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

Universal quasiconformal trees

Efstathios-K. Chrontsios-Garitsis Fotis Ioannidis  and  Vyron Vellis Department of Mathematics
The University of Tennessee
Knoxville, TN 37966
echronts@utk.edu Department of Mathematics
The University of Tennessee
Knoxville, TN 37966
fioannid@vols.utk.edu Department of Mathematics
The University of Tennessee
Knoxville, TN 37966
vvellis@utk.edu
Abstract.

A quasiconformal tree is a doubling (compact) metric tree in which the diameter of each arc is comparable to the distance of its endpoints. We show that for each integer n2n\geq 2, the class of all quasiconformal trees with uniform branch separation and valence at most nn, contains a quasisymmetrically “universal” element, that is, an element of this class into which every other element can be embedded quasisymmetrically. We also show that every quasiconformal tree with uniform branch separation quasisymmetrically embeds into 2\mathbb{R}^{2}. Our results answer two questions of Bonk and Meyer in [BM22], in higher generality, and partially answer one question of Bonk and Meyer in [BM20].

Key words and phrases:
Quasiconformal tree, geodesic tree, quasisymmetric embedding, universal space, Vicsek fractal, continuum self similar tree
2020 Mathematics Subject Classification:
Primary 30L05; Secondary 30L10, 05C05, 28A80
V.V. was partially supported by NSF DMS grant 2154918.

1. Introduction

A collection 𝒳\mathscr{X} of metric spaces is topologically closed if it has the property that if X𝒳X\in\mathscr{X} and YY is a metric space homeomorphic to XX, then Y𝒳Y\in\mathscr{X}. An element X0X_{0} of a topologically closed collection 𝒳\mathscr{X} is universal, if every element of 𝒳\mathscr{X} embeds into X0X_{0}. For example, the standard Cantor set is a universal set for the class of compact metric spaces with topological dimension 0, while the Menger sponge is a universal set for the class of compact metric spaces with topological dimension at most 1. More generally, it is known that for any n0n\in\mathbb{N}_{0}, there exists a universal set for the class of compact metric spaces with topological dimension at most nn [Sta71, Bes84].

Another interesting and well studied topologically closed class is that of metric trees. A metric space XX is called a metric tree (or dendrite) if it does not contain simple closed curves, and it is a Peano continuum, i.e., compact, connected, and locally connected. In other words, any two points x,yx,y can be joined by a unique arc that has x,yx,y as endpoints. On the one hand, trees are among the most simple 1-dimensional connected metric spaces, but on the other hand they have infinitely many topological equivalence classes. Nadler [Nad92] proved that the class of metric trees contains a universal element.

Recently, there has been great interest in a subclass of metric trees, that of quasiconformal trees, which plays an important role in analysis on metric spaces. See, for instance, [BM20, BM22, DV22, DEBV23, FG23a, FG23b] for a non-exhaustive list of references in the topic. A quasiconformal tree is a tree TT that satisfies two simple geometric conditions:

  • TT is doubling: there is a constant N1N\geq 1 such that any ball in TT can be covered by at most NN balls of half its radius, and

  • TT is bounded turning: there is a constant C1C\geq 1 such that each pair of points x,yTx,y\in T can be joined by a continuum whose diameter is at most Cd(x,y)Cd(x,y).

The two conditions above are invariant under quasisymmetric maps, making the class of quasiconformal trees 𝒬𝒞𝒯\mathscr{QCT} quasisymmetrically closed. Quasisymmetric maps are the natural generalization of conformal maps in the abstract metric setting. Roughly speaking, a homeomorphism ff between two connected and doubling metric spaces is quasisymmetric if for three points x,y,zx,y,z in the domain that have comparable mutual distances, the images f(x),f(y),f(z)f(x),f(y),f(z) also have comparable mutual distances; see Section 2 for the precise definition.

Quasiconformal trees were introduced by Kinneberg [Kin17], and the term first appeared in the work of Bonk and Meyer [BM20]. Quasiconformal trees have appeared in several fields of analysis and dynamics. First, if T2T\subset\mathbb{R}^{2} is a tree, then 2T\mathbb{R}^{2}\setminus T is a John domain if and only if TT is a quasiconformal tree [NV91, Theorem 4.5]. Second, Julia sets of semihyperbolic polynomials (e.g., z2+iz^{2}+i) are quasiconformal trees; see for example [CJY94] and [CG93, p.95]. Third, quasiconformal trees TT in 2\mathbb{R}^{2} (often called Gehring trees) were recently characterized by Rohde and Lin [LR18] in terms of the laminations of the conformal map f:𝔻Tf:\mathbb{C}\setminus\mathbb{D}\to\mathbb{C}\setminus T. Fourth, planar quasiconformal trees can be classified in the setting of Kleinian groups [McM98]. Fifth, there exist planar quasiconformal trees (“antenna sets”), whose conformal dimension is not attained by any quasisymmetric map [BT01]; see also [Azz15]. Finally, some planar quasiconformal trees with finitely many branch points are conformally balanced, that is, for every edge, each side has exactly the same harmonic measure with pole at infinity [Bis14].

Quasiconformal trees generalize two well-studied types of spaces. On the one hand, quasiconformal trees that are arcs (i.e., have no branch points) are called quasi-arcs, and have been studied in complex analysis and analysis on metric spaces for decades [GH12]. On the other hand, quasiconformal trees generalize doubling geodesic trees. Geodesic trees are trees in which, for each pair of points x,yx,y, the unique arc joining them has (finite) length equal to the distance of x,yx,y. Thus, in geodesic trees all paths are “straight” (isometric to intervals in the real line), whereas paths in quasiconformal trees may be fractal, like the von Koch snowflake. Geodesic trees (and their non-compact counterparts, the \mathbb{R}-trees) are standard objects of study in geometric group theory [Bes02] and in computer science [LG06]. A theorem of Bonk and Meyer [BM20] states that every quasiconformal tree is quasisymmetrically equivalent to a geodesic doubling tree.

Two of the most well-known and used quasiconformal trees are the continuum self similar tree (CSST) and the Vicsek fractal. See Figure 1. Both are attractors of iterated function systems of similarities in 2\mathbb{R}^{2}, but we omit their formal definitions. The CSST is a trivalent quasiconformal tree that has been studied in [BT21, BM22] and is almost surely homeomorphic to the (Brownian) continuum random tree introduced by Aldous [Ald91a, Ald91b]. The Vicsek fractal, is a 4-valent metric tree of great importance in analysis, probability, and physics; see for instance [Vic83, Met93, HM98, Zho09, CSW11, BC23b, BC23a].

Refer to caption
Refer to caption
Figure 1. The CSST (left) and the Vicsek fractal (right).

Given that quasisymmetric maps “quasi-preserve” many interesting geometric properties, it is a natural question to ask whether there exists a quasiconformal tree which is quasisymmetrically universal for the class 𝒬𝒞𝒯\mathscr{QCT}. In other words, does there exist a quasiconformal tree TT such that every element in 𝒬𝒞𝒯\mathscr{QCT} quasisymmetrically embeds into TT? The answer can easily be seen to be negative. Indeed, if there was a bounded turning tree such that every element of 𝒬𝒞𝒯\mathscr{QCT} quasisymmetrically embeds in, then its valence would be infinite, which implies that it could not have been doubling [BM22, Lemma 3.5]. Consequently, it is natural to pose this question for the more restricted class 𝒬𝒞𝒯(n)\mathscr{QCT}(n), the class of quasiconformal trees of valence at most nn, for some integer n3n\geq 3. Unfortunately, even with this restriction, the answer is still negative already for the smallest class 𝒬𝒞𝒯(3)\mathscr{QCT}(3).

Theorem 1.1.

There exists no quasiconformal tree TT for which every element in 𝒬𝒞𝒯(3)\mathscr{QCT}(3) quasisymmetrically embeds into TT.

The reasoning behind Theorem 1.1 is that, although trees in 𝒬𝒞𝒯(3)\mathscr{QCT}(3) have valence 3, infinitesimally, they can have arbitrarily large valencies. As a result, similarly to the above observation, if there was a bounded turning tree TT in which every element of 𝒬𝒞𝒯(3)\mathscr{QCT}(3) quasisymmetrically embeds, then, infinitesimally, TT would have infinite valence and it could not have been doubling. See Section 3 for the proof.

In [BM22], Bonk and Meyer studied quasiconformal trees where large branches can not be too close. To make this precise, we first define the height of a tree at a branch point. Given a metric tree TT and a point pTp\in T, denote by BT1(p),BT2(p),B_{T}^{1}(p),B_{T}^{2}(p),\dots the components of T{p}T\setminus\{p\}, which we call branches of TT at pp. If there are at least three branches of TT at pp, we say that pp is a branch point. By [BT21, Section 3], only finitely many of the branches of pp can have a diameter exceeding a given positive number. Hence, we can label the branches so that diamBT1(p)diamBT2(p)\operatorname{diam}{B_{T}^{1}(p)}\geq\operatorname{diam}{B_{T}^{2}(p)}\geq\cdots and define the height of pp in TT as

(1.1) HT(p):=diamBT3(p).H_{T}(p):=\operatorname{diam}{B_{T}^{3}(p)}.
Definition 1.2 (Uniform relative separation of branch points).

A metric tree (T,d)(T,d) is said to have uniformly (relatively) separated branch points if there exists C1C\geq 1 such that for all distinct branch points p,qTp,q\in T,

d(p,q)C1min{HT(p),HT(q)}.d(p,q)\geq C^{-1}\min\{H_{T}(p),H_{T}(q)\}.

We denote by 𝒬𝒞𝒯\mathscr{QCT}^{*} the collection of all quasiconformal trees of uniform branch separation, and by 𝒬𝒞𝒯(n)\mathscr{QCT}^{*}(n) the collection of all elements in 𝒬𝒞𝒯\mathscr{QCT}^{*} that have valence at most nn. The class 𝒬𝒞𝒯\mathscr{QCT}^{*} (and each class 𝒬𝒞𝒯(n)\mathscr{QCT}^{*}(n)) is quasisymmetrically closed by [BM22, Lemma 4.3]. Moreover, most examples of quasiconformal trees appearing in analysis [CG93, CJY94, BT01, Bis14, LR18] are of uniform branch separation, making 𝒬𝒞𝒯\mathscr{QCT}^{*} a very natural class.

Bonk and Meyer [BM20, Problem 9.2] asked whether the CSST is a quasisymmetrically universal element for 𝒬𝒞𝒯(3)\mathscr{QCT}^{*}(3). More generally, they asked if 𝒬𝒞𝒯(n)\mathscr{QCT}^{*}(n) contains a quasisymmetrically universal element and whether there exists a bounded turning tree in which we can quasisymmetrically embed every element of 𝒬𝒞𝒯\mathscr{QCT}^{*} [BM22, Problem 9.3]. In our main theorem below, we answer all these questions.

Theorem 1.3.

For each ϵ>0\epsilon>0 there exists a nested family of geodesic, self-similar, and Ahlfors regular metric trees 𝕋2𝕋3𝕋4\mathbb{T}^{2}\subset\mathbb{T}^{3}\subset\mathbb{T}^{4}\subset\cdots with the following properties.

  1. (i)

    Each 𝕋n\mathbb{T}^{n} is quasisymmetrically equivalent to a quasiconvex tree in 2\mathbb{R}^{2}. In particular, 𝕋2\mathbb{T}^{2} is isometric to the segment [0,1][0,1], 𝕋3\mathbb{T}^{3} is quasisymmetrically equivalent to the CSST, and 𝕋4\mathbb{T}^{4} is quasisymmetrically equivalent to the Vicsek fractal.

  2. (ii)

    The sequence of trees (𝕋n)n=2(\mathbb{T}^{n})_{n=2}^{\infty} converges in the Gromov-Hausdorff sense to a geodesic tree 𝕋\mathbb{T}^{\infty} whose Hausdorff dimension is at most 1+ϵ1+\epsilon.

  3. (iii)

    For each n3n\geq 3, 𝕋n\mathbb{T}^{n} is quasisymmetrically universal in 𝒬𝒞𝒯(n)\mathscr{QCT}^{*}(n).

Recall that quasiconvexity is a notion slightly weaker than geodesicity: a metric space is quasiconvex if any two points can be joined by a path with length at most a fixed multiple of the distance of the points. Since Euclidean spaces do not contain geodesic trees (other than line segments), quasiconvexity is the closest to geodesicity in the Euclidean setting.

A few remarks are in order. First, each 𝕋n\mathbb{T}^{n} is universal in the class of all metric trees that have valence at most nn [Cha80]. Second, 𝕋\mathbb{T}^{\infty} is homeomorphic to the universal tree of Nadler [Nad92] and, hence, is itself a universal tree in the class of all metric trees.

Claims (i) and (iii) of Theorem 1.3 imply that the CSST is quasisymmetrically universal in 𝒬𝒞𝒯(3)\mathscr{QCT}^{*}(3) and that the Vicsek fractal is quasisymmetrically universal in 𝒬𝒞𝒯(4)\mathscr{QCT}^{*}(4). The former answers [BM22, Problem 9.2]. Claims (ii) and (iii) of Theorem 1.3 imply that every tree in 𝒬𝒞𝒯(n)\mathscr{QCT}^{*}(n) quasisymmetrically embeds into 𝕋\mathbb{T}^{\infty}, which answers [BM22, Problem 9.3].

Corollary 1.4.

If TT is a quasiconformal tree with uniformly separated branch points, then TT quasisymmetrically embeds into 𝕋\mathbb{T}^{\infty}.

By the doubling property, every quasiconformal tree embeds quasisymmetrically into some n\mathbb{R}^{n} (see for instance [Hei01, Theorem 12.1]). On the other hand, no quasiconformal tree (unless it is an arc) can be embedded into \mathbb{R}. Thus, a very natural question of Bonk and Meyer [BM20, Section 10] asks whether every quasiconformal tree admits a quasisymmetric embedding into 2\mathbb{R}^{2}, and whether the embedded image is quasiconvex. A positive answer to this question could be instrumental in the study of quasiconformal trees as Julia sets of holomorphic or uniformly quasiregular maps of 2\mathbb{R}^{2}. Theorem 1.3 gives a partial answer to this question.

Corollary 1.5.

If TT is a quasiconformal tree with uniformly separated branch points, then TT quasisymmetrically embeds into 2\mathbb{R}^{2} and the image is quasiconvex.

As mentioned in Theorem 1.3, 𝕋2\mathbb{T}^{2} is a line segment. The precise construction of trees 𝕋n\mathbb{T}^{n} with n3n\geq 3 is given in Section 4, but we briefly describe an equivalent way of constructing them, motivated by the methods in [BM22, Section 9]. Fix an integer n3n\geq 3 and numbers 12a3an>0\frac{1}{2}\geq a_{3}\geq\cdots\geq a_{n}>0. Let T1T_{1} be a line segment of length 1 and let pp be the midpoint. We geodesically glue n2n-2 many segments at pp of lengths a3,,ana_{3},\dots,a_{n} (see Definition 11.2), and equip the resulting space T2T_{2} with the path metric. Note that T2T_{2} is a geodesic tree with 1 branch point and nn segments of diameters 12,12,a3,,an\frac{1}{2},\frac{1}{2},a_{3},\dots,a_{n}. We repeat the same process to each of the nn segments to obtain T3T_{3}, and we keep iterating the construction infinitely. It is easy to see that each TiT_{i} is a geodesic tree that isometrically embeds into Ti+1T_{i+1}. The Gromov-Hausdorff limit of the sequence (Ti)i(T_{i})_{i\in\mathbb{N}} is the space 𝕋n\mathbb{T}^{n}. Despite the simplicity of this process, many of the desired properties for 𝕋n\mathbb{T}^{n} seem much harder to establish this way. On the other hand, the equivalent construction of 𝕋n\mathbb{T}^{n} described in Section 4 allows the use of various tools from [DV22], significantly shortening many arguments (see, for instance, Lemma 4.8, Lemma 4.9, Section 5, Section 8).

The proof of Theorem 1.3 comprises of two main steps. The first step is a quasisymmetric uniformization theorem for quasiconformal trees with uniform branching in the spirit of [BM22, Theorem 1.4]. Before we state it, we require certain definitions.

Definition 1.6 (Uniform relative density of branch points).

A metric tree (T,d)(T,d) is said to have uniformly (relatively) dense branch points if there exists C1C\geq 1 such that for all x,yTx,y\in T, xyx\neq y, there exists a branch point p[x,y]p\in[x,y] with

HT(p)C1d(x,y).H_{T}(p)\geq C^{-1}d(x,y).
Definition 1.7 (Uniform branch growth).

A metric tree (T,d)(T,d) is said to have uniform branch growth if there exists C1C\geq 1 such that for all branch points pTp\in T and all i3i\geq 3, we have

diamBTi(p)C1diamBT3(p).\operatorname{diam}{B^{i}_{T}(p)}\geq C^{-1}\operatorname{diam}{B^{3}_{T}(p)}.

Definitions 1.2, 1.6, and 1.7 combined lead to the concept of uniform branching.

Definition 1.8 (Uniformly nn-branching trees).

Given n{3,4,}n\in\{3,4,\dots\}, a tree is called uniformly nn-branching if it has uniform branch growth, uniformly relatively separated branch points, uniformly relatively dense branch points, and all of its branch points have valence equal to nn.

For completeness, we say that a quasiconformal tree TT is uniformly 22-branching if it is an arc. It can be shown that for each n2n\geq 2, the quasiconformal tree 𝕋n\mathbb{T}^{n} is uniformly nn-branching. The main result of [BM22] states that a quasiconformal tree is quasisymmetric to the CSST if, and only if, it is uniformly 3-branching. We also prove the analogue of this result for all valencies, where CSST is replaced by 𝕋n\mathbb{T}^{n}. Note that in the case n=3n=3, the uniform branch growth condition is trivially true, while it is essential for n4n\geq 4, as a quasisymmetric invariant property of 𝕋n\mathbb{T}^{n}.

Theorem 1.9.

Let n{2,3,}n\in\{2,3,\dots\}. A quasiconformal tree TT is uniformly nn-branching if, and only if, it is quasisymmetrically equivalent to 𝕋n\mathbb{T}^{n}.

The second ingredient in the proof of Theorem 1.3 is the construction of three quasisymmetric embeddings in Section 11. First, we show that each quasiconformal tree with uniform branch separation quasisymmetrically embeds into a quasiconformal tree with uniform branch separation and uniform branch growth. Second, we show that each quasiconformal tree with uniform branch separation and uniform branch growth quasisymmetrically embeds into a quasiconformal tree with uniform branch separation, uniform branch growth, and all of its branch points have the same valence nn. Finally, we employ a tiling of quasiconformal trees from [BM20] to show that each quasiconformal tree with uniform branch separation, uniform branch growth, and whose branch points all have valence nn, quasisymmetrically embeds into a uniformly nn-branching tree.

The work of this paper motivates a couple of questions. First, as mentioned earlier, the class 𝒬𝒞𝒯\mathscr{QCT} cannot have a quasisymmetrically universal element, due to such an element not being doubling. One can then drop the doubling requirement and study the more general class of bounded turning trees, which is quasisymmetrically closed, and also contains all geodesic trees. In this class, we ask whether there exists a universal element. It is easy to see that if such an element exists, it can not be 𝕋\mathbb{T}^{\infty}, as 𝕋\mathbb{T}^{\infty} has uniform branch separation, and every metric tree that quasisymmetrically embeds therein must also have uniform branch separation.

Question 1.10.

Does there exist a bounded turning metric tree in which all bounded turning trees quasisymmetrically embed? Does every bounded turning metric tree with uniform branch separation quasisymmetrically embed into 𝕋\mathbb{T}^{\infty}?

In the absence of the doubling property, Question 1.10 can also be studied for the larger class of weakly quasisymmetric maps (see Section 2).

The second question is concerned with the bi-Lipschitz version of universality. It was proved in [DEBV23] by David, Eriksson-Bique, and the third author that every quasiconformal tree bi-Lipschitz embeds in some Euclidean space M\mathbb{R}^{M}, where MM only depends on the doubling and bounded turning constants of the tree (see also [LNP09, GT11]). It is natural to consider the same question, where the Euclidean target space is replaced by a suitable model tree. Note that, contrary to quasisymmetric mappings, bi-Lipschitz mappings preserve all notions of dimension (Hausdorff, Minkowski, Assouad). Therefore, one has to consider a class of spaces of uniformly bounded dimension (see Lemma 2.5).

Question 1.11.

Let 𝒢𝒯(n,c)\mathscr{GT}(n,c) be the class of all geodesic trees that have valencies at most nn and uniform branch separation with constant cc. Does there exist an element T0𝒢𝒯(n,c)T_{0}\in\mathscr{GT}(n,c) in which every element of 𝒢𝒯(n,c)\mathscr{GT}(n,c) bi-Lipschitz embeds? If such a T0T_{0} exists, is it bi-Lipschitz homeomorphic to 𝕋n\mathbb{T}^{n}?

We note that the class 𝒢𝒯(n,c)\mathscr{GT}(n,c) is not bi-Lipschitz closed, but one could instead consider the “bi-Lipschitz closure” of this class, i.e., all metric trees that are bi-Lipschitz equivalent to some tree in 𝒢𝒯(n,c)\mathscr{GT}(n,c). We also note that, in addition to the question of the quasisymmetric embedabbility of all quasiconformal trees in 2\mathbb{R}^{2}, it is unknown and an interesting problem whether all quasiconformal trees with Assouad dimension less than 2 bi-Lipschitz embed into 2\mathbb{R}^{2} [Kin17, Question 1].

Outline of the paper

In Section 2 we present preliminary definitions and results. In Section 3, employing the notions of metric tangents by Gromov, we prove Theorem 1.1.

In Section 4, following the language and techniques from [DV22], we construct for each n{}n\in\mathbb{N}\cup\{\infty\} and for each choice of weights 𝐚=(a1,a2,)\operatorname{\bf{a}}=(a_{1},a_{2},\dots) (with 12=a1=a2a3\frac{1}{2}=a_{1}=a_{2}\geq a_{3}\geq\cdots) a self-similar metric tree 𝕋n,𝐚\mathbb{T}^{n,\operatorname{\bf{a}}}. In Section 5 we show that the trees 𝕋n,𝐚\mathbb{T}^{n,\operatorname{\bf{a}}} are geodesic, and we show that the sequence (𝕋n,𝐚)n2(\mathbb{T}^{n,\operatorname{\bf{a}}})_{n\geq 2} converges in the Gromov-Hausdorff sense to 𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}}.

In Section 6 we study the branch points of trees 𝕋n,𝐚\mathbb{T}^{n,\operatorname{\bf{a}}} and show that 𝕋n,𝐚\mathbb{T}^{n,\operatorname{\bf{a}}} are uniformly nn-branching. In Section 7 we show that the Vicsek fractal is uniformly 44-branching. In Section 8 we show that the trees 𝕋n,𝐚\mathbb{T}^{n,\operatorname{\bf{a}}} are Ahlfors regular (which implies doubling), and estimate the dimension of the limit tree 𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}}. The proof of a quantitative version of Theorem 1.9 is given in Section 9, by following the techniques from [BT21] and [BM22]. In Section 10 we show that trees 𝕋n,𝐚\mathbb{T}^{n,\operatorname{\bf{a}}} quasisymmetrically embed into 2\mathbb{R}^{2}, with the embedded image being a quasiconvex tree.

In Section 11 we prove that every quasiconformal tree with uniformly separated branch points and valence at most nn quasisymmetrically embeds into 𝕋n,𝐚\mathbb{T}^{n,\operatorname{\bf{a}}}. Finally, in Section 12 we prove Theorem 1.3 and Theorem 1.9, by putting together all the results from the earlier sections.

2. Preliminaries

We write 0={0}.\mathbb{N}_{0}=\mathbb{N}\cup\{0\}. Given two non-negative quantities AA and BB, we write ABA\lesssim B if there is a comparability constant C=C()C=C(\lesssim) such that ACBA\leq CB. Similarly, we write ABA\gtrsim B if there is C=C()C=C(\gtrsim) such that AB/CA\geq B/C. If ABA\lesssim B and ABA\gtrsim B we write ABA\simeq B.

Given a set EE in a metric space XX and r>0r>0, we write

Nr(E)={xX:dist(x,E)<r}.N_{r}(E)=\{x\in X:\operatorname{dist}(x,E)<r\}.

2.1. Convergence of metric spaces

Recall that the Hausdorff distance between two subsets X1,X2X_{1},X_{2} of the same space XX is defined by

distH(X1,X2):=inf{r>0:X2Nr(X1) and X1Nr(X2)}\operatorname{dist}_{H}(X_{1},X_{2}):=\inf\{r>0:X_{2}\subset N_{r}(X_{1})\text{ and }X_{1}\subset N_{r}(X_{2})\}

The Gromov-Hausdorff distance of two metric spaces XX and YY is defined by

distGH(X,Y):=inff,g{distH(f(X),g(Y))}\operatorname{dist}_{GH}(X,Y):=\inf_{f,g}\{\operatorname{dist}_{H}(f(X),g(Y))\}

where f:XZf:X\to Z and g:YZg:Y\to Z are isometric embeddings into some ambient metric space ZZ. A sequence of compact metric spaces (Xn)n=1(X_{n})_{n=1}^{\infty} converges in the Gromov-Hausdorff sense to a compact metric space XX (and we write XnGHXX_{n}\xrightarrow{GH}X) if distGH(Xn,X)0\operatorname{dist}_{GH}(X_{n},X)\to 0 as nn\to\infty. Notice that the limit space XX is unique up to isometries. If the spaces XnX_{n} and XX are in the same ambient space for all nn\in\mathbb{N}, then distH(Xn,X)0\operatorname{dist}_{H}(X_{n},X)\to 0 implies XnGHXX_{n}\xrightarrow{GH}X.

A map f:XYf:X\to Y is called an ϵ\epsilon-isometry if

dist(f):=sup{|dX(x1,x2)dY(f(x1),f(x2)|:x1,x2X}ϵ.\operatorname{dist}(f):=\sup\{|d_{X}(x_{1},x_{2})-d_{Y}(f(x_{1}),f(x_{2})|:x_{1},x_{2}\in X\}\leq\epsilon.

A pointed metric space is a triple (X,p,d)(X,p,d), where (X,d)(X,d) is a metric space with a base point pXp\in X. The pointed Gromov-Hausdorff convergence is an analog of Gromov-Hausdorff convergence appropriate for non-compact spaces. A sequence of pointed metric spaces (Xn,pn,dn)(X_{n},p_{n},d_{n}) converges in the pointed Gromov-Hausdorff sense to a complete pointed metric space (X,p,dX)(X,p,d_{X}), and we write

(Xn,pn,dn)GH(X,p,d),(X_{n},p_{n},d_{n})\xrightarrow{GH}(X,p,d),

if for every r>0r>0 and every ϵ(0,r)\epsilon\in(0,r) there exists n0n_{0}\in\mathbb{N} such that for every integer n>n0n>n_{0} there exists an ϵ\epsilon-isometry f:B(pn,r)Xf:B(p_{n},r)\to X with f(pn)=pf(p_{n})=p and B(p,rϵ)Nϵ(f(B(pn,r)))B(p,r-\epsilon)\subset N_{\epsilon}(f(B(p_{n},r))).

Recall that if a metric space XX is doubling with a doubling constant C1C\geq 1, we say that XX is CC-doubling. The next lemma gives a relation between the doubling constants of the metric spaces and their pointed Gromov-Hausdorff limit. For the proof see [BBI01, Theorem 8.1.10].

Lemma 2.1.

Let (Xn,pn,dXn)(X_{n},p_{n},d_{X_{n}}) be a sequence of CC-doubling spaces. Then, the pointed Gromov-Hausdorff limit is also CC-doubling.

2.2. Self-similarity

We say that a compact metric space XX is self-similar if there exists a finite collection of similarity maps {ϕi:XX}i=1,,k\{\phi_{i}:X\to X\}_{i=1,\dots,k} such that X=i=1kϕi(X)X=\bigcup_{i=1}^{k}\phi_{i}(X), and we call XX the attractor of the collection. Moreover, we say that XX satisfies the open set condition if there exists a nonempty open set UXU\subset X such that ϕi(U)U\phi_{i}(U)\subset U and ϕi(U)ϕj(U)=\phi_{i}(U)\cap\phi_{j}(U)=\emptyset for all distinct i,j{1,,k}i,j\in\{1,\dots,k\}. Throughout the paper, all self-similar sets are assumed to satisfy the open set condition, even if not explicitly stated.

2.3. Metric trees

We say a metric space TT is a (metric) tree if TT is compact, connected, and locally connected, with at least two distinct points, and such that for any x,yTx,y\in T there is a unique arc in TT with endpoints x,yx,y. We denote this unique arc by [x,y][x,y]. We say a point pTp\in T has valence equal to m{}m\in\mathbb{N}\cup\{\infty\} and write Val(p,T)=m\operatorname{Val}(p,T)=m if the complement T{p}T\setminus\{p\} has mm connected components, called branches of TT at pp. If Val(p,T)=1\operatorname{Val}(p,T)=1 we say pp is a leaf of TT, if Val(p,T)=2\operatorname{Val}(p,T)=2 we say pp is a double point of TT, and if Val(p,T)3\operatorname{Val}(p,T)\geq 3 we say pp is a branch point of TT. We define the valence of a tree TT by

Val(T):=sup{Val(p,T):pT}.\operatorname{Val}(T):=\sup\{\operatorname{Val}(p,T):\,p\in T\}.

If TT has no branch points, we say that TT is a metric arc. If TT has at least one branch point and there exists m{3,4,}m\in\{3,4,\dots\} such that Val(p,T)=m\operatorname{Val}(p,T)=m for all branch points pp of TT, then we say that TT is mm-valent.

A subset SS of a tree TT is called subtree of TT if SS equipped with the restricted metric of TT is also a tree. By [BT21, Lemma 3.3], SS is a subtree of TT if and only if SS contains at least 2 points and it is closed and connected.

The following lemma is used towards the proof of Theorem 1.9. It is an analogue of [BT21, Theorem 5.4].

Lemma 2.2.

Let SS and SS^{\prime} be mm-valent trees, m3m\geq 3, with dense branch points. Moreover, suppose that p1,p2,p3p_{1},p_{2},p_{3} and q1,q2,q3q_{1},q_{2},q_{3} are three distinct leaves of SS and SS^{\prime} respectively. Then there exists a homeomorphism f:SSf:S\to S^{\prime} such that f(pk)=qkf(p_{k})=q_{k} for k=1,2,3k=1,2,3.

The following proposition is needed in the proof of Lemma 2.2.

Proposition 2.3.

[BT21, Proposition 2.1] Let (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}) be compact metric spaces. Suppose that for each nn\in\mathbb{N}, the spaces X,YX,Y admit decompositions X=i=1MnXn,i,Y=i=1MnYn,iX=\bigcup_{i=1}^{M_{n}}X_{n,i},Y=\bigcup_{i=1}^{M_{n}}Y_{n,i} as finite union of non-empty compact subsets Xn,i,Yn,iX_{n,i},Y_{n,i}, i=1,,Mni=1,\dots,M_{n}\in\mathbb{N}, with the following properties for all n,in,i and jj:

  1. (1)

    Each set Xn+1,jX_{n+1,j} is the subset of some set Xn,iX_{n,i} and each Yn+1,jY_{n+1,j} is the subset of some set Yn,iY_{n,i}.

  2. (2)

    Each set Xn,iX_{n,i} is equal to the union of some of the sets Xn+1,jX_{n+1,j} and each set Yn,iY_{n,i} is equal to the union of some of the sets Yn+1,jY_{n+1,j}.

  3. (3)

    We have that max1iMndiam(Xn,i)0\max_{1\leq i\leq M_{n}}\operatorname{diam}(X_{n,i})\to 0 and max1iMndiam(Yn,i)0\max_{1\leq i\leq M_{n}}\operatorname{diam}(Y_{n,i})\to 0 as nn\to\infty.

Moreover, assume that Xn+1,jXn,iX_{n+1,j}\subset X_{n,i} if and only if Yn+1,jYn,iY_{n+1,j}\subset Y_{n,i} and Xn,iXn,jX_{n,i}\cap X_{n,j}\neq\emptyset if and only if Yn,iYn,jY_{n,i}\cap Y_{n,j}\neq\emptyset for all n,i,jn,i,j.

Then there exists a unique homeomorphism f:XYf:X\to Y such that f(Xn,i)=Yn,if(X_{n,i})=Y_{n,i} for all nn and ii.

Proof of Lemma 2.3.

We need the terminology on marked leaves from [BT21]. Fix a set A={1,,m}A=\{1,\dots,m\} and let TT be an arbitrary metric tree. Suppose we have defined subtrees TuT_{u} of TT for all levels nn\in\mathbb{N} and all uAnu\in A^{n}. The boundary Tu\partial T_{u} of TuT_{u} in TT will consist of one or two points that are leaves of TuT_{u} and branch points of TT. We consider each point in Tu\partial T_{u} as a marked leaf in TuT_{u} and will assign to it an appropriate sign - or ++ so that if there are two marked leaves in TuT_{u}, then they carry different signs. Accordingly, we refer to the points in Tu\partial T_{u} as the signed marked leaves of TuT_{u}. The same point may carry different signs in different subtrees. We write pp^{-} if a marked leaf pp of TuT_{u} carries the sign -, and p+p^{+} if it carries the sign ++. If TuT_{u} has exactly one marked leaf, we call TuT_{u} a leaf-tile, and if there are two marked leaves we call TuT_{u} an arc-tile.

We now describe how to define subtrees of SS of level 11. Let p1p_{1} carry the sign -, while p2p_{2} and p3p_{3} carry the sign ++. We create a decomposition of SS and a decomposition of SS^{\prime} satisfying the conditions of Proposition 2.3, which provides the desired homeomorphism. In particular, choose a branch point cSc\in S so that the leaves p1,p2,p3p_{1},p_{2},p_{3} lie in distinct branches S1,S2,S3S_{1},S_{2},S_{3} of SS respectively. To find such a branch point, one travels from p1p_{1} along [p1,p2][p_{1},p_{2}] until one first meets [p2,p3][p_{2},p_{3}] in a point cc. Then the sets [p1,c),[p2,c),[p3,c)[p_{1},c),\ [p_{2},c),\ [p_{3},c) are pairwise disjoint.

