Asymptotic shape for subadditve processes on groups of polynomial growth
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 graphs2020 Mathematics Subject Classification:
Primary: 52A22, 60F15; Secondary: 60K351. 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 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 . Here, we modify this hypothesis by replacing it with 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 and 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 be a probability space and a finitely generated group with polynomial growth rate. Set to be a -preserving (p.m.p.) ergodic group action. Consider the family of non-negative random variables such that, -a.s.,
(1.1) |
Write for and let . A function satisfying (1.1) is referred to as a subadditive cocycle. Once given a subadditive cocycle , there is a correspondent random pseudo-quasi metric defined by
which is -right equivariant, i.e., for all , and for every ,
The correspondence is one-to-one since given a -right equivariant random pseudoquasimetric , one can easily verify that
(1.2) |
is a subadditive cocycle.
To avoid dealing with unnecessary technicalities, we initially consider as the group of polynomial growth, which is nilpotent and torsion-free. Later, we address the more general case where 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 . We write and for a word length and a word metric, respectively. The following conditions will be needed throughout the paper. We assume the existence of and such that, for all ,
(i) |
where as .
Let be the commutator subgroup of and set . Suppose that there exists such that, for all there is a sequence of positive integers depending on with and, for all and every ,
(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 , which will be defined later.
To obtain the asymptotic result, we will introduce an innerness assumption. Specifically, for each , we require the existence of a finite generating set such that, for -a.s. and for every , we can write with satisfying
(iii) |
When considering First-Passage Percolation models where , condition (iii) is automatically fulfilled (see Section 2.5). Additionally, in the case where 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 be a torsion-free nilpotent finitely generated group with polynomial growth rate and torsion-free abelianization. Consider to be a subadditive cocycle associated with and a p.m.p. ergodic group action .
The limit space is also known as a Carnot group and coincides with the Carnot-Carathéodory metric obtained by the asymptotic cone of as the limit of . More details about its construction and properties will be given in Section 2 along with the definitions of and . The usage of the pointed Gromov-Hausdorff convergence arises naturally from its correspondence with geometric group theory.
Let now 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 with finite index . Set to be the torsion subgroup of and define
Pansu [16] showed that and share the same asymptotic cone. Let us fix as a representative of the coset such that . Consider when . Set to be given by for . Define now to be given by
To refine the first main theorem, let us introduce some new conditions. Suppose that there exists such that, for all there is a sequence of positive integers depending on with as ,
(ii′) |
Let by
(1.4) |
Fix, for each , a and consider given by and (see Sec. 3.3 and Remark 5 for a detailed discussion). We consider a similar innerness assumption to replace (iii). Suppose that, for each , there exists a finite which is a generating set of such that, -a.s., for every , one can write with satisfying
(iii′) |
Similar to (iii), First-Passage Percolation models satisfy (iii′) under specific conditions. In the case where is nilpotent, it suffices to have for an FPP model to satisfy (iii′). The virtually nilpotent case is treated separately in Section 3.4 with additional conditions imposed on and . Moreover, when is abelian, hypothesis (iii′) is not required to verify the theorem below.
Theorem 1.2 (Limiting Shape for Groups with Polynomial Growth).
Let be a finitely generated group with polynomial growth rate and torsion-free. Consider to be a subadditive cocycle associated with and a p.m.p. ergodic group action .
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 , 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 be a group generated by a finite set . The associated Cayley graph represents elements of as vertices, with edges connecting and if and only if for some . Formally, the right-invariant Cayley graph is defined by
Cayley graphs provide a visual representation of the group structure and are fundamental in the study of geometric group theory.
Denote by the relation . Let be the set of self-avoiding paths from to , where each follows with , , , , and for all . We write for , and represents the length of the path.
The word length on with respect to is defined as follows: For any , the length of the shortest word (or self-avoiding path) in that represents is its word length, denoted by . The word metric on is given by .
Throughout this article, various distinct metrics will be considered. Therefore, let us consider a (semi-pseudo-quasi) metric on a non-empty set , where the metric is indexed by . We define as the open -ball centered at . To streamline notation, let and . Here, the set stands for and .
A finitely generated group has polynomial growth with respect to when for a as . The growth is associated with the Cayley graph . The polynomial growth rate of is a constant such that there exists for all satisfying
Thus . Moreover, one can verify that the polynomial growth rate of does not depend on the choice of .
