒳i:={ωΩ:n>m(|1nc(yin,ω)ϕ(yiab)|<εˇ)} and (𝒳i)>1η.\mathcal{X}_{i}:=\left\{\omega\in\operatorname{\Upomega}\colon\forall n>m\left(\left|\frac{1}{n}c(y_{i}^{n},\omega)-\phi(y_{i}^{\operatorname{\operatorname{ab}}})\right|<\check{\varepsilon}\right)\right\}\ \text{ and }\ \operatorname{\mathbbm{P}}(\mathcal{X}_{i})>1-\eta.

Fix 𝒴j:=𝒳j\mathcal{Y}_{j}:=\mathcal{X}_{j} and define inductively 𝒴i1:=𝒳i1Ξmi(𝒴i,yi,ξ)\mathcal{Y}_{i-1}:=\mathcal{X}_{i-1}\cap\Xi_{m_{i}}^{\ast}(\mathcal{Y}_{i},y_{i},\xi) so that, for each i{2,,j}i\in\{2,\dots,j\}, mim_{i}\in\operatorname{\mathbb{N}} is given by Lemma 3.4 satisfying

(Ξmi(𝒴i,yi,ξ))(𝒴i)η.\operatorname{\mathbbm{P}}\big{(}\Xi_{m_{i}}^{\ast}(\mathcal{Y}_{i},y_{i},\xi)\big{)}\geq\operatorname{\mathbbm{P}}(\mathcal{Y}_{i})-\eta.

Therefore,

(𝒴1)>(𝒴2)2η>>(𝒴j)2(j1)η>1(2j1)η.\operatorname{\mathbbm{P}}(\mathcal{Y}_{1})>\operatorname{\mathbbm{P}}(\mathcal{Y}_{2})-2\eta>\cdots>\operatorname{\mathbbm{P}}(\mathcal{Y}_{j})-2(j-1)\eta>1-(2j-1)\eta.

Let now mˇ:=max{m,m1,,mj}\check{m}:=\max\{m,m_{1},\dots,m_{j}\}. Thence, for all ϖi𝒴i\varpi_{i}\in\mathcal{Y}_{i} and every n>mˇn>\check{m} with i{1,,j1}i\in\{1,\dots,j-1\}, if ni<ξnn_{i}<\xi n, then ϑynni(ϖi)=ynniϖi𝒴i+1\operatorname{\upvartheta}_{y^{n-n_{i}}}(\varpi_{i})=y^{n-n_{i}}\cdot\varpi_{i}\in\mathcal{Y}_{i+1}, and

|1nc(yi+1n,ϖi)ϕ(yi+1ab)|<ε.\left|\frac{1}{n}c(y_{i+1}^{n},\varpi_{i})-\phi(y_{i+1}^{\operatorname{\operatorname{ab}}})\right|<\varepsilon^{\prime}.

It follows that, for all n>mˇn>\check{m}, there exist non negative integers n1,,nj1<ξnn_{1},\dots,n_{j-1}<\xi n such that, for all ω𝒴1\omega\in\mathcal{Y}_{1},

|1nc(yjn,yj1nnj1y1nn1ω)ϕ(yjab)|<ε.\left|\frac{1}{n}c(y_{j}^{n},y_{j-1}^{n-n_{j-1}}\cdots y_{1}^{n-n_{1}}\cdot\omega)-\phi(y_{j}^{\operatorname{\operatorname{ab}}})\right|<\varepsilon^{\prime}.

By Lemma 2.3, for each i{2,,j}i\in\{2,\dots,j\} and every n>n n>\accentset{\rule{2.79996pt}{0.5pt}}{n},

dS(yi1ny1n,yi1nni1y1nn1)<nεand\displaystyle d_{S}\left(y_{i-1}^{n}\cdots y_{1}^{n},~{}y_{i-1}^{n-n_{i-1}}\cdots y_{1}^{n-n_{1}}\right)<n\varepsilon^{\prime}\quad\text{and}
dS(yinyi1ny1n,yinyi1nni1y1nn1)<nε.\displaystyle d_{S}\left(y_{i}^{n}y_{i-1}^{n}\cdots y_{1}^{n},~{}y_{i}^{n}y_{i-1}^{n-n_{i-1}}\cdots y_{1}^{n-n_{1}}\right)<n\varepsilon^{\prime}.

Hence, by Lemmas 2.2 and 3.3, there exists Ωj\operatorname{\Upomega}_{j}\in\operatorname{\mathscr{F}} with (Ωj)=1\operatorname{\mathbbm{P}}(\operatorname{\Upomega}_{j})=1, and a random M in \accentset{\rule{2.79996pt}{0.5pt}}{M}_{i}\geq\accentset{\rule{2.79996pt}{0.5pt}}{n} depending on ε\varepsilon^{\prime}, and yiabyi1aby1abyi1aby1ab\|y_{i}^{\operatorname{\operatorname{ab}}}y_{i-1}^{\operatorname{\operatorname{ab}}}\dots y_{1}^{\operatorname{\operatorname{ab}}}\|_{\infty}\vee\|y_{i-1}^{\operatorname{\operatorname{ab}}}\dots y_{1}^{\operatorname{\operatorname{ab}}}\|_{\infty} such that for all n>M in>\accentset{\rule{2.79996pt}{0.5pt}}{M}_{i} and all ωΩj\omega\in\operatorname{\Upomega}_{j},

|c(yin,yi1ny1nω)c(yin,yi1nni1y1nn1ω)|<2βnε.\left|c(y_{i}^{n},~{}y_{i-1}^{n}\cdots y_{1}^{n}\cdot\omega)-c(y_{i}^{n},~{}y_{i-1}^{n-n_{i-1}}\cdots y_{1}^{n-n_{1}}\cdot\omega)\right|<2\upbeta n\varepsilon^{\prime}.

Therefore, for all ωΩj𝒴1\omega\in\operatorname{\Upomega}_{j}\cap\ \mathcal{Y}_{1}, every n>mˇM in>\check{m}\vee\accentset{\rule{2.79996pt}{0.5pt}}{M}_{i} and all i{2,,j}i\in\{2,\dots,j\}

|1nc(yin,yi1ny1nω)ϕ(yiab)|<(2β+1)εˇ,\left|\frac{1}{n}c(y_{i}^{n},~{}y_{i-1}^{n}\cdots y_{1}^{n}\cdot\omega)-\phi(y_{i}^{\operatorname{\operatorname{ab}}})\right|<(2\upbeta+1)\check{\varepsilon}, (3.5)

and (Ωj𝒴1)>1(2j1)η\operatorname{\mathbbm{P}}\left(\operatorname{\Upomega}_{j}\cap\ \mathcal{Y}_{1}\right)>1-(2j-1)\eta. It suffices to consider ηn0\eta_{n}\downarrow 0 replacing η(0,12j)\eta\in(0,\frac{1}{2j}) with nηn<+\sum_{n\in\operatorname{\mathbb{N}}}\eta_{n}<+\infty, then there exists, \operatorname{\mathbbm{P}}-a.s., M0mˇM iM_{0}\geq\check{m}\vee\accentset{\rule{2.79996pt}{0.5pt}}{M}_{i} by Borel-Cantelli Lemma such that (3.5) is satisfied for all n>M0n>M_{0}, which is our assertion with i=ji=j and εˇ=12β+1ε\check{\varepsilon}=\frac{1}{2\upbeta+1}\varepsilon^{\prime}. ∎

3.2. Proof of the first theorem

This subsection is dedicated to proving Theorem 1.1. Therefore, consider all conditions and notations established in the first main theorem for the subsequent results. For instance, here Γ\operatorname{\Gamma} is torsion-free nilpotent with torsion-free abelianization. Before turning to the proof of the theorem, let us refine the techniques of approximation as outlined in the upcoming propositions and lemmas.

Proposition 3.5.

Let gG\mathlcal{g}\in G_{\infty} and ϵ(0,1)\epsilon\in(0,1). Consider {yj}j=1kΓ\{y_{j}\}_{j=1}^{k}\subseteq\operatorname{\Gamma} and p>0p>0 given by Proposition 3.1 for a dd_{\infty}-geodesic curve γ:[0,1]G\upgamma:[0,1]\to G_{\infty} from e\mathlcal{e} to g\mathlcal{g} and ε^=ϵ/2\hat{\varepsilon}=\epsilon/2.

If conditions (i), (ii), and (iii) are satisfied, then there exists, \operatorname{\mathbbm{P}}-a.s., M2>0M_{2}>0 depending on g\mathlcal{g}, ϵ\epsilon, and ωΩ\omega\in\operatorname{\Upomega}, such that, for all n>M2n>M_{2},

|1pnc(ykny1n)ϕ(γ)|<ϵ.\left|\frac{1}{pn}c(y_{k}^{n}\cdots y_{1}^{n})-\ell_{\phi}(\upgamma)\right|<\epsilon.
Proof.

Let us write 𝐲n:=ykny1n\mathbf{y}_{n}:=y_{k}^{n}\dots y_{1}^{n} and consider n0>0n_{0}>0 for ε^=ϵ/2\hat{\varepsilon}=\epsilon/2 given by Proposition 3.1. It follows from subadditivity that

c(𝐲n)j=1kc(yjn)ϑyj1ny1n-a.s.c(\mathbf{y}_{n})\leq\sum_{j=1}^{k}c(y_{j}^{n})\circ\operatorname{\upvartheta}_{y_{j-1}^{n}\dots y_{1}^{n}}\quad\operatorname{\mathbbm{P}}\text{-a.s.}

Then, one has by Proposition 3.2 with εˇp2kϵ\check{\varepsilon}\leq\frac{p}{2k}\epsilon that, \operatorname{\mathbbm{P}}-a.s., for all n>M0n0n>M_{0}\vee n_{0},

1pnc(𝐲n)1pj=1kϕ(ynab)+ϵ2<ϕ(γ)+ϵ.\frac{1}{pn}c(\mathbf{y}_{n})\leq\frac{1}{p}\sum_{j=1}^{k}\phi(y_{n}^{\operatorname{\operatorname{ab}}})+\frac{\epsilon}{2}<\ell_{\phi}(\upgamma)+\epsilon. (3.6)

Set ε(0,1)\upvarepsilon\in(0,1) to be defined later and apply condition (iii) to obtain

j=1kncn,j(1+ε)1pnc(𝐲n)\sum_{j=1}^{k_{n}}c_{n,j}\leq(1+\upvarepsilon)\frac{1}{pn}c(\mathbf{y}_{n})

where cn,j=1pnc(zn,j,zn,j1zn,1ω)c_{n,j}=\frac{1}{pn}c(z_{n,j},{z_{n,j-1}\dots z_{n,1}}\cdot\omega) with zm,iF(ε)z_{m,i}\in F(\upvarepsilon). Define a sequence of piecewise dd_{\infty}-geodesic curves ζn\zeta_{n} between each 1pnzn,jzn,1\frac{1}{pn}\operatorname{{\scriptscriptstyle\bullet}}z_{n,j}\dots z_{n,1} and 1pnzn,j1zn,1\frac{1}{pn}\operatorname{{\scriptscriptstyle\bullet}}z_{n,j-1}\dots z_{n,1} for j{1,,kn}j\in\{1,\dots,k_{n}\} such that zn,0=ez_{n,0}=e and

ζn(τj)=1pnzn,jzn,1for τj=i=1jcn,i/i=1kncn,i.\zeta_{n}(\tau_{j})=\frac{1}{pn}\operatorname{{\scriptscriptstyle\bullet}}z_{n,j}\dots z_{n,1}\quad\text{for }\tau_{j}=\sum_{i=1}^{j}c_{n,i}\Big{/}\sum_{i=1}^{k_{n}}c_{n,i}.

Set 𝚖ε:=minzF(ε)zab>0\mathtt{m}_{\upvarepsilon}:=\min_{z\in F(\upvarepsilon)}\|z^{\operatorname{\operatorname{ab}}}\|>0. It follows from Lemma 2.5 that 𝔼[cn,j]a𝚖ε1pn\operatorname{\mathbbm{E}}[c_{n,j}]\geq a^{\prime}\mathtt{m}_{\upvarepsilon}\frac{1}{pn} and, due the the L1L^{1} convergence in Proposition 2.4, there exists n1>n0n_{1}>n_{0} such that, for all n>M0n1n>M_{0}\vee n_{1},

a𝚖ε1pn𝔼[kn](1+ε)(γ)+kε.a^{\prime}\mathtt{m}_{\upvarepsilon}\frac{1}{pn}\operatorname{\mathbbm{E}}[k_{n}]\leq(1+\upvarepsilon)~{}\ell_{\infty}(\upgamma)+k\upvarepsilon.

Fix 𝙲γ,ε>2pa𝚖ε((1+ε)(γ)+kε)\mathtt{C}_{\upgamma,\upvarepsilon}>\frac{2p}{a^{\prime}\mathtt{m}_{\upvarepsilon}}\big{(}(1+\upvarepsilon)\ell_{\infty}(\upgamma)+k\upvarepsilon\big{)} so that, for all nn\in\operatorname{\mathbb{N}}, 𝔼[kn]𝙲γ,εn/2\operatorname{\mathbbm{E}}[k_{n}]\leq\mathtt{C}_{\upgamma,\upvarepsilon}n/2. By Chernoff bound, (kn𝙲γ,εn)exp(2n)\operatorname{\mathbbm{P}}(k_{n}\geq\mathtt{C}_{\upgamma,\upvarepsilon}n)\leq\exp(-2n). It then follows from an application of Borel-Cantelli Lemma that, \operatorname{\mathbbm{P}}-a.s., there exists M0M0M_{0}^{\prime}\geq M_{0} such that, for every n>M0n>M_{0}^{\prime},

kn𝙲γ,εn.k_{n}\leq\mathtt{C}_{\upgamma,\upvarepsilon}n. (3.7)

Let 𝙼ε:=maxzF(ε)1z\mathtt{M}_{\upvarepsilon}:=\max\limits_{z\in F(\upvarepsilon)}\|1\operatorname{{\scriptscriptstyle\bullet}}z\|_{\infty} Observe now that, for every t,t[0,1]t,t^{\prime}\in[0,1], \operatorname{\mathbbm{P}}-a.s., for n>M0n>M_{0}^{\prime},

d(ζn(t),ζn(t))knpn𝙼ε|tt|1p𝙲γ,ε𝙼ε|tt|.d_{\infty}\big{(}\zeta_{n}(t),\zeta_{n}(t^{\prime})\big{)}\leq\frac{k_{n}}{pn}\mathtt{M}_{\upvarepsilon}|t-t^{\prime}|\leq\frac{1}{p}\mathtt{C}_{\upgamma,\upvarepsilon}\mathtt{M}_{\upvarepsilon}|t-t^{\prime}|.

Hence, one has by Arzelà–Ascoli Theorem that a subsequence of ζn\zeta_{n} converges uniformly to a Lipschitz curve ζ:[0,1]G\zeta:[0,1]\to G_{\infty} such that ζ(0)=e\zeta(0)=\mathlcal{e} and ζ(1)=g\zeta(1)=\mathlcal{g}.

We apply Proposition 3.1 once again fot the curve ζ\zeta with ε^=ε/2\hat{\varepsilon}=\upvarepsilon/2 to obtain p>0p^{\prime}>0, {wi}i=1kΓ\{w_{i}\}_{i=1}^{k^{\prime}}\subseteq\operatorname{\Gamma}, tn=np/pt_{n}=\lfloor np/p^{\prime}\rfloor, and n2>0n_{2}>0 such that, for all n>n2n>n_{2}

i=1kd(1ptnwitnw1tn,ζ(ik))<ε2.\sum_{i=1}^{k^{\prime}}d_{\infty}\left(\frac{1}{p^{\prime}t_{n}}\operatorname{{\scriptscriptstyle\bullet}}w_{i}^{t_{n}}\dots w_{1}^{t_{n}},\zeta\left(\frac{i}{k^{\prime}}\right)\right)<\frac{\upvarepsilon}{2}.