For k,l{1,2,3}k,l\in\{1,2,3\} with klk\neq l the set [pk,c){c}(c,pl][p_{k},c)\cup\{c\}\cup(c,p_{l}] is an arc with endpoints pkp_{k} and plp_{l}, and so it must agree with [pk,pl][p_{k},p_{l}]. In particular, c[pk,pl]c\in[p_{k},p_{l}]. Since each point pkp_{k} is a leaf, it easily follows from [BT21, Lemma 3.2(iii)] that cp1,p2,p3c\neq p_{1},p_{2},p_{3}. We conclude that the connected sets [p1,c),[p2,c),[p3,c)[p_{1},c),\ [p_{2},c),\ [p_{3},c) are non-empty and must lie in different branches S1,S2,S3S_{1},S_{2},S_{3} of SS at cc. Hence, the point cc has at least 3 branches, and since SS is an mm-valent tree, cc has to be a branch point of SS with a total of mm branches. We can choose the labels so that pkSkp_{k}\in S_{k} for k{1,2,3}k\in\{1,2,3\}. We then have the set of marked leaves {p1,c+}\{p_{1}^{-},c^{+}\} in S1S_{1}, {c,p2+}\{c^{-},p_{2}^{+}\} in S2S_{2}, and {c,p3+}\{c^{-},p_{3}^{+}\} in S3S_{3}.

We now continue inductively as in [BT21, Section 5, pp. 178-182]. If we already constructed a subtree SuS_{u} for some nn\in\mathbb{N} and uAnu\in A^{n} with one or two signed marked leaves, then we decompose SuS_{u} into mm branches labeled Su1,,SumS_{u1},\dots,S_{um} by using a suitable branch point cSuc\in S_{u}. Namely, if SuS_{u} is a leaf-tile and has one marked leaf aSua\in S_{u}, we choose a branch point cSu{a}c\in S_{u}\setminus\{a\} with maximal height HS(c)H_{S}(c). If SuS_{u} is an arc-tile with two marked leaves {a,b}Su\{a,b\}\subset S_{u} we choose a branch point cSuc\in S_{u} of maximal height on (a,b)Su(a,b)\subset S_{u}.

A marked leaf xx^{-} of SuS_{u} is passed to Su1S_{u1} with the same sign. Similarly, a marked leaf x+x^{+} of SuS_{u} is passed to Su2S_{u2} with the same sign. If we continue in this manner, we obtain a subtree SuS_{u} with one or two signed marked leaves for all levels nn\in\mathbb{N} and uAnu\in A^{n}.

We apply the same procedure for the tree SS^{\prime} and its leaves q1,q2,q3q_{1},q_{2},q_{3}. Then we apply [BT21, Lemmas 5.1, 5.2, 5.3] to the decompositions SS and SS^{\prime} obtained this way. Hence, by Proposition 2.3 there exists a homeomorphism f:SSf:S\to S^{\prime} such that

(2.1) f(Su)=Sufor all n and uAn.f(S_{u})=S_{u}^{\prime}\qquad\text{for all $n\in\mathbb{N}$ and $u\in A^{n}$.}

Note that p1p_{1} carries the sign - to S1(n)S_{1^{(n)}} for all nn\in\mathbb{N}, and since by [BT21, Lemma 5.3] the diameters of our subtrees SuS_{u}, uAnu\in A^{n}, approach 0 uniformly as nn\to\infty, we have {p1}=n=1S1(n)\{p_{1}\}=\bigcap_{n=1}^{\infty}S_{1^{(n)}}. The same argument shows that {q1}=n=1S1(n)\{q_{1}\}=\bigcap_{n=1}^{\infty}S_{1^{(n)}}^{\prime}. Hence, (2.1) implies that f(p1)=q1f(p_{1})=q_{1}.

Similarly, the points p2,p3,q2,q3p_{2},p_{3},q_{2},q_{3} carry the sign ++ in their respective trees. Therefore, by applying the same arguments we have that f(p2)=q2f(p_{2})=q_{2} and f(p3)=q3f(p_{3})=q_{3}. The proof is complete. ∎

Remark 2.4.

[BT21, Theorem 5.4] is stated only for 33-valent trees and the difference in their inductive decomposition ([BT21, Section 5, pp. 178-182]) is only in the basis step. In our case, the same techniques apply for mm-valent trees, as long as we map three leaves of the tree SS to three leaves of the tree SS^{\prime}.

The next lemma, which is extensively used in Section 11, demonstrates how geodesicity and uniform separation on trees relate to the doubling condition.

Lemma 2.5.

Given N{2,3,}N\in\{2,3,\dots\}, if TT is a geodesic metric tree with Val(T)N\operatorname{Val}(T)\leq N and uniform branch separation with constant C1C\geq 1, then there is D=D(N,C)>0D=D(N,C)>0 such that TT is DD-doubling.

Proof.

Let x0Tx_{0}\in T and let r>0r>0. We construct 3 finite sets 𝒮1,𝒮2,𝒮3B¯(x0,r)\mathcal{S}_{1},\mathcal{S}_{2},\mathcal{S}_{3}\subset\overline{B}(x_{0},r).

For each xTx\in T with d(x,x0)rd(x,x_{0})\geq r, let pxp_{x} be the unique point of [x,x0][x,x_{0}] such that d(px,x0)=3r/4d(p_{x},x_{0})=3r/4. Let 𝒮1={px:xT,d(x,x0)r}\mathcal{S}_{1}=\{p_{x}:x\in T,\,d(x,x_{0})\geq r\}. Let q1,,ql𝒮1q_{1},\dots,q_{l}\in\mathcal{S}_{1} be distinct points and let KK be the convex hull of {x0,q1,,ql}\{x_{0},q_{1},\dots,q_{l}\} in TT (in the geodesic sense). Then x0Kx_{0}\in K and diamK\operatorname{diam}{K} is either 3r/43r/4 or 3r/23r/2; the former being true exactly when x0x_{0} is a leaf of KK. Without loss of generality, we assume for the rest of the proof that diamK=3r/2\operatorname{diam}{K}=3r/2 and, consequently, that d(q1,q2)=3r/2d(q_{1},q_{2})=3r/2. For each i{3,,l}i\in\{3,\dots,l\}, let zi[q1,q2]z_{i}\in[q_{1},q_{2}] be the point with [qi,x0][q1,q2]=[zi,x][q_{i},x_{0}]\cap[q_{1},q_{2}]=[z_{i},x]. Note that each ziz_{i} is a branch point with HT(zi)r/4H_{T}(z_{i})\geq r/4. Since all ziz_{i} are on the geodesic [q1,q2][q_{1},q_{2}] of length 3r/23r/2, by uniform branch separation we have that card{z3,,zl}6C+1\operatorname{card}\{z_{3},\dots,z_{l}\}\leq 6C+1. Note that for i,j{3,,l}i,j\in\{3,\dots,l\} with iji\neq j, we do not necessarily have that zizjz_{i}\neq z_{j}, hence card{z3,,zl}l\operatorname{card}\{z_{3},\dots,z_{l}\}\leq l. However, each ziz_{i} can have at most NN branches, so it follows that

l=card{q1,,ql}(N2)card{z3,,zl}+2(N2)(6C+1)+2.l=\operatorname{card}\{q_{1},\dots,q_{l}\}\leq(N-2)\operatorname{card}\{z_{3},\dots,z_{l}\}+2\leq(N-2)(6C+1)+2.

Since ll was the cardinality of an arbitrary subset of distinct points in 𝒮1\mathcal{S}_{1}, the above proves that

card𝒮1(N2)(6C+1)+2.\operatorname{card}{\mathcal{S}_{1}}\leq(N-2)(6C+1)+2.

For each xTx\in T with d(x,x0)34rd(x,x_{0})\geq\frac{3}{4}r, let pxp_{x}^{\prime} be the unique point in [x0,x][x_{0},x] with d(px,x0)=r/2d(p_{x}^{\prime},x_{0})=r/2. Let also 𝒮2={px:xT,d(x,x0)3r/4}\mathcal{S}_{2}=\{p_{x}^{\prime}:x\in T,\,d(x,x_{0})\geq 3r/4\}. Working as in the preceding paragraph, we can show that

card𝒮2(N2)(6C+1)+2.\operatorname{card}{\mathcal{S}_{2}}\leq(N-2)(6C+1)+2.

Finally, for each xTx\in T with d(x,x0)r/2d(x,x_{0})\geq r/2, let px′′p_{x}^{\prime\prime} be the unique point in [x0,x][x_{0},x] with d(px′′,x0)=r/4d(p_{x}^{\prime\prime},x_{0})=r/4. Let also 𝒮3={px′′:xT,d(x,x0)r/2}\mathcal{S}_{3}=\{p_{x}^{\prime\prime}:x\in T,\,d(x,x_{0})\geq r/2\}. As before,

card𝒮3(N2)(6C+1)+2.\operatorname{card}{\mathcal{S}_{3}}\leq(N-2)(6C+1)+2.

Define now 𝒮={x0}𝒮1𝒮2𝒮3\mathcal{S}=\{x_{0}\}\cup\mathcal{S}_{1}\cup\mathcal{S}_{2}\cup\mathcal{S}_{3}. We claim that for all xB¯(x0,r)x\in\overline{B}(x_{0},r), there exists p𝒮p\in\mathcal{S} such that d(x,p)r/2d(x,p)\leq r/2. To this end, fix xB¯(x0,r)x\in\overline{B}(x_{0},r). If d(x,x0)r/2d(x,x_{0})\leq r/2, then we can choose p=x0p=x_{0}. If r/2<d(x,x0)<3r/4r/2<d(x,x_{0})<3r/4, then d(x,px′′)r/2d(x,p_{x}^{\prime\prime})\leq r/2. If 3r/4d(x,x0)<r3r/4\leq d(x,x_{0})<r, then d(x,px)r/2d(x,p_{x}^{\prime})\leq r/2. Finally, if d(x,x0)=rd(x,x_{0})=r, then d(x,px)r/4d(x,p_{x})\leq r/4. Therefore,

B¯(x,r)p𝒮B¯(p,r/2)\overline{B}(x,r)\subset\bigcup_{p\in\mathcal{S}}\overline{B}(p,r/2)

with 𝒮B¯(x,r)\mathcal{S}\subset\overline{B}(x,r) and card𝒮D:=3(N2)(6C+1)+7\operatorname{card}{\mathcal{S}}\leq D:=3(N-2)(6C+1)+7. ∎

2.4. Quasisymmetric maps

A homeomorphism f:(X,dX)(Y,dY)f:(X,d_{X})\to(Y,d_{Y}) between metric spaces is said to be quasisymmetric (or η\eta-quasisymmetric) if there exists a homeomorphism η:[0,)[0,)\eta\colon[0,\infty)\to[0,\infty) such that for all x,a,bXx,a,b\in X with xbx\neq b

dY(f(x),f(a))dY(f(x),f(b))η(dX(x,a)dX(x,b)).\frac{d_{Y}(f(x),f(a))}{d_{Y}(f(x),f(b))}\leq\eta\left(\frac{d_{X}(x,a)}{d_{X}(x,b)}\right).

The composition of two quasisymmetric maps and the inverse of a quasisymmetric map are quasisymmetric. Two spaces X,YX,Y are quasisymmetrically equivalent if there exists a quasisymmetric map between them.

For doubling connected metric spaces it is known that the quasisymmetric condition is equivalent to a weaker (but simpler) condition known in literature as weak quasisymmetry [Hei01, Theorem 10.19]. A homeomorphism f:(X,dX)(Y,dY)f:(X,d_{X})\to(Y,d_{Y}) between metric spaces is said to be weakly quasisymmetric if there exists H1H\geq 1 such that for all x,a,bXx,a,b\in X,

(2.2) dX(x,a)dX(x,b)impliesdY(f(x),f(a))HdY(f(x),f(b)).d_{X}(x,a)\leq d_{X}(x,b)\qquad\text{implies}\qquad d_{Y}(f(x),f(a))\leq Hd_{Y}(f(x),f(b)).

Quasisymmetric maps were introduced by Tukia and Väisälä [TV80] as the analogue of quasiconformal maps in the abstract metric setting. For more background see [Hei01, Chapters 10-12].

It is well known that the properties “doubling” and “bounded turning” are preserved under quasisymmetric maps quantitatively. Bonk and Meyer showed that the properties “uniformly branch separation” and “uniform branch density” are also preserved [BM22, Lemmas 4.3 and 4.5]. Below we show that the “uniform branch growth” property, which we introduced in the Introduction, is also preserved.

Lemma 2.6.

Let TT be a metric tree with uniform branch growth. If TT^{\prime} is quasisymmetrically homeomorphic to TT, then TT^{\prime} has uniform branch growth.

Proof.

Let f:TTf:T\to T^{\prime} be an η\eta-quasisymmetric homeomorphism. Let pTp\in T^{\prime} be a branch point. If pp has only three branches, then there is nothing to prove. We assume for the rest that pp has at least 4 branches. We have that f1(p)f^{-1}(p) is a branch point of TT with at least 4 branches, which we enumerate B1,B2,B_{1},B_{2},\dots so that diamBidiamBi+1\operatorname{diam}{B_{i}}\geq\operatorname{diam}{B_{i+1}} for all ii.

Fix indices i,ji,j with i>ji>j. Let xiBix_{i}\in B_{i} such that dT(f(xi),p)12diamf(Bi)d_{T^{\prime}}(f(x_{i}),p)\geq\frac{1}{2}\operatorname{diam}{f(B_{i})} and let xjBjx_{j}\in B_{j} such that dT(f1(p),xj)12diamBjd_{T}(f^{-1}(p),x_{j})\geq\frac{1}{2}\operatorname{diam}{B_{j}}. Then,

(2.3) diamf(Bi)diamf(Bj)2dT(f(xi),p)dT(f(xj),p)2η(dT(xi,f1(p))dT(xj,f1(p)))2η(2diamBidiamBj)2η(2).\displaystyle\frac{\operatorname{diam}{f(B_{i})}}{\operatorname{diam}{f(B_{j})}}\leq\frac{2d_{T^{\prime}}(f(x_{i}),p)}{d_{T^{\prime}}(f(x_{j}),p)}\leq 2\eta\left(\frac{d_{T}(x_{i},f^{-1}(p))}{d_{T}(x_{j},f^{-1}(p))}\right)\leq 2\eta\left(2\frac{\operatorname{diam}{B_{i}}}{\operatorname{diam}{B_{j}}}\right)\leq 2\eta(2).

Similarly, if i>j3i>j\geq 3, then

(2.4) diamf(Bj)diamf(Bi)2η(2C),\frac{\operatorname{diam}{f(B_{j})}}{\operatorname{diam}{f(B_{i})}}\leq 2\eta(2C),

where CC is the uniform branch growth constant of TT.

Suppose now that i1,i2,i3,i4i_{1},i_{2},i_{3},i_{4} are distinct indices such that

diamf(Bi1)diamf(Bi2)diamf(Bi3)diamf(Bi4).\operatorname{diam}{f(B_{i_{1}})}\geq\operatorname{diam}{f(B_{i_{2}})}\geq\operatorname{diam}{f(B_{i_{3}})}\geq\operatorname{diam}{f(B_{i_{4}})}.

We proceed with a case study.

Case 1. Assume that i3,i43i_{3},i_{4}\geq 3. Then, uniform branch growth of TT^{\prime} follows immediately from either (2.3) (if i3>i4i_{3}>i_{4}) or (2.4) (if i3<i4i_{3}<i_{4}).

Case 2. Assume that i3{1,2}i_{3}\in\{1,2\}. Then, at least one of i1,i2i_{1},i_{2}, say i1i_{1}, is in {3,4,}\{3,4,\dots\} and we have

diamf(Bi3)diamf(Bi4)diamf(Bi1)diamf(Bi4)2η(2C),\frac{\operatorname{diam}{f(B_{i_{3}})}}{\operatorname{diam}{f(B_{i_{4}})}}\leq\frac{\operatorname{diam}{f(B_{i_{1}})}}{\operatorname{diam}{f(B_{i_{4}})}}\leq 2\eta(2C),

by either (2.3) (if i1>i4i_{1}>i_{4}), or (2.4) (if i1<i4i_{1}<i_{4}).

Case 3. Assume that i4{1,2}i_{4}\in\{1,2\}. Then, at least one of i1,i2i_{1},i_{2}, say i1i_{1}, is in {3,4,}\{3,4,\dots\} and we have

diamf(Bi3)diamf(Bi4)diamf(Bi1)diamf(Bi4)2η(2)\frac{\operatorname{diam}{f(B_{i_{3}})}}{\operatorname{diam}{f(B_{i_{4}})}}\leq\frac{\operatorname{diam}{f(B_{i_{1}})}}{\operatorname{diam}{f(B_{i_{4}})}}\leq 2\eta(2)

by (2.3) since i1>i4i_{1}>i_{4}. ∎

2.5. Quasi-visual subdivisions

The following three definitions and proposition from [BM22] are the key factors for the proof of Theorem 1.9.

Definition 2.7 (Quasi-visual approximations).

Let SS be a bounded metric space. A quasi-visual approximation of SS is a sequence (Xn)n0(\textbf{X}^{n})_{n\in\mathbb{N}_{0}}, where X0={S}\textbf{X}^{0}=\{S\}, and Xn\textbf{X}^{n} is a finite cover of SS, for any nn\in\mathbb{N}, with the following properties. Note that the implicit constants in what follows are independent of n,X,Yn,X,Y:

  1. (i)

    diam(X)diam(Y)\operatorname{diam}(X)\simeq\operatorname{diam}(Y) for all X,YXnX,Y\in\textbf{X}^{n} with XYX\cap Y\neq\emptyset.

  2. (ii)

    dist(X,Y)diam(X)\operatorname{dist}(X,Y)\gtrsim\operatorname{diam}(X) for all X,YXnX,Y\in\textbf{X}^{n} with XY=X\cap Y=\emptyset.

  3. (iii)

    diam(X)diam(Y)\operatorname{diam}(X)\simeq\operatorname{diam}(Y) for all XXn,YXn+1X\in\textbf{X}^{n},Y\in\textbf{X}^{n+1} with XYX\cap Y\neq\emptyset.

  4. (iv)

    For some constants k0k_{0}\in\mathbb{N} and λ(0,1)\lambda\in(0,1) independent of nn we have diam(Y)λdiam(X)\operatorname{diam}(Y)\leq\lambda\operatorname{diam}(X) for all XXnX\in\textbf{X}^{n} and YXn+k0Y\in\textbf{X}^{n+k_{0}} with XYX\cap Y\neq\emptyset.

We write (Xn)(\textbf{X}^{n}) instead of (Xn)n0(\textbf{X}^{n})_{n\in\mathbb{N}_{0}} with the index set 0\mathbb{N}_{0} for nn understood. We call the elements of Xn\textbf{X}^{n} the tiles of level nn, or simply the nn-tiles of the sequence (Xn)(\textbf{X}^{n}).

Definition 2.8 (Subdivisions).

A subdivision of a compact metric space SS is a sequence (Xn)n0(\textbf{X}^{n})_{n\in\mathbb{N}_{0}} with the following properties:

  1. (i)

    X0={S}\textbf{X}^{0}=\{S\}, and Xn\textbf{X}^{n} is a finite collection of compact subsets of SS for each nn\in\mathbb{N}.

  2. (ii)

    For each n0n\in\mathbb{N}_{0} and YXn+1Y\in\textbf{X}^{n+1}, there exists XXnX\in\textbf{X}^{n} with YXY\subset X.

  3. (iii)

    For each n0n\in\mathbb{N}_{0} and XXnX\in\textbf{X}^{n}, we have

    X={YXn+1:YX}.X=\bigcup\{Y\in\textbf{X}^{n+1}:Y\subset X\}.
Definition 2.9 (Quasi-visual subdivisions).

A subdivision (Xn)(\textbf{X}^{n}) of a compact metric space SS is called quasi-visual subdivision of SS if it is a quasi-visual approximation of SS.

Let (Xn)(\textbf{X}^{n}) and (Yn)(\textbf{Y}^{n}) be subdivisions of compact metric spaces SS and TT, respectively. We say (Xn)(\textbf{X}^{n}) and (Yn)(\textbf{Y}^{n}) are isomorphic subdivisions if there exist bijections Fn:XnYnF^{n}:\textbf{X}^{n}\to\textbf{Y}^{n} , n0n\in\mathbb{N}_{0}, such that for all n0n\in\mathbb{N}_{0}, X,YXnX,Y\in\textbf{X}^{n}, and XXn+1X^{\prime}\in\textbf{X}^{n+1} we have

XYif and only ifFn(X)Fn(Y)X\cap Y\neq\emptyset\ \text{if and only if}\ F^{n}(X)\cap F^{n}(Y)\neq\emptyset

and

XXif and only ifFn+1(X)Fn(X).X^{\prime}\subset X\ \text{if and only if}\ F^{n+1}(X^{\prime})\subset F^{n}(X).

We say that the isomorphism between (Xn)(\textbf{X}^{n}) and (Yn)(\textbf{Y}^{n}) is given by the family (Fn)(F^{n}). We say that an isomorphism (Fn)(F^{n}) is induced by a homeomorphism F:STF:S\to T if Fn(X)=F(X)F^{n}(X)=F(X) for all n0n\in\mathbb{N}_{0} and XXnX\in\textbf{X}^{n}.

Proposition 2.10.

[BM22, Proposition 2.13] Let SS and TT be compact metric spaces with quasi-visual subdivisions (Xn)(\textbf{X}^{n}) and (Yn)(\textbf{Y}^{n}) that are isomorphic. Then there exists a unique quasisymmetric homeomorphism F:STF:S\to T that induces the isomorphism between (Xn)(\textbf{X}^{n}) and (Yn)(\textbf{Y}^{n}).

2.6. Combinatorial graphs and trees

An alphabet AA is an at most countable set that is either equal to \mathbb{N} or of the form {1,,M}\{1,\dots,M\} for some integer M2M\geq 2. For each integer k0k\geq 0 denote by AkA^{k} the set of words formed by kk letters of AA, with the convention A0={ϵ}A^{0}=\{\epsilon\}, where ϵ\epsilon denotes the empty word. We set A=k0AkA^{*}=\bigcup_{k\geq 0}A^{k} to be the collection of words of finite length and AA^{\mathbb{N}} to be the collection of words of infinite length. If wAw\in A^{*}, then we denote by |w||w| the length of ww, i.e., the number of letters that ww consists of, with the convention |ϵ|=0|\epsilon|=0. Given two finite words w=i1inw=i_{1}\cdots i_{n} and v=j1jmv=j_{1}\cdots j_{m} in AA^{*}, we denote by wu=i1inj1jmwu=i_{1}\cdots i_{n}j_{1}\cdots j_{m} their concatenation.

Given wAw\in A^{*} and integer k|w|k\geq|w|, denote

Awk={wu:uAk|w|},Aw={wu:uA},Aw={wu:uA}.A^{k}_{w}=\{wu:u\in A^{k-|w|}\},\quad A^{*}_{w}=\{wu:u\in A^{*}\},\quad A^{\mathbb{N}}_{w}=\{wu:u\in A^{\mathbb{N}}\}.

Given nn\in\mathbb{N} and wAw\in A^{\mathbb{N}} denote by w(n)w(n) the unique word uAnu\in A^{n} such that w=uww=uw^{\prime} for some wAw^{\prime}\in A^{\mathbb{N}}. Similarly, if nn\in\mathbb{N} and wAw\in A^{*}, w(n)w(n) denotes the initial subword of ww of length nn, and we set w(n)=ww(n)=w if n|w|n\geq|w|. Given iAi\in A and kk\in\mathbb{N}, we denote by i(k)i^{(k)} the word iiAki\cdots i\in A^{k} and by i()i^{(\infty)} the infinite word iiAii\cdots\in A^{\mathbb{N}}.

A combinatorial graph is a pair G=(V,E)G=(V,E) of a finite or countable vertex set VV and an edge set

E{{v,v}:v,vV and vv}.E\subset\left\{\{v,v^{\prime}\}:v,v^{\prime}\in V\text{ and }v\neq v^{\prime}\right\}.

If {v,v}E\{v,v^{\prime}\}\in E, we say that the vertices vv and vv^{\prime} are adjacent in GG.

A combinatorial graph G=(V,E)G^{\prime}=(V^{\prime},E^{\prime}) is a subgraph of G=(V,E)G=(V,E), and we write GGG\subset G^{\prime}, if VVV^{\prime}\subset V and EEE^{\prime}\subset E. We commonly generate subgraphs of G=(V,E)G=(V,E) by starting with a vertex set VVV^{\prime}\subset V and considering the subgraph of GG induced by VV^{\prime}, i.e., the graph G=(V,E)G^{\prime}=(V^{\prime},E^{\prime}) where EE^{\prime} is the set of all edges between two vertices of VV^{\prime}.

A combinatorial path in GG is an ordered set γ=(v1,v2,,vn)Vn\gamma=(v_{1},v_{2},\dots,v_{n})\in V^{n} such that viv_{i} is adjacent to vi+1v_{i+1} for all i{1,,n1}i\in\{1,\dots,n-1\}; in this case we say that γ\gamma joins v1v_{1} to vnv_{n}. The path γ=(v1,v2,,vn)\gamma=(v_{1},v_{2},\dots,v_{n}) is a combinatorial arc or simple path if for all i,j{1,,n}i,j\in\{1,\dots,n\}, vi=vjv_{i}=v_{j} if and only if i=ji=j; in this case we say that the endpoints of the arc γ\gamma are the points v1,vnv_{1},v_{n}. The length of a path γ\gamma is the number of vertices that it contains. If γ=(v1,,vk)\gamma=(v_{1},\dots,v_{k}) and γ=(u1,,um)\gamma^{\prime}=(u_{1},\dots,u_{m}) are two paths in GG with vk=u1v_{k}=u_{1}, then we denote by γγ\gamma\gamma^{\prime} the concatenation path (v1,,vk,u2,,um)(v_{1},\dots,v_{k},u_{2},\dots,u_{m}). If γ=(v1,,vk)\gamma=(v_{1},\dots,v_{k}) we denote by γ(v1):=(v2,,vk)\gamma\setminus(v_{1}):=(v_{2},\dots,v_{k}) and similarly we define γ(vk)\gamma\setminus(v_{k}).

A combinatorial graph G=(V,E)G=(V,E) is connected, if for any distinct v,vVv,v^{\prime}\in V there exists a path γ\gamma in GG that joins vv with vv^{\prime}. Given vVv\in V, a component of G{v}G\setminus\{v\} is a maximal connected subgraph of G{v}G\setminus\{v\}.

A graph T=(V,E)T=(V,E) is a combinatorial tree if for any distinct v,vv,v^{\prime} there exists unique combinatorial arc γ\gamma whose endpoints are vv and vv^{\prime}. The unique arc in a combinatorial tree TT that starts from vv and ends in vv^{\prime} is denoted by PT(v,v)P_{T}(v,v^{\prime}). Given a combinatorial tree T=(V,E)T=(V,E) and a point vVv\in V, define the valencies

Val(T,v):=card{eE:ve}andVal(T):=supvVVal(T,v)\text{Val}(T,v):=\operatorname{card}\{e\in E:v\in e\}\qquad\text{and}\qquad\text{Val}(T):=\sup_{v\in V}\text{Val}(T,v)

and the set of leaves L(T):={vV:Val(T,v)=1}L(T):=\{v\in V:\text{Val}(T,v)=1\}. Here card\operatorname{card} denotes the cardinality of a finite or countable set, taking values in {}\mathbb{N}\cup\{\infty\}.

Given a combinatorial graph G=(V,E)G=(V,E) and a vertex vVv\in V, we write G{v}G\setminus\{v\} to be the subgraph of GG induced by V{v}V\setminus\{v\}. Note that, if TT is a combinatorial tree, then every component of T{v}T\setminus\{v\} is a combinatorial tree.

3. Weak tangents and the proof of Theorem 1.1

Here we prove Theorem 1.1. The proof uses the notion of weak tangents from [Gro81, Gro99], which we recall here.

Let (X,d)(X,d) be a metric space and let pXp\in X. We call a metric space (Y,d)(Y,d^{\prime}) a weak tangent of XX at pp if there exists qYq\in Y, a sequence (rn)n(r_{n})_{n\in\mathbb{N}} of positive numbers with rn0r_{n}\to 0, and a sequence (pn)n(p_{n})_{n\in\mathbb{N}} of points in XX with pnpp_{n}\to p such that (X,pn,rn1d)GH(Y,q,d)(X,p_{n},r_{n}^{-1}d)\xrightarrow{GH}(Y,q,d^{\prime}). The existence of weak tangents for any doubling metric space at any point is guaranteed (see [BBI01, Theorem 8.1.10]). We denote by Tan(X,p)\operatorname{Tan}(X,p) the collection of all weak tangents of XX at pp.

Recall that a metric space (X,d)(X,d) is an \mathbb{R}-tree if for any x,yXx,y\in X there exists a unique arc from xx to yy, and this arc is a geodesic. In our first lemma we show that pointed Gromov-Hausdorff limits of \mathbb{R}-trees are \mathbb{R}-trees.

Lemma 3.1.

Let (Tm,pm,dm)(T_{m},p_{m},d_{m}) be a sequence of \mathbb{R}-trees that converges to a complete metric space (T,q,d)(T,q,d) in the pointed Gromov-Haudorff sense. Then, TT is an \mathbb{R}-tree.

Proof.

By [BBI01, Theorem 8.1.9] TT is a length space. To show that TT is an \mathbb{R}-tree it suffices to show that TT does not have any simple closed curve.

In this proof we denote by BT(x,r)B_{T}(x,r) (resp. BTm(x,r)B_{T_{m}}(x,r)) the open ball in TT (resp. in TmT_{m}) centered at xx and of radius rr.

Assume for a contradiction that ΓT\Gamma\subset T is a simple closed curve. Let x,yTx,y\in T such that d(x,y)=diamΓd(x,y)=\operatorname{diam}\Gamma. Write Γ{x,y}={α,α}\Gamma\setminus\{x,y\}=\{\alpha,\alpha^{\prime}\} where α,α\alpha,\alpha^{\prime} are arcs that connect x,yx,y. Fix ϵ<101diamΓ\epsilon<10^{-1}\operatorname{diam}\Gamma and let x1,x1x_{1},x^{\prime}_{1} be the last points such that x1αBT(x,ϵ)x_{1}\in\alpha\cap\partial B_{T}(x,\epsilon) and x1αBT(x,ϵ)x_{1}^{\prime}\in\alpha^{\prime}\cap\partial B_{T}(x,\epsilon), and similarly define y1αBT(y,ϵ)y_{1}\in\alpha\cap B_{T}(y,\epsilon) and y1αBT(y,ϵ)y_{1}^{\prime}\in\alpha^{\prime}\cap\partial B_{T}(y,\epsilon). Let βα\beta\subset\alpha and βα\beta^{\prime}\subset\alpha^{\prime} be the subarcs of Γ\Gamma that connect x1,x1x_{1},x_{1}^{\prime} and y1,y1y_{1},y_{1}^{\prime} respectively. Then, dist(β,β)>0\operatorname{dist}(\beta,\beta^{\prime})>0 since Γ\Gamma is a simple closed curve.

Fix ϵ<101dist(β,β)\epsilon^{\prime}<10^{-1}\operatorname{dist}(\beta,\beta^{\prime}). There exist points {zi}i=1N\{z_{i}\}_{i=1}^{N} in β\beta, and points {zj}j=1N\{z_{j}^{\prime}\}_{j=1}^{N^{\prime}} in β\beta^{\prime} (both sets enumerated according to the orientations of β\beta and β\beta^{\prime}) such that z1=x1z_{1}=x_{1}, zN=y1z_{N}=y_{1}, z1=x1z_{1}^{\prime}=x_{1}^{\prime}, zN=y1z_{N^{\prime}}^{\prime}=y_{1}^{\prime}, and for all i{1,,N1}i\in\{1,\dots,N-1\} and j{1,,N1}j\in\{1,\dots,N^{\prime}-1\}

ϵd(xi,xi+1)2ϵandϵd(xj,xj+1)2ϵ.\epsilon^{\prime}\leq d(x_{i},x_{i+1})\leq 2\epsilon^{\prime}\quad\text{and}\quad\epsilon^{\prime}\leq d(x_{j}^{\prime},x_{j+1}^{\prime})\leq 2\epsilon^{\prime}.

Let also x=z0=z0x=z_{0}=z_{0}^{\prime} and y=zN+1=zN+1y=z_{N+1}=z_{N^{\prime}+1}^{\prime}.

Set ϵ′′<19min{dist(β,β)4ϵ,diamΓ4ϵ}\epsilon^{\prime\prime}<\frac{1}{9}\min\{\operatorname{dist}(\beta,\beta^{\prime})-4\epsilon^{\prime},\operatorname{diam}\Gamma-4\epsilon\}. Choose r>0r>0 big enough such that ΓB(q,rϵ′′)\Gamma\subset B(q,r-\epsilon^{\prime\prime}). Then there exists MM\in\mathbb{N} such that for all mMm\geq M we can find f:B(pm,r)Tf:B(p_{m},r)\to T and points {wzi}i=0N+1\{w_{z_{i}}\}_{i=0}^{N+1}, {wzj}j=0N+1\{w_{z_{j}^{\prime}}\}_{j=0}^{N^{\prime}+1} in B(pm,r)B(p_{m},r) such that

d(f(wzi),zi)<ϵ′′andd(f(wzj),zj)<ϵ′′d(f(w_{z_{i}}),z_{i})<\epsilon^{\prime\prime}\quad\text{and}\quad d(f(w_{z_{j}^{\prime}}),z_{j}^{\prime})<\epsilon^{\prime\prime}

for all i{0,,N+1}i\in\{0,\dots,N+1\} and j{0,,N+1}j\in\{0,\dots,N^{\prime}+1\}. Let

γ=[wz1,wz2][wzN1,wzN]andγ=[wz1,wz2][wzN1,wzN].\gamma=[w_{z_{1}},w_{z_{2}}]\cup\dots\cup[w_{z_{N-1}},w_{z_{N}}]\quad\text{and}\quad\gamma^{\prime}=[w_{z_{1}^{\prime}},w_{z_{2}^{\prime}}]\cup\dots\cup[w_{z_{N^{\prime}-1}^{\prime}},w_{z_{N^{\prime}}^{\prime}}].