Recall the definition of the commutator element and the subgroup for any . Set and let for all . Thus forms a lower central series with for all . The group is called nilpotent when there is an such that , i.e., when the sequence stabilizes in the trivial group. More specifically, is nilpotent of class when is the minimal value such that is the trivial group. A group is abelian if and only if it is nilpotent of class . The abelianization of a group is given by .
The group is called virtually nilpotent when there exists a normal subgroup that is nilpotent with finite index . 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 is denoted by and it is defined as the set of all elements with finite order. In other words, . The group is called torsion-free when is the trivial group.
Let to be the -neighborhood of of in a metric space , i.e., the set . The Hausdorff distance detects the largest variations between sets with respect to the given metric
We define the convergence of metric spaces used in the main theorems employing the Hausdorff distance. Let be a sequence of centered, locally compact metric spaces. Consider as a family of isometric embeddings .
The pointed Gromov-Hausdorff convergence of to is denoted by
and it implies, for all ,
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 be a virtually nilpotent group generated by a symmetric and finite . Then
where is a simply connected real graded Lie group (Carnot group). The metric is a right-invariant sub-Riemannian (Carnot-Caratheodory) metric which is homogeneous with respect to a family of homotheties , i.e., for all and .
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 and related results.
2.2. Rescaled distance and asymptotic cone
Consider for now as a nilpotent and torsion-free group, unless stated otherwise. We also assume that its abelianization is torsion-free. In this subsection, we use instead of to simplify notation, but we will subsequently extend the results to the more general case.
Let denote the real Mal’cev completion of . The group can be defined as the smallest simply connected real Lie group such that and, for all and , there exists with . In this case, is nilpotent of the same order of and it is uniquely defined. Furthermore, is simply connected it is associated with the Lie algebra where is cocompact in . We write for the Lie logarithm map.
Define and . It follows from the nilpotency of that threre exists such that . Thus . Since and, in particular, , the Lie bracket on determines a bilinear map
which in turn defines a Lie bracket on
Consider the decomposition given by . Thus, is a graded Lie algebra. Let us define a family of linear maps given by
for each and with . It follows from the definition of that and for all and . Hence, defines a family of automorphisms in the graded Lie group . Here we write and to differentiate the distinct exponential maps of and . Similarly, and stand for their correspondent Lie logarithm maps.
Let be the decomposition given by . Set to be an linear map such that . Consider now to be the linear automorphism on so that for each and . Define the Lie brackets on by
thus is isomorphic to via . Furthermore,
since, given and , one has that the main term belongs to , the other terms of superior order belong to and it makes them insignificant in the rescaled limit (see [7, 16] for a detailed discussion). Set
The convergence established by Theorem 2.1 determines the metric such that
Hence, exactly when converges to in . The corresponding metric statement shows that, given sequences , in , and as with and ,
The abelianized Lie algebras are defined by and . In particular, . By the Frobenius integrability criterion, the integrable curves in , the tangent vectors must belong to at each point of the curve. An admissible (or curve) in is a Lipschitz curve such that the tangent vector for all . Let be a norm in the abelianized algebra. Then the -length of the admissible is
Set to be the inner metric of the length space given by
(2.1) |
In fact, the construction of can be employed to define . The bi-Lipschitz property is a consequence of the results in Section 2.4. One can also verify that the metric is right-invariant and homogeneous with respect to . Let us define the projections
so that, if with , then and . 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 a finitely generated torsion-free nilpotent group, then all of the following hold true:
-
(i)
Let . Then there exists a sequence such that
-
(ii)
Let , , and as be such that and . Then
-
(iii)
Let , then
Remark 1.
The conditions imposed on might appear somewhat restrictive. However, we will subsequently regain many properties by making necessary adjustments for virtually nilpotent through the quotient (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 nilpotent. Let and be given. Then there exist and so that, for all , for all ,
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 be a commutative ring with identity and set to be the set of upper triangular matrices with

The Heisenberg group on is with the matrix multiplication. In particular, is known as discrete Heisenberg group. Let , , , , and . Observe that
Therefore, for all ,
(2.2) |
One can easily see that is a finite generating set of . Furthermore, and . Hence, is nilpotent of class and . Consider the word norm of . It follows from (2.2) that
It highlights how the rescaled norm vanishes as when .
Due to the properties above, one can write . Note that is a finite generating set of the abelianized group which yields an isomorphism of and the square lattice.