Recall that F(ε)F(\upvarepsilon) is a generating set of Γ\operatorname{\Gamma} and 𝒞(Γ,F(ε))\mathcal{C}\big{(}\operatorname{\Gamma},F(\upvarepsilon)\big{)} shares the polynomial growth rate of 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S). Then there exists 𝙲>0\mathtt{C}^{\prime}>0 such that, for a given ε(0,1)\varepsilon^{\prime}\in(0,1),

(supzF(ε){c(z)ϑz:zBF(ε)(e,𝙲γ,εn)}εn)𝙲|F(ε)|nDg(εn)\displaystyle\operatorname{\mathbbm{P}}\left(\sup_{z\in F(\upvarepsilon)}\Big{\{}c(z)\circ\operatorname{\upvartheta}_{z^{\prime}}:z^{\prime}\in B_{F(\upvarepsilon)}(e,\mathtt{C}_{\upgamma,\upvarepsilon}n)\Big{\}}\geq\varepsilon^{\prime}n\right)\leq\mathtt{C}^{\prime}|F(\upvarepsilon)|n^{D}g(\varepsilon^{\prime}n)
𝒪ε(1/nκ)\displaystyle\in\mathcal{O}_{\varepsilon^{\prime}}(1/n^{\upkappa})

as n+n\uparrow+\infty.

It thus follows by an application of Borel-Cantelli Lemma and by (3.7) that for all ε(0,1)\varepsilon^{\prime}\in(0,1), there exist, \operatorname{\mathbbm{P}}-a.s., M0′′M0M_{0}^{\prime\prime}\geq M_{0}^{\prime} and a subdivision function 𝖽n:{0,1,,k}{0,1,,kn}\mathsf{d}_{n}:\{0,1,\dots,k^{\prime}\}\to\{0,1,\dots,k_{n}\} with 𝖽n(0)=0<𝖽n(1)<<𝖽n(k)=kn\mathsf{d}_{n}(0)=0<\mathsf{d}_{n}(1)<\cdots<\mathsf{d}_{n}(k^{\prime})=k_{n} such that, for all n>M0′′n>M_{0}^{\prime\prime},

|1ki=𝖽n(j1)𝖽n(j)1cn,i+1/i=1kncn,i|<ε.\left|\frac{1}{k^{\prime}}-\left.\sum_{i=\mathsf{d}_{n}(j-1)}^{\mathsf{d}_{n}(j)-1}c_{n,i+1}\right/\sum_{i=1}^{k_{n}}c_{n,i}\right|<\varepsilon^{\prime}.

Let

gn,j:=1ptnzn,𝖽n(j)zn,𝖽n(j)1zn,1for j{1,,k}.\mathlcal{g}_{n,j}:=\frac{1}{p^{\prime}t_{n}}\operatorname{{\scriptscriptstyle\bullet}}z_{n,\mathsf{d}_{n}(j)}z_{n,\mathsf{d}_{n}(j)-1}\dots z_{n,1}\quad\text{for }j\in\{1,\dots,k^{\prime}\}.

Then there exist n3n2n_{3}\geq n_{2} and, \operatorname{\mathbbm{P}}-a.s., M2>M0′′M_{2}^{\prime}>M_{0}^{\prime\prime} such that, for all n>M2n3n>M_{2}^{\prime}\vee n_{3}, gn,jζ(j/k)<ε2\|\mathlcal{g}_{n,j}-\zeta(j/k^{\prime})\|_{\infty}<\frac{\upvarepsilon}{2}. Hence, for n>M2n3n>M_{2}^{\prime}\vee n_{3}, j=1kd(1ptnwjtnw1tn,gn,j)<ε\sum_{j=1}^{k^{\prime}}d_{\infty}\left(\frac{1}{p^{\prime}{t_{n}}}\operatorname{{\scriptscriptstyle\bullet}}w_{j}^{t_{n}}\dots w_{1}^{t_{n}},\mathlcal{g}_{n,j}\right)<\upvarepsilon.

It follows that there exists, \operatorname{\mathbbm{P}}-a.s. M2′′>M2n3M_{2}^{\prime\prime}>M_{2}^{\prime}\vee n_{3} so that, for every n>M2′′n>M_{2}^{\prime\prime},

j=1k1ptndS(wjtnw1tn,zn,𝖽n(j)zn,𝖽n(j)1zn,1)<ε\sum_{j=1}^{k^{\prime}}\frac{1}{p^{\prime}t_{n}}d_{S}\left(w_{j}^{t_{n}}\dots w_{1}^{t_{n}},z_{n,\mathsf{d}_{n}(j)}z_{n,\mathsf{d}_{n}(j)-1}\dots z_{n,1}\right)<\upvarepsilon

We thus get from Lemma 3.3 that there exists, \operatorname{\mathbbm{P}}-a.s., M1>M2′′M_{1}^{\prime}>M_{2}^{\prime\prime}, such that, for each n>M1n>M_{1}^{\prime},

j=1k|c(wjtn,wj1tnw1tnω)c(zn,𝖽n(j),zn,𝖽n(j)1zn,1ω)|<4βptnε.\sum_{j=1}^{k^{\prime}}\left|c(w_{j}^{t_{n}},w_{j-1}^{t_{n}}\dots w_{1}^{t_{n}}\cdot\omega)-c(z_{n,\mathsf{d}_{n}(j)},z_{n,\mathsf{d}_{n}(j)-1}\dots z_{n,1}\cdot\omega)\right|<4\upbeta p^{\prime}t_{n}\upvarepsilon.

Set M2M1ppε2M_{2}\geq M_{1}^{\prime}\vee\frac{p^{\prime}}{p\upvarepsilon^{2}}. Therefore, one has, \operatorname{\mathbbm{P}}-a.s., for all n>M2n>M_{2},

1pnc(𝐲n,ω)\displaystyle\frac{1}{pn}c(\mathbf{y}_{n},\omega) >1(1+ε)pnj=1ki=dn(j1)dn(j)c(zn,i,zn,i1zn,1ω)\displaystyle>\frac{1}{(1+\upvarepsilon)pn}\sum_{j=1}^{k^{\prime}}~{}\sum_{i=d_{n}(j-1)}^{d_{n}(j)}c(z_{n,i},z_{n,i-1}\dots z_{n,1}\cdot\omega)
1(1+ε)pnj=1kc(zn,𝖽n(j),zn,𝖽n(j)1zn,1ω)\displaystyle\geq\frac{1}{(1+\upvarepsilon)pn}\sum_{j=1}^{k^{\prime}}c(z_{n,\mathsf{d}_{n}(j)},z_{n,\mathsf{d}_{n}(j)-1}\dots z_{n,1}\cdot\omega)
11+εptnpn(1ptnj=1kc(wjtn,wj1tnw1tnω)4βε)\displaystyle\geq\frac{1}{1+\upvarepsilon}\frac{p^{\prime}t_{n}}{pn}\left(\frac{1}{p^{\prime}t_{n}}\sum_{j=1}^{k^{\prime}}c(w_{j}^{t_{n}},w_{j-1}^{t_{n}}\dots w_{1}^{t_{n}}\cdot\omega)-4\upbeta\upvarepsilon\right)
>(1ε)(ϕ(ζ)(4β+1)ε).\displaystyle>(1-\upvarepsilon)\Big{(}\ell_{\phi}(\zeta)-(4\upbeta+1)\upvarepsilon\Big{)}.

Fix ε=12(ϕ(γ)+4β+1)ϵ\upvarepsilon=\frac{1}{2\big{(}\ell_{\phi}(\upgamma)+4\upbeta+1\big{)}}\epsilon. Hence, since ϕ(ζ)ϕ(γ)ϵ/2\ell_{\phi}(\zeta)\geq\ell_{\phi}(\upgamma)-\epsilon/2, one has, \operatorname{\mathbbm{P}}-a.s., for all n>M2n>M_{2},

1pnc(𝐲n,ω)>ϕ(γ)ϵ.\frac{1}{pn}c(\mathbf{y}_{n},\omega)>\ell_{\phi}(\upgamma)-\epsilon. (3.8)

We complete the proof by combining (3.6) and (3.8). ∎

Lemma 3.6.

Let {xn}n\{x_{n}\}_{n\in\operatorname{\mathbb{N}}} be a sequence in Γ\operatorname{\Gamma} and let {tn}n\{t_{n}\}_{n\in\operatorname{\mathbb{N}}} be an increasing sequence in \operatorname{\mathbb{R}} such that limn+1tnxn=gG\lim_{n\uparrow+\infty}\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{n}=\mathlcal{g}\in G_{\infty}.

Consider a subadditive cocycle c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} satisfying conditions (i) and (ii). If condition (iii) is satisfied or if Γ\operatorname{\Gamma} is abelian, then, for all ϵ(0,1)\upepsilon\in(0,1), there exists, \operatorname{\mathbbm{P}}-a.s., a random M=M(g,ϵ)>0M=M(\mathlcal{g},\epsilon)>0 such that, for tn>Mt_{n}>M,

|1tnc(xn)Φ(g)|<ϵ\left|\frac{1}{t_{n}}c(x_{n})-\Phi(\mathlcal{g})\right|<\upepsilon
Proof.

Set ε>0\varepsilon>0 to be defined later. Consider yk,,y1Γy_{k},\dots,y_{1}\in\operatorname{\Gamma} and pp given by Proposition 3.1 for ε^=ε/2\hat{\varepsilon}=\varepsilon/2 and a dd_{\infty}-geodesic curve γ:[0,1]G\upgamma:[0,1]\to G_{\infty} from e\mathlcal{e} to g\mathlcal{g}. Let tn:=tn/pt_{n}^{\prime}:=\lfloor t_{n}/p\rfloor. Since 1tnxi\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{i} converges to g\mathlcal{g}. It follows from the Borel-Cantelli Lemma applied to (i) that there exists, \operatorname{\mathbbm{P}}-a.s., M>0M^{\prime}>0 so that, for every tn>Mt_{n}>M^{\prime},

(1ptn1tn)c(xn)<p1ptnβxnStn<ϵ4.\left(\frac{1}{pt_{n}^{\prime}}-\frac{1}{t_{n}}\right)c(x_{n})<\frac{p-1}{pt_{n}^{\prime}}\upbeta\frac{\|x_{n}\|_{S}}{t_{n}}<\frac{\upepsilon}{4}. (3.9)

Let us write 𝐲i:=yktny1tn\mathbf{y}_{i}^{\prime}:=y_{k}^{t_{n}^{\prime}}\dots y_{1}^{t_{n}^{\prime}}. Since limn+d(1ptn𝐲n,g)=0\lim_{n\uparrow+\infty}d_{\infty}\left(\frac{1}{pt_{n}^{\prime}}\operatorname{{\scriptscriptstyle\bullet}}\mathbf{y}_{n}^{\prime},\mathlcal{g}\right)=0, one can easily see that there exists n1>0n_{1}^{\prime}>0 such that, for all tn>n1t_{n}^{\prime}>n_{1}^{\prime},

1ptnd(xn,𝐲n)<ε.\frac{1}{pt_{n}^{\prime}}d_{\infty}(x_{n},\mathbf{y}_{n}^{\prime})<\varepsilon.

Then there exists n2n1n_{2}^{\prime}\geq n_{1}^{\prime} such that, for all tn>n2t_{n}^{\prime}>n_{2}^{\prime}, one has xn(𝐲n)1S<ptnε\|x_{n}(\mathbf{y}_{n}^{\prime})^{-1}\|_{S}<pt_{n}^{\prime}\varepsilon.

Let now t=βptnεt=\upbeta pt_{n}^{\prime}\varepsilon in (i). Since c(x)c(x) is identically distributed to c(x)ϑyc(x)\circ\operatorname{\upvartheta}_{y}, we have by Borel-Cantelli Lemma that there exists, \operatorname{\mathbbm{P}}-a.s., M′′Mn2M^{\prime\prime}\geq M^{\prime}\vee n_{2}^{\prime} such that, for tn>M′′t_{n}>M^{\prime\prime},

1ptn|c(xn)c(𝐲n)|\displaystyle\frac{1}{pt_{n}^{\prime}}\big{|}c(x_{n})-c(\mathbf{y}_{n}^{\prime})\big{|} 1ptnmax{c(𝐲nxn1)ϑxn,c(xn(𝐲n)1)ϑ𝐲n}\displaystyle\leq\frac{1}{pt_{n}^{\prime}}\max\big{\{}c(\mathbf{y}_{n}^{\prime}x_{n}^{-1})\circ\operatorname{\upvartheta}_{x_{n}},c(x_{n}(\mathbf{y}_{n}^{\prime})^{-1})\circ\operatorname{\upvartheta}_{\mathbf{y}_{n}^{\prime}}\big{\}}
<2βε.\displaystyle<2\upbeta\varepsilon. (3.10)

Set εϵ8β\varepsilon\leq\frac{\upepsilon}{8\upbeta} and combine (3.9) and (3.10). We thus obtain that, \operatorname{\mathbbm{P}}-a.s., for all tn>M′′t_{n}>M^{\prime\prime},

|1tnc(xn)1ptnc(𝐲n)|<ϵ2.\left|\frac{1}{t_{n}}c(x_{n})-\frac{1}{pt_{n}^{\prime}}c\left(\mathbf{y}_{n}^{\prime}\right)\right|<\frac{\upepsilon}{2}. (3.11)

Consider Γ\operatorname{\Gamma} abelian, then 𝐲n=(yky1)tn\mathbf{y}_{n}^{\prime}=(y_{k}\dots y_{1})^{t_{n}^{\prime}}. In fact, it is straightfoward that k=1k=1 by the standard approach for commutative groups. Then by Proposition 2.4, there is, \operatorname{\mathbbm{P}}-a.s., MM′′M^{\ast}\geq M^{\prime\prime} such that, for all tn>Mt_{n}>M^{\ast},

1p|1tnc(𝐲n)ϕ(y1ab)|<ϵ2.\frac{1}{p}\left|\frac{1}{t_{n}^{\prime}}c\left(\mathbf{y}_{n}^{\prime}\right)-\phi(y_{1}^{\operatorname{\operatorname{ab}}})\right|<\frac{\upepsilon}{2}. (3.12)

Furthermore, we have ϕ(y1ab)/p=ϕ(γ)=Φ(g)\phi(y_{1}^{\operatorname{\operatorname{ab}}})/p=\ell_{\phi}(\upgamma)=\Phi(\mathlcal{g}). Combining the two previous inequalities with (3.11), we can establish the result for the commutative case with M=MM=M^{\ast}. Now, let’s consider the non-abelian case, assuming that (iii) holds true. Notably, by Proposition 3.5 with ϵ/2\epsilon/2 and M=M′′M2M=M^{\prime\prime}\vee M_{2}, for all tn>Mt_{n}>M,

|1ptnc(𝐲n)ϕ(γ)|<ϵ2.\left|\frac{1}{pt_{n}^{\prime}}c\left(\mathbf{y}_{n}^{\prime}\right)-\ell_{\phi}(\upgamma)\right|<\frac{\upepsilon}{2}.

This result, when combined with (3.11), completes the proof. ∎

We now proceed to demonstrate the proof of the first main theorem.

Proof of Theorem 1.1.

We begin by proving the \operatorname{\mathbbm{P}}-a.s. asymptotic equivalence given, which is given by

limx+|c(x)Φ(1x)|xS=0-a.s.\lim_{\|x\|\uparrow+\infty}\frac{|c(x)-\Phi(1\operatorname{{\scriptscriptstyle\bullet}}x)|}{\|x\|_{S}}=0\quad\operatorname{\mathbbm{P}}\text{-a.s.} (3.13)

Suppose, by contradiction, that (3.13) is not true. Consider {vn}nΓ\{v_{n}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\Gamma} to be such that vnS+\|v_{n}\|_{S}\uparrow+\infty as n+n\uparrow+\infty. Let Sr\mathlcal{S}_{r} stand for B(e,r)¯\overline{B_{\infty}(\mathlcal{e},r)}, the closure of the dd_{\infty}-ball or radius r>0r>0 in GG_{\infty}. Due to the compactness of S1\mathlcal{S}_{1} with respect to dd_{\infty}, there exists a subsequence {yn}n{vn}n\{y_{n}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\{v_{n}\}_{n\in\operatorname{\mathbb{N}}} such that, for tn:=ynSt_{n}:=\|y_{n}\|_{S}

limn+1tnyn=hS1.\lim_{n\uparrow+\infty}\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}y_{n}=\mathlcal{h}\in\mathlcal{S}_{1}.