We claim that dist(γ,γ)>0\operatorname{dist}(\gamma,\gamma^{\prime})>0. To see that, fix a1γa_{1}\in\gamma, a2γa_{2}\in\gamma^{\prime} and let i{1,,N1}i\in\{1,\dots,N-1\} and j{1,,N1}j\in\{1,\dots,N^{\prime}-1\} such that a1[wzi,wzi+1]a_{1}\in[w_{z_{i}},w_{z_{i+1}}] and a2[wzj,wzj+1]a_{2}\in[w_{z_{j}^{\prime}},w_{z_{j+1}^{\prime}}]. Then

dm(a1,a2)\displaystyle d_{m}(a_{1},a_{2}) dm(wzi,wzj)dm(a1,wzi)dm(a2,wzj)\displaystyle\geq d_{m}(w_{z_{i}},w_{z_{j}^{\prime}})-d_{m}(a_{1},w_{z_{i}})-d_{m}(a_{2},w_{z_{j}^{\prime}})
d(f(wzi),f(wzj))ϵ′′dm(wzi,wzi+1)dm(wzj,wzj+1)\displaystyle\geq d(f(w_{z_{i}}),f(w_{z_{j}^{\prime}}))-\epsilon^{\prime\prime}-d_{m}(w_{z_{i}},w_{z_{i+1}})-d_{m}(w_{z_{j}^{\prime}},w_{z_{j+1}^{\prime}})
d(zi,zj)3ϵ′′(d(f(wzi),f(wzi+1))+ϵ′′)(d(f(wzi),f(wzi+1))+ϵ′′)\displaystyle\geq d(z_{i},z_{j}^{\prime})-3\epsilon^{\prime\prime}-(d(f(w_{z_{i}}),f(w_{z_{i+1}}))+\epsilon^{\prime\prime})-(d(f(w_{z_{i}^{\prime}}),f(w_{z_{i+1}^{\prime}}))+\epsilon^{\prime\prime})
>dist(β,β)3ϵ′′(d(zi,zi+1)+3ϵ′′)(d(zj,zj+1)+3ϵ′′)\displaystyle>\operatorname{dist}(\beta,\beta^{\prime})-3\epsilon^{\prime\prime}-(d(z_{i},z_{i+1})+3\epsilon^{\prime\prime})-(d(z_{j}^{\prime},z_{j+1}^{\prime})+3\epsilon^{\prime\prime})
dist(β,β)9ϵ′′4ϵ\displaystyle\geq\operatorname{dist}(\beta,\beta^{\prime})-9\epsilon^{\prime\prime}-4\epsilon^{\prime}
>0.\displaystyle>0.

Let ζ=[wz1,wx](wx,wz1]\zeta=[w_{z_{1}},w_{x}]\cup(w_{x},w_{z_{1}^{\prime}}] and ζ=[wzN,wy](wy,wzN]\zeta^{\prime}=[w_{z_{N}},w_{y}]\cup(w_{y},w_{z_{N^{\prime}}^{\prime}}]. Repeating the calculations above, we get that dist(ζ,ζ)>0\operatorname{dist}(\zeta,\zeta^{\prime})>0. To obtain a contradiction, we claim that the curve γζγζ\gamma\cup\zeta^{\prime}\cup\gamma^{\prime}\cup\zeta (which is in the tree TmT_{m}) contains a simple closed curve. Note that

[wz1,wzN]ζ[wzN,wz1]ζγζγζ.[w_{z_{1}},w_{z_{N}}]\cup\zeta^{\prime}\cup[w_{z_{N^{\prime}}^{\prime}},w_{z_{1}^{\prime}}]\cup\zeta\subset\gamma\cup\zeta^{\prime}\cup\gamma^{\prime}\cup\zeta.

Denote with σ\sigma (resp. σ\sigma^{\prime}) the arc in ζ\zeta (resp. ζ)\zeta^{\prime}) that joins [wz1,wzN][w_{z_{1}},w_{z_{N}}] with [wz1,wzN][w_{z^{\prime}_{1}},w_{z^{\prime}_{N^{\prime}}}] at the points w1,w2w_{1},w_{2} respectively (resp. w1,w2w^{\prime}_{1},w^{\prime}_{2}). From the above discussion, points w1,w2,w1,w2w_{1},w_{2},w^{\prime}_{1},w^{\prime}_{2} are all different from each other. It follows that

[w1,w2](w2,w2](w2,w1](w1,w1][w_{1},w_{2}]\cup(w_{2},w^{\prime}_{2}]\cup(w^{\prime}_{2},w^{\prime}_{1}]\cup(w^{\prime}_{1},w_{1}]

is a simple closed curve. ∎

For the remainder of this section, for each NN\in\mathbb{N} define the planar set

TN=0n=2k=1Nn,kT_{N}=\ell_{0}\cup\bigcup_{n=2}^{\infty}\bigcup_{k=1}^{N}\ell_{n,k}

where

0=[0,1]×{0}andn,k={2n+1+k1N4n}×[0,3n].\ell_{0}=[0,1]\times\{0\}\quad\text{and}\quad\ell_{n,k}=\left\{2^{-n+1}+\tfrac{k-1}{N}4^{-n}\right\}\times[0,3^{-n}].

It is easy to see that TNT_{N} is compact (segments n,k\ell_{n,k} accumulate at (0,0)(0,0)), connected, locally connected (the length of n,k\ell_{n,k} goes to 0 as nn goes to infinity) and contains no simple closed curves. Therefore TNT_{N} is a metric tree. Moreover, every branch point has valence 3 (since segments n,k\ell_{n,k} are mutually disjoint). Finally, the total Hausdorff 11-measure of TNT_{N} is

1(TN)=1+Nn=23n<.\mathcal{H}^{1}(T_{N})=1+N\sum_{n=2}^{\infty}3^{-n}<\infty.

Therefore, we can equip TNT_{N} with the geodesic metric dd.

Lemma 3.2.

For each NN\in\mathbb{N}, (TN,d)(T_{N},d) a doubling geodesic tree of valence 3.

Proof.

We only need to show the doubling property. Towards this end, fix xTNx\in T_{N} and r>0r>0. Since TNT_{N} is compact, we may assume that r<1/2r<1/2.

Suppose that xn,kx\in\ell_{n,k} for some n,kn,k and let xx^{\prime} be the unique point in n,k0\ell_{n,k}\cap\ell_{0}. If rd(x,x)r\leq d(x,x^{\prime}), then B(x,r)B(x,r) is a line segment. If rd(x,x)r\geq d(x,x^{\prime}), then B(x,r)B(x,2r)B(x,r)\subset B(x^{\prime},2r). Therefore, we may assume for the rest of the proof that x0x\in\ell_{0}.

Let mm\in\mathbb{N} be the unique integer with 2m1<r2m2^{-m-1}<r\leq 2^{-m}. Let also 0y<z10\leq y<z\leq 1 be such that B(x,r)0=[y,z]×{0}B(x,r)\cap\ell_{0}=[y,z]\times\{0\}. It is easy to see that there exist at most 4N4N many segments n,k\ell_{n,k} that intersect with [y,z]×{0}[y,z]\times\{0\} and have length greater or equal to r/4r/4. Label these segments by 1,,l\ell^{1},\dots,\ell^{l}.

On one hand, 01l\ell_{0}\cup\ell^{1}\cup\cdots\cup\ell^{l} equipped with the geodesic metric is CC-doubling with CC depending only on NN and not on the segments j\ell^{j}. On the other hand, if n,k\ell_{n,k} intersects with [y,z]×{0}[y,z]\times\{0\} and has length less than r/4r/4, then, by design of TNT_{N}, it is contained in B(y,r/2)B(y,r/2). This completes the proof of the doubling property. ∎

The final ingredient in the proof of Theorem 1.1 is a result of Li [Li21] which states that quasisymmetric embeddability is hereditary: if XX quasisymmetrically embeds into YY, then every weak tangent of XX quasisymmetrically embeds into some weak tangent of YY.

Lemma 3.3 ([Li21, Theorem 1.1]).

Let X,YX,Y be proper, doubling metric spaces and f:(X,p,dX)(Y,q,dY)f:(X,p,d_{X})\to(Y,q,d_{Y}) be an η\eta-quasisymmetric map. For any weak tangent TTan(X,p)T\in\operatorname{Tan}(X,p), there exists a weak tangent TTan(Y,q)T^{\prime}\in\operatorname{Tan}(Y,q) such that TT is η\eta-quasisymmetric equivalent to TT^{\prime}.

Proof of Theorem 1.1.

Assume for a contradiction, that there exists a quasiconformal tree TT such that every element of 𝒬𝒞𝒯(3)\mathscr{QCT}(3) quasisymmetrically embeds in TT. Since every quasiconformal tree is quasisymmetric equivalent to a geodesic tree [BM20], we may assume that TT is geodesic and CC-doubling.

Fix NN\in\mathbb{N} such that N+2>CN+2>C, let p=(0,0)TNp=(0,0)\in T_{N}, and let ω1,,ωN+2\omega_{1},\dots,\omega_{N+2} be the (N+2)(N+2)-roots of unity, and let SS be the (N+2)(N+2)-star

j=1N+2{tωj:t0}2\bigcup_{j=1}^{N+2}\{t\omega_{j}:t\geq 0\}\subset\mathbb{R}^{2}

equipped with the geodesic metric dSd_{S} and let q=(0,0)q=(0,0).

We claim that STan(TN,p)S\in\operatorname{Tan}(T_{N},p). Assuming the claim, by Lemma 3.2, TNT_{N} is in 𝒬𝒞𝒯(3)\mathscr{QCT}(3), and by Lemma 3.3 SS quasisymmetrically embeds into a weak tangent SS^{\prime} of TT. By Lemma 2.1 and Lemma 3.1 SS^{\prime} is a CC-doubling \mathbb{R}-tree. However, if BB is a ball with center the embedded image of the branch point qq and radius δ\delta, then any δ/2\delta/2-separated set in BB has N+2N+2 elements (simply by taking the points on the boundary of the ball of radius δ/2\delta/2). It follows that N+2CN+2\leq C which is contradiction.

To prove the claim, set pn=2n+1p_{n}=2^{-n+1} and rn=(7/2)nr_{n}=(7/2)^{-n} for each nn\in\mathbb{N}. Fix ϵ>0\epsilon>0 and R>0R>0. We may assume that R1R\geq 1 since otherwise for large nn the balls B(pn,R)B(p_{n},R) will collapse on the point qq. There exists n0n_{0}\in\mathbb{N} such that for all nn0n\geq n_{0}

R<(7/6)n,andN1N(7/8)n<ϵ.R<(7/6)^{n},\quad\text{and}\quad\tfrac{N-1}{N}(7/8)^{n}<\epsilon.

Note that rn<2n2r_{n}<2^{-n-2} holds for all nn0n\geq n_{0}. Thus, for all n>n0n>n_{0}, if yB(pn,rnR)y\in\partial B(p_{n},r_{n}R), then d(pn,y)=rnR<3nd(p_{n},y)=r_{n}R<3^{-n}. Moreover,

B(pn,rnR)=hn1hn2hn3n,1n,NB(p_{n},r_{n}R)=h_{n}^{1}\cup h_{n}^{2}\cup h_{n}^{3}\cup\ell_{n,1}^{\prime}\cup\cdots\cup\ell_{n,N}^{\prime}

where hn1=(pnrnR,pn]×{0}h_{n}^{1}=(p_{n}-r_{n}R,p_{n}]\times\{0\}, hn2=[pn,pn+N1N4n]×{0}h_{n}^{2}=[p_{n},p_{n}+\frac{N-1}{N}4^{-n}]\times\{0\},

hn3=[pn+N1N4n,pn+rnR)×{0},h_{n}^{3}=[p_{n}+\tfrac{N-1}{N}4^{-n},p_{n}+r_{n}R)\times\{0\},

and

n,j={an,j}×[0,tn,j)=B(pn,rnR)n,j.\ell_{n,j}^{\prime}=\{a_{n,j}\}\times[0,t_{n,j})=B(p_{n},r_{n}R)\cap\ell_{n,j}.

Define a map f:B¯(pn,rnR)Sf:\overline{B}(p_{n},r_{n}R)\to S so that

  1. (1)

    f(an,j,t)=tωjf(a_{n,j},t)=t\omega_{j} for t[0,tn,j)t\in[0,t_{n,j}) and j{1,,N}j\in\{1,\dots,N\},

  2. (2)

    f(pnt,0)=tωN+1f(p_{n}-t,0)=t\omega_{N+1} for t[0,rnR)t\in[0,r_{n}R),

  3. (3)

    f(pn+N1N4n+t,0)=tωN+2f(p_{n}+\frac{N-1}{N}4^{-n}+t,0)=t\omega_{N+2} for t[0,rnRN1N4n)t\in[0,r_{n}R-\frac{N-1}{N}4^{-n}), and

  4. (4)

    f([pn,pn+N1N4n])=(0,0)f([p_{n},p_{n}+\frac{N-1}{N}4^{-n}])=(0,0).

Let x1,x2B(pn,rnR)x_{1},x_{2}\in B(p_{n},r_{n}R). If x1,x2n,jx_{1},x_{2}\in\ell_{n,j}^{\prime} for some j{1,,N}j\in\{1,\dots,N\}, then

|dS(f(x1),f(x2))rn1d(x1,x2)|=0.|d_{S}(f(x_{1}),f(x_{2}))-r_{n}^{-1}d(x_{1},x_{2})|=0.

Similarly, if x1,x2hn1x_{1},x_{2}\in h_{n}^{1} or if x1,x2hn3x_{1},x_{2}\in h^{3}_{n}. If x1,x2hn2x_{1},x_{2}\in h_{n}^{2}, then

|dS(f(x1),f(x2))rn1d(x1,x2)|=|rn1d(x1,x2)|\displaystyle|d_{S}(f(x_{1}),f(x_{2}))-r_{n}^{-1}d(x_{1},x_{2})|=|r_{n}^{-1}d(x_{1},x_{2})| N1N(27)n4n=N1N(78)n<ϵ.\displaystyle\leq\tfrac{N-1}{N}(\tfrac{2}{7})^{n}4^{-n}=\tfrac{N-1}{N}(\tfrac{7}{8})^{n}<\epsilon.

If x1n,jx_{1}\in\ell_{n,j}^{\prime} for some j{1,,N}j\in\{1,\dots,N\} and x2hn1x_{2}\in h_{n}^{1}, then by geodesicity of TNT_{N} and SS we have

|\displaystyle| dS(f(x1),f(x2))rn1d(x1,x2)|\displaystyle d_{S}(f(x_{1}),f(x_{2}))-r_{n}^{-1}d(x_{1},x_{2})|
=|dS(f(x1),q)+dS(q,f(x2))rn1(d(x1,(an,j,0))+d((an,j,0),pn)+d(pn,x2))|\displaystyle=|d_{S}(f(x_{1}),q)+d_{S}(q,f(x_{2}))-r_{n}^{-1}(d(x_{1},(a_{n,j},0))+d((a_{n,j},0),p_{n})+d(p_{n},x_{2}))|
=|rn1d((an,j,0),pn)|\displaystyle=|r_{n}^{-1}d((a_{n,j},0),p_{n})|
N1N(78)n\displaystyle\leq\tfrac{N-1}{N}(\tfrac{7}{8})^{n}
<ϵ.\displaystyle<\epsilon.

All the other cases are similar.

Lastly, we need to verify that B(q,Rϵ)Nϵ(f(B(pn,rnR)))B(q,R-\epsilon)\subset N_{\epsilon}(f(B(p_{n},r_{n}R))). Let x=tωjB(q,Rϵ)x=t\omega_{j}\in B(q,R-\epsilon) for some j{1,,N}j\in\{1,\dots,N\} (the cases where j{N+1,N+2}j\in\{N+1,N+2\} are similar). Let also (an,j,y)(a_{n,j},y) be the unique point in B(pn,rnR)n,j\partial B(p_{n},r_{n}R)\cap\ell_{n,j}. From the construction of ff and geodesicity of TNT_{N},

dS(q,f(an,j,y))\displaystyle d_{S}(q,f(a_{n,j},y)) =rn1(d(pn,(an,j,y))d(pn,(an,j,0)))\displaystyle=r_{n}^{-1}(d(p_{n},(a_{n,j},y))-d(p_{n},(a_{n,j},0)))
rn1(d(pn,(an,j,y))N1N(78)n)\displaystyle\geq r_{n}^{-1}(d(p_{n},(a_{n,j},y))-\tfrac{N-1}{N}(\tfrac{7}{8})^{n})
>Rϵ.\displaystyle>R-\epsilon.

Due to geodesicity of SS, xx is on the arc that joins qq with f(y)f(y). Thus, there exists xB(pn,rnR)x^{\prime}\in B(p_{n},r_{n}R) such that x=f(x)x=f(x^{\prime}). It follows that

B(q,Rϵ)f(B(pn,rnR))Nϵ(f(B(pn,rnR))).B(q,R-\epsilon)\subset f(B(p_{n},r_{n}R))\subset N_{\epsilon}(f(B(p_{n},r_{n}R))).\qed

4. Construction of trees 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}}

For each alphabet AA, and for each choice of weights a1,a2,a_{1},a_{2},\dots with 12=a1=a2a3\frac{1}{2}=a_{1}=a_{2}\geq a_{3}\geq\cdots and liman=0\lim a_{n}=0, we construct an associated metric tree 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}}. These trees are shown to be universal later on.

4.1. Graphs

Given an alphabet AA, for each kk\in\mathbb{N} we construct a graph GkA=(Ak,EkA)G_{k}^{A}=(A^{k},E_{k}^{A}) inductively.

  1. (i)

    Let E1A={{1,i}:iA{1}}E_{1}^{A}=\{\{1,i\}:i\in A\setminus\{1\}\}.

  2. (ii)

    Assume that for some kk\in\mathbb{N} we have defined a graph GkA=(Ak,EkA)G_{k}^{A}=(A^{k},E_{k}^{A}). Set Gk+1A=(Ak+1,Ek+1A)G_{k+1}^{A}=(A^{k+1},E_{k+1}^{A}) where

    Ek+1A={{12(k),i1(k)}:iA{1}}iA{{iw,iu}:w,uAk and {w,u}EkA}.E_{k+1}^{A}=\left\{\{12^{(k)},i1^{(k)}\}:i\in A\setminus\{1\}\right\}\cup\bigcup_{i\in A}\left\{\{iw,iu\}:w,u\in A^{k}\text{ and }\{w,u\}\in E_{k}^{A}\right\}.
Lemma 4.1.

Let k0k\geq 0, w,uAkw,u\in A^{k}.

  1. (i)

    There exist distinct i,jAi,j\in A such that {iw,ju}Ek+1A\{iw,ju\}\in E_{k+1}^{A} if and only if one of the words is of the form 12(k)12^{(k)} and the other of the form l1(k)l1^{(k)} with lA{1}l\in A\setminus\{1\}.

  2. (ii)

    For any vAv\in A^{*} we have that {w,u}EkA\{w,u\}\in E_{k}^{A} if and only if {vw,vu}Ek+|v|A\{vw,vu\}\in E_{k+|v|}^{A}.

  3. (iii)

    The subgraph of Gk+1AG_{k+1}^{A} generated by {wi:iA}\{wi:i\in A\} is connected.

  4. (iv)

    We have that {w,u}EkA\{w,u\}\in E_{k}^{A} if and only if there exist l{0,,k1}l\in\{0,\dots,k-1\}, iA{1}i\in A\setminus\{1\}, and vAkl1v\in A^{k-l-1} such that one of the words is of the form vi1(l)vi1^{(l)} and the other of the form v12(l)v12^{(l)}.

  5. (v)

    If {w,u}Ek\{w,u\}\in E_{k}, then there exist i,j{1,2}i,j\in\{1,2\} such that {wi,uj}Ek+1\{wi,uj\}\in E_{k+1}.

Refer to caption
Refer to caption
Figure 2. Graphs G1A,G2A,G3AG_{1}^{A},G_{2}^{A},G_{3}^{A} in the case that A={1,2,3}A=\{1,2,3\}.
Refer to caption
Refer to caption
Figure 3. Graphs G1A,G2A,G3AG_{1}^{A},G_{2}^{A},G_{3}^{A} in the case that A={1,2,3,4}A=\{1,2,3,4\}.
Proof.

The first claim is clear from the design of graphs GkAG_{k}^{A}.

We prove property (2) by induction on the length of vv. The claim is trivially true for v=ϵv=\epsilon. Inductively, suppose that for some m0m\geq 0 the claim is true for all words vAmv\in A^{m}. Let now v=ivAm+1v=iv^{\prime}\in A^{m+1} with vAmv^{\prime}\in A^{m} and iAi\in A. Assume first that {w,u}EkA\{w,u\}\in E_{k}^{A}. By the inductive hypothesis, {vw,vu}Ek+mA\{v^{\prime}w,v^{\prime}u\}\in E_{k+m}^{A}, and by design of Ek+m+1AE_{k+m+1}^{A} we have that {ivw,ivu}Ek+m+1A\{iv^{\prime}w,iv^{\prime}u\}\in E_{k+m+1}^{A}. Conversely, assume that {ivw,ivu}Ek+m+1A\{iv^{\prime}w,iv^{\prime}u\}\in E_{k+m+1}^{A}. By design of Ek+m+1AE_{k+m+1}^{A} we have that {vw,vu}Ek+mA\{v^{\prime}w,v^{\prime}u\}\in E_{k+m}^{A} and by the inductive hypothesis, {w,u}EkA\{w,u\}\in E_{k}^{A}.

Property (3) follows immediately from property (2) and the fact that G1G_{1} is connected.

For the fourth property, write w=viww=viw^{\prime} and u=vjuu=vju^{\prime} where vAlv\in A^{l} with l{0,,k1}l\in\{0,\dots,k-1\}, u,uAkl1u,u^{\prime}\in A^{k-l-1}, and i,jAi,j\in A with i<ji<j. By property (2) we have that {iw,ju}EklA\{iw^{\prime},ju^{\prime}\}\in E_{k-l}^{A} if and only if {w,u}EkA\{w,u\}\in E_{k}^{A}. The claim now follows from property (1).

Assume that {w,u}EkA\{w,u\}\in E_{k}^{A}. By property (4), there exist l{0,,k1}l\in\{0,\dots,k-1\}, iA{1}i\in A\setminus\{1\}, and vAkl1v\in A^{k-l-1} such that one of the words (say ww) is of the form vi1(l)vi1^{(l)} and the other (say uu) of the form v12(l)v12^{(l)}. Then {w1,u2}Ek+1A\{w1,u2\}\in E_{k+1}^{A}. ∎

Remark 4.2.

Properties (3) and (5) from Lemma 4.1 imply that the collection of graphs (GkA)k(G_{k}^{A})_{k\in\mathbb{N}} are combinatorial data in the sense of [DV22, Definition 1.1].

Lemma 4.3.

For each kk\in\mathbb{N}, the graph GkAG_{k}^{A} is a combinatorial tree.

Proof.

The proof is by induction on kk. For k=1k=1, the claim is clear. Assume the claim to be true for some kk\in\mathbb{N}. Note that for each iAi\in A the subgraph Gk+1,iAG_{k+1,i}^{A} of Gk+1AG_{k+1}^{A} generated by the vertices Aik+1A_{i}^{k+1} is isomorphic to GkAG_{k}^{A}; hence it is a tree.

We first show that Gk+1AG_{k+1}^{A} is connected. Let v1,v2Ak+1v_{1},v_{2}\in A^{k+1}. If there exists iAi\in A such that v1,v2Aik+1v_{1},v_{2}\in A^{k+1}_{i}, then v1,v2v_{1},v_{2} can be joined by path in Gk+1,iAG_{k+1,i}^{A} by the inductive assumption. Suppose now that v1Aik+1v_{1}\in A^{k+1}_{i} and v2Ajk+1v_{2}\in A^{k+1}_{j} with i,jAi,j\in A distinct. Without loss of generality, assume that i,j1i,j\neq 1. By the inductive assumption, there exists a path γ1\gamma_{1} from v1v_{1} to i1(k)i1^{(k)} and a path γ2\gamma_{2} from j1(k)j1^{(k)} to v2v_{2}. Then the concatenation of γ1\gamma_{1}, the edges {i1(k),12(k)}\{i1^{(k)},12^{(k)}\}, {j1(k),12(k)}\{j1^{(k)},12^{(k)}\}, and the path γ2\gamma_{2} is a path joining v1v_{1} to v2v_{2}. Therefore, Gk+1AG_{k+1}^{A} is connected.

To show that Gk+1AG_{k+1}^{A} is a tree, assume for a contradiction, that there exists a simple closed path γ\gamma. Since each subgraph Gk+1,iAG_{k+1,i}^{A} is a tree, there exist distinct i,jAi,j\in A such that the path γ\gamma intersects Gk+1,iA,Gk+1,jAG_{k+1,i}^{A},G_{k+1,j}^{A}. Without loss of generality, assume that i1i\neq 1. Let σ\sigma be a maximal subarc in γGk+1,iA\gamma\cap G_{k+1,i}^{A}. But then the two endpoints of σ\sigma are both i1(k)i1^{(k)} which implies that γ\gamma is not simple. ∎

Lemma 4.4.

If AAA\subset A^{\prime} and nn\in\mathbb{N}, then GnAG_{n}^{A} is a subgraph of GnAG_{n}^{A^{\prime}}.

Proof.

Fix two alphabets A,AA,A^{\prime} with AAA\subset A^{\prime}. For each nn\in\mathbb{N}, it is clear that the vertices of GnAG^{A}_{n} are also vertices of GnAG^{A^{\prime}}_{n}. It suffices to show that for each nn\in\mathbb{N}, EnAEnAE^{A}_{n}\subset E^{A^{\prime}}_{n}. The proof is by induction on nn.

For n=1n=1, E1A={{1,i}:iA{1}}E_{1}^{A}=\{\{1,i\}:i\in A\setminus\{1\}\}\subset {{1,i}:iA{1}}=E1A\{\{1,i\}:i\in A^{\prime}\setminus\{1\}\}=E_{1}^{A^{\prime}} since AAA\subset A^{\prime}.

Assume now that EnAEnAE_{n}^{A}\subset E_{n}^{A^{\prime}} for some nn\in\mathbb{N}. We show that En+1AEn+1AE_{n+1}^{A}\subset E_{n+1}^{A^{\prime}}. Let {w,u}En+1A\{w,u\}\in E_{n+1}^{A}. If {w,u}={12(n),i1(n)}\{w,u\}=\{12^{(n)},i1^{(n)}\}, where iA{1}i\in A\setminus\{1\} then {w,u}En+1A\{w,u\}\in E_{n+1}^{A^{\prime}} since A{1}A{1}A\setminus\{1\}\subset A^{\prime}\setminus\{1\}. Assume now that {w,u}={iw,iu}\{w,u\}=\{iw^{\prime},iu^{\prime}\} for some iAi\in A with w,uAnw^{\prime},u^{\prime}\in A^{n} and {w,u}EnA\{w^{\prime},u^{\prime}\}\in E_{n}^{A}. That means that there exists an iAi\in A^{\prime} such that {w,u}={iw,iu}\{w,u\}=\{iw^{\prime},iu^{\prime}\} with w,uAnw^{\prime},u^{\prime}\in A^{{}^{\prime}n} since AnAnA^{n}\subset A^{{}^{\prime}n}. Also, since {w,u}EnA\{w^{\prime},u^{\prime}\}\in E_{n}^{A^{\prime}} it follows by inductive hypothesis that {w,u}En+1A\{w,u\}\in E_{n+1}^{A^{\prime}}. ∎

4.2. Combinatorial intersection

Given u1,u2Au_{1},u_{2}\in A^{*}, define

Au1Au2\displaystyle A^{\mathbb{N}}_{u_{1}}\wedge A^{\mathbb{N}}_{u_{2}}
:={wAu1:n>max{|u1|,|u2|} there exists uAu2n with {w(n),u}EnA}\displaystyle:=\{w\in A^{\mathbb{N}}_{u_{1}}:\forall n>\max\{|u_{1}|,|u_{2}|\}\text{ there exists $u\in A^{n}_{u_{2}}$ with $\{w(n),u\}\in E_{n}^{A}\}$}
{wAu2:n>max{|u1|,|u2|} there exists uAu1n with {w(n),u}EnA}.\displaystyle\quad\cup\{w\in A^{\mathbb{N}}_{u_{2}}:\forall n>\max\{|u_{1}|,|u_{2}|\}\text{ there exists $u\in A^{n}_{u_{1}}$ with $\{w(n),u\}\in E_{n}^{A}\}$}.

The set Au1Au2A^{\mathbb{N}}_{u_{1}}\wedge A^{\mathbb{N}}_{u_{2}} is called the combinatorial intersection of Au1A^{\mathbb{N}}_{u_{1}} and Au2A^{\mathbb{N}}_{u_{2}}.

Remark 4.5.

It is easy to see that Au1Au2=Au2Au1A^{\mathbb{N}}_{u_{1}}\wedge A^{\mathbb{N}}_{u_{2}}=A^{\mathbb{N}}_{u_{2}}\wedge A^{\mathbb{N}}_{u_{1}}. Also, if uAwu\in A_{w}^{*}, then AuAwAuA_{u}^{\mathbb{N}}\subset A_{w}^{\mathbb{N}}\wedge A^{\mathbb{N}}_{u}

With this notion of combinatorial intersection, we can describe how to move between two infinite words. Given two words w,wAw,w^{\prime}\in A^{\mathbb{N}} we say that {Aw1,,AwN}\{A^{\mathbb{N}}_{w_{1}},\dots,A^{\mathbb{N}}_{w_{N}}\} is a chain joining ww with ww^{\prime} if wAw1w\in A^{\mathbb{N}}_{w_{1}}, wAwNw^{\prime}\in A^{\mathbb{N}}_{w_{N}} and for every i=1,,N1i=1,\dots,N-1, we have AwiAwi+1A^{\mathbb{N}}_{w_{i}}\wedge A^{\mathbb{N}}_{w_{i+1}}\neq\emptyset.

4.3. A metric

A weight is a non-increasing function 𝐚:(0,1/2]\operatorname{\bf{a}}:\mathbb{N}\to(0,1/2] such that 𝐚(1)=𝐚(2)=1/2\operatorname{\bf{a}}(1)=\operatorname{\bf{a}}(2)=1/2 and limi𝐚(i)=0\lim_{i\to\infty}\operatorname{\bf{a}}(i)=0. Given a weight 𝐚\operatorname{\bf{a}} and the alphabet \mathbb{N}, define the associated “diameter function” Δ𝐚:(0,1]\Delta_{\operatorname{\bf{a}}}:\mathbb{N}^{*}\to(0,1] by Δ𝐚(ϵ)=1\Delta_{\operatorname{\bf{a}}}(\epsilon)=1 and

Δ𝐚(i1ik)=𝐚(i1)𝐚(ik).\Delta_{\operatorname{\bf{a}}}(i_{1}\cdots i_{k})=\operatorname{\bf{a}}(i_{1})\cdots\operatorname{\bf{a}}(i_{k}).
Remark 4.6.

For any weight 𝐚\operatorname{\bf{a}},

limnmax{Δ𝐚(w):wn}limn2n=0.\lim_{n\to\infty}\max\{\Delta_{\operatorname{\bf{a}}}(w):w\in\mathbb{N}^{n}\}\leq\lim_{n\to\infty}2^{-n}=0.

Fix now an alphabet AA and a weight 𝐚\operatorname{\bf{a}}. We define a function ρA,𝐚:A×A\rho_{A,\operatorname{\bf{a}}}:A^{\mathbb{N}}\times A^{\mathbb{N}}\to\mathbb{R} by:

(4.1) ρA,𝐚(w,u)=inf{i=1NΔ𝐚(vi):{Av0,,AvN} is a chain joining w with u}.\rho_{A,\operatorname{\bf{a}}}(w,u)=\inf\left\{\sum_{i=1}^{N}\Delta_{\operatorname{\bf{a}}}(v_{i}):\{A^{\mathbb{N}}_{v_{0}},\dots,A^{\mathbb{N}}_{v_{N}}\}\text{ is a chain joining $w$ with $u$}\right\}.

Following the arguments in [DV22, Lemma 3.8], we obtain that ρA,𝐚\rho_{A,\operatorname{\bf{a}}} is a pseudometric.

Taking the quotient space A/A^{\mathbb{N}}/\sim under the equivalence relation www\sim w^{\prime} whenever ρ(w,w)=0\rho(w,w^{\prime})=0, we obtain a metric space

𝕋A,𝐚=(A/,dA,𝐚),dA,𝐚([w],[u])=ρA,𝐚(w,u).\mathbb{T}^{A,\operatorname{\bf{a}}}=(A^{\mathbb{N}}/\sim,d_{A,\operatorname{\bf{a}}}),\qquad d_{A,\operatorname{\bf{a}}}([w],[u])=\rho_{A,\operatorname{\bf{a}}}(w,u).

For any wAw\in A^{*}, write 𝕋wA,𝐚:=Aw/\mathbb{T}^{A,\operatorname{\bf{a}}}_{w}:=A^{\mathbb{N}}_{w}/\sim. Under this metric we have that diam𝕋wAΔ𝐚(w)\operatorname{diam}{\mathbb{T}^{A}_{w}}\leq\Delta_{\operatorname{\bf{a}}}(w) [DV22, Lemma 3.9].