By construction of the asymptotic cone, the Mal’cev completion of is the continuous Heisenberg group with its associated Lie algebra , in this case, and . The Heisenberg algebra is given by with the canonical basis of .
Since for all one has by matrix multiplication, it then follows that with and .
Let , then . Since one can verify by the procedure defined in this section that . It implies that, for all ,
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 be a nilpotent group and consider a finite group. Then the direct product
is a group with the binary operation given by . Note that the commutator is . It follows that, for all and ,
Hence, is a nilpotent group if, and only if, is nilpotent. On the other hand, for all finite group , is virtually nilpotent.
Set and to be finite symmetric generating sets of and , respectively.
is a finite generating set of . We will consider another useful example of generating set of . Let stand for . Then
is also a symmetric generating set of . Furthermore, if is torsion-free, then is analog to while .
Example 2.2.
Let be the of degree two over a field of three elements determined by
A remarkable property of is that it is the smallest group that is not nilpotent. Let the cyclic group with and consider to be the discrete Heisenberg group, as defined in Example 2.1. Set
Then is virtually nilpotent with such that . Hence, considering this notation:
Let us write and fix as representatives for each coset in . Thus,
Now, set
Then
is a finite symmetric generating set of . Moreover, the Cayley graph is homomorphic equivalent to , which is isomorphic to .
More generally, one can also obtain a virtually nilpotent group by the outer semidirect product. Consider a nilpotent and a finite group. Let be a group homomorphism , where is the automorphism group of . Then the semidirect group is
whose elements are the same of but the binary operation is characterized by
Let and be finite symmetric generating sets of and , respectively. Hence, similarly to the direct product,
is a finite symmetric generating set of . Moreover, is also a finite generating set, but not necessarily symmetric. However,
is finite, symmetric, and generates . 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 be a finitely generated abelian group with polynomial growth rate and with . Fix such that and . The generalized virtually nilpotent diheral group is
Consider , then for all ,
Therefore, is non-abelian and . One can easily verify that all elements of are
Hence, for all , one has . We can conclude that is not nilpotent while it is virtually nilpotent since .
2.4. Establishing a Candidate for the Limiting Shape
From this point until Section 3.2, let us once again regard as a torsion-free nilpotent group. Our objective is to characterize the limiting shape using a norm that defines a metric in . 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 because of the constraints imposed by admissible curves.
Set , due to the subadditivity of the cocycle
for all . Thus with . It follows from (ii) that there exists a subsequence of such that -a.s. for sufficiently large .
Recall that and consider , To simplify notation, we also use interchangeably with when it is clear from the context. Let
Since is discrete, there exists such that . Hence, for all , there exist and such that
which implies the subadditivity
Now, regarding whenever , one has for all and , . Let with . Since is a normal subgroup of , is such that . Hence
Therefore, if, and only if, there exists such that . Observe that is a topological lattice of and . Let be an Euclidean norm on and fix such that
Due to the properties of a normed vector space, one has, for all ,
(2.3) |
Set to be given by
It follows immediately from the subadditivity of and the definition of that
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 associated with a p.m.p. ergodic group action be such that for all . Then there exists a unique homogeneous subadditive function such that, for every ,
Moreover, is given by
(2.4) |
Remark 2.
The function obtained above is naturally associated with the abelianized space considering the well-known fact of the convergence of to the projection of onto a subspace isomorphic to . It will allow us to measure distances in with 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 be a subadditive cocycle under the assumptions of Proposition 2.4. Set as in (2.4). Consider satisfying (i) and (ii). Then there exist such that, for all ,
Proof.
We know by Proposition 2.4 that exists and
It follows from (i) and subaditivity that there exists such that, for all ,
Let us fix , then
which is our assertion. ∎
Remark 3.
The Subadditive Ergodic Theorem guarantees the -a.s. existence of the . By combining this fact with previous assertions and the convergence, we obtain the existence of such that
(2.5) |
Recall the definition of in (2.1). Therefore, there is a bi-Lipschitz relation between and . We now define by
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 be a graph and set to be a collection of non-negative random variables. We may regard each as random length (also passage time or weight) of an edge . It turns into a random length space and it motivates the following construction.
The random passage time of a path is given by . Let us now define the first-passage time of with the process starting at by
The random variable is also known as first-hitting time. Observe that is a random intrinsic pseudometric, i.e., does not imply in . We can now consider the group action as a translation such that is a subadditive cocycle (see 1.2) with for all and when .