By construction, Δ:=n(1nΓ)\Delta:=\bigcup_{n\in\operatorname{\mathbb{N}}}\left(\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}\operatorname{\Gamma}\right) is a countable dense subset of GG_{\infty}. Fix, for each gΔ\mathlcal{g}\in\Delta, σ(g)={xn}n\sigma(\mathlcal{g})=\{x_{n}\}_{n\in\operatorname{\mathbb{N}}} such that 1nxn\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}x_{n} converges to g\mathlcal{g} under dd_{\infty} (see Lemma 2.2). Let Ωg\operatorname{\Upomega}_{\mathlcal{g}}\in\operatorname{\mathscr{F}} be the event with (Ωg)=1\operatorname{\mathbbm{P}}(\operatorname{\Upomega}_{\mathlcal{g}})=1 given by Lemma 3.6 for σ(g)\sigma(\mathlcal{g}). Hence, ΩΔ:=gΔΩg\operatorname{\Upomega}_{\Delta}:=\bigcap_{\mathlcal{g}\in\Delta}\operatorname{\Upomega}_{\mathlcal{g}} is such that (ΩΔ)=1\operatorname{\mathbbm{P}}(\operatorname{\Upomega}_{\Delta})=1.

The compactness of Sr\mathlcal{S}_{r} implies the existence of a finite Δr,εSrΔ\Delta_{r,\varepsilon}\subseteq\mathlcal{S}_{r}\cap\Delta such that gΔr,εB(g,ε)\bigcup_{\mathlcal{g}\in\Delta_{r,\varepsilon}}B_{\infty}(\mathlcal{g},\varepsilon) covers Sr\mathlcal{S}_{r}. Thus there exists gΔ1,ε\mathlcal{g}\in\Delta_{1,\varepsilon} so that hB(g,ε)\mathlcal{h}\in B_{\infty}(\mathlcal{g},\varepsilon). Consider σ(g)={xn}n\sigma(\mathlcal{g})=\{x_{n}\}_{n\in\operatorname{\mathbb{N}}} as defined above and let ε>0\varepsilon>0 to be determined later. Then, there exists m(ε)>0m(\varepsilon)>0 so that, for all tn>m(ε)t_{n}>m(\varepsilon),

d(1tnxtn,1tnyn)\displaystyle d_{\infty}\left(\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{t_{n}},\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}y_{n}\right) d(1tnxtn,g)+d(1tnyn,h)+d(g,h)\displaystyle\leq d_{\infty}\left(\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}x_{t_{n}},\mathlcal{g}\right)+d_{\infty}\left(\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}y_{n},\mathlcal{h}\right)+d_{\infty}(\mathlcal{g},\mathlcal{h})
3ε.\displaystyle\leq 3\varepsilon.

and dS(xtn,yn)<7ε=:ηεd_{S}(x_{t_{n}},y_{n})<7\varepsilon=:\eta_{\varepsilon}.

Let M1(g,ηε)>0M_{1}(\mathlcal{g},\eta_{\varepsilon})>0 be given by Lemma 3.3 on Θg\Uptheta{g}\in\operatorname{\mathscr{F}} with (Θg)=1\operatorname{\mathbbm{P}}(\Uptheta_{\mathlcal{g}})=1 satisfying, for all tn>M1(g,ηε)t_{n}>M_{1}(\mathlcal{g},\eta_{\varepsilon}) and unBS(xtn,tnηε)u_{n}\in B_{S}(x_{t_{n}},t_{n}\eta_{\varepsilon}),

|c(xtn)c(un)|<14ynSβε.|c(x_{t_{n}})-c(u_{n})|<14\|y_{n}\|_{S}\upbeta\varepsilon.

Fix, for M(g,ε)M(\mathlcal{g},\varepsilon) given by Lemma 3.6,

M^(ε):=maxgΔS1,ε{M(g,ε),M1(g,ηε)},\widehat{M}(\varepsilon):=\max_{\mathlcal{g}\in\Delta_{\mathlcal{S}_{1},\varepsilon}}\{M(\mathlcal{g},\varepsilon),M_{1}(\mathlcal{g},\eta_{\varepsilon})\},

which is finite on ΘΔ:=gΔ(ΩΔΘg)\Uptheta_{\Delta}:=\bigcap_{\mathlcal{g}\in\Delta}(\operatorname{\Upomega}_{\Delta}\cap\ \Uptheta_{\mathlcal{g}}) with (ΘΔ)=1\operatorname{\mathbbm{P}}(\Uptheta_{\Delta})=1.

Set m^(ε)>m(ε)\hat{m}(\varepsilon)>m(\varepsilon) to be such that |Φ(1tnyn)Φ(h)|<ε\left|\Phi(\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}y_{n})-\Phi(\mathlcal{h})\right|<\varepsilon for all n>m^(ε)n>\hat{m}(\varepsilon). Hence, for all tn>M^(ε)m^(ε)t_{n}>\widehat{M}(\varepsilon)\vee\hat{m}(\varepsilon) on ΘΔ\Uptheta_{\Delta},

|c(yn)Φ(1yn)|ynS\displaystyle\frac{|c(y_{n})-\Phi(1\operatorname{{\scriptscriptstyle\bullet}}y_{n})|}{\|y_{n}\|_{S}} 1tn|c(yn)c(xtn)|+|1tnc(xn)Φ(g)|\displaystyle\leq\frac{1}{t_{n}}|c(y_{n})-c(x_{t_{n}})|+\left|\frac{1}{t_{n}}c(x_{n})-\Phi(\mathlcal{g})\right|
+|Φ(g)Φ(h)|+|Φ(h)Φ(1tnyn)|\displaystyle\hskip 55.0pt+|\Phi(\mathlcal{g})-\Phi(\mathlcal{h})|+\left|\Phi(\mathlcal{h})-\Phi\left(\frac{1}{t_{n}}\operatorname{{\scriptscriptstyle\bullet}}y_{n}\right)\right|
(14β+3)ε,\displaystyle\leq(14\upbeta+3)\varepsilon,

which contradicts the above assumption proving that (3.13) holds true.

It remains to show how 1ndω\frac{1}{n}d_{\omega} converges to dϕd_{\phi} in the asymptotic cone. Recall that dω(x,y)=(c(yx1)ϑx)(ω)d_{\omega}(x,y)=\big{(}c(yx^{-1})\circ\operatorname{\upvartheta}_{x}\big{)}(\omega). Consider now any given h,hG\mathlcal{h},\mathlcal{h}^{\prime}\in G_{\infty} and {un}n\{u_{n}\}_{n\in\operatorname{\mathbb{N}}} a sequence with {tn}n\{t_{n}^{\prime}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\mathbb{N}} such that tn+t_{n}^{\prime}\uparrow+\infty and 1tnunhh1\frac{1}{t_{n}^{\prime}}\operatorname{{\scriptscriptstyle\bullet}}u_{n}\to\mathlcal{h^{\prime}h}^{-1}. Then unS/tn\|u_{n}\|_{S}/t_{n}^{\prime} converges to d(h,h)d_{\infty}(\mathlcal{h},\mathlcal{h}^{\prime}) and 1unSun\frac{1}{\|u_{n}\|_{S}}\operatorname{{\scriptscriptstyle\bullet}}u_{n} converges as above. In particular, one can fix any r>d(h,h)r^{\prime}>d_{\infty}(\mathlcal{h},\mathlcal{h}^{\prime}) to find 𝚔r>0\mathtt{k}_{r^{\prime}}>0 such that unS/tn<r\|u_{n}\|_{S}/t_{n}^{\prime}<r^{\prime} for all tn>𝚔rt_{n}^{\prime}>\mathtt{k}_{r^{\prime}}.

Let us define 𝙺r=(14β+3)r\mathtt{K}_{r^{\prime}}=(14\upbeta+3)r^{\prime} and mr(ε)=m^(ε)𝚔rm_{r^{\prime}}(\varepsilon)=\hat{m}(\varepsilon)\vee\mathtt{k}_{r^{\prime}}. The asymptotic equivalence (3.13) implies the existence of a random M^(ε)>0\widehat{M}(\varepsilon)>0 for ε(0,114β+3)\varepsilon\in(0,\frac{1}{14\upbeta+3}) such that, for all tn>M^(ε)mr(ε)t_{n}^{\prime}>\widehat{M}(\varepsilon)\vee m_{r^{\prime}}(\varepsilon) on ΘΔ\Uptheta_{\Delta},

|1tnc(un)Φ(hh1)|<𝙺rε.\left|\frac{1}{t_{n}^{\prime}}c(u_{n})-\Phi\big{(}\mathlcal{h^{\prime}h}^{-1}\big{)}\right|<\mathtt{K}_{r^{\prime}}\varepsilon.

Due to the fact that ϑ\operatorname{\upvartheta} is a p.m.p. group action, one can repeat all arguments above also in Propositions 3.5 and 3.6 to obtain M^(ε,σ(g))\widehat{M}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)} and ΘΔ(σ(g))\Uptheta_{\Delta}\big{(}\sigma(\mathlcal{g})\big{)} for each σ(g)={xn}n\sigma(\mathlcal{g})=\{x_{n}\}_{n\in\operatorname{\mathbb{N}}} with gG\mathlcal{g}\in G_{\infty} and (ΘΔ(g))=1\operatorname{\mathbbm{P}}\big{(}\Uptheta_{\Delta}(\mathlcal{g})\big{)}=1 so that, for all converging 1tnun\frac{1}{t_{n}^{\prime}}\operatorname{{\scriptscriptstyle\bullet}}u_{n} as above and every tn>M^(ε,σ(g))mr(ε)t_{n}^{\prime}>\widehat{M}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)}\vee m_{r^{\prime}}(\varepsilon) on ΘΔ(σ(g))\Uptheta_{\Delta}\big{(}\sigma(\mathlcal{g})\big{)},

|1tnc(un)ϑxtnΦ(hh1)|<𝙺rε.\left|\frac{1}{t_{n}^{\prime}}c(u_{n})\circ\operatorname{\upvartheta}_{x_{t_{n}^{\prime}}}-\Phi(\mathlcal{h^{\prime}h}^{-1})\right|<\mathtt{K}_{r^{\prime}}\varepsilon. (3.14)

Let now {vn}n\{v_{n}\}_{n\in\operatorname{\mathbb{N}}} be a sequence that 1nvnh\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}v_{n}\to\mathlcal{h} and choose rd(e,h)r\geq d_{\infty}(\mathlcal{e},\mathlcal{h}). Fix gΔr,ε\mathlcal{g}\in\Delta_{r,\varepsilon} so that gB(h,ε)\mathlcal{g}\in B_{\infty}(\mathlcal{h},\varepsilon). By Lemma 3.3, one can find a random M r,r(ε,σ(g))>0\accentset{\rule{2.79996pt}{0.5pt}}{M}_{r,r^{\prime}}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)}>0 and Ξσ(g)\Upxi_{\sigma(\mathlcal{g})} with (Ξσ(g))=1\operatorname{\mathbbm{P}}\big{(}\Upxi_{\sigma(\mathlcal{g})}\big{)}=1 such that, for all n>M r,r(ε,σ(g))n>\accentset{\rule{2.79996pt}{0.5pt}}{M}_{r,r^{\prime}}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)} on Ξσ(g)\Upxi_{\sigma(\mathlcal{g})},

|c(wn)ϑxnc(wn)ϑvn|<2βnε,|c(w_{n})\circ\operatorname{\upvartheta}_{x_{n}}-c(w_{n})\circ\operatorname{\upvartheta}_{v_{n}}|<2\upbeta n\varepsilon, (3.15)

where {wn}n\{w_{n}\}_{n\in\operatorname{\mathbb{N}}} is any convergent sequence 1nwnwB(e,r)\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}w_{n}\to\mathlcal{w}\in B_{\infty}(\mathlcal{e},r^{\prime}). Let us fix

ΞΔ:=gΔ(Ξσ(g)ΘΔ(σ(g))),\Upxi_{\Delta}:=\bigcap_{\mathlcal{g}\in\Delta}\left(\Upxi_{\sigma(\mathlcal{g})}\cap\Uptheta_{\Delta}\big{(}\sigma(\mathlcal{g})\big{)}\right),

and set

Mr,r(ε):=maxgΔr,ε{M r,r(ε,σ(g)),M^(ε,σ(g))}.M_{r,r^{\prime}}(\varepsilon):=\max_{\mathlcal{g}\in\Delta_{r,\varepsilon}}\left\{\accentset{\rule{2.79996pt}{0.5pt}}{M}_{r,r^{\prime}}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)},\widehat{M}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)}\right\}.

Then Mr,r(ε)M_{r,r^{\prime}}(\varepsilon) is finite on ΞΔ\Upxi_{\Delta} and (ΞΔ)=1\operatorname{\mathbbm{P}}\big{(}\Upxi_{\Delta}\big{)}=1. It follows from (3.14) and (3.15) that, for all tn>Mr,r(ε)mr(ε)t_{n}^{\prime}>M_{r,r^{\prime}}(\varepsilon)\vee m_{r^{\prime}}(\varepsilon) on ΞΔ\Upxi_{\Delta},

|1tnc(un)ϑvtnΦ(hh1)|<(𝙺r+2β)ε.\left|\frac{1}{t_{n}^{\prime}}c(u_{n})\circ\operatorname{\upvartheta}_{v_{t_{n}^{\prime}}}-\Phi(\mathlcal{h^{\prime}h}^{-1})\right|<(\mathtt{K}_{r^{\prime}}+2\upbeta)\varepsilon.

This establishes the \operatorname{\mathbbm{P}}-a.s. convergence of 1tndω(vtn,unvtn)\frac{1}{t_{n}^{\prime}}d_{\omega}(v_{t_{n}^{\prime}},u_{n}v_{t_{n}^{\prime}}) to Φ(hh1)=dϕ(h,h)\Phi\big{(}\mathlcal{h^{\prime}h}^{-1}\big{)}=d_{\phi}(\mathlcal{h},\mathlcal{h}^{\prime}) for ωΞΔ\omega\in\Upxi_{\Delta} as n+n\uparrow+\infty. Observe that the bi-Lipschitz equivalence is a straightforward consequence of Lemma 2.5, and this completes the proof. ∎

3.3. Proof of the second theorem

With the first main theorem now established, we have determined the asymptotic shape for finitely generated torsion-free nilpotent groups. The objective of this subsection is to extend this result to a finitely generated virtually nilpotent group Γ\operatorname{\Gamma}.

Recall that the nilpotent subgroup NΓN\unlhd\operatorname{\Gamma} has a finite index κ=[Γ:N]\kappa=[\operatorname{\Gamma}:N], and for each coset N(j)=z(j)NΓ/NN_{(j)}=z_{(j)}N\in\operatorname{\Gamma}/N, we designate a representative z(j)N(j)z_{(j)}\in N_{(j)}. Also, define πN(x)=z(j)1x\uppi_{N}(x)=z_{(j)}^{-1}x for all xN(j)x\in N_{(j)} and j{1,,κ}j\in\{1,\dots,\kappa\}.

We commence by presenting results concerning the properties of p.m.p. ergodic group actions of Γ\operatorname{\Gamma} with respect to NN and Γ\operatorname{\Gamma}^{\prime}. We adopt the notation 𝒜:=A𝒜A\cup\mathcal{A}:=\bigcup_{A\in\mathcal{A}}A.

Lemma 3.7.