Remark 4.7.

Note that for vAv\in A^{*}, {Av12(n),Avj1(n)}\{A_{v12^{(n)}}^{\mathbb{N}},A_{vj1^{(n)}}^{\mathbb{N}}\} is a chain joining v12()v12^{(\infty)} with vj1()vj1^{(\infty)} for j{2,,m}j\in\{2,\dots,m\}. Hence, ρA,𝐚(v12(),vj1())Δ(v)(2(n+1)+𝐚(j)2n)\rho_{A,\operatorname{\bf{a}}}(v12^{(\infty)},vj1^{(\infty)})\leq\Delta(v)(2^{-(n+1)}+\operatorname{\bf{a}}(j)2^{-n}) for all nn\in\mathbb{N}, which yields that [v12()]=[vj1()][v12^{(\infty)}]=[vj1^{(\infty)}] for all j{2,,m}j\in\{2,\dots,m\}.

Lemma 4.8.

The space 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} is a 1-bounded turning metric tree.

Lemma 4.8 is essentially [DV22, Proposition 3.1]. The only difference is that in [DV22] the following condition is assumed when A=A=\mathbb{N}.

(4.2) For any wAw\in A^{*}, Δ𝐚(wi)=0\Delta_{\operatorname{\bf{a}}}(wi)=0 for all but finitely many iAi\in A.

In our setting, property (4.2) is equivalent to the following property.

(4.3) For all but finitely many iAi\in A we have 𝐚(i)=0\operatorname{\bf{a}}(i)=0.

Note that (4.3) is trivial if AA is finite.

Proof of Lemma 4.8.

We start by proving that 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} is compact. Let ([wk])([w_{k}]) be a sequence in 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}}. There are two cases to consider. Suppose first that for any vAv\in A^{*}, wkAviw_{k}\not\in A^{\mathbb{N}}_{vi} for all but finitely many iAi\in A. Then following the arguments in [DV22, Lemma 3.12] we find wAw\in A^{\mathbb{N}} and a subsequence [wkm][w_{k_{m}}] such that [wkm][w][w_{k_{m}}]\to[w]. Suppose now that there exist vAv\in A^{*}, infinitely many distinct i1,i2,Ai_{1},i_{2},\cdots\in A, and infinitely many indices k1<k2<k_{1}<k_{2}<\dots such that wkmAvimw_{k_{m}}\in A^{\mathbb{N}}_{vi_{m}} for all mm\in\mathbb{N}. By fixing mm\in\mathbb{N}, if im=1i_{m}=1 then,

dA,𝐚([wkm],[v12()])diam𝕋vimA,𝐚Δ𝐚(v)𝐚(im)m0.d_{A,\operatorname{\bf{a}}}([w_{km}],[v12^{(\infty)}])\leq\operatorname{diam}{\mathbb{T}^{A,\operatorname{\bf{a}}}_{vi_{m}}}\leq\Delta_{\operatorname{\bf{a}}}(v)\operatorname{\bf{a}}(i_{m})\xrightarrow{m\to\infty}0.

If im1i_{m}\neq 1 then by Remark 4.7 the same argument holds. Therefore, [wkm][v12()][w_{k_{m}}]\to[v12^{(\infty)}].

The fact that 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} is connected, locally connected, and path-connected follows by [DV22, Lemma 3.14]; note that (4.2) is not used in the proof of that part.

The fact that 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} is 1-bounded turning follows from [DV22, Lemma 3.15]; note that (4.2) is not used in the proof of that part.

Finally, the fact that 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} is a metric tree follows from [DV22, Lemma 3.17]. While property (4.2) is mentioned in [DV22, Claim 3.20] to prove the compactness of sets XjX_{j} therein, the compactness of these sets does not play a role in the proof of the said claim. ∎

4.4. Self-similarity

In the next lemma we define similarity maps on the trees 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}}, which are essential for various arguments in following sections.

Lemma 4.9.

For each alphabet AA, each weight 𝐚\operatorname{\bf{a}}, and each iAi\in A, the map

ϕiA,𝐚:𝕋A,𝐚𝕋iA,𝐚,ϕiA,𝐚([w])=[iw]\phi^{A,\operatorname{\bf{a}}}_{i}:\mathbb{T}^{A,\operatorname{\bf{a}}}\to\mathbb{T}^{A,\operatorname{\bf{a}}}_{i},\qquad\phi^{A,\operatorname{\bf{a}}}_{i}([w])=[iw]

is a similarity with scaling factor 𝐚(i)\operatorname{\bf{a}}(i).

Proof.

By Lemma 4.1(ii), for any iAi\in A, any kk\in\mathbb{N}, and any two words w,uAkw,u\in A^{k} we have that ww is adjacent to uu in GkG_{k} if and only if iwiw is adjacent to iuiu in Gk+1G_{k+1}. Therefore, AwAuA^{\mathbb{N}}_{w}\wedge A^{\mathbb{N}}_{u}\neq\emptyset if and only if AiwAiuA^{\mathbb{N}}_{iw}\wedge A^{\mathbb{N}}_{iu}\neq\emptyset.

Fix now w,uAw,u\in A^{\mathbb{N}} and iAi\in A. We show that

ρA,𝐚(iw,iu)=𝐚(i)ρA,𝐚(w,u).\rho_{A,\operatorname{\bf{a}}}(iw,iu)=\operatorname{\bf{a}}(i)\rho_{A,\operatorname{\bf{a}}}(w,u).

To see that, let first Av1,,AvNA^{\mathbb{N}}_{v_{1}},\dots,A^{\mathbb{N}}_{v_{N}} be a chain joining ww with uu. Then Aiv1,,AivNA^{\mathbb{N}}_{iv_{1}},\dots,A^{\mathbb{N}}_{iv_{N}} is a chain joining iwiw with iuiu. Therefore, ρA,𝐚(iw,iu)𝐚(i)ρA,𝐚(w,u)\rho_{A,\operatorname{\bf{a}}}(iw,iu)\leq\operatorname{\bf{a}}(i)\rho_{A,\operatorname{\bf{a}}}(w,u).

For the reverse inequality, we show that for any chain Av1,,AvNA^{\mathbb{N}}_{v_{1}},\dots,A^{\mathbb{N}}_{v_{N}} joining iwiw with iuiu, there exists another chain Aiu1,AiukA^{\mathbb{N}}_{iu_{1}},\dots A^{\mathbb{N}}_{iu_{k}} joining them that satisfies

Δ𝐚(iu1)++Δ𝐚(iuk)Δ𝐚(v1)++Δ𝐚(vN).\Delta_{\operatorname{\bf{a}}}(iu_{1})+\cdots+\Delta_{\operatorname{\bf{a}}}(iu_{k})\leq\Delta_{\operatorname{\bf{a}}}(v_{1})+\cdots+\Delta_{\operatorname{\bf{a}}}(v_{N}).

Noting that Au1,,AukA^{\mathbb{N}}_{u_{1}},\dots,A^{\mathbb{N}}_{u_{k}} is also a chain joining ww, with uu, assuming the claim, we readily have that ρA,𝐚(iw,iu)𝐚(i)ρA,𝐚(w,u)\rho_{A,\operatorname{\bf{a}}}(iw,iu)\geq\operatorname{\bf{a}}(i)\rho_{A,\operatorname{\bf{a}}}(w,u).

To prove the claim, assume without loss of generality that i=1i=1 and fix a chain Av1,,AvNA^{\mathbb{N}}_{v_{1}},\dots,A^{\mathbb{N}}_{v_{N}} joining 1w1w with 1u1u. If there exists j{1,,N}j\in\{1,\dots,N\} such that |vj|1|v_{j}|\leq 1, then vj{1,ϵ}v_{j}\in\{1,\epsilon\} and it is clear that the chain consisting only of A1A^{\mathbb{N}}_{1} satisfies the conclusions of the claim. Assume now that |vj|2|v_{j}|\geq 2 for all j{1,,N}j\in\{1,\dots,N\}. Let l,sl,s be the first and last, respectively, indices jj such that vjA1v_{j}\not\in A^{*}_{1}. We know that 2lsN12\leq l\leq s\leq N-1. Write vj=1ujv_{j}=1u_{j} for j{1,,l1,s+1,,N}j\in\{1,\dots,l-1,s+1,\dots,N\}. Since A1ul1AvlA^{\mathbb{N}}_{1u_{l-1}}\wedge A^{\mathbb{N}}_{v_{l}}\neq\emptyset, there exist iA{1}i^{\prime}\in A\setminus\{1\}, integer mmax{1+|ul1|,|vl|}m\geq\max\{1+|u_{l-1}|,|v_{l}|\}, uAm|ul1|u^{\prime}\in A^{m-|u_{l-1}|}, and u′′Am|vl|u^{\prime\prime}\in A^{m-|v_{l}|} such that {1ul1u,ivlu′′}Em+1A\{1u_{l-1}u^{\prime},i^{\prime}v_{l}u^{\prime\prime}\}\in E_{m+1}^{A}. By Lemma 4.1(1), 1ul1u=12(m)1u_{l-1}u^{\prime}=12^{(m)} which implies that 1ul1=12(m1)1u_{l-1}=12^{(m_{1})} for some integer m10m_{1}\geq 0. Similarly, 1us+1=12(m2)1u_{s+1}=12^{(m_{2})} for some integer m20m_{2}\geq 0. Therefore, by Remark 4.5, we have that A1ul1A1us+1A^{\mathbb{N}}_{1u_{l-1}}\wedge A^{\mathbb{N}}_{1u_{s+1}}\neq\emptyset and the collection A1u1,,A1ul1,A1us+1,,A1uNA^{\mathbb{N}}_{1u_{1}},\dots,A^{\mathbb{N}}_{1u_{l-1}},A^{\mathbb{N}}_{1u_{s+1}},\dots,A^{\mathbb{N}}_{1u_{N}} is a chain satisfying the conclusions of the claim. ∎

Note that if A={1,,m}A=\{1,\dots,m\} is a finite alphabet, then Lemma 4.9 is enough to show that all 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} are self-similar in the sense of §2.2. Indeed, 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} is the attractor of {ϕiA,𝐚:𝕋A,𝐚𝕋A,𝐚}iA\{\phi^{A,\operatorname{\bf{a}}}_{i}:\mathbb{T}^{A,\operatorname{\bf{a}}}\to\mathbb{T}^{A,\operatorname{\bf{a}}}\}_{i\in A}, and U:=𝕋A,𝐚{[1(),2()]}U:=\mathbb{T}^{A,\operatorname{\bf{a}}}\setminus\{[1^{(\infty)},2^{(\infty)}]\} is the open set required for the open set condition.

5. Geodesicity of trees 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}}

The focus of this section is to show that metric trees 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} defined above are all geodesic.

Proposition 5.1.

For each alphabet AA and each weight 𝐚\operatorname{\bf{a}} the metric tree 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} is geodesic.

Fix for the rest of this section an alphabet AA and a weight 𝐚\operatorname{\bf{a}}. Given kk\in\mathbb{N} and w,uAkw,u\in A^{k}, we denote by PA,k(w,u)P_{A,k}(w,u) the unique combinatorial arc in GkAG_{k}^{A} with endpoints ww, uu.

Remark 5.2.

By design of graphs (GkA)k(G_{k}^{A})_{k\in\mathbb{N}} we have that for every iAi\in A, kk\in\mathbb{N}, and w,uAkw,u\in A^{k}, if PA,k(w,u)=(v1,,vn)P_{A,k}(w,u)=(v_{1},\dots,v_{n}), then (iv1,,ivn)(iv_{1},\dots,iv_{n}) is also an arc which we denote by iPA,k(w,u)iP_{A,k}(w,u). Therefore,

PA,k+1(iw,iu)=iPA,k(w,u).P_{A,k+1}(iw,iu)=iP_{A,k}(w,u).

Recall the definition of length (γ)\ell(\gamma) of a combinatorial path γ\gamma, and the concatenation γγ\gamma\cdot\gamma^{\prime} of paths γ,γ\gamma,\gamma^{\prime} in a graph GG from §2.6.

Lemma 5.3.

For each nn\in\mathbb{N},

PA,n(1(n1)3,2(n))=PA,n(1(n1)3,12(n1))PA,n(12(n1),2(n)).P_{A,n}(1^{(n-1)}3,2^{(n)})=P_{A,n}(1^{(n-1)}3,12^{(n-1)})\cdot P_{A,n}(12^{(n-1)},2^{(n)}).

Moreover, for any arc γ\gamma in GnAG_{n}^{A},

(γ)(PA,n(1(n1)3,2(n)))=2n+1.\ell(\gamma)\leq\ell(P_{A,n}(1^{(n-1)}3,2^{(n)}))=2^{n}+1.
Proof.

For the first claim, fix nn\in\mathbb{N}. By Lemma 4.1(i) we have that if 1w1w is adjacent to 2u2u in some graph GnAG^{A}_{n} (here w,uAn1w,u\in A^{n-1}), then w=2(n1)w=2^{(n-1)} and u=1(n1)u=1^{(n-1)}. Since GnAG_{n}^{A} is a combinatorial tree, every path in GnAG^{A}_{n} that goes from 1(n1)31^{(n-1)}3 to 2(n)2^{(n)} must contain the (directed) edge (12(n1),21(n1))(12^{(n-1)},21^{(n-1)}). Therefore,

PA,n(1(n1)3,2(n))=PA,n(1(n1)3,12(n1))PA,n(12(n1),2(n)).P_{A,n}(1^{(n-1)}3,2^{(n)})=P_{A,n}(1^{(n-1)}3,12^{(n-1)})\cdot P_{A,n}(12^{(n-1)},2^{(n)}).

The proof of the second claim is by induction on nn. For n=1n=1, PA,1(3,2)=(3,1,2)P_{A,1}(3,2)=(3,1,2) is the arc of longest length in G1AG_{1}^{A} and (PA,1(1,2))=3\ell(P_{A,1}(1,2))=3. Assume now that the second claim is true for some nn\in\mathbb{N}. First note that the only point adjacent to 1(n)31^{(n)}3 in Gn+1AG^{A}_{n+1} is 1(n+1)1^{(n+1)}. Therefore, by the first claim and by Remark 5.2,

PA,n+1(\displaystyle P_{A,n+1}( 1(n)3,2(n+1))\displaystyle 1^{(n)}3,2^{(n+1)})
=PA,n+1(1(n)3,12(n))(12(n),21(n))PA,n+1(21(n),2(n+1))\displaystyle=P_{A,n+1}(1^{(n)}3,12^{(n)})\cdot(12^{(n)},21^{(n)})\cdot P_{A,n+1}(21^{(n)},2^{(n+1)})
=(1PA,n(1(n1)3,2(n)))(12(n),21(n))(2PA,n(1(n),2(n)))\displaystyle=(1P_{A,n}(1^{(n-1)}3,2^{(n)}))\cdot(12^{(n)},21^{(n)})\cdot(2P_{A,n}(1^{(n)},2^{(n)}))
=(1PA,n(1(n1)3,2(n)))(12(n),21(n))(2PA,n(1(n1)3,2(n)){1(n1)3})\displaystyle=(1P_{A,n}(1^{(n-1)}3,2^{(n)}))\cdot(12^{(n)},21^{(n)})\cdot(2P_{A,n}(1^{(n-1)}3,2^{(n)})\setminus\{1^{(n-1)}3\})

which by the inductive assumption implies that (PA,n+1(1(n)3,2(n+1)))=2n+1+1\ell(P_{A,n+1}(1^{(n)}3,2^{(n+1)}))=2^{n+1}+1.

Let now PP be an arc in Gn+1G_{n+1}. If PP crosses vertices contained entirely in the vertices of jGnjG_{n} for some jAj\in A, then P=jPP=jP^{\prime} for some arc PP^{\prime} in GnG_{n}. By the inductive hypothesis,

(P)=(P)(PA,n(1(n1)3,2(n)))=2n+1.\ell(P)=\ell(P^{\prime})\leq\ell(P_{A,n}(1^{(n-1)}3,2^{(n)}))=2^{n}+1.

Assume for the rest of the proof that PP crosses vertices of at least two sub-graphs jGnjG_{n}, kGnkG_{n} of Gn+1G_{n+1} for distinct j,kAj,k\in A. Note that PP does not cross vertices lying on three or more distinct sub-graphs, unless one of them is 1Gn1G_{n}, which it only crosses at 12(n)12^{(n)}. Indeed, suppose it contains vertices from the sub-graphs jGnjG_{n}, kGnkG_{n}, lGnlG_{n} of Gn+1G_{n+1} with distinct j,k,lAj,k,l\in A. By the inductive definition of Gn+1G_{n+1}, any path joining vertices of two distinct sub-graphs has to contain 12(n)12^{(n)}. This implies that one of the sub-graphs is 1Gn1G_{n}. Without loss of generality, assume that l=1l=1. Moreover, if vi1v_{i_{1}}, vi2v_{i_{2}}, vi3v_{i_{3}} are vertices of jGnjG_{n}, kGnkG_{n}, 1Gn{12(n)}1G_{n}\setminus\{12^{(n)}\}, respectively, then the sub-paths of PP joining vi1v_{i_{1}} with vi2v_{i_{2}} and joining vi2v_{i_{2}} with vi3v_{i_{3}} both contain 12(n)12^{(n)}, since Gn+1G_{n+1} is a combinatorial tree. Therefore,

P=(v1,,vi1,,12(n),,vi2,,12(n),,vi3,vM),P=(v_{1},\dots,v_{i_{1}},\dots,12^{(n)},\dots,v_{i_{2}},\dots,12^{(n)},\dots,v_{i_{3}},\dots v_{M}),

which is a contradiction. As a result, PP crosses vertices of either exactly two sub-graphs jGnjG_{n}, 1Gn1G_{n} of Gn+1G_{n+1} for jA{1}j\in A\setminus\{1\}, or exactly three sub-graphs jGnjG_{n}, kGnkG_{n}, 1Gn1G_{n} of Gn+1G_{n+1} for distinct j,kA{1}j,k\in A\setminus\{1\} and the only vertex of 1Gn1G_{n} it contains is 12(n)12^{(n)}. In the latter case, note that 12(n)12^{(n)} cannot be an end-point of PP, since then there is a vertex q1(n)q1^{(n)} of Gn+1G_{n+1} with qA{1,j,k}q\in A\setminus\{1,j,k\}, i.e. which is not a vertex of jGnjG_{n}, kGnkG_{n}, and {12(n),q1(n)}En+1\{12^{(n)},q1^{(n)}\}\in E_{n+1}, implying that the length of PP is not maximal in Gn+1G_{n+1}.

Suppose PP intersects exactly three sub-graphs (the case where it intersects exactly two is similar). We decompose P=(jP1)((12(n)))(kP2)P=(jP_{1})\cdot((12^{(n)}))\cdot(kP_{2}), where j,kA{1}j,k\in A\setminus\{1\} are distinct and P1,P2P_{1},P_{2} are paths in GnG_{n}. If none of P1P_{1}, P2P_{2} is maximal in GnG_{n}, then by the inductive hypothesis

(P)=(jP1)+1+(kP2)=(P1)+1+(P2)2n+1+2n=2n+1+1.\ell(P)=\ell(jP_{1})+1+\ell(kP_{2})=\ell(P_{1})+1+\ell(P_{2})\leq 2^{n}+1+2^{n}=2^{n+1}+1.

Note that {v,12(n)}En+1\{v,12^{(n)}\}\in E_{n+1} if, and only if, v=i1(n)v=i1^{(n)} for some iA{1}i\in A\setminus\{1\}. As a result, both P1P_{1} and P2P_{2} have 1(n)1^{(n)} as an end-point. However, we have {1,2},{1,3}E1\{1,2\},\{1,3\}\in E_{1}, which implies {1(n),1(n1)2},{1(n),1(n1)3}En\{1^{(n)},1^{(n-1)}2\},\{1^{(n)},1^{(n-1)}3\}\in E_{n} by Lemma 4.1(ii). Therefore, not all endpoints of P1P_{1} and P2P_{2} are leaves of GnG_{n}, implying that they are both not maximal in GnG_{n}. Hence, for every path PP in Gn+1G_{n+1}, (P)2n+1+1\ell(P)\leq 2^{n+1}+1. ∎

Corollary 5.4.

For every nn\in\mathbb{N}, (PA,n(1(n),2(n)))=2n\ell(P_{A,n}(1^{(n)},2^{(n)}))=2^{n}.

Proof.

By Lemma 4.1(ii), for each nn\in\mathbb{N} we have {1(n1)3,1(n)}EnA\{1^{(n-1)}3,1^{(n)}\}\in E_{n}^{A} and

PA,n(1(n1)3,2(n))=(1(n1)3,1(n))PA,n(1(n),2(n)).P_{A,n}(1^{(n-1)}3,2^{(n)})=(1^{(n-1)}3,1^{(n)})\cdot P_{A,n}(1^{(n)},2^{(n)}).

Hence, (PA,n(1(n),2(n)))=2n\ell(P_{A,n}(1^{(n)},2^{(n)}))=2^{n}. ∎

Given w,uAw,u\in A^{\mathbb{N}}, we denote by 𝒞A,n(w,u)\mathcal{C}_{A,n}(w,u) the unique chain {Av1,,Avk}\{A^{\mathbb{N}}_{v_{1}},\dots,A^{\mathbb{N}}_{v_{k}}\} where PA,n(w(n),u(n))=(v1,,vk)P_{A,n}(w(n),u(n))=(v_{1},\cdots,v_{k}). Moreover, given a chain C={Aw1,,Awm}C=\{A^{\mathbb{N}}_{w_{1}},\dots,A^{\mathbb{N}}_{w_{m}}\} we set

𝐚(C)=Δ𝐚(w1)++Δ𝐚(wm).\ell_{\operatorname{\bf{a}}}(C)=\Delta_{\operatorname{\bf{a}}}(w_{1})+\cdots+\Delta_{\operatorname{\bf{a}}}(w_{m}).
Corollary 5.5.

Let wAw\in A^{*}, n>|w|n>|w|, and u,uAwu,u^{\prime}\in A^{\mathbb{N}}_{w}. Then,

𝐚(𝒞A,n(u,u))Δ𝐚(w)(1+2|w|n).\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(u,u^{\prime}))\leq\Delta_{\operatorname{\bf{a}}}(w)(1+2^{|w|-n}).
Proof.

Since the subgraph of GnAG_{n}^{A} induced by vertices AwnA^{n}_{w} is connected, the combinatorial arc PA,n(u(n),u(n))P_{A,n}(u(n),u^{\prime}(n)) has all its vertices in AwnA^{n}_{w}. Therefore, we can write 𝒞A,n(u,u)={Awv1,,Awvk}\mathcal{C}_{A,n}(u,u^{\prime})=\{A^{\mathbb{N}}_{wv_{1}},\dots,A^{\mathbb{N}}_{wv_{k}}\} where (v1,,vk)(v_{1},\dots,v_{k}) is a combinatorial arc in An|w|A^{n-|w|}. By Lemma 5.3, k2n|w|+1k\leq 2^{n-|w|}+1. Since Δ𝐚(vi)2|w|n\Delta_{\operatorname{\bf{a}}}(v_{i})\leq 2^{|w|-n} for all ii

𝐚(𝒞A,n(u,u))=Δ𝐚(w)(Δ𝐚(v1)++Δ𝐚(vk))Δ𝐚(w)(1+2|w|n).\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(u,u^{\prime}))=\Delta_{\operatorname{\bf{a}}}(w)\left(\Delta_{\operatorname{\bf{a}}}(v_{1})+\cdots+\Delta_{\operatorname{\bf{a}}}(v_{k})\right)\leq\Delta_{\operatorname{\bf{a}}}(w)(1+2^{|w|-n}).\qed
Lemma 5.6.

For all w,uAw,u\in A^{\mathbb{N}}, ρA,𝐚(w,u)=limn𝐚(𝒞A,n(w,u))\rho_{A,\operatorname{\bf{a}}}(w,u)=\lim_{n\to\infty}\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(w,u)).

Proof.

Fix w,uAw,u\in A^{\mathbb{N}}. It suffices to show that for any chain CC connecting ww and uu and for any ϵ(0,1)\epsilon\in(0,1) there exists NN\in\mathbb{N} such that for all nNn\geq N,

(5.1) 𝐚(𝒞A,n(w,u))𝐚(C)+ϵ.\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(w,u))\leq\ell_{\operatorname{\bf{a}}}(C)+\epsilon.

To this end, fix a chain C={Aw1,,Awk}C=\{A^{\mathbb{N}}_{w_{1}},\dots,A^{\mathbb{N}}_{w_{k}}\} connecting ww and uu and ϵ(0,1)\epsilon\in(0,1). Choose NN\in\mathbb{N} large enough such that

N>log(k/ϵ)log2+maxi=1,,k|wi|.N>\frac{\log(k/\epsilon)}{\log{2}}+\max_{i=1,\dots,k}|w_{i}|.

For each i{2,,k}i\in\{2,\dots,k\} there exists vi1Awi1Nv_{i-1}^{\prime}\in A^{N}_{w_{i-1}} and viAwiNv_{i}\in A^{N}_{w_{i}} such that vi1v_{i-1}^{\prime} is adjacent to viv_{i} in GNAG^{A}_{N}. Set also v1=w(N)v_{1}=w(N) and vk=u(N)v_{k}^{\prime}=u(N). Note that

𝒞A,N(v1,v1)𝒞A,N(vk,vk)\mathcal{C}_{A,N}(v_{1},v_{1}^{\prime})\cup\cdots\cup\mathcal{C}_{A,N}(v_{k},v_{k}^{\prime})

is a chain joining ww with uu and it contains 𝒞A,N(w,u)\mathcal{C}_{A,N}(w,u). By Corollary 5.5 and the choice of NN,

𝐚(𝒞A,N(w,u))\displaystyle\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,N}(w,u)) 𝐚(𝒞A,N(v1,v1))++𝐚(𝒞A,N(vk,vk))\displaystyle\leq\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,N}(v_{1},v_{1}^{\prime}))+\cdots+\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,N}(v_{k},v_{k}^{\prime}))
Δ𝐚(w1)(1+2|w1|N)++Δ𝐚(wk)(1+2|wk|N)\displaystyle\leq\Delta_{\operatorname{\bf{a}}}(w_{1})(1+2^{|w_{1}|-N})+\cdots+\Delta_{\operatorname{\bf{a}}}(w_{k})(1+2^{|w_{k}|-N})
Δ𝐚(w1)(1+ϵ/k)++Δ𝐚(wk)(1+ϵ/k)\displaystyle\leq\Delta_{\operatorname{\bf{a}}}(w_{1})(1+\epsilon/k)+\cdots+\Delta_{\operatorname{\bf{a}}}(w_{k})(1+\epsilon/k)
𝐚(C)+ϵ.\displaystyle\leq\ell_{\operatorname{\bf{a}}}(C)+\epsilon.

By (5.1), we have that lim supn𝐚(𝒞A,n(w,u))ρA,𝐚(w,u)\limsup_{n\to\infty}\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(w,u))\leq\rho_{A,\operatorname{\bf{a}}}(w,u). On the other hand, 𝐚(𝒞A,n(w,u))ρA,𝐚(w,u)\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(w,u))\geq\rho_{A,\operatorname{\bf{a}}}(w,u) for all nn so

lim infn𝐚(𝒞A,n(w,u))ρA,𝐚(w,u).\liminf_{n\to\infty}\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(w,u))\geq\rho_{A,\operatorname{\bf{a}}}(w,u).\qed

We are now ready to prove Proposition 5.1.

Proof of Proposition 5.1.

Fix [w],[u]𝕋A,𝐚[w],[u]\in\mathbb{T}^{A,\operatorname{\bf{a}}} and let vAv\in A^{\mathbb{N}} so that [v][v] lies on the unique arc that connects [w][w] with [u][u]. Following the proof of [DV22, Lemma 3.17], we may assume that v(n)PA,n(w(n),u(n))v(n)\in P_{A,n}(w(n),u(n)) for all nn\in\mathbb{N}. Recall that PA,n(w(n),u(n))=PA,n(w(n),v(n))PA,n(v(n),u(n))P_{A,n}(w(n),u(n))=P_{A,n}(w(n),v(n))\cdot P_{A,n}(v(n),u(n)). By Lemma 5.6,

dA,𝐚([w],[v])+dA,𝐚([v],[u])\displaystyle d_{A,\operatorname{\bf{a}}}([w],[v])+d_{A,\operatorname{\bf{a}}}([v],[u]) =limn(𝐚(𝒞A,n(w,v))+𝐚(𝒞A,n(v,u)))\displaystyle=\lim_{n\to\infty}(\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(w,v))+\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(v,u)))
=limn(𝐚(𝒞A,n(w,u))Δ𝐚(v(n)))\displaystyle=\lim_{n\to\infty}(\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(w,u))-\Delta_{\operatorname{\bf{a}}}(v(n)))
=dA,𝐚([w],[u])\displaystyle=d_{A,\operatorname{\bf{a}}}([w],[u])

which proves the geodesicity of 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}}. ∎

Corollary 5.7.

We have that dA,𝐚([1()],[2()])=1d_{A,\operatorname{\bf{a}}}([1^{(\infty)}],[2^{(\infty)}])=1.

Proof.

Note that for each nn\in\mathbb{N}, every vertex in PA,n(1(n),2(n))P_{A,n}(1^{(n)},2^{(n)}) is in {1,2}n\{1,2\}^{n} due to Lemma 4.4. Therefore, every vertex vv in PA,n(1(n),2(n))P_{A,n}(1^{(n)},2^{(n)}) satisfies Δ𝐚(v)=2n\Delta_{\operatorname{\bf{a}}}(v)=2^{-n}. By Lemma 5.6 and Corollary 5.4,

dA,𝐚([1()],[2()])=limn𝐚(𝒞A,n(1(n),2(n)))=1.d_{A,\operatorname{\bf{a}}}([1^{(\infty)}],[2^{(\infty)}])=\lim_{n\to\infty}\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(1^{(n)},2^{(n)}))=1.\qed
Corollary 5.8.

For any wAw\in A^{*}, diam𝕋wA,𝐚=Δ𝐚(w)\operatorname{diam}\mathbb{T}_{w}^{A,\operatorname{\bf{a}}}=\Delta_{\operatorname{\bf{a}}}(w).

Proof.

Fix wAw\in A^{*}. On the one hand, diam𝕋wAΔ𝐚(w)\operatorname{diam}{\mathbb{T}^{A}_{w}}\leq\Delta_{\operatorname{\bf{a}}}(w) [DV22, Lemma 3.9]. On the other hand, by Lemma 4.9 and Corollary 5.7,

diam𝕋wA,𝐚dA,𝐚([w1()],[w2()])=Δ𝐚(w)ρA,𝐚(1(),2())=Δ𝐚(w).\operatorname{diam}\mathbb{T}_{w}^{A,\operatorname{\bf{a}}}\geq d_{A,\operatorname{\bf{a}}}([w1^{(\infty)}],[w2^{(\infty)}])=\Delta_{\operatorname{\bf{a}}}(w)\rho_{A,\operatorname{\bf{a}}}(1^{(\infty)},2^{(\infty)})=\Delta_{\operatorname{\bf{a}}}(w).\qed
Remark 5.9.

Corollary 5.7 and Corollary 5.8, along with the fact that 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} is geodesic, imply that both [1()][1^{(\infty)}] and [2()][2^{(\infty)}] are leaves of 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}}.

5.1. Isometric embeddings

Here we show that if AAA\subset A^{\prime} and 𝐚\operatorname{\bf{a}} is a weight, then 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} is essentially a subset of 𝕋A,𝐚\mathbb{T}^{A^{\prime},\operatorname{\bf{a}}}.

Lemma 5.10.

Let A,AA,A^{\prime} be two alphabets with AAA\subset A^{\prime}, and let 𝐚,𝐚\operatorname{\bf{a}},\operatorname{\bf{a}}^{\prime} be two weights with 𝐚|A=𝐚|A\operatorname{\bf{a}}^{\prime}|_{A}=\operatorname{\bf{a}}|_{A}. Then the map

Φ:𝕋A,𝐚𝕋A,𝐚,Φ([w])=[w]\Phi:\mathbb{T}^{A,\operatorname{\bf{a}}}\to\mathbb{T}^{A^{\prime},\operatorname{\bf{a}}^{\prime}},\qquad\Phi([w])=[w]

is an isometric embedding. Moreover, if AAA\subsetneq A^{\prime} then

distH(Φ(𝕋A,𝐚),𝕋A,𝐚)maxjAA𝐚(j).\operatorname{dist}_{H}(\Phi(\mathbb{T}^{A,\operatorname{\bf{a}}}),\mathbb{T}^{A^{\prime},\operatorname{\bf{a}}^{\prime}})\leq\max_{j\in A^{\prime}\setminus A}\operatorname{\bf{a}}^{\prime}(j).
Proof.