By requiring to be ergodic, we obtain for the FPP model that, for all and ,
It also follows that . Therefore, each direction of 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 . 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 , then describes the growth of the process, and
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 is a subadditive cocycle, and is finitely generated by the symmetric set with polynomial growth rate . To set the stage, we begin by stating Proposition 3.1 from [8].
Proposition 3.1.
Let be a Lipschitz curve and let . Then there exists so that one can find, for all , and such that, for all ,
Moreover, for a subadditive homogeneous function bi-Lipschitz with respect to , one has that
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 be a torsion-free nilpotent finitely generated group with torsion-free abelianization. Consider a subadditive cocycle associated with an ergodic group action satisfying (i). Then for all integer and ,
In particular, if we let , then there exists, -a.s., a random depending on and on such that, for all ,
Before proving Proposition 3.2 we show the following lemma.
Lemma 3.3.
Let and consider a subadditive cocycle that satisfies condition (i). There exists, -a.s., such that if , , , and are sequences in satisfying, for a and all :
-
(i)
There exist elements and such that
and ;
-
(ii)
.
Then
for all .
Proof.
Fix and . Then and . Observe that and . We thus obtain the -almost surely inequalities below:
(3.1) | ||||
(3.2) |
Observe now that, by items (i) and (ii), for ,
Remark 4.
If are such that and in , 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 , , , , and ,
Set
We now state Lemma 3.6 of [8] without proof before the proving Proposition 3.2.
Lemma 3.4.
Let , , and . Then, for all , there is such that, for ,
We proceed below with the proof of ergodic subadditive approximation via polygonal paths.
Proof of Proposition 3.2.
Consider and fixed. Let and be given by Lemma 2.3 for . Set and sufficiently large so that, for each , one has by Proposition 2.4,
Fix and define inductively so that, for each , is given by Lemma 3.4 satisfying
Therefore,
Let now . Thence, for all and every with , if , then , and
It follows that, for all , there exist non negative integers such that, for all ,
By Lemma 2.3, for each and every ,
Hence, by Lemmas 2.2 and 3.3, there exists with , and a random depending on , and such that for all and all ,
Therefore, for all , every and all
(3.5) |
and . It suffices to consider replacing with , then there exists, -a.s., by Borel-Cantelli Lemma such that (3.5) is satisfied for all , which is our assertion with and . ∎
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 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 and . Consider and given by Proposition 3.1 for a -geodesic curve from to and .
Proof.
Let us write and consider for given by Proposition 3.1. It follows from subadditivity that
Then, one has by Proposition 3.2 with that, -a.s., for all ,
(3.6) |
Set to be defined later and apply condition (iii) to obtain
where with . Define a sequence of piecewise -geodesic curves between each and for such that and
Set . It follows from Lemma 2.5 that and, due the the convergence in Proposition 2.4, there exists such that, for all ,
Fix so that, for all , . By Chernoff bound, . It then follows from an application of Borel-Cantelli Lemma that, -a.s., there exists such that, for every ,
(3.7) |
Let Observe now that, for every , -a.s., for ,
Hence, one has by Arzelà–Ascoli Theorem that a subsequence of converges uniformly to a Lipschitz curve such that and .
We apply Proposition 3.1 once again fot the curve with to obtain , , , and such that, for all
Recall that is a generating set of and shares the polynomial growth rate of . Then there exists such that, for a given ,
as .
It thus follows by an application of Borel-Cantelli Lemma and by (3.7) that for all , there exist, -a.s., and a subdivision function with such that, for all ,
Let
Then there exist and, -a.s., such that, for all , . Hence, for , .
It follows that there exists, -a.s. so that, for every ,
We thus get from Lemma 3.3 that there exists, -a.s., , such that, for each ,
Set . Therefore, one has, -a.s., for all ,
Lemma 3.6.
Let be a sequence in and let be an increasing sequence in such that .
Proof.
Set to be defined later. Consider and given by Proposition 3.1 for and a -geodesic curve from to . Let . Since converges to . It follows from the Borel-Cantelli Lemma applied to (i) that there exists, -a.s., so that, for every ,
(3.9) |
Let us write . Since , one can easily see that there exists such that, for all ,
Then there exists such that, for all , one has .