Let Γ\operatorname{\Gamma} be a discrete group Γ\operatorname{\Gamma} and NΓN\unlhd\operatorname{\Gamma} a finite normal subgroup with finite index [Γ:N]=κ[\operatorname{\Gamma}:N]=\kappa. Consider that ϑ:Γ(Ω,,)\operatorname{\upvartheta}:\operatorname{\Gamma}\curvearrowright(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}) is a p.m.p. ergodic group action. Then there exists a finite 𝔅N\mathfrak{B}_{N}\subseteq\operatorname{\mathscr{F}} such that, for all B𝔅NB\in\mathfrak{B}_{N}, (B)1/κ\operatorname{\mathbbm{P}}(B)\geq 1/\kappa and ϑ|N,\operatorname{\upvartheta}\big{|}_{N}, the restriction of ϑ\operatorname{\upvartheta} on NN, induces a p.m.p. ergodic group action on (B,B,(B))\big{(}B,\operatorname{\mathscr{F}}_{\cap B},\operatorname{\mathbbm{P}}(~{}\cdot\mid B)\big{)}. Furthermore, |𝔅N|κ|\mathfrak{B}_{N}|\leq\kappa and (𝔅N)=1\operatorname{\mathbbm{P}}(\cup\mathfrak{B}_{N})=1.

Proof.

Set 𝔄N\mathfrak{A}_{N}\subseteq\operatorname{\mathscr{F}} to be the family of all non-empty NN-invariant events under ϑ\operatorname{\upvartheta}. Then, for all A𝔄NA\in\mathfrak{A}_{N},

(j=1κz(j)A)=1\operatorname{\mathbbm{P}}\left(\bigcup_{j=1}^{\kappa}z_{(j)}\cdot A\right)=1

which implies (A)1/κ\operatorname{\mathbbm{P}}(A)\geq 1/\kappa. Observe that 𝔄N\mathfrak{A}_{N} is closed under countable unions and non-empty countable intersections. Let us fix A0𝔄NA_{0}\in\mathfrak{A}_{N} such that (A0)=infA𝔄(A)\operatorname{\mathbbm{P}}(A_{0})=\inf_{A\in\mathfrak{A}}\operatorname{\mathbbm{P}}(A). Define 𝔅N={z(j)A0}j=1κ\mathfrak{B}_{N}=\{z_{(j)}\cdot A_{0}\}_{j=1}^{\kappa}.

Since NN is a normal subgroup of Γ\operatorname{\Gamma}, NN acts ergodically on (B,B,(B))\big{(}B,\operatorname{\mathscr{F}}_{\cap B},\operatorname{\mathbbm{P}}(~{}\cdot\mid B)\big{)} for all B𝔅NB\in\mathfrak{B}_{N} and it inherits the measure preserving property. ∎

We use Lemma 3.7 to write (B,B,B)(B,\operatorname{\mathscr{F}}_{B},\operatorname{\mathbbm{P}}_{B}) with B:=B={EB:E}\operatorname{\mathscr{F}}_{B}:=\operatorname{\mathscr{F}}_{\cap B}=\{E\cap B:E\in\operatorname{\mathscr{F}}\} and B(E):=(EB)\operatorname{\mathbbm{P}}_{B}(E):=\operatorname{\mathbbm{P}}(E\mid B) for each B𝔅NB\in\mathfrak{B}_{N}. Let us denote by [ω]=torNω[\omega]=\operatorname{\operatorname{tor}}N\cdot\omega, the orbit of ωB\omega\in B under the action on torN\operatorname{\operatorname{tor}}N. Set

([B],B,B):=(B,B,B)/torN\big{(}[B],\operatorname{\mathscr{F}}_{B}^{\prime},\operatorname{\mathbbm{P}}_{B}^{\prime}\big{)}:=(B,\operatorname{\mathscr{F}}_{B},\operatorname{\mathbbm{P}}_{B})/\operatorname{\operatorname{tor}}N

where B={[E]:EB}\operatorname{\mathscr{F}}_{B}^{\prime}=\big{\{}[E]:E\in\operatorname{\mathscr{F}}_{B}\big{\}} and B([E])\operatorname{\mathbbm{P}}_{B}^{\prime}\big{(}[E]\big{)} is the induced probability measure (torN)B(E)=B([E])(\operatorname{\operatorname{tor}}N)_{\ast}\operatorname{\mathbbm{P}}_{B}(E)=\operatorname{\mathbbm{P}}_{B}\left(\cup[E]\right). Let us fix υx=υxx\upupsilon_{x}=\upupsilon_{\llbracket x\rrbracket}\in\llbracket x\rrbracket for each xΓ\llbracket x\rrbracket\in\operatorname{\Gamma}^{\prime}. Define θ:Γ([B],B,B)\theta:\operatorname{\Gamma}^{\prime}\curvearrowright\big{(}[B],\operatorname{\mathscr{F}}_{B}^{\prime},\operatorname{\mathbbm{P}}_{B}^{\prime}\big{)} so that

θx([ω])=[ϑυx(ω)].\theta_{\llbracket x\rrbracket}\big{(}[\omega]\big{)}=\big{[}\operatorname{\upvartheta}_{\upupsilon_{x}}(\omega)\big{]}.
Lemma 3.8.

Let 𝔅N\mathfrak{B}_{N} be the set obtained in Lemma 3.7. Then, for each B𝔅NB\in\mathfrak{B}_{N}, θ:Γ([B],B,B)\theta:\operatorname{\Gamma}^{\prime}\curvearrowright\big{(}[B],\operatorname{\mathscr{F}}_{B}^{\prime},\operatorname{\mathbbm{P}}_{B}^{\prime}\big{)} is a p.m.p. ergodic group action.

Proof.

The measure preserving property is immediately inherited from ϑ\operatorname{\upvartheta}. Let ϑv(ω)=vω\operatorname{\upvartheta}_{v}(\omega)=v\cdot\omega. Due to the normality of torNN\operatorname{\operatorname{tor}}N\unlhd N, for all ABA\in\operatorname{\mathscr{F}}_{B} and each vv.torNv^{\prime}\in v.\operatorname{\operatorname{tor}}N,

[vA]=v([A]).\cup[v\cdot A]=v^{\prime}\cdot\left(\cup[A]\right).

Hence, if for all v.torNΓv.\operatorname{\operatorname{tor}}N\in\operatorname{\Gamma}^{\prime}, one has [vA]=[A][v\cdot A]=[A]. Then, for all xNx\in N

x([A])=[A].x\cdot\left(\cup[A]\right)=\cup[A].

It follows from the ergodicity of ϑ:N(B,B,B)\operatorname{\upvartheta}:N\curvearrowright(B,\operatorname{\mathscr{F}}_{B},\operatorname{\mathbbm{P}}_{B}) that B([A]){0,1}\operatorname{\mathbbm{P}}_{B}^{\prime}\big{(}[A]\big{)}\in\{0,1\}, which is the desired conclusion. ∎

Remark 5.

Recall that definition (1.4) determines

c(x):=maxyxztorNc(y)ϑz.c^{\prime}\big{(}\llbracket x\rrbracket\big{)}:=\max_{\begin{subarray}{c}y\in\llbracket x\rrbracket\\ z\in\operatorname{\operatorname{tor}}N\end{subarray}}c(y)\circ\operatorname{\upvartheta}_{z}.

It is straightforward to see that cc^{\prime} is compatible with the probability space ([B],b,B)\big{(}[B],\operatorname{\mathscr{F}}_{b}^{\prime},\operatorname{\mathbbm{P}}_{B}^{\prime}\big{)} for each B𝔅NB\in\mathfrak{B}_{N}. Futhermore, it is a subadditive cocycle associated with θ\theta. Additionally, cc^{\prime} is well defined on (B,B,B)(B,\operatorname{\mathscr{F}}_{B},\operatorname{\mathbbm{P}}_{B}). Let Ω:=𝔅N\operatorname{\Upomega}^{\prime}:=\cup\mathfrak{B}_{N} and (Ω)=1\operatorname{\mathbbm{P}}(\operatorname{\Upomega}^{\prime})=1. Consequently, one can investigate cc^{\prime} on ([B],B,B)([B],\operatorname{\mathscr{F}}_{B}^{\prime},\operatorname{\mathbbm{P}}_{B}^{\prime}), and the results can be naturally extended \operatorname{\mathbbm{P}}-a.s. to (Ω,,)(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}).

In preparation for the asymptotic comparison between cocycles cc and cc^{\prime}, the following lemmas provide essential insights into their respective properties and relationships.

Lemma 3.9.

Let ε,r>0\varepsilon,r>0 and consider a subadditive cocycle cc that satisfies condition (i). Then there exists, \operatorname{\mathbbm{P}}-a.s., MN=MN(ε,r)>0M_{N}=M_{N}(\varepsilon,r)>0 such that, for all n>MNn>M_{N} and every xBS(e,rn)x\in B_{S}(e,rn),

|c(x)c(πN(x))|<εn.\big{|}c(x)-c\big{(}\uppi_{N}(x)\big{)}\big{|}<\varepsilon n.
Proof.

It follows from subadditivity that, for xN(j)x\in N_{(j)},

|c(x)c(z(j)1x)|max{c(z(j))ϑz(j)1x,c(z(j)1)ϑx}-a.s.|c(x)-c(z_{(j)}^{-1}x)|\leq\max\left\{c(z_{(j)})\circ\operatorname{\upvartheta}_{z_{(j)}^{-1}x},\;c(z_{(j)}^{-1})\circ\operatorname{\upvartheta}_{x}\right\}\quad\operatorname{\mathbbm{P}}\text{-a.s.}

for every j{1,,κ}j\in\{1,\dots,\kappa\}. Let 𝚖κ=max{z(j)S:1jκ}\mathtt{m}_{\kappa}=\max\left\{\|z_{(j)}\|_{S}:1\leq j\leq\kappa\right\}. Hence, one has by (i) and a 𝙲>0\mathtt{C}>0 that

(maxxBS(e,rn){|c(x)c(πN(x))|}εn)\displaystyle\operatorname{\mathbbm{P}}\left(\max_{x\in B_{S}(e,rn)}\big{\{}|c(x)-c(\uppi_{N}(x))|\big{\}}\geq\varepsilon n\right) |BS(e,rn)|j=1κ(c(z(j)±1)εn)\displaystyle\leq|B_{S}(e,rn)|\sum_{j=1}^{\kappa}\operatorname{\mathbbm{P}}\left(c(z_{(j)}^{\pm 1})\geq\varepsilon n\right)
𝙲rDnDg(nε)𝒪ε,r(1/nD+κ)\displaystyle\leq\mathtt{C}r^{D}n^{D}g(n\varepsilon)\in\mathcal{O}_{\varepsilon,r}\big{(}1/n^{D+\upkappa}\big{)}

for n>β𝚖κ/εn>\upbeta\mathtt{m}_{\kappa}/\varepsilon. The result is derived through the application of the Borel-Cantelli Lemma. ∎

Lemma 3.10.

Let ε,r>0\varepsilon,r>0 and consider a subadditive cocycle cc that satisfies condition (i). Then there exists, \operatorname{\mathbbm{P}}-a.s., Mq=Mq(ε,r)>0M_{q}=M_{q}(\varepsilon,r)>0 such that, for all n>Mqn>M_{q} and every xBS(e,rn)x\in B_{S}(e,rn),

|c(x1)ϑy1c(x2)ϑy2|<εn\left|c\left(x_{1}\right)\circ\operatorname{\upvartheta}_{y_{1}}-c(x_{2})\circ\operatorname{\upvartheta}_{y_{2}}\right|<\varepsilon n

where x1,x2xx_{1},x_{2}\in\llbracket x\rrbracket and y1,y2torNy_{1},y_{2}\in\operatorname{\operatorname{tor}}N.

Proof.

Since torN\operatorname{\operatorname{tor}}N is a normal subgroup of NN, the exists v2torNv_{2}\in\operatorname{\operatorname{tor}}N such that x1=v2x2y3x_{1}=v_{2}x_{2}y_{3} with y3=y2y11y_{3}=y_{2}y_{1}^{-1}. Thus

c(x1)ϑy1\displaystyle c(x_{1})\circ\operatorname{\upvartheta}_{y_{1}} c(y3)ϑy1+c(v2x2)ϑy2\displaystyle\leq c(y_{3})\circ\operatorname{\upvartheta}_{y_{1}}+c(v_{2}x_{2})\circ\operatorname{\upvartheta}_{y_{2}}
c(y3)ϑy1+c(x2)ϑy2+c(v2)ϑx2y2-a.s.\displaystyle\leq c(y_{3})\circ\operatorname{\upvartheta}_{y_{1}}+c(x_{2})\circ\operatorname{\upvartheta}_{y_{2}}+c(v_{2})\circ\operatorname{\upvartheta}_{x_{2}y_{2}}~{}~{}~{}\operatorname{\mathbbm{P}}\text{-a.s.}

We apply the same reasoning for c(x2)ϑy2c(x_{2})\circ\operatorname{\upvartheta}_{y_{2}} obtaining that

|c(x1)ϑy1c(x2)ϑy2|maxy,ztorN{c(y)ϑz}+maxy,ztorN{c(y)ϑx1z}-a.s.\left|c(x_{1})\circ\operatorname{\upvartheta}_{y_{1}}-c(x_{2})\circ\operatorname{\upvartheta}_{y_{2}}\right|\leq\max_{y,z\in\operatorname{\operatorname{tor}}N}\{c(y)\circ\operatorname{\upvartheta}_{z}\}+\max_{y,z\in\operatorname{\operatorname{tor}}N}\{c(y)\circ\operatorname{\upvartheta}_{x_{1}z}\}\quad\operatorname{\mathbbm{P}}\text{-a.s.}

By (i) and the finitness of torN\operatorname{\operatorname{tor}}N, there exists a constant C>0C^{\prime}>0 such that

(supxBS(e,rn)x1,x2xy1,y2torN|c(x1)ϑy1c(x2)ϑy2|εn)\displaystyle\mathbb{P}\left(\sup\limits_{\begin{subarray}{c}x\in B_{S}(e,rn)\\ x_{1},x_{2}\in\llbracket x\rrbracket\\ y_{1},y_{2}\in\operatorname{\operatorname{tor}}N\end{subarray}}\big{|}c(x_{1})\circ\operatorname{\upvartheta}_{y_{1}}-c(x_{2})\circ\operatorname{\upvartheta}_{y_{2}}\big{|}\geq\varepsilon n\right) \displaystyle\leq 2|torN|4|BS(e,rn)|2g(εn)\displaystyle 2|\operatorname{\operatorname{tor}}N|^{4}{|B_{S}(e,rn)|^{2}}g(\varepsilon n)
\displaystyle\leq C(rn)2Dg(εn)𝒪ε,r(1/nκ)\displaystyle C^{\prime}{(rn)^{2D}}g(\varepsilon n)\in\mathcal{O}_{\varepsilon,r}(1/n^{\upkappa})

for n>βmax{zS:ztorN}/εn>\upbeta\max\{\|z\|_{S}:z\in\operatorname{\operatorname{tor}}N\}/\varepsilon. The desired conclusion follows from an application of Borel-Cantelli Lemma. ∎

Let us define, for all xΓ\llbracket x\rrbracket\in\operatorname{\Gamma}^{\prime},

|x|Sinf:=min1i,jκminy(z(j).x.z(i)1)yS,\big{|}\llbracket x\rrbracket\big{|}_{S}^{\inf}:=\min_{1\leq i,j\leq\kappa}~{}\min_{y\in(z_{(j)}.\llbracket x\rrbracket.z_{(i)}^{-1})}\|y\|_{S},

and

|x|Ssup:=max1i,jκmaxy(z(j).x.z(i)1)yS.\big{|}\llbracket x\rrbracket\big{|}_{S}^{\sup}:=\max_{1\leq i,j\leq\kappa}~{}\max_{y\in(z_{(j)}.\llbracket x\rrbracket.z_{(i)}^{-1})}\|y\|_{S}.

Set

𝚖κ,q:=max1i,jκmaxz(z(j).e.z(i)1)zS.\mathtt{m}_{\kappa,q}:=\max_{1\leq i,j\leq\kappa}~{}\max_{z\in(z_{(j)}.\llbracket e\rrbracket.z_{(i)}^{-1})}\|z\|_{S}.