For the first claim of the lemma, fix w,uA(A)w,u\in A^{\mathbb{N}}\subset(A^{\prime})^{\mathbb{N}}. We show that ρA,𝐚(w,u)=ρA,𝐚(w,u)\rho_{A^{\prime},\operatorname{\bf{a}}^{\prime}}(w,u)=\rho_{A,\operatorname{\bf{a}}}(w,u). To this end, fix nn\in\mathbb{N} and write PA,n(w(n),u(n))={v0,,vk}P_{A,n}(w(n),u(n))=\{v_{0},\dots,v_{k}\} and PA,n(w(n),u(n))={v0,,vk}P_{A^{\prime},n}(w(n),u(n))=\{v^{\prime}_{0},\dots,v^{\prime}_{k^{\prime}}\}. By Lemma 4.4, GnAG_{n}^{A} is a subgraph of GnAG_{n}^{A^{\prime}}, so implies that PA,n(w(n),u(n))P_{A,n}(w(n),u(n)) is a combinatorial arc in GnAG_{n}^{A^{\prime}}. Since GnAG_{n}^{A^{\prime}} is a combinatorial tree, it follows that PA,n(w(n),u(n))=PA,n(w(n),u(n))P_{A,n}(w(n),u(n))=P_{A^{\prime},n}(w(n),u(n)); that is, k=kk=k^{\prime} and vi=viv_{i}=v^{\prime}_{i} for all i{1,,k}i\in\{1,\dots,k\}. Since, viAv^{\prime}_{i}\in A^{\mathbb{N}}, we have Δ𝐚(vi)=Δ𝐚(vi)\Delta_{\operatorname{\bf{a}}}(v_{i})=\Delta_{\operatorname{\bf{a}}^{\prime}}(v_{i}). By Lemma 5.6,

ρA,𝐚(w,u)=limn𝐚(𝒞A,n(w,u))=limn𝐚(𝒞A,n(w,u))=ρA,𝐚(w,u).\rho_{A^{\prime},\operatorname{\bf{a}}^{\prime}}(w,u)=\lim_{n\to\infty}\ell_{\operatorname{\bf{a}}^{\prime}}(\mathcal{C}_{A^{\prime},n}(w,u))=\lim_{n\to\infty}\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(w,u))=\rho_{A,\operatorname{\bf{a}}}(w,u).

For the second claim, assume that AAA\subsetneq A^{\prime}. Since Φ(𝕋A,𝐚)𝕋A,𝐚\Phi(\mathbb{T}^{A,\operatorname{\bf{a}}})\subset\mathbb{T}^{A^{\prime},\operatorname{\bf{a}}^{\prime}},

distH(Φ(𝕋A,𝐚),𝕋A,𝐚)=supv(A)AinfwAρA,𝐚(w,v).\operatorname{dist}_{H}(\Phi(\mathbb{T}^{A,\operatorname{\bf{a}}}),\mathbb{T}^{A^{\prime},\operatorname{\bf{a}}^{\prime}})=\sup_{v\in(A^{\prime})^{\mathbb{N}}\setminus A^{\mathbb{N}}}\inf_{w\in A^{\mathbb{N}}}\rho_{A^{\prime},\operatorname{\bf{a}}^{\prime}}(w,v).

Let v(A)Av\in(A^{\prime})^{\mathbb{N}}\setminus A^{\mathbb{N}} and write v=u1juv=u_{1}ju where u1Au_{1}\in A^{*}, jAAj\in A^{\prime}\setminus A and u(A)u\in(A^{\prime})^{\mathbb{N}}. By Lemma 4.9, Remark 4.7, and Corollary 5.8,

infwAρA,𝐚(w,v)\displaystyle\inf_{w\in A^{\mathbb{N}}}\rho_{A^{\prime},\operatorname{\bf{a}}^{\prime}}(w,v) ρA,𝐚(u112(),u1ju)\displaystyle\leq\rho_{A^{\prime},\operatorname{\bf{a}}^{\prime}}(u_{1}12^{(\infty)},u_{1}ju)
=Δ𝐚(u1)ρA,𝐚(12(),ju)\displaystyle=\Delta_{\operatorname{\bf{a}}^{\prime}}(u_{1})\rho_{A^{\prime},\operatorname{\bf{a}}^{\prime}}(12^{(\infty)},ju)
=Δ𝐚(u1)ρA,𝐚(j1(),ju)\displaystyle=\Delta_{\operatorname{\bf{a}}^{\prime}}(u_{1})\rho_{A^{\prime},\operatorname{\bf{a}}^{\prime}}(j1^{(\infty)},ju)
Δ𝐚(u1)𝐚(j)\displaystyle\leq\Delta_{\operatorname{\bf{a}}^{\prime}}(u_{1})\operatorname{\bf{a}}^{\prime}(j)
maxjAA𝐚(j).\displaystyle\leq\max_{j\in A^{\prime}\setminus A}\operatorname{\bf{a}}^{\prime}(j).\qed

Based on Lemma 5.10, we make the following convention for the rest of the paper.

Convention.

If AAA\subset A^{\prime} are two alphabets and 𝐚\operatorname{\bf{a}} is a weight, we write from now on 𝕋A,𝐚𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}}\subset\mathbb{T}^{A^{\prime},\operatorname{\bf{a}}} identifying 𝕋A,𝐚\mathbb{T}^{A,\operatorname{\bf{a}}} with its isometric embedded image in 𝕋A,𝐚\mathbb{T}^{A^{\prime},\operatorname{\bf{a}}}.

If A={1,,n}A=\{1,\dots,n\} for some n{2,3,}n\in\{2,3,\dots\} and 𝐚\operatorname{\bf{a}} is a weight, we write 𝕋n,𝐚:=𝕋A,𝐚\mathbb{T}^{n,\operatorname{\bf{a}}}:=\mathbb{T}^{A,\operatorname{\bf{a}}}. If A=A=\mathbb{N}, then we write 𝕋,𝐚:=𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}}:=\mathbb{T}^{\mathbb{N},\operatorname{\bf{a}}}. Under the above notation and convention, we have for a weight 𝐚\operatorname{\bf{a}},

𝕋2,𝐚𝕋3,𝐚𝕋,𝐚\mathbb{T}^{2,\operatorname{\bf{a}}}\subset\mathbb{T}^{3,\operatorname{\bf{a}}}\subset\cdots\subset\mathbb{T}^{\infty,\operatorname{\bf{a}}}

and 𝕋n,𝐚\mathbb{T}^{n,\operatorname{\bf{a}}} converge in the Hausdorff sense to 𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}}.

6. Branch points and valence of trees 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}

Recall the notation 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} from the end of Section 5. In this section we study the valence of trees 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

Proposition 6.1.

Let 𝐚\operatorname{\bf{a}} be a weight. The space 𝕋2,𝐚\mathbb{T}^{2,\operatorname{\bf{a}}} is a metric arc. Moreover, if m{2,3,}m\in\{2,3,\dots\}, then 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is uniformly mm-branching.

Fix for the rest of this section an integer m{2,3,}m\in\{2,3,\dots\} and a weight 𝐚\operatorname{\bf{a}}, and let A={1,,m}A=\{1,\dots,m\}. The case where m=2m=2 is given in [DV22, Lemma 6.1]. Therefore, we may assume for the rest of this section that m3m\geq 3.

Recall the similarity maps ϕiA,𝐚\phi^{A,\operatorname{\bf{a}}}_{i} from Lemma 4.9. To simplify the notation below, we drop the superindexes A,𝐚A,\operatorname{\bf{a}} and simply write ϕi\phi_{i}. Given w=i1inAw=i_{1}\cdots i_{n}\in A^{*} and uAu\in A^{\mathbb{N}}, we write

ϕw([u]):=(ϕi1ϕin)([u])=[wu]\phi_{w}([u]):=(\phi_{i_{1}}\circ\cdots\circ\phi_{i_{n}})([u])=[wu]

with the convention that ϕϵ\phi_{\epsilon} is the identity map. We also write for each wAw\in A^{*}

𝕋wm,𝐚=ϕw(𝕋m,𝐚).\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}=\phi_{w}(\mathbb{T}^{m,\operatorname{\bf{a}}}).

Note that 𝕋wm,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w} are subtrees of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

We split the proof of Proposition 6.1 into several steps. First we show how different 𝕋im,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{i}, 𝕋jm,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{j} intersect in the following Lemma.

Lemma 6.2.
  1. (i)

    For distinct i,jAi,j\in A, we have 𝕋im,𝐚𝕋jm,𝐚={[12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}_{i}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{j}=\{[12^{(\infty)}]\}.

  2. (ii)

    We have [w]=[12()][w]=[12^{(\infty)}] if, and only if, w{12,21(),31(),,m1()}w\in\{12^{\infty},21^{(\infty)},31^{(\infty)},\dots,m1^{(\infty)}\}.

  3. (iii)

    We have that [1()]𝕋1m,𝐚(j2𝕋jm,𝐚)[1^{(\infty)}]\in\mathbb{T}^{m,\operatorname{\bf{a}}}_{1}\setminus(\bigcup_{j\geq 2}\mathbb{T}^{m,\operatorname{\bf{a}}}_{j}) and [2()]𝕋2m,𝐚(j2𝕋jm,𝐚)[2^{(\infty)}]\in\mathbb{T}^{m,\operatorname{\bf{a}}}_{2}\setminus(\bigcup_{j\neq 2}\mathbb{T}^{m,\operatorname{\bf{a}}}_{j}).

Proof.

For (i), note that by Remark 4.7 [12()]=[j1()]𝕋jm,𝐚[12^{(\infty)}]=[j1^{(\infty)}]\in\mathbb{T}^{m,\operatorname{\bf{a}}}_{j}. Assume for a contradiction that there is p𝕋im,𝐚𝕋jm,𝐚{[12()]}p\in\mathbb{T}^{m,\operatorname{\bf{a}}}_{i}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{j}\setminus\{[12^{(\infty)}]\} for some distinct i,jAi,j\in A. Suppose p=[u]=[v]p=[u]=[v] where uAiu\in A_{i}^{\mathbb{N}} and vAjv\in A_{j}^{\mathbb{N}}. By Lemma 5.6,

dA,𝐚([u],[v])=ρA,𝐚(u,v)=limn𝐚(𝒞A,n(u,v))=0.d_{A,\operatorname{\bf{a}}}([u],[v])=\rho_{A,\operatorname{\bf{a}}}(u,v)=\lim_{n\rightarrow\infty}\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(u,v))=0.

However, u(n)Ainu(n)\in A_{i}^{n} and v(n)Ajnv(n)\in A_{j}^{n}, so the combinatorial arc PA,n(u(n),v(n))P_{A,n}(u(n),v(n)) needs to cross the vertex 12(n1)12^{(n-1)}, implying that

PA,n(u(n),v(n))=PA,n(u(n),12(n1))PA,n(12(n1),v(n))).P_{A,n}(u(n),v(n))=P_{A,n}(u(n),12^{(n-1)})\cdot P_{A,n}(12^{(n-1)},v(n))).

Hence,

ρ(u,12())=limn𝐚(𝒞A,n(u,12()))limn𝐚(𝒞A,n(u,v))=0,\rho(u,12^{(\infty)})=\lim_{n\rightarrow\infty}\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(u,12^{(\infty)}))\leq\lim_{n\rightarrow\infty}\ell_{\operatorname{\bf{a}}}(\mathcal{C}_{A,n}(u,v))=0,

which contradicts p=[u][12()]p=[u]\neq[12^{(\infty)}].

For (ii), the left inclusion follows by Remark 4.7. For the right inclusion let w~=i1i2A{12(),21(),31(),,m1()}\tilde{w}=i_{1}i_{2}\cdots\in A^{\mathbb{N}}\setminus\{12^{(\infty)},21^{(\infty)},31^{(\infty)},\dots,m1^{(\infty)}\} with [w~]=[12()][\tilde{w}]=[12^{(\infty)}]. Assume that i1=1i_{1}=1; the case i11i_{1}\neq 1 can be treated similarly, by replacing 12()12^{(\infty)} with j1()j1^{(\infty)} in the following arguments. We show that ρA,𝐚(w~,12())>0\rho_{A,\operatorname{\bf{a}}}(\tilde{w},12^{(\infty)})>0. Let n2n\geq 2 be the smallest integer such that in2i_{n}\neq 2. Then

ρA,𝐚(w~,12())=ρA,𝐚(1i2i3,12())=𝐚(1)𝐚(i2)𝐚(in1)ρA,𝐚(w,2()),\rho_{A,\operatorname{\bf{a}}}(\tilde{w},12^{(\infty)})=\rho_{A,\operatorname{\bf{a}}}(1i_{2}i_{3}\cdots,12^{(\infty)})=\operatorname{\bf{a}}(1)\operatorname{\bf{a}}(i_{2})\cdots\operatorname{\bf{a}}(i_{n-1})\rho_{A,\operatorname{\bf{a}}}(w,2^{(\infty)}),

where w=inin+1A2w=i_{n}i_{n+1}\cdots\notin A_{2}^{\mathbb{N}}. Since wAA2w\in A^{\mathbb{N}}\setminus A^{\mathbb{N}}_{2}, by (i) and the geodesicity of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}},

dA,𝐚([w],2())dA,𝐚(12(),2())=𝐚(2)ρA,𝐚(1(),2())=12.d_{A,\operatorname{\bf{a}}}([w],2^{(\infty)})\geq d_{A,\operatorname{\bf{a}}}(12^{(\infty)},2^{(\infty)})=\operatorname{\bf{a}}(2)\rho_{A,\operatorname{\bf{a}}}(1^{(\infty)},2^{(\infty)})=\tfrac{1}{2}.

Therefore, ρA,𝐚(w~,12())>0\rho_{A,\operatorname{\bf{a}}}(\tilde{w},12^{(\infty)})>0.

Claim (iii) follows by (i) and (ii). ∎

We need the following general result about connected components of subsets of a metric space.

Lemma 6.3 ([Sea07, Theorem 11.5.3]).

Let XX be a metric space and EE a non-empty subset of XX. If A1,,AnEA_{1},\dots,A_{n}\subset E, nn\in\mathbb{N}, are non-empty, pairwise disjoint, relatively closed, and connected sets with E=A1AnE=A_{1}\cup\dots\cup A_{n}, then these are the components of EE.

The next lemma shows that all points {[u12()]:uA}\{[u12^{(\infty)}]:u\in A^{*}\}, have valence mm.

Lemma 6.4.
  1. (i)

    The components of 𝕋m,𝐚{[12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[12^{(\infty)}]\} are exactly the sets 𝕋im,𝐚{[12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}_{i}\setminus\{[12^{(\infty)}]\}, iAi\in A.

  2. (ii)

    If uAu\in A^{*}, then 𝕋m,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\} has exactly mm components. Each set 𝕋uim,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}_{ui}\setminus\{[u12^{(\infty)}]\}, for iAi\in A, lies in a different component of 𝕋m,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\}.

Proof.

For (i), by Remark 5.9, the sets 𝕋m,𝐚{[1()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[1^{(\infty)}]\} and 𝕋m,𝐚{[2()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[2^{(\infty)}]\} are non-empty, and connected. Note that 𝕋1m,𝐚{[12()]}=ϕ1(𝕋m,𝐚{[2()]})\mathbb{T}^{m,\operatorname{\bf{a}}}_{1}\setminus\{[12^{(\infty)}]\}=\phi_{1}(\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[2^{(\infty)}]\}) and 𝕋jm,𝐚{[12()]}=ϕj(𝕋{[1()]})\mathbb{T}^{m,\operatorname{\bf{a}}}_{j}\setminus\{[12^{(\infty)}]\}=\phi_{j}(\mathbb{T}\setminus\{[1^{(\infty)}]\}) for all jA{1}j\in A\setminus\{1\} by Lemma 6.2(ii). Hence, the sets 𝕋im,𝐚{[12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}_{i}\setminus\{[12^{(\infty)}]\}, iAi\in A are non-empty and connected. They are also relatively closed in 𝕋m,𝐚{[12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[12^{(\infty)}]\} and pairwise disjoint by Lemma 6.2(i). Then

𝕋m,𝐚{[12()]}=jA𝕋jm,𝐚{[12()]}.\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[12^{(\infty)}]\}=\bigcup_{j\in A}\mathbb{T}^{m,\operatorname{\bf{a}}}_{j}\setminus\{[12^{(\infty)}]\}.

and by Lemma 6.3, the sets 𝕋jm,𝐚{[12()]}\mathbb{T}_{j}^{m,\operatorname{\bf{a}}}\setminus\{[12^{(\infty)}]\}, jAj\in A, are the components of 𝕋m,𝐚{[12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[12^{(\infty)}]\}.

For (ii), we apply induction on the length n0n\geq 0 of uAu\in A^{*}. If n=0n=0, then u=ϵu=\epsilon and the statement follows by (i).

Suppose the statement is true for all words of length n1n-1 and let u=iuAnu=iu^{\prime}\in A^{n} for some uAn1u^{\prime}\in A^{n-1}. Assume as we may that i=1i=1; the cases iA{1}i\in A\setminus\{1\} follow similarly. Note that 1u12(){12(),21(),,m1()}1u^{\prime}12^{(\infty)}\not\in\{12^{(\infty)},21^{(\infty)},\dots,m1^{(\infty)}\} so by Lemma 6.2(ii) [u12()][12()][u12^{(\infty)}]\neq[12^{(\infty)}]. Since the first letter of uu is 11, [u12()]𝕋1m,𝐚{[12()]}[u12^{(\infty)}]\in\mathbb{T}^{m,\operatorname{\bf{a}}}_{1}\setminus\{[12^{(\infty)}]\}.

By induction hypothesis, 𝕋m,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u^{\prime}12^{(\infty)}]\} has exactly mm connected components V1,,VmV_{1},\dots,V_{m}. After re-ordering the components, we may assume that 𝕋ukm,𝐚{[u12()]}Vk\mathbb{T}^{m,\operatorname{\bf{a}}}_{u^{\prime}k}\setminus\{[u^{\prime}12^{(\infty)}]\}\subset V_{k} for all kAk\in A. It follows that

ϕ1(𝕋m,𝐚{[u12()]})=𝕋1m,𝐚{[u12()]}\phi_{1}(\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u^{\prime}12^{(\infty)}]\})=\mathbb{T}^{m,\operatorname{\bf{a}}}_{1}\setminus\{[u12^{(\infty)}]\}

has exactly mm connected components Uk=ϕ1(Vk)𝕋1m,𝐚U_{k}=\phi_{1}(V_{k})\subset\mathbb{T}^{m,\operatorname{\bf{a}}}_{1}, kAk\in A, with

𝕋ukm,𝐚{[u12()]}=ϕ1(𝕋ukm,𝐚{[u12()]})ϕ1(Vk)=Uk.\mathbb{T}^{m,\operatorname{\bf{a}}}_{uk}\setminus\{[u12^{(\infty)}]\}=\phi_{1}(\mathbb{T}^{m,\operatorname{\bf{a}}}_{u^{\prime}k}\setminus\{[u^{\prime}12^{(\infty)}]\})\subset\phi_{1}(V_{k})=U_{k}.

Fix kAk\in A. Since VkV_{k} is a component of 𝕋m,𝐚{ϕu([12()])}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{\phi_{u^{\prime}}([12^{(\infty)}])\}, it is relatively closed in 𝕋m,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u^{\prime}12^{(\infty)}]\}, which implies Vk=Vk¯(𝕋m,𝐚{[u12()]})V_{k}=\overline{V_{k}}\cap(\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u^{\prime}12^{(\infty)}]\}). As a result,

Uk\displaystyle U_{k} =ϕ1(Vk)=ϕ1(Vk)¯(𝕋1m,𝐚{[u12()]})=Uk¯(𝕋1m,𝐚{[u12()]}).\displaystyle=\phi_{1}(V_{k})=\overline{\phi_{1}(V_{k})}\cap(\mathbb{T}_{1}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\})=\overline{U_{k}}\cap(\mathbb{T}_{1}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\}).

Since 𝕋1m,𝐚𝕋m,𝐚\mathbb{T}_{1}^{m,\operatorname{\bf{a}}}\subset\mathbb{T}^{m,\operatorname{\bf{a}}} is compact, Uk𝕋1m,𝐚U_{k}\subset\mathbb{T}_{1}^{m,\operatorname{\bf{a}}} implies Uk¯𝕋1m,𝐚\overline{U_{k}}\subset\mathbb{T}_{1}^{m,\operatorname{\bf{a}}}. As a result, all limit points of UkU_{k} other than [u12()][u12^{(\infty)}] lie in UkU_{k}, implying that UkU_{k} is relatively closed in 𝕋m,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\}.

Exactly one of the components of 𝕋1m,𝐚{[u12()]}\mathbb{T}_{1}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\} needs to contain [12()]𝕋1m,𝐚{[u12()]}[12^{(\infty)}]\in\mathbb{T}_{1}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\}. Suppose [12()]U1[12^{(\infty)}]\in U_{1} and the proof is analogous otherwise. Then U1U1j=2m𝕋jm,𝐚U^{\prime}_{1}\coloneqq U_{1}\cup\bigcup_{j=2}^{m}\mathbb{T}_{j}^{m,\operatorname{\bf{a}}} is a relatively closed subset of 𝕋m,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\}. This set is also connected, because U1U_{1}, 𝕋2m,𝐚,,𝕋mm,𝐚\mathbb{T}_{2}^{m,\operatorname{\bf{a}}},\dots,\mathbb{T}_{m}^{m,\operatorname{\bf{a}}} are connected and intersect at [12()][12^{(\infty)}]. Hence the connected sets U1U^{\prime}_{1}, U2,,UmU_{2},\dots,U_{m} are pairwise disjoint, relatively closed in 𝕋m,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\}, and

𝕋m,𝐚{[u12()]}=(𝕋1m,𝐚{[u12()]})(j=2m𝕋jm,𝐚)=U1(j=2mUj).\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\}=(\mathbb{T}_{1}^{m,\operatorname{\bf{a}}}\setminus\ \{[u12^{(\infty)}]\})\cup\left(\bigcup_{j=2}^{m}\mathbb{T}_{j}^{m,\operatorname{\bf{a}}}\right)=U^{\prime}_{1}\cup\left(\bigcup_{j=2}^{m}U_{j}\right).

This implies by Lemma 6.3 that 𝕋m,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\} has exactly mm connected components that are U1U^{\prime}_{1}, U2,,UmU_{2},\dots,U_{m}. Moreover, the sets 𝕋ujm,𝐚{[u12()]}\mathbb{T}_{uj}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\}, jAj\in A, lie in the different components U1,U2,,UmU_{1}^{\prime},U_{2},\dots,U_{m} of 𝕋m,𝐚{[u12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[u12^{(\infty)}]\}, respectively. This completes the proof. ∎

We need the following property of relative boundaries in 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}, in order to show that all branch points of 𝕋\mathbb{T} have a specific form.

Lemma 6.5.

For all nn\in\mathbb{N} and uAnu\in A^{n}, 𝕋um,𝐚{[u1()],[u2()]}\partial\mathbb{T}^{m,\operatorname{\bf{a}}}_{u}\subset\{[u1^{(\infty)}],[u2^{(\infty)}]\}. Moreover, 𝕋um,𝐚\partial\mathbb{T}^{m,\operatorname{\bf{a}}}_{u}\neq\emptyset and if p𝕋um,𝐚p\in\partial\mathbb{T}^{m,\operatorname{\bf{a}}}_{u}, there exists wk=0n1Akw\in\bigcup_{k=0}^{n-1}A^{k} such that p=[w12()]p=[w12^{(\infty)}].

Proof.

The proof is by induction on nn. For n=1n=1, let uAu\in A. Then 𝕋um,𝐚{[12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}_{u}\setminus\{[12^{(\infty)}]\} is a component of 𝕋m,𝐚{[12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[12^{(\infty)}]\} by Lemma 6.4(i). Since 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is a metric tree, this implies that 𝕋um,𝐚{[12()]}\mathbb{T}^{m,\operatorname{\bf{a}}}_{u}\setminus\{[12^{(\infty)}]\} is a relatively open set, its points lie in the relative interior of 𝕋um,𝐚\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}, do not lie in 𝕋um,𝐚\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}, and 𝕋um,𝐚={[12()]}\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}=\{[12^{(\infty)}]\} (see, for instance, [BT21, Lemma 3.2]). Since A0={ϵ}A^{0}=\{\epsilon\} and

[12()]=ϕϵ([12()])=ϕ1([2()])=ϕ2([1()])==ϕm([1()]),[12^{(\infty)}]=\phi_{\epsilon}([12^{(\infty)}])=\phi_{1}([2^{(\infty)}])=\phi_{2}([1^{(\infty)}])=\dots=\phi_{m}([1^{(\infty)}]),

the lemma is true for n=1n=1.

Assume the statement of the lemma to be true for all vAnv\in A^{n} for fixed nn\in\mathbb{N}. Let uAn+1u\in A^{n+1} with u=vju=vj for vAnv\in A^{n} and jAj\in A. Without loss of generality, we assume that j=1j=1. Since 𝕋1m,𝐚{[12()]}\mathbb{T}_{1}^{m,\operatorname{\bf{a}}}\setminus\{[12^{(\infty)}]\} is open in 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}, the set

ϕv(𝕋1m,𝐚{[12()]})=𝕋um,𝐚{[v12()]}\phi_{v}(\mathbb{T}_{1}^{m,\operatorname{\bf{a}}}\setminus\{[12^{(\infty)}]\})=\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}\setminus\{[v12^{(\infty)}]\}

is a relatively open subset of 𝕋vm,𝐚\mathbb{T}_{v}^{m,\operatorname{\bf{a}}}. Suppose p𝕋um,𝐚p\in\mathbb{T}_{u}^{m,\operatorname{\bf{a}}} is not an interior point of 𝕋um,𝐚\mathbb{T}_{u}^{m,\operatorname{\bf{a}}} in 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}. Then either p=[v12()]p=[v12^{(\infty)}] or pp lies on the boundary of 𝕋vm,𝐚\mathbb{T}_{v}^{m,\operatorname{\bf{a}}}. By the inductive hypothesis for nn,

(6.1) 𝕋um,𝐚{[v12()]}𝕋vm,𝐚{[v12()],[v1()],[v2()]}.\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}\subset\{[v12^{(\infty)}]\}\cup\partial\mathbb{T}_{v}^{m,\operatorname{\bf{a}}}\subset\{[v12^{(\infty)}],[v1^{(\infty)}],[v2^{(\infty)}]\}.

By the inductive hypothesis, every element of 𝕋vm,𝐚\partial\mathbb{T}_{v}^{m,\operatorname{\bf{a}}} can be written in the form [w12()][w12^{(\infty)}] with wk=1n1Akw\in\bigcup_{k=1}^{n-1}A^{k}. Thus, every element of 𝕋um,𝐚\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}} can be written in the form [w12()][w12^{(\infty)}] with wk=1nAkw\in\bigcup_{k=1}^{n}A^{k}.

Since 𝕋um,𝐚\mathbb{T}_{u}^{m,\operatorname{\bf{a}}} is compact, it is closed in 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}, implying 𝕋um,𝐚𝕋um,𝐚\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}\subset\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}. As a result, by (6.1) it is enough to show that 𝕋um,𝐚\mathbb{T}_{u}^{m,\operatorname{\bf{a}}} contains at most two of [v12()][v12^{(\infty)}], [v1()][v1^{(\infty)}], [v2()][v2^{(\infty)}]. Since [2()]𝕋1m,𝐚[2^{(\infty)}]\not\in\mathbb{T}_{1}^{m,\operatorname{\bf{a}}} by Lemma 6.2(iii), we have [v2()]ϕv(𝕋1m,𝐚)=𝕋um,𝐚[v2^{(\infty)}]\not\in\phi_{v}(\mathbb{T}_{1}^{m,\operatorname{\bf{a}}})=\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}. It follows that 𝕋um,𝐚{[v12()],[v1()]}\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}\subset\{[v12^{(\infty)}],[v1^{(\infty)}]\}.

Assume towards contradiction that [v12()][v12^{(\infty)}], [v1()][v1^{(\infty)}] lie in the interior of 𝕋um,𝐚\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}, i.e., 𝕋um,𝐚=\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}=\emptyset. Then there are relatively open neighborhoods Nv12(),Nv1()𝕋um,𝐚N_{v12^{(\infty)}},N_{v1^{(\infty)}}\subset\mathbb{T}_{u}^{m,\operatorname{\bf{a}}} of [v12()],[v1()][v12^{(\infty)}],[v1^{(\infty)}], respectively. If v=v1vnv=v_{1}\dots v_{n}, this implies that

[v2vn12()]ϕv11(Nv12())𝕋v2vn1m,𝐚[v_{2}\dots v_{n}12^{(\infty)}]\in\phi_{v_{1}}^{-1}(N_{v12^{(\infty)}})\subset\mathbb{T}_{v_{2}\dots v_{n}1}^{m,\operatorname{\bf{a}}}

and [v2vn1()]ϕv11(Nv1())𝕋v2vn1m,𝐚[v_{2}\dots v_{n}1^{(\infty)}]\in\phi_{v_{1}}^{-1}(N_{v1^{(\infty)}})\subset\mathbb{T}_{v_{2}\dots v_{n}1}^{m,\operatorname{\bf{a}}}. Since ϕv1\phi_{v_{1}} is continuous, the two points [v2vn12()],[v2vn1()][v_{2}\dots v_{n}12^{(\infty)}],[v_{2}\dots v_{n}1^{(\infty)}] lie in the interior of 𝕋v2vn1m,𝐚\mathbb{T}_{v_{2}\dots v_{n}1}^{m,\operatorname{\bf{a}}}, which contradicts the inductive hypothesis. Therefore, 𝕋um,𝐚\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}\neq\emptyset and the inductive step is complete. ∎

The next lemma gives us all the branch points of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

Lemma 6.6.

Every branch point of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is of the form [u12()][u12^{(\infty)}] for some uAu\in A^{*}.

Proof.

Assume towards a contradiction that [w][w] is a branch point of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} with [w]{[u12()]:uA}[w]\not\in\{[u12^{(\infty)}]:u\in A^{*}\}. Let U1U_{1}, U2U_{2}, U3U_{3} be three distinct components of 𝕋m,𝐚{[w]}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[w]\} and fix xiUix_{i}\in U_{i} for all i{1,2,3}i\in\{1,2,3\}. Since the diameter diam𝕋w(n)m,𝐚\operatorname{diam}{\mathbb{T}_{w(n)}^{m,\operatorname{\bf{a}}}} decreases to 0 as nn increases, we can find some nn\in\mathbb{N} such that

min{dA,𝐚(xi,[w])|:i=1,2,3}>diam𝕋w(n)m,𝐚.\min\{d_{A,\operatorname{\bf{a}}}(x_{i},[w])|:\,i=1,2,3\}>\operatorname{diam}\mathbb{T}_{w(n)}^{m,\operatorname{\bf{a}}}.

Hence, xi𝕋w(n)m,𝐚x_{i}\notin\mathbb{T}_{w(n)}^{m,\operatorname{\bf{a}}} for all i{1,2,3}i\in\{1,2,3\}. Since [w]{[u12()]:uA}[w]\not\in\{[u12^{(\infty)}]:u\in A^{*}\} Lemma 6.5 implies that [w][w] lies in the relative interior of 𝕋w(n)m,𝐚\mathbb{T}_{w(n)}^{m,\operatorname{\bf{a}}} in 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}. Let γi\gamma_{i} be the unique arc in 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} joining xix_{i} with [w][w] and let yi𝕋w(n)m,𝐚y_{i}\in\partial\mathbb{T}_{w(n)}^{m,\operatorname{\bf{a}}} be the first boundary point of 𝕋w(n)m,𝐚\mathbb{T}_{w(n)}^{m,\operatorname{\bf{a}}} on γi\gamma_{i} going from xix_{i} towards [w][w]. Since [w][w] lies in the interior of 𝕋w(n)m,𝐚\mathbb{T}_{w(n)}^{m,\operatorname{\bf{a}}}, we have yi[w]y_{i}\neq[w] for all ii. Let γi\gamma_{i}^{\prime} be the subarc of γi\gamma_{i} with endpoints xix_{i}, yiy_{i}. Since γi(𝕋m,𝐚{[w]})\gamma_{i}^{\prime}\cap(\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{[w]\}) is connected and since xiγix_{i}\in\gamma_{i}^{\prime}, we have that γiUi\gamma_{i}^{\prime}\subset U_{i}, and, hence, yiUiy_{i}\in U_{i} for all i{1,2,3}i\in\{1,2,3\}. However, the points y1,y2,y3𝕋w(n)m,𝐚y_{1},y_{2},y_{3}\in\partial\mathbb{T}_{w(n)}^{m,\operatorname{\bf{a}}} cannot be distinct, since by Lemma 6.5 the set 𝕋w(n)m,𝐚\partial\mathbb{T}_{w(n)}^{m,\operatorname{\bf{a}}} consists of at most two points. This contradiction completes the proof. ∎

In the next lemma we calculate the height of the branch points of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

Lemma 6.7.

If wAw\in A^{*} and B1,,BmB_{1},\dots,B_{m} are the branches of the branch point p=[w12()]p=[w12^{(\infty)}] of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}, then with suitable labeling we have

𝕋w1m,𝐚B1,𝕋w2m,𝐚B2,𝕋wjm,𝐚=Bj,\mathbb{T}_{w1}^{m,\operatorname{\bf{a}}}\subset B_{1},\ \mathbb{T}_{w2}^{m,\operatorname{\bf{a}}}\subset B_{2},\ \mathbb{T}_{wj}^{m,\operatorname{\bf{a}}}=B_{j},

for all j{3,,m}j\in\{3,\dots,m\}. In particular, H𝕋m,𝐚(p)=𝐚(3)Δ(w)H_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p)=\operatorname{\bf{a}}(3)\Delta(w).

Proof.

Let wAw\in A^{*}, then p=[w12()]p=[w12^{(\infty)}] is a branch point of 𝕋wm,𝐚\mathbb{T}_{w}^{m,\operatorname{\bf{a}}} with mm distinct branches 𝕋w1m,𝐚,,𝕋wmm,𝐚\mathbb{T}_{w1}^{m,\operatorname{\bf{a}}},\dots,\mathbb{T}_{wm}^{m,\operatorname{\bf{a}}}. Hence there exist unique distinct branches B1,,BmB_{1},\dots,B_{m} of pp in 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} such that 𝕋wkm,𝐚Bk\mathbb{T}_{wk}^{m,\operatorname{\bf{a}}}\subset B_{k} for k{1,,m}k\in\{1,\dots,m\} by [BM22, Lemma 3.2(i)]. By Lemma 6.5 we have 𝕋wm,𝐚{[w1()],[w2()]}\partial\mathbb{T}_{w}^{m,\operatorname{\bf{a}}}\subset\{[w1^{(\infty)}],[w2^{(\infty)}]\}. Moreover, [w1()],[w2()]𝕋w3m,𝐚[w1^{(\infty)}],[w2^{(\infty)}]\notin\mathbb{T}_{w3}^{m,\operatorname{\bf{a}}}, since [1()],[2()]𝕋jm,𝐚[1^{(\infty)}],[2^{(\infty)}]\notin\mathbb{T}_{j}^{m,\operatorname{\bf{a}}} for all j{3,,m}j\in\{3,\dots,m\}. Thus 𝕋w3𝕋wm,𝐚=\mathbb{T}_{w3}\cap\partial\mathbb{T}_{w}^{m,\operatorname{\bf{a}}}=\emptyset. Hence it follows by [BM22, Lemma 3.2 (iii)] that 𝕋wjm,𝐚\mathbb{T}_{wj}^{m,\operatorname{\bf{a}}} is actually a branch of pp in 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} for all j{3,,m}j\in\{3,\dots,m\}, and so 𝕋wjm,𝐚=Bj\mathbb{T}_{wj}^{m,\operatorname{\bf{a}}}=B_{j}.