Let now in (i). Since is identically distributed to , we have by Borel-Cantelli Lemma that there exists, -a.s., such that, for ,
(3.10) |
Consider abelian, then . In fact, it is straightfoward that by the standard approach for commutative groups. Then by Proposition 2.4, there is, -a.s., such that, for all ,
(3.12) |
Furthermore, we have . Combining the two previous inequalities with (3.11), we can establish the result for the commutative case with . Now, let’s consider the non-abelian case, assuming that (iii) holds true. Notably, by Proposition 3.5 with and , for all ,
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 -a.s. asymptotic equivalence given, which is given by
(3.13) |
Suppose, by contradiction, that (3.13) is not true. Consider to be such that as . Let stand for , the closure of the -ball or radius in . Due to the compactness of with respect to , there exists a subsequence such that, for
By construction, is a countable dense subset of . Fix, for each , such that converges to under (see Lemma 2.2). Let be the event with given by Lemma 3.6 for . Hence, is such that .
The compactness of implies the existence of a finite such that covers . Thus there exists so that . Consider as defined above and let to be determined later. Then, there exists so that, for all ,
and .
Let be given by Lemma 3.3 on with satisfying, for all and ,
Set to be such that for all . Hence, for all on ,
which contradicts the above assumption proving that (3.13) holds true.
It remains to show how converges to in the asymptotic cone. Recall that . Consider now any given and a sequence with such that and . Then converges to and converges as above. In particular, one can fix any to find such that for all .
Let us define and . The asymptotic equivalence (3.13) implies the existence of a random for such that, for all on ,
Due to the fact that is a p.m.p. group action, one can repeat all arguments above also in Propositions 3.5 and 3.6 to obtain and for each with and so that, for all converging as above and every on ,
(3.14) |
Let now be a sequence that and choose . Fix so that . By Lemma 3.3, one can find a random and with such that, for all on ,
(3.15) |
where is any convergent sequence . Let us fix
and set
Then is finite on and . It follows from (3.14) and (3.15) that, for all on ,
This establishes the -a.s. convergence of to for as . 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 .
Recall that the nilpotent subgroup has a finite index , and for each coset , we designate a representative . Also, define for all and .
We commence by presenting results concerning the properties of p.m.p. ergodic group actions of with respect to and . We adopt the notation .
Lemma 3.7.
Let be a discrete group and a finite normal subgroup with finite index . Consider that is a p.m.p. ergodic group action. Then there exists a finite such that, for all , and the restriction of on , induces a p.m.p. ergodic group action on . Furthermore, and .
Proof.
Set to be the family of all non-empty -invariant events under . Then, for all ,
which implies . Observe that is closed under countable unions and non-empty countable intersections. Let us fix such that . Define .
Since is a normal subgroup of , acts ergodically on for all and it inherits the measure preserving property. ∎
We use Lemma 3.7 to write with and for each . Let us denote by , the orbit of under the action on . Set
where and is the induced probability measure . Let us fix for each . Define so that
Lemma 3.8.
Let be the set obtained in Lemma 3.7. Then, for each , is a p.m.p. ergodic group action.
Proof.
The measure preserving property is immediately inherited from . Let . Due to the normality of , for all and each ,
Hence, if for all , one has . Then, for all
It follows from the ergodicity of that , which is the desired conclusion. ∎
Remark 5.
Recall that definition (1.4) determines
It is straightforward to see that is compatible with the probability space for each . Futhermore, it is a subadditive cocycle associated with . Additionally, is well defined on . Let and . Consequently, one can investigate on , and the results can be naturally extended -a.s. to .
In preparation for the asymptotic comparison between cocycles and , the following lemmas provide essential insights into their respective properties and relationships.
Lemma 3.9.
Let and consider a subadditive cocycle that satisfies condition (i). Then there exists, -a.s., such that, for all and every ,
Proof.
It follows from subadditivity that, for ,
for every . Let . Hence, one has by (i) and a that
for . The result is derived through the application of the Borel-Cantelli Lemma. ∎
Lemma 3.10.
Let and consider a subadditive cocycle that satisfies condition (i). Then there exists, -a.s., such that, for all and every ,
where and .
Proof.
Since is a normal subgroup of , the exists such that with . Thus
We apply the same reasoning for obtaining that
By (i) and the finitness of , there exists a constant such that
for . The desired conclusion follows from an application of Borel-Cantelli Lemma. ∎
Let us define, for all ,
and
Set
Thus, one has, for all with ,
(3.16) |
By the same arguments employed in Section 2.4, the discrete norm
(3.17) |
exhibits the same properties as when is a generating set of .