Thus, one has, for all yz(j).xy\in z_{(j)}.\llbracket x\rrbracket with j{1,,κ}j\in\{1,\dots,\kappa\},

|x|SinfyS|x|Ssup|x|Sinf+2𝚖κ,q.\big{|}\llbracket x\rrbracket\big{|}_{S}^{\inf}\leq\|y\|_{S}\leq\big{|}\llbracket x\rrbracket\big{|}_{S}^{\sup}\leq\big{|}\llbracket x\rrbracket\big{|}_{S}^{\inf}+2\cdot\mathtt{m}_{\kappa,q}. (3.16)

By the same arguments employed in Section 2.4, the discrete norm

|x|Sab:=infy(x.[Γ,Γ])|y|Sinf\big{|}\llbracket x\rrbracket\big{|}_{S}^{\operatorname{\operatorname{ab}}}:=\inf_{\llbracket y\rrbracket\in\big{(}\llbracket x\rrbracket.[\operatorname{\Gamma}^{\prime},\operatorname{\Gamma}^{\prime}]\big{)}}\big{|}\llbracket y\rrbracket\big{|}_{S}^{\inf} (3.17)

exhibits the same properties as Sab\|-\|_{S}^{\operatorname{\operatorname{ab}}} when S\llbracket S\rrbracket is a generating set of Γ\operatorname{\Gamma}^{\prime}.

Consider σ(g)={xn}nΓ\sigma(\mathlcal{g})=\big{\{}\llbracket x\rrbracket_{n}\big{\}}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\Gamma}^{\prime} to be the sequences fixed for each gG\mathlcal{g}\in G_{\infty} in the proof of Theorem 1.1. Set xn:=υxnx_{n}:=\upupsilon_{\llbracket x\rrbracket_{n}} with υ\upupsilon defined by the group action θ\theta. Then

xn=xn\llbracket x_{n}\rrbracket=\llbracket x\rrbracket_{n}

when σ(g)\sigma(\mathlcal{g}) is given. Let us write υσ(g)={xn}n\upupsilon_{\sigma}(\mathlcal{g})=\{x_{n}\}_{n\in\operatorname{\mathbb{N}}} for each σ(g)={xn}nΓ\sigma(\mathlcal{g})=\big{\{}\llbracket x\rrbracket_{n}\big{\}}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\Gamma}^{\prime}. Also, one can easily verify that

limn+xnSn=limn+|xn|Sinfn=limn+|xn|Ssupn=d(e,g).\lim_{n\uparrow+\infty}\frac{\|x_{n}\|_{S}}{n}=\lim_{n\uparrow+\infty}\frac{\big{|}\llbracket x_{n}\rrbracket\big{|}_{S}^{\inf}}{n}=\lim_{n\uparrow+\infty}\frac{\big{|}\llbracket x_{n}\rrbracket\big{|}_{S}^{\sup}}{n}=d_{\infty}(\mathlcal{e},\mathlcal{g}).

The proposition below shows us that cc and cc^{\prime} share the same linear asymptotic behaviour.

Proposition 3.11.

Let Γ\operatorname{\Gamma} be a virtually nilpotent group, and let c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} be a subadditive cocycle associated with ϑ\operatorname{\upvartheta}.

If condition (i) is satisfied, then cc and cc^{\prime} are asymptotically equivalent, i.e., there exists, \operatorname{\mathbbm{P}}-a.s., M(ε)>0M^{\prime}(\varepsilon)>0 such that, for all xΓx\in\operatorname{\Gamma} with xS>M(ε)\|x\|_{S}>M^{\prime}(\varepsilon),

|c(x)c(x)|<εxS.\big{|}c(x)-c^{\prime}(\llbracket x\rrbracket)\big{|}<\varepsilon\|x\|_{S}. (3.18)

In particular, (i) implies the \operatorname{\mathbbm{P}}-a.s. existence of M(ε,r,υσ(g))>0M^{\prime}\big{(}\varepsilon,r,\upupsilon_{\sigma}(\mathlcal{g})\big{)}>0 so that, for all n>M(ε,r,υσ(g))n>M^{\prime}\big{(}\varepsilon,r,\upupsilon_{\sigma}(\mathlcal{g})\big{)} and every yBS(e,rn)y\in B_{S}(e,rn),

|c(y)c(y)|ϑxn<nε.\big{|}c(y)-c^{\prime}(\llbracket y\rrbracket)\big{|}\circ\operatorname{\upvartheta}_{x_{n}}<n\varepsilon. (3.19)
Proof.

From Lemmas 3.9 and 3.10, we can deduce that, for every ε>0\varepsilon>0, one can fix M(ε)=MN(ε2,1)Mq(ε2,1)M^{\prime}(\varepsilon)=M_{N}(\frac{\varepsilon}{2},1)\vee M_{q}(\frac{\varepsilon}{2},1) so that, \operatorname{\mathbbm{P}}-a.s., for all n>M(ε)n>M^{\prime}(\varepsilon) and every xBS(e,n+1)BS(e,n)x\in B_{S}(e,n+1)\setminus B_{S}(e,n),

|c(x)c(x)|xS<|c(x)c(π(x))|n+|c(π(x))c(x)|n<ε.\frac{|c(x)-c^{\prime}(x)|}{\|x\|_{S}}<\frac{\left|c(x)-c\big{(}\uppi(x)\big{)}\right|}{n}+\frac{\left|c\big{(}\uppi(x)\big{)}-c^{\prime}(x)\right|}{n}<\varepsilon.

The inequality above implies the asymptotic equivalence of cc and cc^{\prime} on Γ\operatorname{\Gamma}.

Since ϑ\operatorname{\upvartheta} is p.m.p. group action, one can obtain from Lemmas 3.9 and 3.10 the random variables MN>0M_{N}>0 and Mq>0M_{q}>0 depending on υσ(g))>0\upupsilon_{\sigma}(\mathlcal{g})\big{)}>0 determining

M(ε,r,υσ(g))=MN(ε/2,r,υσ(g))Mq(ε/2,r,υσ(g))M^{\prime}\big{(}\varepsilon,r,\upupsilon_{\sigma}(\mathlcal{g})\big{)}=M_{N}\big{(}{\varepsilon}/{2},r,\upupsilon_{\sigma}(\mathlcal{g})\big{)}\vee M_{q}\big{(}{\varepsilon}/{2},r,\upupsilon_{\sigma}(\mathlcal{g})\big{)}

so that (3.19) holds true. ∎

The following result extends the subadditive ergodic theorem to cc^{\prime} with respect to ||Sab|-|_{S}^{\operatorname{\operatorname{ab}}}.

Lemma 3.12.

Consider Γ\operatorname{\Gamma} to be a virtually nilpotent group generated by a finite symmetric set SΓS\subseteq\operatorname{\Gamma} with S\llbracket S\rrbracket a generating set of Γ\operatorname{\Gamma}^{\prime}.

If the subadditive cocycle cc satisfies (i) and (ii) with respect to the word norm S\|-\|_{S}, then cc^{\prime} satisfies (i) and (ii) with respect to ||Sinf|-|_{S}^{\inf}. In particular, Lemma 2.5 is still valid with xab=xabx^{\operatorname{\operatorname{ab}}}=\llbracket x\rrbracket^{\operatorname{\operatorname{ab}}} and

ϕ(xab)=infn𝔼[c(xn)]n.\phi(x^{\operatorname{\operatorname{ab}}})=\inf_{n\in\operatorname{\mathbb{N}}}\frac{\operatorname{\mathbbm{E}}[c^{\prime}(\llbracket x\rrbracket^{n})]}{n}.
Proof.

First, observe that (i) and (ii) imply, for all xΓx\in\operatorname{\Gamma} and

(c(x)t)κ|torN|g(t),for all t>β|x|Ssup,\operatorname{\mathbbm{P}}\Big{(}c^{\prime}\big{(}\llbracket x\rrbracket\big{)}\geq t\Big{)}\leq\kappa~{}|\operatorname{\operatorname{tor}}N|~{}g(t),\quad\text{for all }t>\upbeta\big{|}\llbracket x\rrbracket\big{|}_{S}^{\sup},

and

𝔼[c(x)]a|x|Sinf.\operatorname{\mathbbm{E}}\Big{[}c^{\prime}\big{(}\llbracket x\rrbracket\big{)}\Big{]}\geq a\big{|}\llbracket x\rrbracket\big{|}_{S}^{\inf}.

Therefore, it follows from (3.16) that cc^{\prime} satisfy (i) and (ii) with respect to ||Sinf|-|_{S}^{\inf} for a new g(t)𝒪(t2D+κ)g^{\prime}(t)\in\mathcal{O}\big{(}t^{2D+\upkappa}\big{)} and β>0\upbeta^{\prime}>0. The proof is complete by replacing S\|-\|_{S} with ||Sinf|-|_{S}^{\inf} and applying (3.17) in the proof of Lemma 2.5.

Having established the aforementioned results, we now move forward to prove the second theorem.

Proof of Theorem 1.2.

Observe that it follows from Lemmas 3.7, 3.8, 5 and 3.12 that, for each B𝔅NB\in\mathfrak{B}_{N}, Theorem 1.1 holds true for cc^{\prime} on (B,B,B)(B,\operatorname{\mathscr{F}}_{B},\operatorname{\mathbbm{P}}_{B}). Therefore, it suffices to extend the results to (Ω,,)(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}) and compare cc with cc^{\prime}.

The asymptotic equivalence is an immediate consequence of (3.13) and (3.18), we focus on the second part of the proof of Theorem 1.1. Recall de definition of Δ\Delta as a dense subset of GG_{\infty}, the finite Δr,ε\Delta_{r,\varepsilon}. Similarly, we consider {un}nΓ\{u_{n}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\Gamma} and {tn}n\{t_{n}^{\prime}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\mathbb{N}} with tn+t_{n}\uparrow+\infty as n+n\uparrow+\infty and 1tnunhh1\frac{1}{t_{n}^{\prime}}\operatorname{{\scriptscriptstyle\bullet}}u_{n}\to\mathlcal{h}^{\prime}\mathlcal{h}^{-1}. Note that we may regard un=un\llbracket u\rrbracket_{n}=\llbracket u_{n}\rrbracket to replace the orifinal sequence in the proof of Thm. 1.1 and let 𝙺r\mathtt{K}_{r^{\prime}} and mr(ε)m_{r^{\prime}}(\varepsilon) be defined as before with r>d(h,h)r^{\prime}>d_{\infty}(\mathlcal{h},\mathlcal{h}^{\prime}).

Set M^(ε,σ(g),B)\widehat{M}\big{(}\varepsilon,\sigma(\mathlcal{g}),B\big{)} and ΘΔ(σ(g),B)\Uptheta_{\Delta}\big{(}\sigma(\mathlcal{g}),B\big{)} to be defined by (3.14) for each B𝔅NB\in\mathfrak{B}_{N} with (ΘΔ(σ(g)),B)B)=1\operatorname{\mathbbm{P}}\left(\Uptheta_{\Delta}\big{(}\sigma(\mathlcal{g})\big{)},B\big{)}\mid B\right)=1 so that, for all tn>M^(ε,σ(g),B)mr(ε)t_{n}^{\prime}>\widehat{M}\big{(}\varepsilon,\sigma(\mathlcal{g}),B\big{)}\vee m_{r^{\prime}}(\varepsilon),

|1tnc(un)ϑxtnΦ(hh1)|<𝙺rε.\left|\frac{1}{t_{n}^{\prime}}c^{\prime}\big{(}\llbracket u_{n}\rrbracket\big{)}\circ\operatorname{\upvartheta}_{x_{t_{n}}}-\Phi(\mathlcal{h}^{\prime}\mathlcal{h}^{-1})\right|<\mathtt{K}_{r^{\prime}}\varepsilon. (3.20)

on ΘΔ(σ(g)),B)\Uptheta_{\Delta}\big{(}\sigma(\mathlcal{g})\big{)},B\big{)} with υσ(g)={xn}n\upupsilon_{\sigma}(\mathlcal{g})=\{x_{n}\}_{n\in\operatorname{\mathbb{N}}}. Fix

M^(ε,σ(g)):=B𝔅NM^(ε,σ(g),B)𝟙B+𝟙Ω(𝔅N).\widehat{M}^{\prime}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)}:=\sum_{B\in\mathfrak{B}_{N}}\widehat{M}\big{(}\varepsilon,\sigma(\mathlcal{g}),B\big{)}\mathbbm{1}_{B}+\mathbbm{1}_{\operatorname{\Upomega}\setminus(\cup\mathfrak{B}_{N})}.

Consider {yn}n\{y_{n}\}_{n\in\operatorname{\mathbb{N}}} with ynS/n<r\|y_{n}\|_{S}/n<r^{\prime} for every n>mr(ε)n>m_{r^{\prime}}(\varepsilon). Then Proposition 3.11 ensures the existence of M(ε,r,υσ(g))>0M^{\prime}\big{(}\varepsilon,r^{\prime},\upupsilon_{\sigma}(\mathlcal{g})\big{)}>0 and Λσ(g)\Uplambda_{\sigma(\mathlcal{g})}\in\operatorname{\mathscr{F}} with (Λσ(g))=1\operatorname{\mathbbm{P}}(\Uplambda_{\sigma(\mathlcal{g})})=1 so that, for all n>M(ε,r,υσ(g))n>M^{\prime}\big{(}\varepsilon,r^{\prime},\upupsilon_{\sigma}(\mathlcal{g})\big{)} on Λσ(g)\Uplambda_{\sigma(\mathlcal{g})},

1n|c(yn)c(yn)|ϑxn<ε.\frac{1}{n}\left|c(y_{n})-c^{\prime}\big{(}\llbracket y_{n}\rrbracket\big{)}\right|\circ\operatorname{\upvartheta}_{x_{n}}<\varepsilon. (3.21)

Let now {vn}nΓ\{v_{n}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\Gamma} be a sequence such that 1nvnh\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}v_{n}\to\mathlcal{h} and choose rd(e,h)r\geq d_{\infty}(\mathlcal{e},\mathlcal{h}). Fix gΔr,ε\mathlcal{g}\in\Delta_{r,\varepsilon} so that gB(h,ε)\mathlcal{g}\in B_{\infty}(\mathlcal{h},\varepsilon). Observe that (3.15) is still valid for cc. Hence, by Lemma 3.3, one can find M r,r(ε,σ(g))>0\accentset{\rule{2.79996pt}{0.5pt}}{M}_{r,r^{\prime}}^{\prime}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)}>0 and Ξσ(g)\Upxi_{\sigma(\mathlcal{g})}^{\prime} with (Ξσ(g))=1\operatorname{\mathbbm{P}}\big{(}\Upxi_{\sigma(\mathlcal{g})}^{\prime}\big{)}=1 such that, for all n>M r,r(ε,σ(g))n>\accentset{\rule{2.79996pt}{0.5pt}}{M}_{r,r^{\prime}}^{\prime}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)} on Ξσ(g)\Upxi_{\sigma(\mathlcal{g})}^{\prime},

|c(wn)ϑxnc(wn)ϑvn|<2βnε,|c(w_{n})\circ\operatorname{\upvartheta}_{x_{n}}-c(w_{n})\circ\operatorname{\upvartheta}_{v_{n}}|<2\upbeta n\varepsilon, (3.22)

where {wn}nΓ\{w_{n}\}_{n\in\operatorname{\mathbb{N}}}\subseteq\operatorname{\Gamma} is any convergent sequence 1nwnwB(e,r)\frac{1}{n}\operatorname{{\scriptscriptstyle\bullet}}w_{n}\to\mathlcal{w}\in B_{\infty}(\mathlcal{e},r^{\prime}). Let us fix

ΛΔ:=gΔ(Λσ(g)Ξσ(g)(B𝔅NΘΔ(σ(g),B))),\Uplambda_{\Delta}:=\bigcap_{\mathlcal{g}\in\Delta}\left(\Uplambda_{\sigma(\mathlcal{g})}\cap\Upxi_{\sigma(\mathlcal{g})}\cap\left(\bigcup_{B\in\mathfrak{B}_{N}}\Uptheta_{\Delta}\big{(}\sigma(\mathlcal{g}),B\big{)}\right)\right),

and set

Mr,r(ε):=maxgΔr,ε{M(ε,r,υσ(g)),M r,r(ε,σ(g)),M^(ε,σ(g))}.M_{r,r^{\prime}}^{\prime}(\varepsilon):=\max_{\mathlcal{g}\in\Delta_{r,\varepsilon}}\left\{M^{\prime}\big{(}\varepsilon,r^{\prime},\upupsilon_{\sigma}(\mathlcal{g})\big{)},~{}\accentset{\rule{2.79996pt}{0.5pt}}{M}_{r,r^{\prime}}^{\prime}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)},~{}\widehat{M}^{\prime}\big{(}\varepsilon,\sigma(\mathlcal{g})\big{)}\right\}.