For the second part of the statement, we have diamBkdiam𝕋wkm,𝐚=21Δ(w)\operatorname{diam}B_{k}\geq\operatorname{diam}\mathbb{T}_{wk}^{m,\operatorname{\bf{a}}}=2^{-1}\Delta(w) for k=1,2k=1,2. But for j{3,,m}j\in\{3,\dots,m\}, diamBj=diam𝕋wjm,𝐚=𝐚(j)Δ(w)\operatorname{diam}B_{j}=\operatorname{diam}\mathbb{T}_{wj}^{m,\operatorname{\bf{a}}}=\operatorname{\bf{a}}(j)\Delta(w). Hence, H𝕋m,𝐚(p)=diamB3=𝐚(3)Δ(w)H_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p)=\operatorname{diam}B_{3}=\operatorname{\bf{a}}(3)\Delta(w). ∎

The next lemma is a useful characterization of the words that belong in the same class of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

Lemma 6.8.

Let v,wAv,w\in A^{\mathbb{N}} with vwv\neq w. Then [v]=[w][v]=[w] if and only if there exists uAu\in A^{*} such that v,w{u12(),uj1()}v,w\in\{u12^{(\infty)},uj1^{(\infty)}\}, for j2j\geq 2. In this case, [v]=[w]=[u12()][v]=[w]=[u12^{(\infty)}].

Proof.

Suppose that [v]=[w][v]=[w] for some v,wAv,w\in A^{\mathbb{N}} with vwv\neq w. Let uAu\in A^{*} be the longest common initial word of vv and ww i.e. v=uvv=uv^{\prime} and w=uww=uw^{\prime} with v,wAv^{\prime},w^{\prime}\in A^{\mathbb{N}} and v(1)w(1)Av^{\prime}(1)\neq w^{\prime}(1)\in A. We have

0=ρA,𝐚(v,w)=ρA,𝐚(uv,uw)=Δ(u)ρA,𝐚(v,w),0=\rho_{A,\operatorname{\bf{a}}}(v,w)=\rho_{A,\operatorname{\bf{a}}}(uv^{\prime},uw^{\prime})=\Delta(u)\rho_{A,\operatorname{\bf{a}}}(v^{\prime},w^{\prime}),

which means that ρA,𝐚(v,w)=0\rho_{A,\operatorname{\bf{a}}}(v^{\prime},w^{\prime})=0. Hence, [v]=[w][v^{\prime}]=[w^{\prime}]. But [v]𝕋v(1)m,𝐚[v^{\prime}]\in\mathbb{T}_{v^{\prime}(1)}^{m,\operatorname{\bf{a}}} and [w]𝕋w(1)m,𝐚[w^{\prime}]\in\mathbb{T}_{w^{\prime}(1)}^{m,\operatorname{\bf{a}}}. So, due to v(1)w(1)v^{\prime}(1)\neq w^{\prime}(1) and Lemma 6.2(i), [v]=[12()]=[w][v^{\prime}]=[12^{(\infty)}]=[w^{\prime}]. Therefore, [v]=[u12()]=[w][v]=[u12^{(\infty)}]=[w], which by Lemma 6.6 are also branch points of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}. The reverse implication follows from 6.2(ii) and the above shrinking property of ρA,𝐚\rho_{A,\operatorname{\bf{a}}}. ∎

We now estimate the distance between branch points.

Lemma 6.9.

If v,wAv,w\in A^{*}, vwv\neq w then

dA,𝐚([v12()],[w12()])12min{Δ𝐚(v),Δ𝐚(w)}.d_{A,\operatorname{\bf{a}}}([v12^{(\infty)}],[w12^{(\infty)}])\geq\frac{1}{2}\min\{\Delta_{\operatorname{\bf{a}}}(v),\Delta_{\operatorname{\bf{a}}}(w)\}.
Proof.

We start the proof by making a useful observation. By Lemma 4.9, Corollary 5.7, and Lemma 6.2, if iAi\in A and uAu\in A^{*}, then

(6.2) dA,𝐚([u12()],[ui12()])=Δ𝐚(u)ρA,𝐚(12(),i12())\displaystyle d_{A,\operatorname{\bf{a}}}([u12^{(\infty)}],[ui12^{(\infty)}])=\Delta_{\operatorname{\bf{a}}}(u)\rho_{A,\operatorname{\bf{a}}}(12^{(\infty)},i12^{(\infty)}) =Δ𝐚(u)ρA,𝐚(i1(),i12())\displaystyle=\Delta_{\operatorname{\bf{a}}}(u)\rho_{A,\operatorname{\bf{a}}}(i1^{(\infty)},i12^{(\infty)})
=12Δ𝐚(ui).\displaystyle=\frac{1}{2}\Delta_{\operatorname{\bf{a}}}(ui).

Let v,wAv,w\in A^{*} with vwv\neq w, and let uAu\in A^{*} be the longest common initial word of vv and ww. Then v=uvv=uv^{\prime} and w=uww=uw^{\prime} with v,wAv^{\prime},w^{\prime}\in A^{*}, where either vv^{\prime} or ww^{\prime} is the empty word, or both vv^{\prime} and ww^{\prime} are non-empty, but have different initial letter. Accordingly, we consider two cases.

Case 1: v=ϵv^{\prime}=\epsilon or w=ϵw^{\prime}=\epsilon. We may assume that w=ϵw^{\prime}=\epsilon. Then vϵv^{\prime}\neq\epsilon, and so |v||w||v|\geq|w|. Write v=i1inv^{\prime}=i_{1}\cdots i_{n}. By triangle inequality and (6.2)

dA,𝐚([v12()],[w12()])\displaystyle d_{A,\operatorname{\bf{a}}}([v12^{(\infty)}],[w12^{(\infty)}])
=Δ𝐚(u)dA,𝐚([v12()],[12()])\displaystyle=\Delta_{\operatorname{\bf{a}}}(u)d_{A,\operatorname{\bf{a}}}([v^{\prime}12^{(\infty)}],[12^{(\infty)}])
Δ𝐚(u)(dA,𝐚([v(1)12()]),[12()])j=1n1dA,𝐚([v(j)12()],[v(j+1)12()]))\displaystyle\geq\Delta_{\operatorname{\bf{a}}}(u)\left(d_{A,\operatorname{\bf{a}}}([v^{\prime}(1)12^{(\infty)}]),[12^{(\infty)}])-\sum_{j=1}^{n-1}d_{A,\operatorname{\bf{a}}}([v^{\prime}(j)12^{(\infty)}],[v^{\prime}(j+1)12^{(\infty)}])\right)
=12Δ𝐚(u)Δ𝐚(v(1))Δ𝐚(u)12j=1n1Δ𝐚(v(j+1))\displaystyle=\frac{1}{2}\Delta_{\operatorname{\bf{a}}}(u)\Delta_{\operatorname{\bf{a}}}(v^{\prime}(1))-\Delta_{\operatorname{\bf{a}}}(u)\frac{1}{2}\sum_{j=1}^{n-1}\Delta_{\operatorname{\bf{a}}}(v^{\prime}(j+1))
12𝐚(i1)Δ𝐚(u)Δ𝐚(u)12𝐚(i1)j=1n12j\displaystyle\geq\frac{1}{2}\operatorname{\bf{a}}(i_{1})\Delta_{\operatorname{\bf{a}}}(u)-\Delta_{\operatorname{\bf{a}}}(u)\frac{1}{2}\operatorname{\bf{a}}(i_{1})\sum_{j=1}^{n-1}2^{-j}
12𝐚(i1)Δ𝐚(u)21n\displaystyle\geq\frac{1}{2}\operatorname{\bf{a}}(i_{1})\Delta_{\operatorname{\bf{a}}}(u)2^{1-n}
12Δ𝐚(u)Δ𝐚(v)\displaystyle\geq\frac{1}{2}\Delta_{\operatorname{\bf{a}}}(u)\Delta_{\operatorname{\bf{a}}}(v^{\prime})
=12Δ𝐚(v).\displaystyle=\frac{1}{2}\Delta_{\operatorname{\bf{a}}}(v).

Case 2: v,wϵv^{\prime},w^{\prime}\neq\epsilon. Then vv^{\prime} and ww^{\prime} necessarily start with different letters, that is, v=iv′′v^{\prime}=iv^{\prime\prime} and w=jw′′w^{\prime}=jw^{\prime\prime} with i,jAi,j\in A, iji\neq j, and v′′,w′′Av^{\prime\prime},w^{\prime\prime}\in A^{*}. Then [v12()]𝕋im,𝐚[v^{\prime}12^{(\infty)}]\in\mathbb{T}_{i}^{m,\operatorname{\bf{a}}} and [w12()]𝕋jm,𝐚[w^{\prime}12^{(\infty)}]\in\mathbb{T}_{j}^{m,\operatorname{\bf{a}}}. By Proposition 5.1 and Case 1 we have that

dA,𝐚([v12()],\displaystyle d_{A,\operatorname{\bf{a}}}([v12^{(\infty)}], [w12()])\displaystyle[w12^{(\infty)}])
=Δ𝐚(u)dA,𝐚([v12()],[w12()])\displaystyle=\Delta_{\operatorname{\bf{a}}}(u)d_{A,\operatorname{\bf{a}}}([v^{\prime}12^{(\infty)}],[w^{\prime}12^{(\infty)}])
=Δ𝐚(u)(dA,𝐚([v12()],[12()])+dA,𝐚([12()],[w12()]))\displaystyle=\Delta_{\operatorname{\bf{a}}}(u)\left(d_{A,\operatorname{\bf{a}}}([v^{\prime}12^{(\infty)}],[12^{(\infty)}])+d_{A,\operatorname{\bf{a}}}([12^{(\infty)}],[w^{\prime}12^{(\infty)}])\right)
12Δ𝐚(v)+12Δ𝐚(w).\displaystyle\geq\frac{1}{2}\Delta_{\operatorname{\bf{a}}}(v)+\frac{1}{2}\Delta_{\operatorname{\bf{a}}}(w).\qed
Proof of Proposition 6.1.

By Lemma 6.4 and Lemma 6.6, 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is an mm-valent metric tree.

Fix a branch point p𝕋m,𝐚p\in\mathbb{T}^{m,\operatorname{\bf{a}}}. By Lemma 6.6, p=[w12()]p=[w12^{(\infty)}] for some wAw\in A^{*} and 𝕋m,𝐚{p}\mathbb{T}^{m,\operatorname{\bf{a}}}\setminus\{p\} has exactly mm branches. Therefore, 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is an mm-valent metric tree and by Lemma 6.7 for all j,j{3,,m}j,j^{\prime}\in\{3,\dots,m\}

diamB𝕋m,𝐚j(p)diamB𝕋m,𝐚j(p)𝐚(3)𝐚(m)\frac{\operatorname{diam}{B^{j}_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p)}}{\operatorname{diam}{B^{j^{\prime}}_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p)}}\leq\frac{\operatorname{\bf{a}}(3)}{\operatorname{\bf{a}}(m)}

which shows the uniform branch growth of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

Let p,q𝕋m,𝐚p,q\in\mathbb{T}^{m,\operatorname{\bf{a}}} be two distinct branch points of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}. Then p=[v12()]p=[v12^{(\infty)}] and q=[w12()]q=[w12^{(\infty)}] with v,wAv,w\in A^{*}, vwv\neq w. By Lemmas 6.7,6.9,

dA,𝐚(p,q)12min{Δ𝐚(v),Δ𝐚(w)}\displaystyle d_{A,\operatorname{\bf{a}}}(p,q)\geq\frac{1}{2}\min\{\Delta_{\operatorname{\bf{a}}}(v),\Delta_{\operatorname{\bf{a}}}(w)\} 𝐚(3)min{Δ𝐚(v),Δ𝐚(w)}\displaystyle\geq\operatorname{\bf{a}}(3)\min\{\Delta_{\operatorname{\bf{a}}}(v),\Delta_{\operatorname{\bf{a}}}(w)\}
=min{H𝕋m,𝐚(p),H𝕋m,𝐚(q)}.\displaystyle=\min\{H_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p),H_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(q)\}.

This shows that 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} has branch points that are uniformly relatively separated.

To establish uniform density, let [w],[v]𝕋m,𝐚[w],[v]\in\mathbb{T}^{m,\operatorname{\bf{a}}} with [w][v][w]\neq[v]. Let uAu\in A^{*} be the longest common initial word of vv and ww, that is, w=uww=uw^{\prime} and v=uvv=uv^{\prime} with w,vAw^{\prime},v^{\prime}\in A^{\mathbb{N}} and w(1)v(1)w^{\prime}(1)\neq v^{\prime}(1). Hence [w]𝕋uw(1)m,𝐚[w]\in\mathbb{T}_{uw^{\prime}(1)}^{m,\operatorname{\bf{a}}} and [v]𝕋uv(1)m,𝐚[v]\in\mathbb{T}_{uv^{\prime}(1)}^{m,\operatorname{\bf{a}}}. Note that by Remark 4.7 and Lemma 6.2(i) the unique arc in 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} that joins [w][w] with [v][v] goes through the branch point p=[u12()]p=[u12^{(\infty)}] which satisfies

H𝕋m,𝐚(p)=𝐚(3)Δ𝐚(u)\displaystyle H_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p)=\operatorname{\bf{a}}(3)\Delta_{\operatorname{\bf{a}}}(u) 𝐚(3)Δ𝐚(u)dA,𝐚([w],[v])\displaystyle\geq\operatorname{\bf{a}}(3)\Delta_{\operatorname{\bf{a}}}(u)d_{A,\operatorname{\bf{a}}}([w^{\prime}],[v^{\prime}])
=𝐚(3)dA,𝐚([w],[v]).\displaystyle=\operatorname{\bf{a}}(3)d_{A,\operatorname{\bf{a}}}([w],[v]).

This completes the proof of the proposition. ∎

Remark 6.10.

Note that 𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}} has also uniformly separated and uniformly dense branch points (but 𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}} doesn’t have uniform branch growth). If xx is a branch point of 𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}} then it is a branch point of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} for some mm\in\mathbb{N} and H𝕋,𝐚(x)=H𝕋m,𝐚(x)H_{\mathbb{T}^{\infty,\operatorname{\bf{a}}}}(x)=H_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(x) by Lemma 5.10. Observe that the uniform separation and uniform density constant of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is 𝐚(3)\operatorname{\bf{a}}(3) and the uniform branch growth constant is 𝐚(3)/𝐚(m)\operatorname{\bf{a}}(3)/\operatorname{\bf{a}}(m).

For wAnw\in A^{n}, we call the sets of the form 𝕋wm,𝐚=ϕwA,𝐚(𝕋m,𝐚)\mathbb{T}_{w}^{m,\operatorname{\bf{a}}}=\phi_{w}^{A,\operatorname{\bf{a}}}(\mathbb{T}^{m,\operatorname{\bf{a}}}) the nn-tiles of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}. We finish this section with two lemmas that are useful in later sections.

Lemma 6.11.

The tile 𝕋11m,𝐚\mathbb{T}_{11}^{m,\operatorname{\bf{a}}} is the only tile XX of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} with [1()]X[1^{(\infty)}]\in X and [112()]X[112^{(\infty)}]\in\partial X. The tile 𝕋22m,𝐚\mathbb{T}_{22}^{m,\operatorname{\bf{a}}} is the only tile XX of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} with [2()]X[2^{(\infty)}]\in X and [212()]X[212^{(\infty)}]\in\partial X.

Proof.

First, by Lemma 6.5 [1()]𝕋11m,𝐚[1^{(\infty)}]\in\mathbb{T}_{11}^{m,\operatorname{\bf{a}}} and 𝕋11m,𝐚{[1()],[112()]}\partial\mathbb{T}_{11}^{m,\operatorname{\bf{a}}}\subset\{[1^{(\infty)}],[112^{(\infty)}]\}. By Remark 5.9 [1()][1^{(\infty)}] is a leaf hence 𝕋11m,𝐚={[112()]}\partial\mathbb{T}_{11}^{m,\operatorname{\bf{a}}}=\{[112^{(\infty)}]\} (note that by Lemma 6.5 the tile 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is the only tile with empty boundary). For the uniqueness let 𝕋um,𝐚\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}, uAnu\in A^{n} with n0n\in\mathbb{N}_{0}, be a tile such that [1()]𝕋um,𝐚[1^{(\infty)}]\in\mathbb{T}_{u}^{m,\operatorname{\bf{a}}} and [112()]𝕋um,𝐚[112^{(\infty)}]\in\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}. That implies that u=1(n)u=1^{(n)}. Also, by Lemma 6.5 𝕋um,𝐚{[u1()],[u2()]}={[1()],[1(n)2()]}\partial\mathbb{T}_{u}^{m,\operatorname{\bf{a}}}\subset\{[u1^{(\infty)}],[u2^{(\infty)}]\}=\{[1^{(\infty)}],[1^{(n)}2^{(\infty)}]\}. Hence either [112()]=[1()][112^{(\infty)}]=[1^{(\infty)}] or [112()]=[1(n)2()][112^{(\infty)}]=[1^{(n)}2^{(\infty)}]. For the first case, that would imply that

0=ρA,𝐚(112(),1())=𝐚(1)2ρA,𝐚(1(),2())>0,0=\rho_{A,\operatorname{\bf{a}}}(112^{(\infty)},1^{(\infty)})={\operatorname{\bf{a}}(1)}^{2}\rho_{A,\operatorname{\bf{a}}}(1^{(\infty)},2^{(\infty)})>0,

where the last follows from Corollary 5.7. For the second case we have

0=ρA,𝐚(112(),1(n)2())=𝐚(1)2ρA,𝐚(2(),1(n2)2())>00=\rho_{A,\operatorname{\bf{a}}}(112^{(\infty)},1^{(n)}2^{(\infty)})=\operatorname{\bf{a}}(1)^{2}\rho_{A,\operatorname{\bf{a}}}(2^{(\infty)},1^{(n-2)}2^{(\infty)})>0

where the last inequality holds for n>2n>2 since the only words that belong to the same class are of the form u12(),uj1()u12^{(\infty)},uj1^{(\infty)} from Lemma 6.8. Hence, n=2n=2 and u=11u=11. The second statement is similar and we omit the details. ∎

Lemma 6.12.

Let 𝐚\operatorname{\bf{a}} be a weight, let nn\in\mathbb{N}, and let w,uAnw,u\in A^{n} be distinct with 𝕋wm,𝐚𝕋um,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{u}\neq\emptyset. Then

(6.3) 2mini=1,,m𝐚(i)Δ𝐚(w)Δ𝐚(u)(2mini=1,,m𝐚(i))1.2\min_{i=1,\dots,m}\operatorname{\bf{a}}(i)\leq\frac{\Delta_{\operatorname{\bf{a}}}(w)}{\Delta_{\operatorname{\bf{a}}}(u)}\leq\left(2\min_{i=1,\dots,m}\operatorname{\bf{a}}(i)\right)^{-1}.
Proof.

The proof is by induction on the length nn. If n=1n=1, then 𝕋im,𝐚𝕋jm,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{i}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{j}\neq\emptyset for all i,jAi,j\in A and (6.3) is clear. Assume now that for some nn\in\mathbb{N}, (6.3) is true whenever w,uAnw,u\in A^{n} are distinct and 𝕋wm,𝐚𝕋um,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{u}\neq\emptyset. Fix now distinct w,uAn+1w,u\in A^{n+1} such that 𝕋wm,𝐚𝕋um,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{u}\neq\emptyset. If there exists iAi\in A such that w=iww=iw^{\prime} and u=iuu=iu^{\prime}, then (6.3) follows by the inductive hypothesis and the fact that Δ𝐚(w)/Δ𝐚(u)=Δ𝐚(w)/Δ𝐚(u)\Delta_{\operatorname{\bf{a}}}(w)/\Delta_{\operatorname{\bf{a}}}(u)=\Delta_{\operatorname{\bf{a}}}(w^{\prime})/\Delta_{\operatorname{\bf{a}}}(u^{\prime}). Assume now that there exist distinct i,jAi,j\in A such that wAin+1w\in A_{i}^{n+1} and uAjn+1u\in A_{j}^{n+1}. Without loss of generality, we may assume that i=1i=1. By Lemma 6.2(i),

𝕋wm,𝐚𝕋um,𝐚𝕋1m,𝐚𝕋jm,𝐚={[12()]}.\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{u}\subset\mathbb{T}^{m,\operatorname{\bf{a}}}_{1}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{j}=\{[12^{(\infty)}]\}.

Since [12()]=[j1()][12^{(\infty)}]=[j1^{(\infty)}] (by Remark 4.7) and wA1n+1w\in A_{1}^{n+1}, we have that w=12(n)w=12^{(n)} and u=j1(n)u=j1^{(n)} and

Δ𝐚(w)Δ𝐚(u)=2n1𝐚(j)2n=12𝐚(j)\frac{\Delta_{\operatorname{\bf{a}}}(w)}{\Delta_{\operatorname{\bf{a}}}(u)}=\frac{2^{-n-1}}{\operatorname{\bf{a}}(j)2^{-n}}=\frac{1}{2\operatorname{\bf{a}}(j)}

and (6.3) follows. We work similarly if j=1j=1, or if neither of i,ji,j is equal to 1. ∎

7. Uniform 4-branching of the Vicsek fractal

Recall that the Vicsek fractal 𝕍\mathbb{V} is defined as the attractor of the iterated function system ={ϕ1,,ϕ5}\mathcal{F}=\{\phi_{1},\dots,\phi_{5}\} on \mathbb{C} where

ϕ1(z)=13(z22i),ϕ2(z)=13(z2+2i),ϕ3(z)=13(z+2+2i),\phi_{1}(z)=\tfrac{1}{3}(z-2-2i),\quad\phi_{2}(z)=\tfrac{1}{3}(z-2+2i),\quad\phi_{3}(z)=\tfrac{1}{3}(z+2+2i),
ϕ4(z)=13(z+22i),ϕ5(z)=13z.\phi_{4}(z)=\tfrac{1}{3}(z+2-2i),\quad\phi_{5}(z)=\tfrac{1}{3}z.

The purpose of this section is to show the following for 𝕍\mathbb{V}.

Proposition 7.1.

The Vicsek fractal 𝕍\mathbb{V} is a uniformly 44-branching quasiconformal tree.

Fix for the rest of this section A={1,,5}A=\{1,\dots,5\}. Given w=i1ikAkw=i_{1}\cdots i_{k}\in A^{k} define

ϕw=ϕi1ϕik,and𝕍w=ϕw(𝕍).\phi_{w}=\phi_{i_{1}}\circ\cdots\circ\phi_{i_{k}},\quad\text{and}\quad\mathbb{V}_{w}=\phi_{w}(\mathbb{V}).

In [DV22, §6.2], it was shown that 𝕍\mathbb{V} is a quasiconformal tree. Moreover, there exist an equivalence relation \sim on AA^{\mathbb{N}}, and a bi-Lipschitz homeomorphism

f:(A/,d)𝕍,f(Aw/)=𝕍w for all wA.f:(A^{\mathbb{N}}/\sim,d)\to\mathbb{V},\quad f(A^{\mathbb{N}}_{w}/\sim)=\mathbb{V}_{w}\text{ for all $w\in A^{*}$}.

Henceforth, we identify points in A/A^{\mathbb{N}}/\sim with their images in 𝕍\mathbb{V} under ff. Also, the equivalence relation in AA^{\mathbb{N}} implies that

[13()]=[51()],[24()]=[52()],[35()]=[53()],[42()]=[45()].[13^{(\infty)}]=[51^{(\infty)}],\quad[24^{(\infty)}]=[52^{(\infty)}],\quad[35^{(\infty)}]=[53^{(\infty)}],\quad[42^{(\infty)}]=[45^{(\infty)}].

The proof of the following is almost identical with that in [BT21, Proposition 4.2], hence we omit the details.

Lemma 7.2.

For each nn\in\mathbb{N}, let Jn=wAnϕw(X)J_{n}=\bigcup_{w\in A^{n}}\phi_{w}(X) and Kn=wAnϕw([1,1]2)K_{n}=\bigcup_{w\in A^{n}}\phi_{w}([-1,1]^{2}), where X2X\subset\mathbb{R}^{2} is the union of the diagonals of [1,1]2[-1,1]^{2}. Then, for all nn\in\mathbb{N}, JnJ_{n} and KnK_{n} are compact, and

JnJn+1𝕍Kn+1Kn.J_{n}\subset J_{n+1}\subset\mathbb{V}\subset K_{n+1}\subset K_{n}.

Moreover, n0Jn¯=𝕍=n0Kn\overline{\bigcup_{n\in\mathbb{N}_{0}}J_{n}}=\mathbb{V}=\bigcap_{n\in\mathbb{N}_{0}}K_{n}.

The above lemma implies that diam𝕍=2\operatorname{diam}\mathbb{V}=\sqrt{2}. We also need to classify the branch points of 𝕍\mathbb{V}, and the boundary points of the subtrees 𝕍w\mathbb{V}_{w}.

Lemma 7.3.

Let wAw\in A^{*} .

  1. (i)

    The set 𝕍𝕍w5\mathbb{V}\setminus\mathbb{V}_{w5} has exactly 44 components, and

    (𝕍𝕍w5¯)𝕍w5={[w51()],[w52()],[w53()],[w54()]}.(\overline{\mathbb{V}\setminus\mathbb{V}_{w5}})\cap\mathbb{V}_{w5}=\{[w51^{(\infty)}],[w52^{(\infty)}],[w53^{(\infty)}],[w54^{(\infty)}]\}.
  2. (ii)

    If i{1,2,3,4}i\in\{1,2,3,4\}, then 𝕍𝕍wi\mathbb{V}\setminus\mathbb{V}_{wi} has either 11 or 22 components.

    1. (a)

      If 𝕍𝕍wi\mathbb{V}\setminus\mathbb{V}_{wi} has one component, then there exists j{1,2,3,4}j\in\{1,2,3,4\} such that (𝕍𝕍wi¯)𝕍wi={[wij()]}(\overline{\mathbb{V}\setminus\mathbb{V}_{wi}})\cap\mathbb{V}_{wi}=\{[wij^{(\infty)}]\}.

    2. (b)

      If 𝕍𝕍wi\mathbb{V}\setminus\mathbb{V}_{wi} has two components, then (𝕍𝕍wi¯)𝕍wi(\overline{\mathbb{V}\setminus\mathbb{V}_{wi}})\cap\mathbb{V}_{wi} is either the set {[wi1()],[wi3()]}\{[wi1^{(\infty)}],[wi3^{(\infty)}]\}, or the set {[wi2()],[wi4()]}\{[wi2^{(\infty)}],[wi4^{(\infty)}]\}.

  3. (iii)

    If w is not ending with 5, then

    1. (a)

      either the closures of the components of 𝕍𝕍w5\mathbb{V}\setminus\mathbb{V}_{w5} are 𝕍w1,𝕍w3\mathbb{V}_{w1},\mathbb{V}_{w3} and 𝕍w2\mathbb{V}_{w2},𝕍w4\mathbb{V}_{w4} lie in different connected components,

    2. (b)

      or the closure of the components of 𝕍𝕍w5\mathbb{V}\setminus\mathbb{V}_{w5} are 𝕍w2,𝕍w4\mathbb{V}_{w2},\mathbb{V}_{w4} and 𝕍w1,𝕍w3\mathbb{V}_{w1},\mathbb{V}_{w3} lie in different connected components.

Proof.

For (i), note first that

𝕍𝕍5¯=𝕍ϕ5([1,1]2)¯=𝕍i=14ϕi([1,1]2)=𝕍1𝕍4.\overline{\mathbb{V}\setminus\mathbb{V}_{5}}=\overline{\mathbb{V}\setminus\phi_{5}([-1,1]^{2})}=\mathbb{V}\cap\bigcup_{i=1}^{4}\phi_{i}([-1,1]^{2})=\mathbb{V}_{1}\cup\cdots\cup\mathbb{V}_{4}.

Moreover, if i{1,2,3,4}i\in\{1,2,3,4\}, (say i=1i=1), then

{[13()]}={[51()]}𝕍1𝕍5ϕ1([1,1]2)ϕ5([1,1]2),\{[13^{(\infty)}]\}=\{[51^{(\infty)}]\}\subset\mathbb{V}_{1}\cap\mathbb{V}_{5}\subset\phi_{1}([-1,1]^{2})\cap\phi_{5}([-1,1]^{2}),

where the latter set has exactly one point. Therefore,

(𝕍𝕍5¯)𝕍5={[51()],[52()],[53()],[54()]}.(\overline{\mathbb{V}\setminus\mathbb{V}_{5}})\cap\mathbb{V}_{5}=\{[51^{(\infty)}],[52^{(\infty)}],[53^{(\infty)}],[54^{(\infty)}]\}.

Fix now wAw\in A^{*}. On the one hand,

𝕍w𝕍w5=ϕw(𝕍𝕍5),\mathbb{V}_{w}\setminus\mathbb{V}_{w5}=\phi_{w}(\mathbb{V}\setminus\mathbb{V}_{5}),

which has 4 components. On the other hand, 𝕍\mathbb{V} is a tree and 𝕍w\mathbb{V}_{w} is a subtree, so 𝕍𝕍w5\mathbb{V}\setminus\mathbb{V}_{w5} has as many components as those of 𝕍w𝕍w5\mathbb{V}_{w}\setminus\mathbb{V}_{w5}. Moreover,

(𝕍𝕍w5¯)𝕍w5=(𝕍w𝕍w5¯)𝕍w5={[w51()],[w52()],[w53()],[w54()]}.(\overline{\mathbb{V}\setminus\mathbb{V}_{w5}})\cap\mathbb{V}_{w5}=(\overline{\mathbb{V}_{w}\setminus\mathbb{V}_{w5}})\cap\mathbb{V}_{w5}=\{[w51^{(\infty)}],[w52^{(\infty)}],[w53^{(\infty)}],[w54^{(\infty)}]\}.

The proof of (ii) is by induction on |w||w|. If w=ϵw=\epsilon, and if i{1,2,3,4}i\in\{1,2,3,4\}, then

𝕍𝕍i¯=𝕍5j{1,2,3,4}{i}𝕍j,\overline{\mathbb{V}\setminus\mathbb{V}_{i}}=\mathbb{V}_{5}\cup\bigcup_{j\in\{1,2,3,4\}\setminus\{i\}}\mathbb{V}_{j},

which is a connected set. For the second part of (ii), assume without loss of generality that i=1i=1, and the proof is similar in other cases. Then,

{[13()]}={[51()]}(𝕍𝕍1¯)𝕍1ϕ1([1,1]2)ϕ5([1,1]2).\{[13^{(\infty)}]\}=\{[51^{(\infty)}]\}\subset(\overline{\mathbb{V}\setminus\mathbb{V}_{1}})\cap\mathbb{V}_{1}\subset\phi_{1}([-1,1]^{2})\cap\phi_{5}([-1,1]^{2}).

with the latter set having only one element. Therefore, (𝕍𝕍1¯)𝕍1={13()}(\overline{\mathbb{V}\setminus\mathbb{V}_{1}})\cap\mathbb{V}_{1}=\{13^{(\infty)}\}.

Suppose now that (ii) is true for some wAw\in A^{*} and let jAj\in A. By the inductive hypothesis and by (i), there are three possible cases.

Case 1: Suppose that j=5j=5. Then, by (i),

(𝕍𝕍w5¯)𝕍w5={[w51()],[w52()],[w53()],[w54()]}.(\overline{\mathbb{V}\setminus\mathbb{V}_{w5}})\cap\mathbb{V}_{w5}=\{[w51^{(\infty)}],[w52^{(\infty)}],[w53^{(\infty)}],[w54^{(\infty)}]\}.

Assume that i=1i=1, and the case i{2,3,4}i\in\{2,3,4\} is similar. Note that

𝕍w51(𝕍w5𝕍w51¯)={[w513()]},and𝕍w51(𝕍𝕍w5¯)={[w51()]}.\mathbb{V}_{w51}\cap(\overline{\mathbb{V}_{w5}\setminus\mathbb{V}_{w51}})=\{[w513^{(\infty)}]\},\quad\text{and}\quad\mathbb{V}_{w51}\cap(\overline{\mathbb{V}\setminus\mathbb{V}_{w5}})=\{[w51^{(\infty)}]\}.

Therefore, 𝕍𝕍w51\mathbb{V}\setminus\mathbb{V}_{w51} has 22 components.

Case 2: Suppose that j{1,2,3,4}j\in\{1,2,3,4\} and 𝕍𝕍wj\mathbb{V}\setminus\mathbb{V}_{wj} has 22 components. By the inductive hypothesis, either

(7.1) 𝕍𝕍wj¯𝕍wj={[wj1()],[wj3()]},\overline{\mathbb{V}\setminus\mathbb{V}_{wj}}\cap\mathbb{V}_{wj}=\{[wj1^{(\infty)}],[wj3^{(\infty)}]\},

or

(7.2) 𝕍𝕍wj¯𝕍wj={[wj2()],[wj4()]}.\overline{\mathbb{V}\setminus\mathbb{V}_{wj}}\cap\mathbb{V}_{wj}=\{[wj2^{(\infty)}],[wj4^{(\infty)}]\}.