Consider to be the sequences fixed for each in the proof of Theorem 1.1. Set with defined by the group action . Then
when is given. Let us write for each . Also, one can easily verify that
The proposition below shows us that and share the same linear asymptotic behaviour.
Proposition 3.11.
Let be a virtually nilpotent group, and let be a subadditive cocycle associated with .
If condition (i) is satisfied, then and are asymptotically equivalent, i.e., there exists, -a.s., such that, for all with ,
(3.18) |
In particular, (i) implies the -a.s. existence of so that, for all and every ,
(3.19) |
Proof.
From Lemmas 3.9 and 3.10, we can deduce that, for every , one can fix so that, -a.s., for all and every ,
The inequality above implies the asymptotic equivalence of and on .
Since is p.m.p. group action, one can obtain from Lemmas 3.9 and 3.10 the random variables and depending on determining
so that (3.19) holds true. ∎
The following result extends the subadditive ergodic theorem to with respect to .
Lemma 3.12.
Consider to be a virtually nilpotent group generated by a finite symmetric set with a generating set of .
Proof.
Therefore, it follows from (3.16) that satisfy (i) and (ii) with respect to for a new and . The proof is complete by replacing with 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 , Theorem 1.1 holds true for on . Therefore, it suffices to extend the results to and compare with .
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 as a dense subset of , the finite . Similarly, we consider and with as and . Note that we may regard to replace the orifinal sequence in the proof of Thm. 1.1 and let and be defined as before with .
Consider with for every . Then Proposition 3.11 ensures the existence of and with so that, for all on ,
(3.21) |
Let now be a sequence such that and choose . Fix so that . Observe that (3.15) is still valid for . Hence, by Lemma 3.3, one can find and with such that, for all on ,
(3.22) |
where is any convergent sequence . Let us fix
and set
Then is finite on and . It follows from (3.20), (3.21), and (3.22) with that, for all on ,
This establishes the -a.s. convergence of to for as . ∎
3.4. An additional result for FPP models
In the preceding sections, we delved into the asymptotic behavior of and . The definition of depends only on the action of restricted to , ensuring that we can systematically investigate the group action of within a fixed .
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 ,
and consider . Note that restricted to is not well-defined when there exists another set distinct from . This inherent limitation prompts the necessity for specific conditions in the subsequent result.
The following lemma outlines the criteria under which inherits the FPP property from . Before presenting this result, we establish the notation:
Lemma 3.13.
Let be a virtually nilpotent group generated by a finite symmetric set with a generating set of . Consider a subadditive cocycle determining a FPP model on which satisfies (i). Suppose that the restriction is a p.m.p. ergodic group action.
If, for all , , then determines a FPP model on and condition (iii′) is satisfied when .
Proof.
Define, for each and every ,
and note that preserves the symmetry
Condition (i) imply that is -a.s. finite and there exists of a (finite) geodesic path. Observe that for all induces a graph homomorphism of and . In other words, for all and , and if in , then in . Hence, one can easily verify by the minimax property that
This is a direct consequence of the graph homomorphism. Property (iii′) arises naturally from the given definition when . ∎
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 with .
Proof.
Notice that
Consequently, , -a.s., as . Therefore, defining is a suitable choice for investigating the asymptotic cone of in comparison to .
The arguments in the proofs of Lemmas 3.9, 3.10, 3.11 and 3.12 can be repeated for , yielding the same properties up to a constant factor. ∎
Corollary 3.15.
Let be a finitely generated group with polynomial growth rate and torsion-free. Consider to be a subadditive cocycle associated with and a p.m.p. ergodic group action .
Suppose that describes a FPP model which satisfies conditions (i) and (ii′) for a finite symmetric generating set so that
-
(i)
For all , , and
-
(ii)
generates .
Then
where is a simply connected graded Lie group, and is a quasimetric homogeneous with respect to a family of homotheties . Moreover, is bi-Lipschitz equivalent to on .
Proof.
First, according to Proposition 3.14, the random variables and share similar properties. Observe that , ensuring that well-defined and a suitable replacement of in the proof of Theorem 1.2, which establishes the result. ∎
The next example highlights a case where 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 exhibits equivariance. Recall properties fiscussed in Section 2.5 for FPP models and notice that, for all and ,
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 and set for , i.e., the function fixes one element of each .