Then Mr,r(ε)M_{r,r^{\prime}}^{\prime}(\varepsilon) is finite on ΛΔ\Uplambda_{\Delta} and (ΛΔ)=1\operatorname{\mathbbm{P}}\big{(}\Uplambda_{\Delta}\big{)}=1. It follows from (3.20), (3.21), and (3.22) with un=wtn=ytnu_{n}=w_{t_{n}^{\prime}}=y_{t_{n}^{\prime}} that, for all tn>Mr,r(ε)mr(ε)t_{n}^{\prime}>M_{r,r^{\prime}}^{\prime}(\varepsilon)\vee m_{r^{\prime}}(\varepsilon) on ΛΔ\Uplambda_{\Delta},

|1tnc(un)ϑvtnΦ(hh1)|<(𝙺r+2β+1)ε.\left|\frac{1}{t_{n}^{\prime}}c(u_{n})\circ\operatorname{\upvartheta}_{v_{t_{n}^{\prime}}}-\Phi(\mathlcal{h^{\prime}h}^{-1})\right|<(\mathtt{K}_{r^{\prime}}+2\upbeta+1)\varepsilon.

This establishes the \operatorname{\mathbbm{P}}-a.s. convergence of 1tndω(vtn,unvtn)\frac{1}{t_{n}^{\prime}}d_{\omega}(v_{t_{n}^{\prime}},u_{n}v_{t_{n}^{\prime}}) to Φ(hh1)=dϕ(h,h)\Phi\big{(}\mathlcal{h^{\prime}h}^{-1}\big{)}=d_{\phi}(\mathlcal{h},\mathlcal{h}^{\prime}) for ωΛΔ\omega\in\Uplambda_{\Delta} as n+n\uparrow+\infty. ∎

3.4. An additional result for FPP models

In the preceding sections, we delved into the asymptotic behavior of cc and cc^{\prime}. The definition of cc^{\prime} depends only on the action of ϑ\operatorname{\upvartheta} restricted to NΓN\unlhd\operatorname{\Gamma}, ensuring that we can systematically investigate the group action of Γ\operatorname{\Gamma}^{\prime} within a fixed B𝔅NB\in\mathfrak{B}_{N}.

To broaden the scope of our findings and establish the validity of (iii) for FPP models on virtually nilpotent groups, we will introduce a new random variable induced by a graph homomorphism. Let us now define, for all xΓ{e}\llbracket x\rrbracket\in\operatorname{\Gamma}^{\prime}\setminus\{\llbracket e\rrbracket\},

c′′(x):=max1i,jκmaxyz(j).xzz(i).torNc(y)ϑzc^{\prime\prime}\big{(}\llbracket x\rrbracket\big{)}:=\max_{1\leq i,j\leq\kappa}\max_{\begin{subarray}{c}y\in z_{(j)}.\llbracket x\rrbracket\\ z\in z_{(i)}.\operatorname{\operatorname{tor}}N\end{subarray}}c(y)\circ\operatorname{\upvartheta}_{z}

and consider c′′(e):=0c^{\prime\prime}\big{(}\llbracket e\rrbracket\big{)}:=0. Note that c′′c^{\prime\prime} restricted to B𝔅NB\in\mathfrak{B}_{N} is not well-defined when there exists another set B𝔅NB^{\prime}\in\mathfrak{B}_{N} distinct from BB. This inherent limitation prompts the necessity for specific conditions in the subsequent result.

The following lemma outlines the criteria under which c′′c^{\prime\prime} inherits the FPP property from cc. Before presenting this result, we establish the notation:

S±:={s±1:sS}.\llbracket S\rrbracket^{\pm}:=\big{\{}\llbracket s\rrbracket^{\pm 1}\colon s\in S\big{\}}.
Lemma 3.13.

Let (Γ,.)(\operatorname{\Gamma},.) be a virtually nilpotent group generated by a finite symmetric set SΓS\subseteq\operatorname{\Gamma} with S\llbracket S\rrbracket a generating set of Γ\operatorname{\Gamma}^{\prime}. Consider a subadditive cocycle c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} determining a FPP model on 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) which satisfies (i). Suppose that the restriction ϑ|N:N(Ω,,)\operatorname{\upvartheta}\big{|}_{N}:N\curvearrowright(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}) is a p.m.p. ergodic group action.

If, for all sSs\in S, s1=s1\llbracket s^{-1}\rrbracket=\llbracket s\rrbracket^{-1}, then c′′c^{\prime\prime} determines a FPP model on 𝒞(Γ,S±)\mathcal{C}(\operatorname{\Gamma},\llbracket S\rrbracket^{\pm}) and condition (iii) is satisfied when S±Γ[Γ,Γ]\llbracket S\rrbracket^{\pm}\subseteq\operatorname{\Gamma}^{\prime}\setminus[\operatorname{\Gamma}^{\prime},\operatorname{\Gamma}^{\prime}].

Proof.

Define, for each xΓx\in\operatorname{\Gamma} and every sS±\llbracket s\rrbracket\in\llbracket S\rrbracket^{\pm},

τ(x,sx):=max1i,jκmaxyz(j).xhz(i).sτ(y,hy)\tau\big{(}\llbracket x\rrbracket,\llbracket s\rrbracket\llbracket x\rrbracket\big{)}:=\max_{1\leq i,j\leq\kappa}\max_{\begin{subarray}{c}~{}y\in z_{(j)}.\llbracket x\rrbracket~{}\\ h\in z_{(i)}.\llbracket s\rrbracket\end{subarray}}\tau(y,hy)

and note that τ\tau preserves the symmetry

τ(x,sx)=τ(sx,s1(sx))=τ(sx,x).\tau\big{(}\llbracket x\rrbracket,\llbracket s\rrbracket\llbracket x\rrbracket\big{)}=\tau\Big{(}\llbracket s\rrbracket\llbracket x\rrbracket,\llbracket s\rrbracket^{-1}\big{(}\llbracket s\rrbracket\llbracket x\rrbracket\big{)}\Big{)}=\tau\big{(}\llbracket s\rrbracket\llbracket x\rrbracket,\llbracket x\rrbracket\big{)}.

Condition (i) imply that c′′c^{\prime\prime} is \operatorname{\mathbbm{P}}-a.s. finite and there exists of a (finite) geodesic path. Observe that s1=s1\llbracket s^{-1}\rrbracket=\llbracket s\rrbracket^{-1} for all sSs\in S induces a graph homomorphism of 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) and 𝒞(Γ,S±)\mathcal{C}(\operatorname{\Gamma}^{\prime},\llbracket S\rrbracket^{\pm}). In other words, for all w1,w2ww_{1},w_{2}\in\llbracket w\rrbracket and i,j{1,,κ}i,j\in\{1,\dots,\kappa\}, z(j).w1≁z(i).w2z_{(j)}.w_{1}\not\sim z_{(i)}.w_{2} and if xyx\sim y in 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S), then xy\llbracket x\rrbracket\sim\llbracket y\rrbracket in 𝒞(Γ,S±)\mathcal{C}\big{(}\operatorname{\Gamma}^{\prime},\llbracket S\rrbracket^{\pm}\big{)}. Hence, one can easily verify by the minimax property that

c′′(x)\displaystyle c^{\prime\prime}(\llbracket x\rrbracket) :=max1i,jκmaxyz(j).xzz(i).torN(infγ𝒫(e,y){u,v}γτ(u,v))ϑz\displaystyle:=\max_{1\leq i,j\leq\kappa}\max_{\begin{subarray}{c}y\in z_{(j)}.\llbracket x\rrbracket\\ z\in z_{(i)}.\operatorname{\operatorname{tor}}N\end{subarray}}\left(\inf_{\gamma\in\mathscr{P}(e,y)}\sum_{\{u,v\}\in\gamma}\tau(u,v)\right)\circ\operatorname{\upvartheta}_{z}
=infγ𝒫(e,x)({u,v}γmax1i,jκmaxuz(j).usz(i).vu1τ(u,su))-a.s.\displaystyle\phantom{:}=\inf_{\gamma\in\mathscr{P}(\llbracket e\rrbracket,\llbracket x\rrbracket)}\left(\sum_{\{\llbracket u\rrbracket,\llbracket v\rrbracket\}\in\gamma}\max_{1\leq i,j\leq\kappa}\max_{\begin{subarray}{c}~{}u^{\prime}\in z_{(j)}.\llbracket u\rrbracket~{}\\ s^{\prime}\in z_{(i)}.\llbracket vu^{-1}\rrbracket\end{subarray}}\tau(u^{\prime},s^{\prime}u^{\prime})\right)\quad\operatorname{\mathbbm{P}}\text{-a.s.}

This is a direct consequence of the graph homomorphism. Property (iii) arises naturally from the given definition when S±Γ[Γ,Γ]\llbracket S\rrbracket^{\pm}\subseteq\operatorname{\Gamma}^{\prime}\setminus[\operatorname{\Gamma}^{\prime},\operatorname{\Gamma}^{\prime}]. ∎

Proposition 3.14.

Under the same hypotheses stated in Lemma 3.13, it follows that the results in Lemmas 3.9, 3.10, 3.11, 3.12 and 5 also hold when replacing cc^{\prime} with c′′c^{\prime\prime}.

Proof.

Notice that

(maxxBS(e,n)maxyj=1κz(j).torNc(y)ϑx>n)𝒪(1/nκ).\operatorname{\mathbbm{P}}\left(\max_{x\in B_{S}(e,n)}\max_{~{}y\in\bigcup_{j=1}^{\kappa}z_{(j)}.\operatorname{\operatorname{tor}}N}c(y)\circ\operatorname{\upvartheta}_{x}>\sqrt{n}\right)\in\mathcal{O}(1/n^{\upkappa}).

Consequently, maxxBS(e,n)maxyj=1κz(j).torNc(y)ϑxo(n)\max\limits_{x\in B_{S}(e,n)}\max\limits_{~{}y\in\bigcup_{j=1}^{\kappa}z_{(j)}.\operatorname{\operatorname{tor}}N}c(y)\circ\operatorname{\upvartheta}_{x}\in o(n), \operatorname{\mathbbm{P}}-a.s., as n+n\uparrow+\infty. Therefore, defining c′′(e)=0c^{\prime\prime}\big{(}\llbracket e\rrbracket\big{)}=0 is a suitable choice for investigating the asymptotic cone of c′′c^{\prime\prime} in comparison to cc.

The arguments in the proofs of Lemmas 3.9, 3.10, 3.11 and 3.12 can be repeated for c′′c^{\prime\prime}, yielding the same properties up to a constant factor. ∎

Corollary 3.15.

Let (Γ,.)(\operatorname{\Gamma},.) be a finitely generated group with polynomial growth rate D1D\geq 1 and Γ/[Γ,Γ]\operatorname{\Gamma}^{\prime}/[\operatorname{\Gamma}^{\prime},\operatorname{\Gamma}^{\prime}] torsion-free. Consider c:Γ×Ω0c:\operatorname{\Gamma}\times\operatorname{\Upomega}\to\operatorname{\mathbb{R}}_{\geq 0} to be a subadditive cocycle associated with dωd_{\omega} and a p.m.p. ergodic group action ϑ|N:N(Ω,,)\operatorname{\upvartheta}\big{|}_{N}:N\curvearrowright(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}).

Suppose that cc describes a FPP model which satisfies conditions (i) and (ii) for a finite symmetric generating set SΓS\subseteq\operatorname{\Gamma} so that

  • (i)

    For all sSs\in S, s1=s1\llbracket s^{-1}\rrbracket=\llbracket s\rrbracket^{-1}, and

  • (ii)

    S±Γ[Γ,Γ]\llbracket S\rrbracket^{\pm}\subseteq\operatorname{\Gamma}^{\prime}\setminus[\operatorname{\Gamma}^{\prime},\operatorname{\Gamma}^{\prime}] generates Γ\operatorname{\Gamma}^{\prime}.

Then

(Γ,1ndω,e)GH(G,dϕ,e)-a.s.\quad\quad\left(\operatorname{\Gamma},\frac{1}{n}d_{\omega},e\right)\operatorname{\xrightarrow{\text{GH}}}\left(G_{\infty},d_{\phi},\mathlcal{e}\right)\quad\quad\operatorname{\mathbbm{P}}\text{-a.s.}

where GG_{\infty} is a simply connected graded Lie group, and dϕd_{\phi} is a quasimetric homogeneous with respect to a family of homotheties {δt}t>0\{\delta_{t}\}_{t>0}. Moreover, dϕd_{\phi} is bi-Lipschitz equivalent to dd_{\infty} on GG_{\infty}.

Proof.

First, according to Proposition 3.14, the random variables cc^{\prime} and c′′c^{\prime\prime} share similar properties. Observe that |𝔅N|=1|\mathfrak{B}_{N}|=1, ensuring that c′′c^{\prime\prime} well-defined and a suitable replacement of cc^{\prime} in the proof of Theorem 1.2, which establishes the result. ∎

The next example highlights a case where ϑ|N\operatorname{\upvartheta}\big{|}_{N} acts ergodically on the probability space followed by an example of virtually nilpotent group with generating set satisfying items (i) and (ii) of Corollary 3.15.

Example 3.1 (Independent FPP models).

The subadditive cocycle cc exhibits equivariance. Recall properties fiscussed in Section 2.5 for FPP models and notice that, for all x,yΓx,y\in\operatorname{\Gamma} and sSs\in S,

τ(x,sx)τ(y,s±1y).\tau(x,sx)\sim\tau\big{(}y,s^{\pm 1}y\big{)}.

Consider that the random weights are independent, but not necessarily identically distributed (see [4] for FPP with i.i.d. random variables). Let us define S:={{s,s1}:sS}S^{\prime}:=\big{\{}\{s,s^{-1}\}:s\in S\big{\}} and set ς(s):=ss\varsigma(s^{\prime}):=s\in s^{\prime} for sSs^{\prime}\in S^{\prime}, i.e., the function ς\varsigma fixes one element of each sSs^{\prime}\in S^{\prime}.

Suppose that, for all sSs\in S, s2es^{2}\neq e and consider ν(s)\nu^{(s^{\prime})} to be the law of τ(x,ς(s)x)\tau(x,\varsigma(s^{\prime})x) with xΓx\in\operatorname{\Gamma} and ssSs\in s^{\prime}\in S^{\prime}. Thus, one can write

(sSν(s))Γ=(j=1κsSν(s))NxNν(x),\operatorname{\mathbbm{P}}\equiv\left(\bigotimes_{s^{\prime}\in S^{\prime}}\nu^{(s^{\prime})}\right)^{\otimes\operatorname{\Gamma}}=\left(\bigotimes_{j=1}^{\kappa}\bigotimes_{s^{\prime}\in S^{\prime}}\nu^{(s^{\prime})}\right)^{\otimes N}\equiv\bigotimes_{x\in N}\nu^{(x)},

where, for each xNx\in N, ν(x)j=1κsSν(s)\nu^{(x)}\equiv\bigotimes_{j=1}^{\kappa}\bigotimes_{s^{\prime}\in S^{\prime}}\nu^{(s^{\prime})}. Let EE\in\operatorname{\mathscr{F}} be such that, for all xNx\in N, ϑx(E)=E\operatorname{\upvartheta}_{x}(E)=E. Then, for all x,yNx,y\in N,

ν(x)(E)=ν(y)(E)=:𝚔E[0,1].\nu^{(x)}(E)=\nu^{(y)}(E)=:\mathtt{k}_{E}\in[0,1].