Assume that i=1i=1 (the case i{2,3,4}i\in\{2,3,4\} is similar). In the case of (7.1), we have

𝕍wj1(𝕍𝕍wj¯)={[wj1()]},\mathbb{V}_{wj1}\cap(\overline{\mathbb{V}\setminus\mathbb{V}_{wj}})=\{[wj1^{(\infty)}]\},

and 𝕍𝕍wj1\mathbb{V}\setminus\mathbb{V}_{wj1} has 22 components. In the case of (7.2), we have 𝕍wj1(𝕍𝕍wj¯)=\mathbb{V}_{wj1}\cap(\overline{\mathbb{V}\setminus\mathbb{V}_{wj}})=\emptyset, which implies

𝕍wj1(𝕍wj𝕍wj1¯)={[wj13()]},\mathbb{V}_{wj1}\cap(\overline{\mathbb{V}_{wj}\setminus\mathbb{V}_{wj1}})=\{[wj13^{(\infty)}]\},

and 𝕍𝕍wj1\mathbb{V}\setminus\mathbb{V}_{wj1} has 11 component.

Case 3: Suppose that j{1,2,3,4}j\in\{1,2,3,4\} and 𝕍𝕍wj\mathbb{V}\setminus\mathbb{V}_{wj} has 11 component. By the inductive hypothesis, there exists k{1,2,3,4}k\in\{1,2,3,4\} such that 𝕍𝕍wj¯𝕍wj={[wk1()]}\overline{\mathbb{V}\setminus\mathbb{V}_{wj}}\cap\mathbb{V}_{wj}=\{[wk1^{(\infty)}]\}.

Assume that i=1i=1 (the case i{2,3,4}i\in\{2,3,4\} is similar). If k=1k=1, then

𝕍wj1(𝕍𝕍wj¯)={[wj1()]},\mathbb{V}_{wj1}\cap(\overline{\mathbb{V}\setminus\mathbb{V}_{wj}})=\{[wj1^{(\infty)}]\},

while if k=2,3,4k=2,3,4, then

𝕍w1(𝕍𝕍w¯)=,\mathbb{V}_{w^{\prime}1}\cap(\overline{\mathbb{V}\setminus\mathbb{V}_{w^{\prime}}})=\emptyset,

and as before we have

𝕍wj1(𝕍wj𝕍wj1¯)={[wj13()]}.\mathbb{V}_{wj1}\cap(\overline{\mathbb{V}_{wj}\setminus\mathbb{V}_{wj1}})=\{[wj13^{(\infty)}]\}.

Hence, 𝕍𝕍wj1\mathbb{V}\setminus\mathbb{V}_{wj1} has 22 components when k=1k=1, and 11 component when k=2,3,4k=2,3,4.

We now show (iii). Suppose that w=uiw=ui, with uAu\in A^{*} and i{1,2,3,4}i\in\{1,2,3,4\}. By (ii), 𝕍𝕍w\mathbb{V}\setminus\mathbb{V}_{w} has either 11 or 22 components, and there are two cases to consider.

Case 1: Suppose that 𝕍𝕍w\mathbb{V}\setminus\mathbb{V}_{w} has 1 component. Let U=𝕍𝕍w¯U=\overline{\mathbb{V}\setminus\mathbb{V}_{w}}, and let B1,,B4B_{1},\dots,B_{4} be the closures of the components of 𝕍𝕍w5\mathbb{V}\setminus\mathbb{V}_{w5}. On the one hand, since 𝕍w𝕍w5¯=i=14𝕍wi\overline{\mathbb{V}_{w}\setminus\mathbb{V}_{w5}}=\bigcup_{i=1}^{4}\mathbb{V}_{wi}, we may assume that 𝕍wiBi\mathbb{V}_{wi}\subset B_{i} for all ii, potentially by relabeling if necessary.

On the other hand, there exists i{1,2,3,4}i\in\{1,2,3,4\} such that 𝕍𝕍w¯𝕍w={wi()}\overline{\mathbb{V}\setminus\mathbb{V}_{w}}\cap\mathbb{V}_{w}=\{wi^{(\infty)}\}. Since 𝕍𝕍w¯𝕍𝕍w5¯\overline{\mathbb{V}\setminus\mathbb{V}_{w}}\subset\overline{\mathbb{V}\setminus\mathbb{V}_{w5}}, we have that U𝕍wiBiU\cup\mathbb{V}_{wi}\subset B_{i}. Moreover, note that

𝕍w𝕍w5¯=(𝕍𝕍w5)(𝕍𝕍w)¯,\overline{\mathbb{V}_{w}\setminus\mathbb{V}_{w5}}=\overline{(\mathbb{V}\setminus\mathbb{V}_{w5})\setminus(\mathbb{V}\setminus\mathbb{V}_{w})},

which implies that

𝕍w1𝕍w2𝕍w3𝕍w4=BiU¯l{1,2,3,4}{i}Bl.\mathbb{V}_{w1}\cup\mathbb{V}_{w2}\cup\mathbb{V}_{w3}\cup\mathbb{V}_{w4}=\overline{B_{i}\setminus U}\cup\bigcup_{l\in\{1,2,3,4\}\setminus\{i\}}B_{l}.

The fact that 𝕍wiBi\mathbb{V}_{wi}\subset B_{i} and that 𝕍wl\mathbb{V}_{wl}, BlB_{l} are connected components, imply that, after relabeling, Bi=𝕍wiUB_{i}=\mathbb{V}_{wi}\cup U and 𝕍wl=Bl\mathbb{V}_{wl}=B_{l}, for l{1,2,3,4}{i}l\in\{1,2,3,4\}\setminus\{i\}.

Case 2: Suppose that 𝕍𝕍w\mathbb{V}\setminus\mathbb{V}_{w} has 22 components. Denote these components by U1U_{1}, U2U_{2}. Denote again by BiB_{i}, i{1,2,3,4}i\in\{1,2,3,4\}, the closures of the components of 𝕍𝕍w5\mathbb{V}\setminus\mathbb{V}_{w5}, and assume without loss of generality that 𝕍𝕍w¯𝕍w={[w1()],[w3()]}\overline{\mathbb{V}\setminus\mathbb{V}_{w}}\cap\mathbb{V}_{w}=\{[w1^{(\infty)}],[w3^{(\infty)}]\}. We may further assume that [w1()]U1[w1^{(\infty)}]\in U_{1} and that [w3()]U2[w3^{(\infty)}]\in U_{2}. Working as in Case 1, it follows that U1𝕍w1B1U_{1}\cup\mathbb{V}_{w1}\subset B_{1} and U2𝕍w3B2U_{2}\cup\mathbb{V}_{w3}\subset B_{2}. Furthermore, note again that

𝕍w𝕍w5¯=(𝕍𝕍w5)(𝕍𝕍w)¯,\overline{\mathbb{V}_{w}\setminus\mathbb{V}_{w5}}=\overline{(\mathbb{V}\setminus\mathbb{V}_{w5})\setminus(\mathbb{V}\setminus\mathbb{V}_{w})},

which implies that

𝕍w1𝕍w2𝕍w3𝕍w4=B1U1¯B2U2¯B3B4.\mathbb{V}_{w1}\cup\mathbb{V}_{w2}\cup\mathbb{V}_{w3}\cup\mathbb{V}_{w4}=\overline{B_{1}\setminus U_{1}}\cup\overline{B_{2}\setminus U_{2}}\cup B_{3}\cup B_{4}.

Since 𝕍w1B1\mathbb{V}_{w1}\subset B_{1}, 𝕍w3B2\mathbb{V}_{w3}\subset B_{2}, and all 𝕍wi\mathbb{V}_{wi}, BiB_{i} are connected components, after relabeling, we may assume that B1=𝕍w1U1B_{1}=\mathbb{V}_{w1}\cup U_{1}, B2=𝕍w3U2B_{2}=\mathbb{V}_{w3}\cup U_{2} and B3=𝕍w2B_{3}=\mathbb{V}_{w2}, B4=𝕍w4B_{4}=\mathbb{V}_{w4}. ∎

The next lemma provides us with a complete classification of the branch points of 𝕍\mathbb{V} and their heights.

Lemma 7.4.

A point [w][w] is a branch point of 𝕍\mathbb{V} if, and only if, w=u5()w=u5^{(\infty)} for some uAu\in A^{*}. Moreover, every branch point [u5()][u5^{(\infty)}] has valence 44, and

H𝕍([u5()])=diamB𝕍4([u5()])=123|u|.H_{\mathbb{V}}([u5^{(\infty)}])=\operatorname{diam}B_{\mathbb{V}}^{4}([u5^{(\infty)}])=\tfrac{1}{\sqrt{2}}3^{-|u|}.
Proof.

Assume first that w=u5()w=u5^{(\infty)}. We may assume that uu is not ending with 5. Note that [w][w] is the unique element in n=1𝕍u5(n)\bigcap_{n=1}^{\infty}\mathbb{V}_{u5^{(n)}}. Therefore, 𝕍{[w]}=n=1(𝕍𝕍u5(n)¯)\mathbb{V}\setminus\{[w]\}=\bigcup_{n=1}^{\infty}(\overline{\mathbb{V}\setminus\mathbb{V}_{u5^{(n)}}}). For each nn\in\mathbb{N}, denote by BinB_{i}^{n}, i{1,2,3,4}i\in\{1,2,3,4\}, the closures of the components of 𝕍𝕍u5(n)\mathbb{V}\setminus\mathbb{V}_{u5^{(n)}} given by Lemma 7.3(i). Note that (𝕍𝕍u5(n)¯)n(\overline{\mathbb{V}\setminus\mathbb{V}_{u5^{(n)}}})_{n\in\mathbb{N}} is an increasing sequence, so we may assume that Bin+1BinB_{i}^{n+1}\subset B_{i}^{n}. It follows that

𝕍{[w]}=i{1,2,3,4}n=1Bin.\mathbb{V}\setminus\{[w]\}=\bigcup_{i\in\{1,2,3,4\}}\bigcup_{n=1}^{\infty}B_{i}^{n}.

Since the sets n=1Bin\bigcup_{n=1}^{\infty}B_{i}^{n} are connected for each i{1,2,3,4}i\in\{1,2,3,4\} and disjoint for iji\neq j, they are the components of 𝕍[w]\mathbb{V}\setminus[w]. Thus, [w][w] is a branch point of valence 44.

For the converse, let w=u1i1u2i2Aw=u_{1}i_{1}u_{2}i_{2}\dots\in A^{\mathbb{N}} where ukAu_{k}\in A^{*} and ik{1,2,3,4}i_{k}\in\{1,2,3,4\} for kk\in\mathbb{N}. Set wn=u1i1uninw_{n}=u_{1}i_{1}\dots u_{n}i_{n}. By Lemma 7.3(ii), 𝕍𝕍wn\mathbb{V}\setminus\mathbb{V}_{w_{n}} has either 11 or 22 components. We consider two cases.

Case 1. Suppose that for all nn\in\mathbb{N}, 𝕍𝕍wn\mathbb{V}\setminus\mathbb{V}_{w_{n}} has 11 component denoted by KnK_{n}. Since 𝕍wn+1𝕍wn\mathbb{V}_{w_{n+1}}\subset\mathbb{V}_{w_{n}} it follows that KnKn+1K_{n}\subset K_{n+1}. Note, also, that {[w]}=n=1𝕍wn\{[w]\}=\bigcap_{n=1}^{\infty}\mathbb{V}_{w_{n}}. Thus, 𝕍{[w]}=n=1Kn\mathbb{V}\setminus\{[w]\}=\bigcup_{n=1}^{\infty}K_{n}, and the latter set is connected, hence is a leaf.

Case 2. Suppose that there exists n0n_{0}\in\mathbb{N} such that 𝕍𝕍wn0\mathbb{V}\setminus\mathbb{V}_{w_{n_{0}}} has 22 components, denoted by Kn01,Kn02K_{n_{0}}^{1},K_{n_{0}}^{2}. We claim that for all nn0n\geq n_{0}, 𝕍𝕍wn\mathbb{V}\setminus\mathbb{V}_{w_{n}} has 22 components Kn1K_{n}^{1}, Kn2K_{n}^{2}, such that Kn01Kn1K_{n_{0}}^{1}\subset K_{n}^{1} and Kn02Kn2K_{n_{0}}^{2}\subset K_{n}^{2}. Assume towards contradiction that there exists mn0m\geq n_{0} such that the claim doesn’t hold, and let Um=𝕍𝕍wm¯U_{m}=\overline{\mathbb{V}\setminus\mathbb{V}_{w_{m}}}. By Lemma 7.3(ii), UmU_{m} has 11 or 22 components. If it has 11 component, then Kn01Kn02UmK_{n_{0}}^{1}\cup K_{n_{0}}^{2}\subset U_{m}, since (𝕍𝕍wn¯)n(\overline{\mathbb{V}\setminus\mathbb{V}_{w_{n}}})_{n\in\mathbb{N}} is an increasing sequence. This is a contradiction, since UmU_{m} is a connected set, and Kn01K_{n_{0}}^{1}, Kn02K_{n_{0}}^{2} are disjoint. If UmU_{m} has 22 components, say Um1,Um2U_{m}^{1},U_{m}^{2}, then, similarly, Kn01Kn02K_{n_{0}}^{1}\cup K_{n_{0}}^{2} would have to lie in only one of them. But then, UmU_{m} would have at least three components, which is contradiction. Moreover, let Uk=𝕍𝕍wk¯U_{k}=\overline{\mathbb{V}\setminus\mathbb{V}_{w_{k}}} for k{1,,n01}k\in\{1,\dots,n_{0}-1\}. Note that (Uk)k(U_{k})_{k\in\mathbb{N}} is an increasing sequence of connected sets, since n0n_{0} is minimal. Hence the set k=1n01Uk\bigcup_{k=1}^{n_{0}-1}U_{k} lies in either Kn01K_{n_{0}}^{1} or Kn02K_{n_{0}}^{2}. It follows that

𝕍{[w]}=n=n0Kn1n=n0Kn2.\mathbb{V}\setminus\{[w]\}=\bigcup_{n=n_{0}}^{\infty}K_{n}^{1}\cup\bigcup_{n=n_{0}}^{\infty}K_{n}^{2}.

Since Ki:=n=n0KniK^{i}:=\bigcup_{n=n_{0}}^{\infty}K_{n}^{i}, i{1,2}i\in\{1,2\}, are connected and disjoint, it follows that K1,K2K^{1},K^{2} are all the components of 𝕍{[w]}\mathbb{V}\setminus\{[w]\}, which implies that [w][w] is a double point.

For the last claim, we show that two of the components (in fact the ones with the smallest diameter) are n=0𝕍u5(n)i\bigcup_{n=0}^{\infty}\mathbb{V}_{u5^{(n)}i} and n=0𝕍u5(n)j\bigcup_{n=0}^{\infty}\mathbb{V}_{u5^{(n)}j}, where (i,j){(1,3),(2,4)}(i,j)\in\{(1,3),(2,4)\}. By Lemma 7.3(iii), one of the components of 𝕍𝕍w5¯\overline{\mathbb{V}\setminus\mathbb{V}_{w5}} is 𝕍wi\mathbb{V}_{wi} for i{1,2,3,4}i\in\{1,2,3,4\}. Suppose 𝕍wi=𝕍w1\mathbb{V}_{wi}=\mathbb{V}_{w1}, and other cases are similar. Let BB be the component of 𝕍𝕍w55¯\overline{\mathbb{V}\setminus\mathbb{V}_{w55}} for which 𝕍w1B\mathbb{V}_{w1}\subset B. Moreover, the fact that the set 𝕍w51\mathbb{V}_{w51} is a component of 𝕍w5𝕍w55¯\overline{\mathbb{V}_{w5}\setminus\mathbb{V}_{w55}}, along with 𝕍w1𝕍w51={[w13()]}\mathbb{V}_{w1}\cap\mathbb{V}_{w51}=\{[w13^{(\infty)}]\}, imply that 𝕍w1𝕍w51B\mathbb{V}_{w1}\cup\mathbb{V}_{w51}\subset B. On the other hand, note that

𝕍w5𝕍w55¯=(𝕍𝕍w55)(𝕍𝕍w5)¯,\overline{\mathbb{V}_{w5}\setminus\mathbb{V}_{w55}}=\overline{(\mathbb{V}\setminus\mathbb{V}_{w55})\setminus(\mathbb{V}\setminus\mathbb{V}_{w5})},

and that one of the components of the latter set is B𝕍w1¯\overline{B\setminus\mathbb{V}_{w1}}. Since 𝕍w51B\mathbb{V}_{w51}\subset B, it follows that B=𝕍w1𝕍w51B=\mathbb{V}_{w1}\cup\mathbb{V}_{w51}. The rest of the claim follows similarly by applying Lemma 7.3(i) and the fact that for each nn\in\mathbb{N}, 𝕍u5(n)𝕍u5(n+1)=i{1,2,3,4}𝕍u5(n)i\mathbb{V}_{u5^{(n)}}\setminus\mathbb{V}_{u5^{(n+1)}}=\bigcup_{i\in\{1,2,3,4\}}\mathbb{V}_{u5^{(n)}i}. Note that by Lemma 7.3(iii), it can be similarly shown that the rest of the components have larger diameter. Hence,

H𝕍(p)=diamB𝕍4(p)=2n=03|u|1n=3|u|12n=03n=123|u|.H_{\mathbb{V}}(p)=\operatorname{diam}B_{\mathbb{V}}^{4}(p)=\sqrt{2}\sum_{n=0}^{\infty}3^{-|u|-1-n}=3^{-|u|-1}\sqrt{2}\sum_{n=0}^{\infty}3^{-n}=\tfrac{1}{\sqrt{2}}3^{-|u|}.\qed
Proof of Proposition 7.1.

By Lemma 7.4, it follows that every branch point of 𝕍\mathbb{V} has valence 44, and that 𝕍\mathbb{V} has uniform branch growth. It remains to show uniform branch separation and uniform density.

To show uniform branch separation, fix v,wAv,w\in A^{*} with vwv\neq w, and let [v5()][v5^{(\infty)}], [w5()][w5^{(\infty)}] be two branch points of 𝕍\mathbb{V}. We may assume that v,wv,w are not ending with 5. Let uAu\in A^{*} be the longest common initial word of vv and ww. Then v=uvv=uv^{\prime} and w=uww=uw^{\prime} with w,vAw^{\prime},v^{\prime}\in A^{*} and w(1)v(1)w^{\prime}(1)\neq v^{\prime}(1). Without loss of generality, assume that |w||v||w^{\prime}|\geq|v^{\prime}|, and that wϵw^{\prime}\neq\epsilon. Since ϕw([5()])ϕw([1,1]2)\phi_{w^{\prime}}([5^{(\infty)}])\in\phi_{w^{\prime}}([-1,1]^{2}) and ϕv([5)])ϕv([1,1]2)\phi_{v^{\prime}}([5^{\infty)}])\in\phi_{v^{\prime}}([-1,1]^{2}), an elementary geometric estimate shows that

|ϕv([5()])ϕw([5()])|123|w|.|\phi_{v^{\prime}}([5^{(\infty)}])-\phi_{w^{\prime}}([5^{(\infty)}])|\geq\tfrac{1}{2}3^{-|w^{\prime}|}.

Hence, we have

|ϕv([5()])ϕw([5()])|\displaystyle|\phi_{v}([5^{(\infty)}])-\phi_{w}([5^{(\infty)}])| =|ϕu(ϕv([5()])ϕu(ϕw([5()]))|\displaystyle=|\phi_{u}(\phi_{v^{\prime}}([5^{(\infty)}])-\phi_{u}(\phi_{w^{\prime}}([5^{(\infty)}]))|
=3|u||ϕv([5()])ϕw([5()])|\displaystyle=3^{-|u|}|\phi_{v^{\prime}}([5^{(\infty)}])-\phi_{w^{\prime}}([5^{(\infty)}])|
123|u|3|w|\displaystyle\geq\tfrac{1}{2}3^{-|u|}3^{-|w^{\prime}|}
min{H𝕍([v5()]),H𝕍([w5()])}.\displaystyle\gtrsim\min\{H_{\mathbb{V}}([v5^{(\infty)}]),H_{\mathbb{V}}([w5^{(\infty)}])\}.

To show uniform density, let [w],[v]𝕍[w],[v]\in\mathbb{V} with [w][v][w]\neq[v]. Let uAu\in A^{*} be the longest common initial word of vv and ww, that is, w=uww=uw^{\prime} and v=uvv=uv^{\prime} with w,vAw^{\prime},v^{\prime}\in A^{\mathbb{N}} and w(1)v(1)w^{\prime}(1)\neq v^{\prime}(1). Hence, [w]𝕍uw(1)[w]\in\mathbb{V}_{uw^{\prime}(1)} and [v]𝕍uv(1)[v]\in\mathbb{V}_{uv^{\prime}(1)}. Note that for each nn\in\mathbb{N}, the unique combinatorial arc that connects uw(n)uw^{\prime}(n) with uv(n)uv^{\prime}(n) passes through u5(n)u5^{(n)}. Therefore, by [DV22, Claim 3.19], the unique arc in 𝕍\mathbb{V} that joins [w][w] with [v][v] contains the branch point p=[u5()]p=[u5^{(\infty)}], which satisfies

H𝕍(p)=123|u|3|u||[w][v]|=|[w][v]|.\displaystyle H_{\mathbb{V}}(p)=\tfrac{1}{\sqrt{2}}3^{-|u|}\gtrsim 3^{-|u|}|[w^{\prime}]-[v^{\prime}]|=|[w]-[v]|.

This completes the proof of the proposition. ∎

8. Ahlfors regularity and dimensions

In this section we study the dimensions of trees 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} where m{2,3,}{}m\in\{2,3,\dots\}\cup\{\infty\}. Recall that a metric space XX is ss-Ahlfors regular if there exists a measure μ\mu on XX and a constant C>1C>1 such that

C1rsμ(B(x,r))Crs,for all xX and r(0,diamX).C^{-1}r^{s}\leq\mu(B(x,r))\leq Cr^{s},\qquad\text{for all $x\in X$ and $r\in(0,\operatorname{diam}{X})$.}
Proposition 8.1.

Let 𝐚\operatorname{\bf{a}} be a weight. For each m{2,3,}m\in\{2,3,\dots\} define Ψm,𝐚(t)=𝐚(1)t++𝐚(m)t\Psi_{m,\operatorname{\bf{a}}}(t)=\operatorname{\bf{a}}(1)^{t}+\cdots+\operatorname{\bf{a}}(m)^{t}.

  1. (i)

    If m{2,3,}m\in\{2,3,\dots\}, then 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is Ahlfors ss-regular where ss is the unique solution of the equation Ψm,𝐚(s)=1\Psi_{m,\operatorname{\bf{a}}}(s)=1.

  2. (ii)

    Suppose that there exists s>0s>0 such that limmΨm,𝐚(s)<1\lim_{m\to\infty}\Psi_{m,\operatorname{\bf{a}}}(s)<1. Then the Hausdorff dimension of 𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}} is at most ss.

We start with the second claim of the proposition.

Proof of Proposition 8.1(ii).

Fix δ>0\delta>0. By Lemma 5.10 and our assumption, there exist integers m,n2m,n\geq 2 such that 2n<δ/42^{-n}<\delta/4, distH(𝕋m,𝐚,𝕋,𝐚)<δ/2\operatorname{dist}_{H}(\mathbb{T}^{m,\operatorname{\bf{a}}},\mathbb{T}^{\infty,\operatorname{\bf{a}}})<\delta/2, and Ψm,𝐚(s)<1\Psi_{m,\operatorname{\bf{a}}}(s)<1. Fix for each w{1,,m}nw\in\{1,\dots,m\}^{n} a point xw𝕋m,𝐚x_{w}\in\mathbb{T}^{m,\operatorname{\bf{a}}}. Since 𝕋wm,𝐚B(xw,2Δ𝐚(w))\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}\subset B(x_{w},2\Delta_{\operatorname{\bf{a}}}(w)) we have that

{B(xw,2Δ𝐚(w)):w{1,,m}n}\{B(x_{w},2\Delta_{\operatorname{\bf{a}}}(w)):w\in\{1,\dots,m\}^{n}\}

is an open cover of 𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}} by balls with diameters 4Δ𝐚(w)4(1/2)n<δ4\Delta_{\operatorname{\bf{a}}}(w)\leq 4(1/2)^{n}<\delta. Therefore,

δs(𝕋,𝐚)w{1,,m}n(4Δ𝐚(w))s\displaystyle\mathcal{H}^{s}_{\delta}(\mathbb{T}^{\infty,\operatorname{\bf{a}}})\leq\sum_{w\in\{1,\dots,m\}^{n}}(4\Delta_{\operatorname{\bf{a}}}(w))^{s} =4si1,,in{1,,m}𝐚(i1)s𝐚(in)s\displaystyle=4^{s}\sum_{i_{1},\dots,i_{n}\in\{1,\dots,m\}}\operatorname{\bf{a}}(i_{1})^{s}\cdots\operatorname{\bf{a}}(i_{n})^{s}
=4s(Ψm,𝐚(s))n\displaystyle=4^{s}\left(\Psi_{m,\operatorname{\bf{a}}}(s)\right)^{n}
<4s.\displaystyle<4^{s}.

Thus, as δ0\delta\to 0, we get that s(𝕋,𝐚)4s<\mathcal{H}^{s}(\mathbb{T}^{\infty,\operatorname{\bf{a}}})\leq 4^{s}<\infty, so the Hausdorff dimension of 𝕋,𝐚\mathbb{T}^{\infty,\operatorname{\bf{a}}} is at most ss. ∎

We now turn to the proof of Proposition 8.1(i). Fix for the rest of this section an integer m{2,3,}m\in\{2,3,\dots\} and set A={1,,m}A=\{1,\dots,m\}. Let ss be the unique solution of the equation Ψm,𝐚(t)=1\Psi_{m,\operatorname{\bf{a}}}(t)=1. Then for each nn\in\mathbb{N}

wAnΔ𝐚(w)s=i1,,in{1,,m}𝐚(i1)s𝐚(in)s=(Ψm,𝐚(s))n=1.\sum_{w\in A^{n}}\Delta_{\operatorname{\bf{a}}}(w)^{s}=\sum_{i_{1},\dots,i_{n}\in\{1,\dots,m\}}\operatorname{\bf{a}}(i_{1})^{s}\cdots\operatorname{\bf{a}}(i_{n})^{s}=\left(\Psi_{m,\operatorname{\bf{a}}}(s)\right)^{n}=1.

Denote by Σ\Sigma the σ\sigma-algebra generated by the cylinders AwA^{\mathbb{N}}_{w} where wAw\in A^{*}. There exists a unique probability measure μ:Σ[0,1]\mu:\Sigma\to[0,1] such that for all wAw\in A^{*},

μ(Aw)=Δ𝐚(w)s,\mu(A^{\mathbb{N}}_{w})=\Delta_{\operatorname{\bf{a}}}(w)^{s},

(e.g. see [Str93, §3.1]). Define also the projection map π:A𝕋m,𝐚\pi:A^{\mathbb{N}}\to\mathbb{T}^{m,\operatorname{\bf{a}}} given by π(w)=[w]\pi(w)=[w].

Lemma 8.2.

The σ\sigma-algebra generated by the sets 𝕋wm,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w} is the same as the Borel σ\sigma-algebra on 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

Proof.

Let 𝒜\mathcal{A} denote the σ\sigma-algebra generated by the sets 𝕋wm,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}, and let (𝕋m,𝐚)\mathcal{B}(\mathbb{T}^{m,\operatorname{\bf{a}}}) be the Borel σ\sigma-algebra on 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}. That 𝒜(𝕋m,𝐚)\mathcal{A}\subset\mathcal{B}(\mathbb{T}^{m,\operatorname{\bf{a}}}) is clear, as each 𝕋wm,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w} is a Borel set. For the reverse inclusion, fix x𝕋m,𝐚x\in\mathbb{T}^{m,\operatorname{\bf{a}}} and r>0r>0. Let W(x,r)={wA:𝕋wm,𝐚B(x,r)}W(x,r)=\{w\in A^{*}:\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}\subset B(x,r)\} and note that W(x,r)W(x,r) is a countable set since AA^{*} is countable, and so wW(x,r)𝕋wm,𝐚𝒜\bigcup_{w\in W(x,r)}\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}\in\mathcal{A}. Furthermore for each point yB(x,r)𝕋m,𝐚y\in B(x,r)\cap\mathbb{T}^{m,\operatorname{\bf{a}}} there is a word wAw\in A^{\mathbb{N}} with [w]=y[w]=y, and since B(x,r)𝕋m,𝐚B(x,r)\cap\mathbb{T}^{m,\operatorname{\bf{a}}} is an open set and limndiam(𝕋w(n)m,𝐚)=0\lim_{n\to\infty}\operatorname{diam}(\mathbb{T}^{m,\operatorname{\bf{a}}}_{w(n)})=0 there is some nn\in\mathbb{N} large enough that w(n)W(x,r)w(n)\in W(x,r). Therefore, wW(x,r)𝕋wm,𝐚=B(x,r)𝕋m,𝐚\bigcup_{w\in W(x,r)}\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}=B(x,r)\cap\mathbb{T}^{m,\operatorname{\bf{a}}}. ∎

We say a word wAw\in A^{\mathbb{N}} is non-injective if there exists uA{w}u\in A^{\mathbb{N}}\setminus\{w\} such that [w]=[u][w]=[u].

Lemma 8.3.

The set 𝒩\mathcal{N} of non-injective words is a μ\mu-null set.

Proof.

Suppose that ww is a non-injective word. Then there exists uA{w}u\in A^{\mathbb{N}}\setminus\{w\} such that [w]=[u][w]=[u]. There exist vAv\in A^{*} and j,jAj,j^{\prime}\in A such that jjj\neq j^{\prime}, wAvjw\in A^{\mathbb{N}}_{vj}, and uAvju\in A^{\mathbb{N}}_{vj^{\prime}}. Since [w]=[u][w]=[u], it follows that [w]𝕋vjm,𝐚𝕋vjm,𝐚[w]\in\mathbb{T}^{m,\operatorname{\bf{a}}}_{vj}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{vj^{\prime}}\neq\emptyset. However, by Lemma 6.2(i) we have that [w]=[v12()][w]=[v12^{(\infty)}] and by Lemma 6.2(ii) we have that w{v12(),v21(),,vm1()}w\in\{v12^{(\infty)},v21^{(\infty)},\dots,vm1^{(\infty)}\}. Therefore,

𝒩{v12():vA}{vj1():vA,jA}\mathcal{N}\subset\{v12^{(\infty)}:v\in A^{*}\}\cup\{vj1^{(\infty)}:v\in A^{*},\,j\in A\}

which implies that 𝒩\mathcal{N} is countable. Therefore, μ(𝒩)=0\mu(\mathcal{N})=0. ∎

We now prove the first claim of Proposition 8.1.

Proof of Proposition 8.1(i).

Set c=mini=1,,m𝐚(i)c=\min_{i=1,\dots,m}\operatorname{\bf{a}}(i). We claim that there exists C>1C>1 depending only on mm, cc, ss such that the pushforward measure π#μ\pi\#\mu defined on (𝕋m,𝐚)\mathcal{B}(\mathbb{T}^{m,\operatorname{\bf{a}}}) by Lemma 8.2 satisfies

(8.1) C1rsπ#μ(B(x,r))Crs,for all x𝕋m,𝐚 and r(0,2c).C^{-1}r^{s}\leq\pi\#\mu(B(x,r))\leq Cr^{s},\qquad\text{for all $x\in\mathbb{T}^{m,\operatorname{\bf{a}}}$ and $r\in(0,2c)$.}

First, for any wAw\in A^{*}, we have Awπ1(𝕋wm,𝐚)𝒩AwA^{\mathbb{N}}_{w}\subset\pi^{-1}(\mathbb{T}^{m,\operatorname{\bf{a}}}_{w})\subset\mathcal{N}\cup A^{\mathbb{N}}_{w}, where 𝒩\mathcal{N} is the set of non-injective words in AA^{\mathbb{N}}. Hence, by Lemma 8.3 for any wAw\in A^{*},

π#μ(𝕋wm,𝐚)=Δ𝐚(w)s.\pi\#\mu(\mathbb{T}^{m,\operatorname{\bf{a}}}_{w})=\Delta_{\operatorname{\bf{a}}}(w)^{s}.

Fix for the rest of the proof a point x𝕋m,𝐚x\in\mathbb{T}^{m,\operatorname{\bf{a}}}, a radius r(0,2c)r\in(0,2c), and a word wAw\in A^{\mathbb{N}} such that [w]=x[w]=x.

For the upper bound of (8.1), let nn\in\mathbb{N} such that

Δ𝐚(w(n+1))(2c)1r<Δ𝐚(w(n)).\Delta_{\operatorname{\bf{a}}}(w(n+1))\leq(2c)^{-1}r<\Delta_{\operatorname{\bf{a}}}(w(n)).

By Lemma 6.2(ii) and Lemma 6.5, if w1,,wkw_{1},\dots,w_{k} are exactly the words in An{w(n)}A^{n}\setminus\{w(n)\} such that 𝕋wm,𝐚𝕋wim,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}\cap\mathbb{T}^{m,\operatorname{\bf{a}}}_{w_{i}}\neq\emptyset, then k2mk\leq 2m. Moreover, by (6.3), for each i=1,,ki=1,\dots,k,

r<2cdiam𝕋w(n)m,𝐚diam𝕋wim,𝐚(2c)1diam𝕋w(n)m,𝐚.r<2c\operatorname{diam}{\mathbb{T}^{m,\operatorname{\bf{a}}}_{w(n)}}\leq\operatorname{diam}{\mathbb{T}^{m,\operatorname{\bf{a}}}_{w_{i}}}\leq(2c)^{-1}\operatorname{diam}{\mathbb{T}^{m,\operatorname{\bf{a}}}_{w(n)}}.