Suppose that, for all , and consider to be the law of with and . Thus, one can write
where, for each , . Let be such that, for all , . Then, for all ,
The condition of polynomial growth rate ensures that is countably infinite. Consequently,
Therefore, 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 a torsion-free nilpotent group with torsion-free abelianization and a symmetric finite generating set . Set to be a finite group. Recall the properties highlighted in Section 2.3. Let us define
Then is a symmetric finite generating set of . Fix for all and . One can easily see that with .
Furthermore, for any , the inverse is given by , leading to
As a consequence, both items (i) and (ii) of Corollary 3.15 hold when is the direct product equipped with the generating set 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 . 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 , 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 . 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 ).
Let us now consider a dependent Bernoulli FPP model based on the random coloring studied by Fontes and Newman [12]. Set to be a family of i.i.d. random variables taking values in a finite set of colors . The model generates color clusters by assigning weight to edges between sites with same color and weight otherwise. We define for every edge
Set for each self-avoinding path the random length . The first-passage time is
Let then one can verify that is a FPP model with dependent identically distributed passage times . One can easily see that is a subadditive cocycle and the translations are ergodic due to the fact that are i.i.d. random variables .
Observe that is bounded above by the word norm , items (i) and (iii) are immediately satisfied. Consider for all . Set
The lemma below establishes a sufficient condition for (ii) and (ii′).
Proof.
Let with the set of all self-avoiding paths in of graph length starting at . Fix , then
where with respect to . Let us regard , thus
It is a well-known fact that . Therefore, there exists such that . By Chernoff bound, one can obtain
(4.2) |
Observe that the base of (4.2) converges to as . Hence, there exist such that with when satisfies (4.1). It then follows that there exists such that
Similarly to Example 3.1, let be the law of the random coloring of a vertex. Then
with . By the same reasoning employed for in Example 3.1, we verify that acts ergodically on .
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 nilpotent with a finite generating set or in the case where 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 is virtually nilpotent.
Remark 6.
Observe that (4.1) provides a lower bound for the critical probability of site percolation on (see for instance [14]). To verify that, fix a color and we say that a site is open when . Therefore, one can write
Note that it stochastically dominates with -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, -a.s., there is no infinite open cluster in when .
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 . Let be the set of directions of . Consider a set of strictly positive random variables that are independent over a probability measure . Set to be a collection of independent random variables over such that
Let us regard as a fixed realization of the random environment. The growth process is defined by the family of independent random variables such that
(4.3) |
Set to be the quenched probability law of (4.3). We write, for each path with , its random length .
The first-passage time is
It is straightforward to see that is subadditive. However, the group action is not ergodic over for a given . Let be the annealed probability. It then follows that preserves the measure and it is ergodic.
Note that 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 such that, for all ,
for all .
Proof.
Let be a -geodesic with . Then one has by Chernoff bound and the independence of that
where
Let and set . Thus, by the Dominated Convergence Theorem,
Therefore, it suffices to choose to complete the proof. ∎
Lemma 4.3.
Consider the RGTRE defined as above. The there exists such that, for all ,
Proof.
It is a well-known fact that, for all and every , that and, therefore, . By the right continuity of the cumulative distribution function, one can find and such that, for every ,
We use similar arguments as those employed in the proof of Lemma 4.1, we may consider over . Then there exists such that, for any ,
It follows that there exists such that, for ,
Since , 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 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 nilpotent with a finite generating set or in scenarios where is abelian. Additionally, when 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 .
Set to be the simple random walk on of a particle originally placed at and let be the first time the random walk visits , i.e., it defined the random variable . Note that with strictly positive probability.
The activation time of the particle originally positioned at is given by the random variable
Observe that and are not necessarily neighbours. We proved in [9] that is a subadditive cocycle with respect to the translation , which is p.m.p. and an ergodic group action. Futhermore, is not identically distributes as in the FPP models (see Section 2.5).
Due to the discrete time random walks, 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 be a group of polynomial growth rate with a symmetric finite generating set . Then there exists and such that, for all and every , one has
Consider now as a group with polynomial growth rate generated by a symmetric finite set . According to Theorem 1.2 and the preceding results, it can be inferred that the Frog Model on exhibits a limiting shape when generates an abelian group . This phenomenon can be exemplified by the generalized dihedral group when is a finitely generated abelian group with polynomial growth rate (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 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 . 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.