The condition of polynomial growth rate D1D\geq 1 ensures that NN is countably infinite. Consequently,

(E)=xN𝚔E{0,1}.\operatorname{\mathbbm{P}}(E)=\prod_{x\in N}\mathtt{k}_{E}\in\{0,1\}.

Therefore, ϑ|N\operatorname{\upvartheta}\big{|}_{N} as defined in Section 2.5 constitutes a probability measure-preserving (p.m.p.) ergodic group action for independent FPP models.

Example 3.2 (Direct product).

Consider L\mathrm{L} a torsion-free nilpotent group with torsion-free abelianization and a symmetric finite generating set SLL[L,L]S_{\mathrm{L}}\subseteq\mathrm{L}\setminus[\mathrm{L},\mathrm{L}]. Set MM to be a finite group. Recall the properties highlighted in Section 2.3. Let us define

Γ=L×M,andS=SL×M.\operatorname{\Gamma}=\mathrm{L}\times M,\quad\text{and}\quad S=S_{\mathrm{L}}\times M.

Then SS is a symmetric finite generating set of Γ\operatorname{\Gamma}. Fix πN(x,m)=(x,e)\uppi_{N}(x,m)=(x,e) for all xLx\in\mathrm{L} and mMm\in M. One can easily see that ΓL\operatorname{\Gamma}^{\prime}\cong\mathrm{L} with S=S±SL\llbracket S\rrbracket=\llbracket S\rrbracket^{\pm}\cong S_{\mathrm{L}}.

Furthermore, for any (x,m)Γ(x,m)\in\operatorname{\Gamma}, the inverse (x,m)1(x,m)^{-1} is given by (x1,m1)(x^{-1},m^{-1}), leading to

(x,m)1x1(x,m)1.\llbracket(x,m)^{-1}\rrbracket\cong x^{-1}\cong\llbracket(x,m)\rrbracket^{-1}.

As a consequence, both items (i) and (ii) of Corollary 3.15 hold when Γ\operatorname{\Gamma} is the direct product equipped with the generating set SS defined above.

4. Applications to random growth models

In this section, we delve into three distinct examples that serve as applications of the main results outlined in this article for a random growth on 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S). These examples have been deliberately chosen to address scenarios that fall outside the scope of previous works, thereby offering a nuanced examination of the versatility and robustness of our established theorems.

The first example considers a First-Passage Percolation (FPP) model with dependent random variables, challenging the assumption of LL^{\infty}, since we allow random weights to be zero with a strict positive probability. Transitioning to the second example, we investigate a FPP model with independent random variables that are not identically distributed and also not LL^{\infty}. The third example shifts focus to an interacting particle system that extends is not a FPP model. Notably, this model fails to meet the conditions found in the literature.

Example 4.1 (First-Passage Percolation for a Random Coloring of Γ\operatorname{\Gamma}).

Let us now consider a dependent Bernoulli FPP model based on the random coloring studied by Fontes and Newman [12]. Set {Xx}xΓ\{X_{x}\}_{x\in\operatorname{\Gamma}} to be a family of i.i.d. random variables taking values in a finite set of colors F\mathlcal{F}. The model generates color clusters by assigning weight 0 to edges between sites with same color and weight 11 otherwise. We define for every edge uvu\sim v

τ(u,v)=𝟙(XuXv),\tau(u,v)=\mathbbm{1}(X_{u}\neq X_{v}),

Set for each self-avoinding path γ𝒫(x,y)\gamma\in\mathscr{P}(x,y) the random length T(γ)=𝚎γτ(𝚎)T(\gamma)=\sum_{\mathtt{e}\in\gamma}\tau(\mathtt{e}). The first-passage time is

T(x,y):=infγ𝒫(x,y)T(γ)T(x,y):=\inf_{\gamma\in\mathscr{P}(x,y)}T(\gamma)

Let ps:=(Xx=s)p_{\mathlcal{s}}:=\operatorname{\mathbbm{P}}(X_{x}=\mathlcal{s}) then one can verify that T(x,y)T(x,y) is a FPP model with dependent identically distributed passage times τ(x,y)Ber(1sFps2)\tau(x,y)\sim\operatorname{Ber}\left(1-\sum_{\mathlcal{s}\in\mathlcal{F}}p_{\mathlcal{s}}^{2}\right). One can easily see that c(x):=T(e,x)c(x):=T(e,x) is a subadditive cocycle and the translations ϑ\operatorname{\upvartheta} are ergodic due to the fact that {Xx}xΓ\{X_{x}\}_{x\in\operatorname{\Gamma}} are i.i.d. random variables .

Observe that c(x)c(x) is bounded above by the word norm xS\|x\|_{S}, items (i) and (iii) are immediately satisfied. Consider ps(0,1)p_{\mathlcal{s}}\in(0,1) for all sF\mathlcal{s}\in\mathlcal{F}. Set

p:=maxsFps,q:=maxsF(1ps), and p:=pp+q.p:=\max_{\mathlcal{s}\in\mathlcal{F}}~{}p_{\mathlcal{s}}~{},\quad q:=\max_{\mathlcal{s}\in\mathlcal{F}}~{}(1-p_{\mathlcal{s}}),\quad\text{ and }\quad p^{\prime}:=\frac{p}{p+q}.

The lemma below establishes a sufficient condition for (ii) and (ii).

Lemma 4.1.

Consider the Random Coloring Model of Γ\operatorname{\Gamma} on 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) satisfying

p<1|S|1,p<\frac{1}{|S|-1}, (4.1)

then (ii) and (ii) hold true.

Proof.

Let γ=(x0=e,x1,,xn)𝒫n\gamma=(x_{0}=e,x_{1},\dots,x_{n})\in\mathscr{P}_{n} with 𝒫n\mathscr{P}_{n} the set of all self-avoiding paths in 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) of graph length nn starting at ee. Fix [n]:={1,,n}[n]:=\{1,\dots,n\}, then

(T(γ)=m)\displaystyle\operatorname{\mathbbm{P}}\big{(}T(\gamma)=m\big{)} A[n]|A|=mi[n]A(Xxi=Xxi1|Xxi1)jA(XxiXxi1|Xxi1)\displaystyle\leq\sum_{\begin{subarray}{c}A\subseteq[n]\\ |A|=m\end{subarray}}\prod_{i\in[n]\setminus A}\mathbb{P}(X_{x_{i}}=X_{x_{i-1}}|X_{x_{i-1}})\prod_{j\in A}\mathbb{P}(X_{x_{i}}\neq X_{x_{i-1}}|X_{x_{i-1}})
(nm)pnmqm=(p+q)nP(Y=m)\displaystyle\leq\binom{n}{m}p^{n-m}q^{m}=(p+q)^{n}P(Y=m)

where YBinomial(n,1p)Y\sim\operatorname{Binomial}(n,1-p^{\prime}) with respect to PP. Let us regard xS=n\|x\|_{S}=n, thus

(c(x)αxS)\displaystyle\operatorname{\mathbbm{P}}\big{(}c(x)\leq\upalpha\|x\|_{S}\big{)} (γ𝒫n:T(γ)αn)\displaystyle\leq\operatorname{\mathbbm{P}}\big{(}\exists\gamma\in\mathscr{P}_{n}:T(\gamma)\leq\upalpha n\big{)}
|𝒫n|(p+q)nP(Yαn).\displaystyle\leq\big{|}\mathscr{P}_{n}\big{|}(p+q)^{n}\cdot P(Y\leq\upalpha n).

It is a well-known fact that |𝒫n||S|(|S|1)n1|\mathscr{P}_{n}|\leq|S|(|S|-1)^{n-1}. Therefore, there exists 𝙲>0\mathtt{C}>0 such that |𝒫n|𝙲(|S|1)n|\mathscr{P}_{n}|\leq\mathtt{C}(|S|-1)^{n}. By Chernoff bound, one can obtain

P(Yαn)\displaystyle P(Y\leq\upalpha n) exp(n(α1)log1α1pαlogαp))\displaystyle\leq\exp\left(n\left(\upalpha-1)\log\frac{1-\upalpha}{1-p^{\prime}}-\upalpha\log\frac{\upalpha}{p^{\prime}}\right)\right)
=((p)α(1p)(1α)αα(1α)α1)n.\displaystyle=\left((p^{\prime})^{\upalpha}(1-p^{\prime})^{(1-\upalpha)}\upalpha^{-\upalpha}(1-\upalpha)^{\alpha-1}\right)^{n}. (4.2)

Observe that the base of (4.2) converges to pp^{\prime} as α0\upalpha\downarrow 0. Hence, there exist α,p′′>0\upalpha,p^{\prime\prime}>0 such that P(Yαn)(p′′)nP(Y\leq\upalpha n)\leq\left(p^{\prime\prime}\right)^{n} with p<p′′<1/((p+1)(|S|1))p^{\prime}<p^{\prime\prime}<1/((p+1)(|S|-1)) when pp satisfies (4.1). It then follows that there exists 𝙲>0\mathtt{C}^{\prime}>0 such that

(c(x)αxS)𝙲(p′′(p+q)(|S|1))n=𝙲exp(𝙲n).\operatorname{\mathbbm{P}}\big{(}c(x)\leq\upalpha\|x\|_{S}\big{)}\leq\mathtt{C}\big{(}p^{\prime\prime}(p+q)(|S|-1)\big{)}^{n}=\mathtt{C}\exp(-\mathtt{C}^{\prime}n).

Let now a:=α/2a:=\upalpha/2 and choose xS1\|x\|_{S}\gg 1 so that (c(x)αxS)1/2\operatorname{\mathbbm{P}}\big{(}c(x)\leq\upalpha\|x\|_{S}\big{)}\leq 1/2, then axS𝔼[c(x)]a\|x\|_{S}\leq\operatorname{\mathbbm{E}}[c(x)], which yields (ii) and (ii) as a consequence. ∎

Similarly to Example 3.1, let ν\nu be the law of the random coloring of a vertex. Then

νΓ=(j=1κν)NxNν(x)\operatorname{\mathbbm{P}}\equiv\nu^{\otimes\operatorname{\Gamma}}=\Bigg{(}\bigotimes_{j=1}^{\kappa}\nu\Bigg{)}^{\otimes N}\equiv\bigotimes_{x\in N}\nu^{(x)}

with ν(x)j=1κν\nu^{(x)}\equiv\bigotimes_{j=1}^{\kappa}\nu. By the same reasoning employed for ν(x)\nu^{(x)} in Example 3.1, we verify that ϑ|N\operatorname{\upvartheta}\big{|}_{N} acts ergodically on (Ω,,)(\operatorname{\Upomega},\operatorname{\mathscr{F}},\operatorname{\mathbbm{P}}).

Hence, under the assumption of (4.1) and based on the aforementioned results, the Shape Theorems 1.1 and 1.2 are applicable to the random coloring of Γ=N\operatorname{\Gamma}=N nilpotent with a finite generating set SN([N,N]torN)S\subseteq N\setminus\big{(}[N,N]\cup\operatorname{\operatorname{tor}}N\big{)} or in the case where Γ\operatorname{\Gamma}^{\prime} is abelian. Moreover, under the fulfillment of conditions (i) and (ii) in Corollary 3.15, the existence of the limiting shape is also guaranteed when Γ\operatorname{\Gamma} is virtually nilpotent.

Remark 6.

Observe that (4.1) provides a lower bound for the critical probability of site percolation on 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) (see for instance [14]). To verify that, fix a color sF\mathlcal{s}\in\mathlcal{F} and we say that a site xΓx\in\operatorname{\Gamma} is open when Xx=sX_{x}=\mathlcal{s}. Therefore, one can write

τsite(s)(x,y)=𝟙(Xxs or Xys).\tau_{\operatorname{site}}^{(\mathlcal{s})}(x,y)=\mathbbm{1}(X_{x}\neq\mathlcal{s}\text{ or }X_{y}\neq\mathlcal{s}).

Note that it stochastically dominates with ττsite(s)\tau\leq\tau_{\operatorname{site}}^{(\mathlcal{s})} \operatorname{\mathbbm{P}}-a.s. The open edges are the new edges of length zero. By Lemma 4.1, we can apply Theorem 1.1 to obtain that, \operatorname{\mathbbm{P}}-a.s., there is no infinite open cluster in 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) when ps<1|S|1p_{\mathlcal{s}}<\frac{1}{|S|-1}.

Example 4.2 (Richardson’s Growth Model in a Translation Invariant Random Environment).

In this example, we define a variant of the Richardson’s Growth Model which is commonly employed to describe the spread of infectious diseases. This version of the model involves independent random variables that are not identically distributed (see [13, 18] for similar models). Specifically, we consider that the transmission rate of the disease between neighboring sites is randomly chosen. The distribution of this variable will vary depending on the directions of the Cayley graph.

Consider that the infection rates between neighbors are determined by a random environment taking values in Λ:=(0,+)E\Uplambda:=(0,+\infty)^{E}. Let S:={{s,s1}:sS}S^{\prime}:=\big{\{}\{s,s^{-1}\}:s\in S\big{\}} be the set of directions of 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S). Consider {λs}sS\{\uplambda_{s^{\prime}}\}_{s^{\prime}\in S^{\prime}} a set of strictly positive random variables that are independent over a probability measure ν\nu. Set (λ(𝚎))𝚎E\big{(}\uplambda(\mathtt{e})\big{)}_{\mathtt{e}\in E} to be a collection of independent random variables over ν\nu such that

λ(x,sx)λswith s={s±1}.\uplambda(x,sx)\sim\uplambda_{s^{\prime}}\quad\text{with }s^{\prime}=\{s^{\pm 1}\}.

Let us regard λΛ\uplambda\in\Uplambda as a fixed realization of the random environment. The growth process is defined by the family of independent random variables {τ(x,sx):xΓ,sS}{\{\tau(x,sx):x\in\operatorname{\Gamma},s\in S\}} such that

τ(x,sx)Exp(λ(x,sx)).\tau(x,sx)\sim\text{Exp}\big{(}\uplambda(x,sx)\big{)}. (4.3)

Set λ\operatorname{\mathbbm{P}}_{\uplambda} to be the quenched probability law of (4.3). We write, for each path γ𝒫(x,y)\upgamma\in\mathscr{P}(x,y) with x,yΓx,y\in\operatorname{\Gamma}, its random length T(γ):=𝚎γτ(𝚎)T(\upgamma):=\sum_{\mathtt{e}\in\upgamma}\tau(\mathtt{e}).

The first-passage time is

c(x):=infγ𝒫(e,x)T(γ).c(x):=\inf_{\upgamma\in\mathscr{P}(e,x)}T(\upgamma).

It is straightforward to see that c(x)c(x) is subadditive. However, the group action ϑ\operatorname{\upvartheta} is not ergodic over λ\operatorname{\mathbbm{P}}_{\uplambda} for a given λΛ\uplambda\in\Uplambda. Let ()=Λγ()𝑑ν(λ)\operatorname{\mathbbm{P}}(\cdot)=\int_{\Uplambda}\operatorname{\mathbbm{P}}_{\upgamma}(\cdot)d\nu(\uplambda) be the annealed probability. It then follows that ϑ\operatorname{\upvartheta} preserves the measure \operatorname{\mathbbm{P}} and it is ergodic.

Note that c(x)c(x) defines a First-Passage Percolation (FPP) model, which we refer to as Richardson’s Growth Model in a Translation Random Environment (RGTRE). In the following, we establish that conditions (i), (ii), and (ii) are met.

Lemma 4.2.

Consider the RGTRE defined as above. Then there exist β,𝙲,𝙲>0\upbeta,\mathtt{C},\mathtt{C}^{\prime}>0 such that, for all xΓx\in\operatorname{\Gamma},

(c(x)t)𝙲exp(𝙲t)\operatorname{\mathbbm{P}}\big{(}c(x)\geq t\big{)}\leq\mathtt{C}\exp\big{(}-\mathtt{C}^{\prime}t\big{)}

for all tβxSt\geq\upbeta\|x\|_{S}.

Proof.

Let γ𝒫(e,x)\upgamma\in\mathscr{P}(e,x) be a dSd_{S}-geodesic with xS=n\|x\|_{S}=n. Then one has by Chernoff bound and the independence of {τ(𝚎)}𝚎E\{\tau(\mathtt{e})\}_{\mathtt{e}\in E} that

λ(c(x)t)λ(T(γ)t)𝚎γ𝔼λ[eατ(𝚎)]eαt.\operatorname{\mathbbm{P}}_{\uplambda}\big{(}c(x)\geq t\big{)}\leq\operatorname{\mathbbm{P}}_{\uplambda}\big{(}T(\upgamma)\geq t\big{)}\leq\frac{\prod_{\mathtt{e}\in\gamma}\operatorname{\mathbbm{E}}_{\uplambda}\left[e^{\alpha\tau(\mathtt{e})}\right]}{e^{\alpha t}}.

where

𝔼λ[eατ(𝚎)]=m=0+(αλ(e))m\operatorname{\mathbbm{E}}_{\uplambda}\left[e^{\alpha\tau(\mathtt{e})}\right]=\sum_{m=0}^{+\infty}\left(\frac{\alpha}{\uplambda(e)}\right)^{m}

Let λ min:=minsS𝔼[λs]\accentset{\rule{2.79996pt}{0.5pt}}{\uplambda}_{\min}:=\min_{s^{\prime}\in S^{\prime}}\operatorname{\mathbbm{E}}[\uplambda_{s^{\prime}}] and set α=λ min/2\alpha=\accentset{\rule{2.79996pt}{0.5pt}}{\uplambda}_{\min}/2. Thus, by the Dominated Convergence Theorem,

λ(c(x)t)2neλ mint/2.\operatorname{\mathbbm{P}}_{\uplambda}\big{(}c(x)\geq t\big{)}\leq\frac{2^{n}}{e^{\accentset{\rule{2.79996pt}{0.5pt}}{\uplambda}_{\min}t/2}}.

Therefore, it suffices to choose β>2log(2)/λ min\upbeta>2\log(2)/\accentset{\rule{2.79996pt}{0.5pt}}{\uplambda}_{\min} to complete the proof. ∎

Lemma 4.3.

Consider the RGTRE defined as above. The there exists a>0a>0 such that, for all xΓx\in\operatorname{\Gamma},

axS𝔼[c(x)].a\|x\|_{S}\leq\operatorname{\mathbbm{E}}[c(x)].
Proof.

It is a well-known fact that, for all λΛ\uplambda\in\Uplambda and every 𝚎E\mathtt{e}\in E, that λ(τ(𝚎)=0)=0\operatorname{\mathbbm{P}}_{\uplambda}\big{(}\tau(\mathtt{e})=0\big{)}=0 and, therefore, (τ(𝚎)=0)=0\operatorname{\mathbbm{P}}\big{(}\tau(\mathtt{e})=0\big{)}=0. By the right continuity of the cumulative distribution function, one can find δ>0\delta>0 and p0p\geq 0 such that, for every 𝚎E\mathtt{e}\in E,

(τ(𝚎)<δ)=p<1|S|1.\operatorname{\mathbbm{P}}\big{(}\tau(\mathtt{e})<\delta\big{)}=p<\frac{1}{|S|-1}.

We use similar arguments as those employed in the proof of Lemma 4.1, we may consider YBinomial(n,1p)Y\sim\operatorname{Binomial}(n,1-p) over PP. Then there exists α>0\upalpha>0 such that, for any γ𝒫n\upgamma\in\mathscr{P}_{n},

(T(γ)αn)P(Yαn/δ)pn.\operatorname{\mathbbm{P}}\big{(}T(\upgamma)\leq\upalpha n\big{)}\leq P\big{(}Y\leq\upalpha n/\delta\big{)}\leq p^{n}.

It follows that there exists C>0C>0 such that, for xS=n\|x\|_{S}=n,

(c(x)αxS)|𝒫n|P(Yαn/δ)C((|S|1)p)n.\operatorname{\mathbbm{P}}\big{(}c(x)\leq\alpha\|x\|_{S}\big{)}\leq\big{|}\mathscr{P}_{n}\big{|}\cdot P\big{(}Y\leq\upalpha n/\delta\big{)}\leq C\big{(}(|S|-1)p\big{)}^{n}.

Since (|S|1)p<1(|S|-1)p<1, we can complete the proof by following the same steps as in Lemma 4.1. ∎

It follows from Lemmas 4.2 and 4.3 that conditions (i), (ii), and (ii) are satisfied. Observe that ϑ|N\operatorname{\upvartheta}\big{|}_{N} acts ergodically on the probability space (see Example 3.1).

Therefore, building upon the preceding results, the Limiting Shape Theorems 1.1 and 1.2 apply to the RGTRE with Γ=N\operatorname{\Gamma}=N nilpotent with a finite generating set SN([N,N]torN)S\subseteq N\setminus\big{(}[N,N]\cup\operatorname{\operatorname{tor}}N\big{)} or in scenarios where Γ\operatorname{\Gamma}^{\prime} is abelian. Additionally, when Γ\operatorname{\Gamma} is virtually nilpotent and conditions (i) and (ii) from Corollary 3.15 are satisfied, the existence of the limiting shape is also assured.

Example 4.3 (The Frog Model).

The Frog Model, originally introduced by Alves et al. [1] and previously featured as an example in [19], is a discrete-time interacting particle system determined by the intersection of random walks on a graph. In this model, particles, often representing individuals, are distributed across the vertices and they can be in either active (awake) or inactive (sleeping) states. At discrete time steps, active particles perform simple random walks, while inactive ones remain stationary. The activation of an inactive particle occurs when its vertex is visited by an active counterpart, thereby characterizing an awakening process. This straightforward yet potent model serves as a valuable tool for analyzing diverse dynamic processes, such as the spread of information and disease transmission.

In our previous study Coletti and de Lima [9], we investigated the Frog Model on finitely generated groups. We can now extend our findings to virtually nilpotent groups as a consequence of Theorem 1.2. Let us define the model in detail. The initial configuration of the process at time zero begins with one particle at each vertex and the only active particle lies on the origin eΓe\in\operatorname{\Gamma}.

Set SnxS_{n}^{x} to be the simple random walk on 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) of a particle originally placed at xΓx\in\operatorname{\Gamma} and let t(x,y)t(x,y) be the first time the random walk SnxS_{n}^{x} visits yΓy\in\operatorname{\Gamma}, i.e., it defined the random variable t(x,y)=inf{n0:Snx=y}t(x,y)=\inf\{n\in\operatorname{\mathbb{N}}_{0}:S_{n}^{x}=y\}. Note that t(x,y)=+t(x,y)=+\infty with strictly positive probability.

The activation time of the particle originally positioned at xx is given by the random variable

T(x)=inf{i=1mt(xi1,xi):m,{xi}i=1mΓ,x0=e}.T(x)=\inf\left\{\sum_{i=1}^{m}t(x_{i-1},x_{i})\colon m\in\operatorname{\mathbb{N}},~{}\{x_{i}\}_{i=1}^{m}\subseteq\operatorname{\Gamma},~{}x_{0}=e\right\}.

Observe that xi1x_{i-1} and xix_{i} are not necessarily neighbours. We proved in [9] that c(x)=T(x)c(x)=T(x) is a subadditive cocycle with respect to the translation ϑ\operatorname{\upvartheta}, which is p.m.p. and an ergodic group action. Futhermore, τ(e)(x,sx)=|T(x)T(sx)|\tau^{(e)}(x,sx)=|T(x)-T(sx)| is not identically distributes as in the FPP models (see Section 2.5).

Due to the discrete time random walks, T(x)xST(x)\geq\|x\|_{S} and therefore (ii) is immediately satisfied. The at least linear growth in virtually nilpotent group was already investigated in [9]. Hence, condition (i) is a consequence of the following result.

Lemma 4.4 (Prop. 2.10 of [9]).

Let Γ\operatorname{\Gamma} be a group of polynomial growth rate D3D\geq 3 with a symmetric finite generating set SΓ{e}S\subseteq\operatorname{\Gamma}\setminus\{e\}. Then there exists 𝙲,ϰ>0\mathtt{C},\varkappa>0 and β>1\upbeta>1 such that, for all xΓx\in\operatorname{\Gamma} and every t>βxSt>\upbeta\|x\|_{S}, one has

(T(x)t)𝙲exp(tϰ).\operatorname{\mathbbm{P}}\big{(}T(x)\geq t\big{)}\leq\mathtt{C}\exp(-t^{\varkappa}).

Consider now Γ\operatorname{\Gamma} as a group with polynomial growth rate D3D\geq 3 generated by a symmetric finite set SΓ{e}S\subseteq\operatorname{\Gamma}\setminus\{e\}. According to Theorem 1.2 and the preceding results, it can be inferred that the Frog Model on 𝒞(Γ,S)\mathcal{C}(\operatorname{\Gamma},S) exhibits a limiting shape when S\big{\langle}\llbracket S\rrbracket\big{\rangle} generates an abelian group Γ\operatorname{\Gamma}^{\prime}. This phenomenon can be exemplified by the generalized dihedral group Γ=Dih(N)\operatorname{\Gamma}=\operatorname{Dih}(N) when NN is a finitely generated abelian group with polynomial growth rate D3D\geq 3 (see Example 2.3).

5. Final remarks

In this work, we have successfully established the Asymptotic Shape Theorem for random subadditive processes on both nilpotent and virtually nilpotent groups. By extending existing results in the literature, we have achieved a comprehensive understanding of the behavior of these processes under more relaxed growth conditions—both at least and at most linear growth. This broadening of applicability enhances the utility of our results in diverse mathematical contexts.

A noteworthy contribution of our work lies in the exploration of FPP models, a crucial class of processes meeting the considered conditions, especially the innerness property. Leveraging this, we were able to derive a corollary for the limiting shape in FPP models, thereby extending the reach of our results to encompass this important and widely studied class of random processes.

Moreover, our presentation of examples generalizes previously known results in shape theorems. These examples illustrate scenarios where the strong restriction of LL^{\infty} cocycles is alleviated, emphasizing the versatility of our established theorems in capturing a broader range of applications.

Looking forward, possible future research may involve the exploration of other types of random variables exhibiting almost subadditive behavior. Additionally, considering point processes on nilpotent Lie groups to define random graphs opens up intriguing possibilities for further investigation. An interesting direction for future works could involve refining our theorems based on the generating set, recognizing the crucial role it plays in certain key aspects. Such refinements could leverage quasi-isometric properties, offering a more nuanced understanding of the interplay between the generating set and the behavior of random subadditive processes.

References

  • Alves et al. [2002] O. S. M. Alves, F. P. Machado, and S. Y. Popov. The shape theorem for the frog model. Ann. Appl. Probab., 12(2):533–546, 2002. ISSN 1050-5164. doi: 10.1214/aoap/1026915614. URL https://doi.org/10.1214/aoap/1026915614.
  • Auffinger and Gorski [2023] A. Auffinger and C. Gorski. Asymptotic shapes for stationary first passage percolation on virtually nilpotent groups. Probab. Theory Related Fields, 186(1-2):285–326, 2023. ISSN 0178-8051. doi: 10.1007/s00440-023-01196-7. URL https://doi.org/10.1007/s00440-023-01196-7.
  • Austin [2016] T. Austin. Integrable measure equivalence for groups of polynomial growth. Groups Geom. Dyn., 10(1):117–154, 2016. ISSN 1661-7207. doi: 10.4171/GGD/345. URL https://doi.org/10.4171/GGD/345. With Appendix B by Lewis Bowen.
  • Benjamini and Tessera [2015] I. Benjamini and R. Tessera. First passage percolation on nilpotent Cayley graphs. Electron. J. Probab., 20:no. 99, 20, 2015. doi: 10.1214/EJP.v20-3940. URL https://doi.org/10.1214/EJP.v20-3940.
  • Björklund [2010] M. Björklund. The asymptotic shape theorem for generalized first passage percolation. Ann. Probab., 38(2):632–660, 2010. ISSN 0091-1798. doi: 10.1214/09-AOP491. URL https://doi.org/10.1214/09-AOP491.
  • Boivin [1990] D. Boivin. First passage percolation: the stationary case. Probab. Theory Related Fields, 86(4):491–499, 1990. ISSN 0178-8051. doi: 10.1007/BF01198171. URL https://doi.org/10.1007/BF01198171.
  • Breuillard [2014] E. Breuillard. Geometry of locally compact groups of polynomial growth and shape of large balls. Groups Geom. Dyn., 8(3):669–732, 2014. ISSN 1661-7207. doi: 10.4171/GGD/244. URL https://doi.org/10.4171/GGD/244.
  • Cantrell and Furman [2017] M. Cantrell and A. Furman. Asymptotic shapes for ergodic families of metrics on nilpotent groups. Groups Geom. Dyn., 11(4):1307–1345, 2017. ISSN 1661-7207. doi: 10.4171/GGD/430. URL https://doi.org/10.4171/GGD/430.
  • Coletti and de Lima [2021] C. F. Coletti and L. R. de Lima. The asymptotic shape theorem for the frog model on finitely generated abelian groups. ESAIM Probab. Stat., 25:204–219, 2021. ISSN 1292-8100. doi: 10.1051/ps/2021007. URL https://doi.org/10.1051/ps/2021007.
  • Cornulier and de la Harpe [2016] Y. Cornulier and P. de la Harpe. Metric geometry of locally compact groups, volume 25 of EMS Tracts in Mathematics. European Mathematical Society (EMS), Zürich, 2016. ISBN 978-3-03719-166-8. doi: 10.4171/166. URL https://doi.org/10.4171/166. Winner of the 2016 EMS Monograph Award.
  • de Cornulier [2011] Y. de Cornulier. Asymptotic cones of Lie groups and cone equivalences. Illinois J. Math., 55(1):237–259 (2012), 2011. ISSN 0019-2082. URL http://projecteuclid.org/euclid.ijm/1355927035.
  • Fontes and Newman [1993] L. Fontes and C. M. Newman. First passage percolation for random colorings of 𝐙d{\bf Z}^{d}. Ann. Appl. Probab., 3(3):746–762, 1993. ISSN 1050-5164. URL https://doi.org/10.1214/aoap/1177005361.
  • Garet and Marchand [2012] O. Garet and R. Marchand. Asymptotic shape for the contact process in random environment. Ann. Appl. Probab., 22(4):1362–1410, 2012. ISSN 1050-5164. doi: 10.1214/11-AAP796. URL https://doi.org/10.1214/11-AAP796.
  • Grimmett and Stacey [1998] G. R. Grimmett and A. M. Stacey. Critical probabilities for site and bond percolation models. Ann. Probab., 26(4):1788–1812, 1998. ISSN 0091-1798. doi: 10.1214/aop/1022855883. URL https://doi.org/10.1214/aop/1022855883.
  • Gromov [1981] M. Gromov. Groups of polynomial growth and expanding maps. Inst. Hautes Études Sci. Publ. Math., 53:53–73, 1981. ISSN 0073-8301. URL http://www.numdam.org/item?id=PMIHES_1981__53__53_0.
  • Pansu [1983] P. Pansu. Croissance des boules et des géodésiques fermées dans les nilvariétés. Ergodic Theory Dynam. Systems, 3(3):415–445, 1983. ISSN 0143-3857. doi: 10.1017/S0143385700002054. URL https://doi.org/10.1017/S0143385700002054.
  • Raghunathan [1972] M. S. Raghunathan. Discrete subgroups of Lie groups. Springer-Verlag, New York-Heidelberg, 1972. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 68.
  • Richardson [1973] D. Richardson. Random growth in a tessellation. Proc. Cambridge Philos. Soc., 74:515–528, 1973. ISSN 0008-1981. doi: 10.1017/s0305004100077288. URL https://doi.org/10.1017/s0305004100077288.
  • Telcs and Wormald [1999] A. Telcs and N. C. Wormald. Branching and tree indexed random walks on fractals. J. Appl. Probab., 36(4):999–1011, 1999. ISSN 0021-9002. doi: 10.1017/s0021900200017812. URL https://doi.org/10.1017/s0021900200017812.