Therefore, since r>(2c)Δ𝐚(w(n+1))>(2c2)Δ𝐚(w(n))r>(2c)\Delta_{\operatorname{\bf{a}}}(w(n+1))>(2c^{2})\Delta_{\operatorname{\bf{a}}}(w(n)),

π#μ(B(x,r))π#μ(𝕋w(n)m,𝐚)+i=1kπ#μ(𝕋wim,𝐚)\displaystyle\pi\#\mu(B(x,r))\leq\pi\#\mu(\mathbb{T}^{m,\operatorname{\bf{a}}}_{w(n)})+\sum_{i=1}^{k}\pi\#\mu(\mathbb{T}^{m,\operatorname{\bf{a}}}_{w_{i}}) (1+2m(2c)s)Δ𝐚(w(n))s\displaystyle\leq(1+2m(2c)^{-s})\Delta_{\operatorname{\bf{a}}}(w(n))^{s}
(1+2m(2c)s)(2c2)srs.\displaystyle\leq(1+2m(2c)^{-s})(2c^{2})^{-s}r^{s}.

For the lower bound of (8.1), let nn\in\mathbb{N} such that

Δ𝐚(w(n+1))<rΔ𝐚(w(n)).\Delta_{\operatorname{\bf{a}}}(w(n+1))<r\leq\Delta_{\operatorname{\bf{a}}}(w(n)).

Then,

π#μ(B(x,r))π#μ(𝕋w(n+1)m,𝐚)=Δ𝐚(w(n+1))scsrs.\pi\#\mu(B(x,r))\geq\pi\#\mu(\mathbb{T}^{m,\operatorname{\bf{a}}}_{w(n+1)})=\Delta_{\operatorname{\bf{a}}}(w(n+1))^{s}\geq c^{s}r^{s}.\qed

9. Quasisymmetric uniformization of uniformly mm-branching trees

In this section we prove the following quantitative version of Theorem 1.9.

Theorem 9.1.

Let m{2,3,}m\in\{2,3,\dots\} and let 𝐚\operatorname{\bf{a}} be a weight. A metric space TT is a uniformly mm-branching quasiconformal tree if and only if it is quasisymmetrivally equivalent to 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

By Proposition 5.1 and Proposition 6.1, 𝕋2,𝐚\mathbb{T}^{2,\operatorname{\bf{a}}} is a geodesic metric arc, hence isometric to the (Euclidean) unit interval [0,1][0,1]. Therefore, the case m=2m=2 in Theorem 9.1 follows by the quasisymmetric uniformization of quasi-arcs by Tukia and Väisälä [TV80]. Thus, we may assume for the rest of this section that m3m\geq 3.

One direction of Theorem 9.1 is simple and we record it as a lemma.

Lemma 9.2.

Suppose that TT is quasisymmetrivally equivalent to 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} for some m{3,4,}m\in\{3,4,\dots\} and some weight 𝐚\operatorname{\bf{a}}. Then, TT is a uniformly mm-branching quasiconformal tree.

Proof.

By Lemma 4.8 and Proposition 6.1, 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is an mm-valent metric tree. Since TT is homeomorphic to 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}, it follows that TT is an mm-valent metric tree as well.

Moreover, by Lemma 4.8 and Proposition 8.1, 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is doubling and bounded turning, and since both these properties are quasisymmetrically invariant, it follows that TT is bounded turning and doubling, hence a quasiconformal tree.

By Proposition 6.1, 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is uniformly mm-branching. By [BM22, Lemma 4.3], TT has uniformly relatively separated branches and by [BM22, Lemma 4.5] TT has uniformly dense branch points. To show that TT is uniformly mm-branching is to show that TT has uniform branch growth, which follows by Lemma 2.6. ∎

Note that it is enough to show Theorem 9.1 for a fixed weight 𝐚\operatorname{\bf{a}} since 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is uniformly mm-branching for all choices of weights 𝐚\operatorname{\bf{a}} by Proposition 6.1.

For the rest of Section 9, we fix m{3,4,}m\in\{3,4,\dots\}, we fix a uniformly mm-branching tree TT, we set A={1,,m}A=\{1,\dots,m\}, and we fix a weight 𝐚\operatorname{\bf{a}} such that 𝐚(i)=1/2\operatorname{\bf{a}}(i)=1/2 for all iAi\in A. By rescaling the metric on TT if necessary, we may assume that diam(T)=1\operatorname{diam}(T)=1 for convenience. We show below that TT is quasisymmetrically equivalent to 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}. The proof follows closely the proof of [BM22, Theorem 1.4], which is the special case m=3m=3. Since the proof and techniques are almost identical, we mostly sketch the steps.

9.1. Quasi-visual subdivision of TT

By [BM22, Section 3] a finite set VT\textbf{V}\subset T that does not contain leaves of TT decomposes TT into a set of tiles X (in the sense of Definition 2.7). These tiles are subtrees of TT. We want to map them to tiles of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}, i.e., sets of the form 𝕋wm,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}, wAw\in A^{*}. By Lemma 6.5 each tile 𝕋wm,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w} with |w|>0|w|>0 has one or two boundary points. For this reason, we are interested in decompositions such that every tile XXX\in\textbf{X} has one or two boundary points.

We call X an edge-like decomposition of TT if V=\textbf{V}=\emptyset and X={T}\textbf{X}=\{T\} (as a degenerate case), or if V\textbf{V}\neq\emptyset and card(X)2\operatorname{card}(\partial X)\leq 2 for each XXX\in\textbf{X}. Note that in the latter case 1card(X)21\leq\operatorname{card}(\partial X)\leq 2 by [BM22, Lemma 3.3(iii)]. We say that a tile XXX\in\textbf{X} in an edge-like decomposition X of TT is a leaf-tile if card(X)=1\operatorname{card}(\partial X)=1 and an edge-tile if card(X)=2\operatorname{card}(\partial X)=2.

We now set V0=\textbf{V}^{0}=\emptyset, fix δ(0,1)\delta\in(0,1), and for nn\in\mathbb{N} define

(9.1) Vn={pT:pis a branch point ofTwithHT(p)δn},\textbf{V}^{n}=\{p\in T:p\ \text{is a branch point of}\ T\ \text{with}\ H_{T}(p)\geq\delta^{n}\},

where the height HTH_{T} is as defined in (1.1). Note that (Vn)n0(\textbf{V}^{n})_{n\in\mathbb{N}_{0}} is an increasing nested sequence and none of the sets Vn\textbf{V}^{n} contains a leaf of TT. Denote by Xn\textbf{X}^{n} the set of tiles in the decomposition of TT induced by Vn\textbf{V}^{n}. Then the sequence (Xn)(\textbf{X}^{n}) forms a subdivision of TT in the sense of Definition 2.8 (see also the discussion in [BM22] after (3.3)). The next proposition is a direct analogue of [BM22, Proposition 6.1].

Proposition 9.3.

Let δ(0,1)\delta\in(0,1), and (Vn)n0(\textbf{V}^{n})_{n\in\mathbb{N}_{0}} be as in (9.1). Then the following statements are true:

  1. (i)

    Vn\textbf{V}^{n} is a finite set for each integer n0n\geq 0.

  2. (ii)

    (Xn)n0(\textbf{X}^{n})_{n\in\mathbb{N}_{0}} is a quasi-visual subdivision of TT.

Let n0n\geq 0, XXnX\in\textbf{X}^{n}, and VXn+1:=Vn+1int(X)\textbf{V}_{X}^{n+1}:=\textbf{V}^{n+1}\cap\mathrm{int}(X). Then we have:

  1. (iii)

    card(X)2\operatorname{card}(\partial X)\leq 2, and if we denote by XXn+1\textbf{X}_{X}^{n+1} the decomposition of XX induced by VXn+1\textbf{V}_{X}^{n+1}, then XXn+1\textbf{X}_{X}^{n+1} is edge-like.

  2. (iv)

    There exists NN\in\mathbb{N} depending on δ\delta, but independent of nn and XX, such that card(VXn+1)N\operatorname{card}(\textbf{V}_{X}^{n+1})\leq N.

If δ(0,1)\delta\in(0,1) is sufficiently small (independent of nn and XX), then we also have:

  1. (v)

    card(VXn+1)2\operatorname{card}(\textbf{V}_{X}^{n+1})\geq 2.

  2. (vi)

    If card(X)=2\operatorname{card}(\partial X)=2 and X={u,v}T\partial X=\{u,v\}\subset T, then (u,v)VXn+1(u,v)\cap\textbf{V}_{X}^{n+1} contains at least three elements.

Proof.

The proof of this proposition is almost identical to the proof of [BM22, Proposition 6.1] so we only sketch it. The specific value m=3m=3 is only implicitly used in the proof of (ii). Hence, we only focus on that part, since (i) and (iii)-(vi) above follow in the exact same way as in the m=3m=3 case. Note that (iii) is proven before (ii) in [BM22, Proposition 6.1], so we may assume card(X)2\operatorname{card}(X)\leq 2 for all XXnX\in\textbf{X}^{n}.

To show that Xn\textbf{X}^{n} is a quasi-visual subdivision of TT, by [BM22, Lemma 2.4], it suffices to verify the following two conditions:

  1. (a)

    diamXδn\operatorname{diam}{X}\simeq\delta^{n} for all n0n\geq 0 and XXnX\in\textbf{X}^{n},

  2. (b)

    dist(X,Y)δn\operatorname{dist}(X,Y)\gtrsim\delta^{n} for all n0n\geq 0 and X,YXnX,Y\in\textbf{X}^{n} with XY=X\cap Y=\emptyset.

Relations diam(X)δn\operatorname{diam}(X)\lesssim\delta^{n} and (b) can be proved for m3m\geq 3 verbatim as in the case m=3m=3 in [BM22, Proposition 6.1(ii)].

For the relation diam(X)δn\operatorname{diam}(X)\gtrsim\delta^{n}, assume that X\partial X\neq\emptyset, otherwise X=TX=T and so diam(X)=1δn\operatorname{diam}(X)=1\geq\delta^{n}. If X\partial X consists of one point pTp\in T, then pXVnp\in\partial X\subset\textbf{V}^{n} (due to [BM22, Lemma 3.3(ii)] and so pp is a branch point of TT with HT(p)δnH_{T}(p)\geq\delta^{n}. In this case, XX is a branch of pp in TT (see [BM22, Lemma 3.3(vi)]). Hence, either X=BjX=B_{j}, j{1,2}j\in\{1,2\}, which implies diam(X)HT(p)\operatorname{diam}(X)\geq H_{T}(p), or

diam(X)HT(p)δn,\operatorname{diam}(X)\simeq H_{T}(p)\geq\delta^{n},

since TT has uniform branch growth. If X\partial X consists of two distinct points u,vVnu,v\in\textbf{V}^{n}, then we have

diam(X)|uv|min{HT(u),HT(v)}δn\operatorname{diam}(X)\geq|u-v|\gtrsim\min\{H_{T}(u),H_{T}(v)\}\geq\delta^{n}

where the second inequality follows from the fact that TT has uniformly relatively separated branch points. ∎

9.2. Quasi-visual subdivision of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}

The following proposition, which is the direct analogue of [BM22, Proposition 7.1], gives us a quasi-visual subdivision of the metric tree 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} that is isomorphic to that of TT constructed in §9.1.

Proposition 9.4.

Let (Vn)n0(\textbf{V}^{n})_{n\in\mathbb{N}_{0}} be as in (9.1), with δ(0,1)\delta\in(0,1) small enough so that, for all n0n\in\mathbb{N}_{0}, all the statements in Proposition 9.3 are true for the decomposition Xn\textbf{X}^{n} of TT induced by the set Vn\textbf{V}^{n}. Then there exists a quasi-visual subdivision (Yn)n0(\textbf{Y}^{n})_{n\in\mathbb{N}_{0}} of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} that is isomorphic to the quasi-visual subdivision (Xn)n0(\textbf{X}^{n})_{n\in\mathbb{N}_{0}} of TT.

The proof of Proposition 9.4 follows that of [BM22, Proposition 7.1] very closely.

By Proposition 9.3(iii) the decompositions Xn\textbf{X}^{n} of TT are edge-like. The points in X\partial X are leaves of XX and the only points where XX intersects other tiles of the same level (see [BM22, Lemma 3.3(iii), (iv)]). We follow the definition of marked leaves as in the proof of Lemma 2.2 with the difference that points in X\partial X do not carry a sign here. Moreover, XXnX\in\textbf{X}^{n} (viewed as a tree) has an edge-like decomposition XXn+1\textbf{X}_{X}^{n+1} induced by VXn+1=Vn+1int(X)\textbf{V}_{X}^{n+1}=\textbf{V}^{n+1}\cap\mathrm{int}(X).

We want to find a decomposition of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} into tiles 𝕋wm,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}_{w} that is isomorphic to XXn+1\textbf{X}_{X}^{n+1} and respects the marked leaves. As in [BM22], we consider [1()][1^{(\infty)}] or [2()][2^{(\infty)}] as a marked leaf of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} if card(X)=1\operatorname{card}(\partial X)=1, or both [1()][1^{(\infty)}] and [2()][2^{(\infty)}] as marked leaves of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} if card(X)=2\operatorname{card}(\partial X)=2. Since 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is uniformly mm-branching, for fixed leaves p1,p2Tp_{1},p_{2}\in T, by Lemma 2.2 there exists a homeomorphism f:T𝕋m,𝐚f:T\to\mathbb{T}^{m,\operatorname{\bf{a}}} such that f(p1)=[1()]f(p_{1})=[1^{(\infty)}] and f(p2)=[2()]f(p_{2})=[2^{(\infty)}].

If wAw\in A^{*}, then we say that the level of the tile 𝕋wm,𝐚\mathbb{T}_{w}^{m,\operatorname{\bf{a}}} (denoted by L(𝕋wm,𝐚)\texttt{L}(\mathbb{T}_{w}^{m,\operatorname{\bf{a}}})) is equal |w||w|. We also say that a homeomorphism F:T𝕋m,𝐚F:T\to\mathbb{T}^{m,\operatorname{\bf{a}}} is a tile-homeomorphism (for X)\textbf{X}) if F(X)F(X) is of the form 𝕋wm,𝐚\mathbb{T}_{w}^{m,\operatorname{\bf{a}}} for each XXX\in\textbf{X}. Then FF maps tiles in TT to tiles 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} and so the level L(F(X))\texttt{L}(F(X)) (of F(X)F(X) as a tile of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}) is defined for each XXX\in\textbf{X}.

The following lemma is a useful characterization of decompositions of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} of the form 𝕋wm,𝐚\mathbb{T}_{w}^{m,\operatorname{\bf{a}}} where wAw\in A^{*}.

Lemma 9.5.

Let V𝕋m,𝐚V\subset\mathbb{T}^{m,\operatorname{\bf{a}}} be a finite set of branch points of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}, and let 𝕏\mathbb{X} be the set of tiles in the decomposition of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} induced by VV. If 𝕏\mathbb{X} is of the form 𝕋wm,𝐚\mathbb{T}_{w}^{m,\operatorname{\bf{a}}} for wAw\in A^{*}, then the levels of tiles in 𝕏\mathbb{X} satisfy

L(X)card(V)for eachX𝕏.\texttt{L}(X)\leq\operatorname{card}(V)\ \text{for each}\ X\in\mathbb{X}.

If, in addition, VV\neq\emptyset, then we have L(X)1\texttt{L}(X)\geq 1 for each X𝕏X\in\mathbb{X}.

Proof.

The proof is identical to that of [BM22, Lemma 5.5] for m=3m=3. We solely point out that the point 00\in\mathbb{C} of the CSST corresponds in our setup to the point [12()]𝕋m,𝐚[12^{(\infty)}]\in\mathbb{T}^{m,\operatorname{\bf{a}}}. ∎

The following lemma is an analogue of [BM22, Lemma 7.3].

Lemma 9.6.

Let SS be an mm-valent metric tree whose branch points are dense in SS. Suppose VS\textbf{V}\subset S is a finite set of branch points of SS that induces an edge-like decomposition of SS into the set of tiles X.

Assume that either SS has no marked leaves, or SS has exactly one marked leaf pp, or SS has exactly two marked leaves p,qSp,q\in S. If V\textbf{V}\neq\emptyset, we also assume that the marked leaf pp (if it exists) lies in a leaf-tile PXP\in\textbf{X}, and the other marked leaf qq (if it exists) lies in a leaf-tile QXQ\in\textbf{X} distinct from PP.

Then there exists a tile-homeomorphism F:S𝕋m,𝐚F:S\to\mathbb{T}^{m,\operatorname{\bf{a}}} for X such that the following statements are true.

  1. (i)

    F(p)=[1()]F(p)=[1^{(\infty)}], or alternatively, F(p)=[2()]F(p)=[2^{(\infty)}], if SS has one marked leaf pp; or F(p)=[1()]F(p)=[1^{(\infty)}] and F(q)=[2()]F(q)=[2^{(\infty)}], if SS has two marked leaves pp and qq.

  2. (ii)

    If SS has one marked leaf pp and card(V)2\operatorname{card}(\textbf{V})\geq 2, then we may also assume that FF satisfies L(F(P))=2\texttt{L}(F(P))=2.

  3. (iii)

    If SS has two marked leaves p,qSp,q\in S and [p,q]V[p,q]\cap\textbf{V} contains at least three points, then we may also assume that FF satisfies L(F(P))=L(F(Q))=2\texttt{L}(F(P))=\texttt{L}(F(Q))=2.

  4. (iv)

    For each XXX\in\textbf{X} we have L(F(X))card(V)\texttt{L}(F(\textbf{X}))\leq\operatorname{card}(\textbf{V}), and if V\textbf{V}\neq\emptyset, then L(F(X))1\texttt{L}(F(X))\geq 1.

Proof.

The proof is almost identical to that of [BM22, Lemma 7.3] (the case m=3m=3) with slight adjustments to notation and lemmas employed. Namely, the points 0,1,10,-1,1\in\mathbb{C} of the CSST in the proof of [BM22, Lemma 7.3] correspond to [12()],[1()],[2()]𝕋m,𝐚[12^{(\infty)}],[1^{(\infty)}],[2^{(\infty)}]\in\mathbb{T}^{m,\operatorname{\bf{a}}}, respectively. Lemma 2.2 and Lemma 9.5 are used for the case m3m\geq 3 in place of [BM22, Theorem 7.2] and [BM22, Lemma 5.5], respectively. Lastly, [BM22, Lemma 3.3(viii)] (which states that tiles of mm-valent trees with dense branch points are also mm-valent trees with dense branch points) is stated and proved for m=3m=3, but the proof for m3m\geq 3 is verbatim. ∎

We are now ready to prove Proposition 9.4. The desired quasi-visual subdivision (Yn)(\textbf{Y}^{n}) will be constructed inductively from auxiliary tile-homeomorphisms Fn:T𝕋m,𝐚F^{n}:T\to\mathbb{T}^{m,\operatorname{\bf{a}}} for Xn\textbf{X}^{n}, n0n\in\mathbb{N}_{0}. Set

Yn:=Fn(Xn)={Fn(X):XXn}.\textbf{Y}^{n}:=F^{n}(\textbf{X}^{n})=\{F^{n}(X):X\in\textbf{X}^{n}\}.
Proof of Proposition 9.4.

Suppose TT is an uniformly mm-branching quasiconformal tree, and let the set Vn\textbf{V}^{n} be as in (9.1), for n0n\in\mathbb{N}_{0}, with δ(0,1)\delta\in(0,1) small enough for Proposition 9.3 to hold for edge-like decompositions Xn\textbf{X}^{n} of TT induced by Vn\textbf{V}^{n}. Note that V0=\textbf{V}^{0}=\emptyset, the sequence of sets (Vn)(\textbf{V}^{n}) consists of sets of branch points, is increasing, and the induced sequence (Xn)(\textbf{X}^{n}) is a subdivision of TT. Proposition 9.3(v) implies that Vn\textbf{V}^{n}\neq\emptyset for nn\in\mathbb{N}.

Following the proof of [BM22, Proposition 7.1], we can construct homeomorphisms Fn:T𝕋m,𝐚F^{n}:T\to\mathbb{T}^{m,\operatorname{\bf{a}}} for all n0n\in\mathbb{N}_{0} with the following properties:

  1. (A)

    For each n0n\in\mathbb{N}_{0} the map Fn:T𝕋m,𝐚F^{n}:T\to\mathbb{T}^{m,\operatorname{\bf{a}}} is a tile-homeomorphism for Xn\textbf{X}^{n}, i.e., FnF^{n} is a homeomorphism from TT onto 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} such that Fn(X)F^{n}(X) is of the form 𝕋wm,𝐚\mathbb{T}_{w}^{m,\operatorname{\bf{a}}}, with wAw\in A^{*}, for XXnX\in\textbf{X}^{n}.

  2. (B)

    For all nn\in\mathbb{N}, the maps Fn1F^{n-1} and FnF^{n} are compatible, in the sense that

    Fn1(X)=Fn(X),F^{n-1}(X)=F^{n}(X),

    for all XXn1X\in\textbf{X}^{n-1}.

  3. (C)

    For all nn\in\mathbb{N}, if XXn1X\in\textbf{X}^{n-1} and YXnY\in\textbf{X}^{n} with YXY\subset X, then

    L(Fn1(X))+1L(Fn(Y))L(Fn1(X))+N,\texttt{L}(F^{n-1}(X))+1\leq\texttt{L}(F^{n}(Y))\leq\texttt{L}(F^{n-1}(X))+N,

    where NN\in\mathbb{N} is independent of nn and XX.

  4. (D)

    For all nn\in\mathbb{N}, if XXn1X\in\textbf{X}^{n-1}, YXnY\in\textbf{X}^{n} with YXY\subset X and YXY\cap\partial X\neq\emptyset, then

    L(Fn(Y))=L(Fn1(X))+2.\texttt{L}(F^{n}(Y))=\texttt{L}(F^{n-1}(X))+2.

The construction of FnF^{n} for m3m\geq 3 is identical to that for m=3m=3 and we note that Lemma 2.2, Proposition 9.3 and Lemma 9.6 are used for the former.

It is enough to show that (Fn(Xn))(F^{n}(\textbf{X}^{n})) is a quasi-visual subdivision of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} isomorphic to (Xn)(\textbf{X}^{n}). Due to the choice of weights 𝐚(i)=12\operatorname{\bf{a}}(i)=\frac{1}{2} for all iAi\in A, this follows by the above properties similarly to the case m=3m=3 in [BM22, Proposition 7.1]. In particular, the fact that (Fn(Xn))(F^{n}(\textbf{X}^{n})) is a subdivision of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} and that the subdivisions (Xn)(\textbf{X}^{n}) and (Fn(Xn))(F^{n}(\textbf{X}^{n})) are isomorphic follow from properties (A) and (B) of the homeomorphisms Fn(Xn)F^{n}(\textbf{X}^{n}). To show that (Fn(Xn))(F^{n}(\textbf{X}^{n})) is a quasi-visual approximation of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}, the conditions (i)-(iv) in Definition 2.7 need to be verified. This follows by properties (C) and (D) of the homeomorphisms Fn(Xn)F^{n}(\textbf{X}^{n}), as well as the geodesicity of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} (instead of the quasiconvexity of the CSST used in[BM22, Proposition 7.1]). ∎

We finish this section by proving the second direction of Theorem 9.1.

Proof of Theorem 9.1.

Recall that we have fixed m{3,4,}m\in\{3,4,\dots\}, a uniformly mm-branching tree TT with diamT=1\operatorname{diam}{T}=1, and a weight 𝐚\operatorname{\bf{a}} such that 𝐚(i)=1/2\operatorname{\bf{a}}(i)=1/2 for all iAi\in A.

There exists δ>0\delta>0 small enough so that the subdivision (Xn)(\textbf{X}^{n}) induced by the sets Vn\textbf{V}^{n} as defined in (9.1) satisfies properties of Proposition 9.3. In particular, (Xn)(\textbf{X}^{n}) is a quasi-visual subdivision of TT. By Proposition 9.4 there exists a quasi-visual subdivision (Yn)(\textbf{Y}^{n}) of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} isomorphic to (Xn)(\textbf{X}^{n}). By Proposition 2.10 there exists a quasisymmetric homeomorphism F:T𝕋m,𝐚F:T\to\mathbb{T}^{m,\operatorname{\bf{a}}} that induces the isomorphism between (Xn)(\textbf{X}^{n}) and (Yn)(\textbf{Y}^{n}). ∎

10. Bi-Lipschitz and quasisymmetric embedding into 2\mathbb{R}^{2}

In this section we show the following proposition.

Proposition 10.1.

For all m2m\geq 2 there exists a weight 𝐚\operatorname{\bf{a}} such that 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} bi-Lipschitz embeds into 2\mathbb{R}^{2} with the embedded image being a quasiconvex subset of 2\mathbb{R}^{2}.

In conjunction with Theorem 9.1, Proposition 10.1 implies that all trees 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} quasisymmetrically embed in 2\mathbb{R}^{2} for all weights 𝐚\operatorname{\bf{a}}.

Corollary 10.2.

For any m{2,3,}m\in\{2,3,\dots\} and any weight 𝐚\operatorname{\bf{a}}, there exists a quasiconvex tree in 2\mathbb{R}^{2} that is quasisymmetric equivalent to 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

If m=2m=2, then by Proposition 6.1, 𝕋2,𝐚\mathbb{T}^{2,\operatorname{\bf{a}}} is a metric arc and by Proposition 5.1 and Corollary 5.8, it is a geodesic space of diameter 1. Therefore, 𝕋2,𝐚\mathbb{T}^{2,\operatorname{\bf{a}}} is isometric to the unit interval [0,1][0,1].

Fix for the rest of this section an integer m3m\geq 3. Let A={1,,m}A=\{1,\dots,m\}, let θ=π3m3\theta=\frac{\pi}{3m-3} and let 𝐚\operatorname{\bf{a}} be a weight with 𝐚(1)=𝐚(2)=1/2\operatorname{\bf{a}}(1)=\operatorname{\bf{a}}(2)=1/2, 𝐚(j)=12sinθ\operatorname{\bf{a}}(j)=\frac{1}{2}\sin\theta for all j{3,,m}j\in\{3,\dots,m\} and 𝐚(j)=0\operatorname{\bf{a}}(j)=0 for jmj\geq m.

Let R2R\subset\mathbb{R}^{2} be the convex hull of the 4 points {±12,±i2tanθ}\{\pm\frac{1}{2},\pm\frac{i}{2}\tan\theta\} and let {ψj:}j=1m\{\psi_{j}:\mathbb{C}\to\mathbb{C}\}_{j=1}^{m} be contracting similarities given (in complex notation) by

ψ1(z)=𝐚(1)(z¯12),ψ2(z)=𝐚(2)(z+12),\psi_{1}(z)=\operatorname{\bf{a}}(1)(\overline{z}-\tfrac{1}{2}),\quad\psi_{2}(z)=\operatorname{\bf{a}}(2)(z+\tfrac{1}{2}),

and

ψj(z)=𝐚(j)(z+12)ei(3j6)θfor j{3,,m}.\psi_{j}(z)=\operatorname{\bf{a}}(j)\left(z+\tfrac{1}{2}\right)e^{i(3j-6)\theta}\quad\text{for $j\in\{3,\dots,m\}$}.

See Figure 4 for the case m=4m=4.

Refer to caption
Figure 4. The images of RR under similarities ψ1,,ψ4\psi_{1},\dots,\psi_{4}. Here we have chosen m=4m=4.

For each w=j1jnAw=j_{1}\cdots j_{n}\in A^{*}, define

ψw:=ψj1ψj2ψjn\psi_{w}:=\psi_{j_{1}}\circ\psi_{j_{2}}\circ\cdots\circ\psi_{j_{n}}

with the convention that ψϵ\psi_{\epsilon} is the identity map on \mathbb{C}. Note that the scaling factor of ψw\psi_{w} is Δ𝐚(w)=𝐚(j1)𝐚(jn)\Delta_{\operatorname{\bf{a}}}(w)=\operatorname{\bf{a}}(j_{1})\cdots\operatorname{\bf{a}}(j_{n}). For each wAw\in A^{*}, set Rw=ψw(R)R_{w}=\psi_{w}(R). It is elementary to show that the choice of the angle θ\theta implies that the rhombuses RjR_{j}, jAj\in A, lie in RR and are well separated in the following sense.

Lemma 10.3.

For each distinct j,jAj,j^{\prime}\in A, RjRR_{j}\subset R, RjRj={0}R_{j}\cap R_{j^{\prime}}=\{0\}, and there exists a double cone centered at 0 and of angle θ\theta that separates the interior of RjR_{j} from the interior of RjR_{j^{\prime}}.

By Lemma 10.3, ψw(R)R\psi_{w}(R)\subset R for all wAw\in A^{*}. By [Hut81], the iterated function system {ψj}jA\{\psi_{j}\}_{j\in A} has an attractor 𝒯m\mathcal{T}^{m}. That is, 𝒯m\mathcal{T}^{m} is the unique nonempty compact set such that 𝒯m=jAψj(𝒯m)\mathcal{T}^{m}=\bigcup_{j\in A}\psi_{j}(\mathcal{T}^{m}). For each wAw\in A^{*}, we set 𝒯wm=ψw(𝒯m)\mathcal{T}^{m}_{w}=\psi_{w}(\mathcal{T}^{m}). See Figure 5 for the case m=4m=4.111Figure 5 was generated using the IFS construction kit in https://adelapo.github.io/ifs.html.

Refer to caption
Figure 5. The attractor 𝒯4\mathcal{T}^{4} of the iterated function system {ψ1,ψ2,ψ3,ψ4}\{\psi_{1},\psi_{2},\psi_{3},\psi_{4}\}.

The proof of the following is almost identical with that in [BT21, Proposition 4.2], hence, we omit the details.

Lemma 10.4.

Let J0=[12,12]J_{0}=[-\frac{1}{2},\frac{1}{2}]\subset\mathbb{C}. For n0n\geq 0 define

Jn=wAnψw(J0)andKn=wAnψw(R).J_{n}=\bigcup_{w\in A^{n}}\psi_{w}(J_{0})\quad\text{and}\quad K_{n}=\bigcup_{w\in A^{n}}\psi_{w}(R).

Then the sets JnJ_{n} and KnK_{n} are compact for all integers n0n\geq 0 and

JnJn+1𝒯mKn+1Kn.J_{n}\subset J_{n+1}\subset\mathcal{T}^{m}\subset K_{n+1}\subset K_{n}.

Since J0𝒯mRJ_{0}\subset\mathcal{T}^{m}\subset R and diam(J0)=diam(R)=1\operatorname{diam}(J_{0})=\operatorname{diam}(R)=1, we have diam(𝒯m)=1\operatorname{diam}(\mathcal{T}^{m})=1. Therefore, diam𝒯wm=Δ(w)\operatorname{diam}{\mathcal{T}_{w}^{m}}=\Delta(w) for all wAw\in A^{*}.

Observe that 0=ψ1(12)=ψ2(12)==ψm(12)0=\psi_{1}(\frac{1}{2})=\psi_{2}(-\frac{1}{2})=\cdots=\psi_{m}(-\frac{1}{2}), so 0𝒯kmRk0\in\mathcal{T}_{k}^{m}\subset R_{k} for all kAk\in A. Therefore, if j,j{1,,m}j,j^{\prime}\in\{1,\dots,m\} are distinct, Lemma 10.3 implies that 𝒯km𝒯lm={0}\mathcal{T}_{k}^{m}\cap\mathcal{T}_{l}^{m}=\{0\}.

Lemma 10.5.

For all n0n\in\mathbb{N}_{0}, JnJ_{n} is connected.

Proof.

The proof is done by induction. For n=0n=0 it’s clear. Assume that JnJ_{n} is connected for n1n\geq 1. Then,

Jn+1=wAn+1ψw(J0)=i=1mwAnψiw(J0)=i=1mψi(wAnψw(J0))=i=1mψi(Jn).J_{n+1}=\bigcup_{w\in A^{n+1}}\psi_{w}(J_{0})=\bigcup_{i=1}^{m}\bigcup_{w^{\prime}\in A^{n}}\psi_{iw^{\prime}}(J_{0})=\bigcup_{i=1}^{m}\psi_{i}\big{(}\bigcup_{w^{\prime}\in A^{n}}\psi_{w^{\prime}}(J_{0})\big{)}=\bigcup_{i=1}^{m}\psi_{i}(J_{n}).

The sets ψi(Jn)\psi_{i}(J_{n}) are connected by continuity of ψi\psi_{i} for each i{1,,m}i\in\{1,\dots,m\}. Moreover, 0ψi(Jn)0\in\psi_{i}(J_{n}) for all ii, since ψi(J0)ψi(Jn)\psi_{i}(J_{0})\subset\psi_{i}(J_{n}) by Lemma 10.4 and 0ψi(J0)0\in\psi_{i}(J_{0}) by the above observation. Hence, Jn+1J_{n+1} is connected. ∎

Lemma 10.6.

The set 𝒯m\mathcal{T}^{m} is a a quasiconvex metric tree in \mathbb{C}.

Proof.

That 𝒯m\mathcal{T}^{m} is a metric tree, can be proved using verbatim the same techniques as in [BT21, Proposition 1.4] (see also [BT21, Lemma 4.3, Lemma 4.4 and Lemma 4.5]). The fact that 𝒯m\mathcal{T}^{m} is quasiconvex follows from the existence of double cones separating distinct Rj,RjR_{j},R_{j^{\prime}} in Lemma 10.3, and applying the techniques of [BT21, Proposition 1.4] verbatim. ∎

Define a geodesic metric ρ\rho on 𝒯m\mathcal{T}^{m} by setting ρ(a,b)=length(γ)\rho(a,b)=\text{length}(\gamma) for a,b𝒯ma,b\in\mathcal{T}^{m}, where γ\gamma is the unique arc in 𝒯m\mathcal{T}^{m} joining aa and bb. Then, by Lemma 10.6, the metric space (𝒯m,ρ)(\mathcal{T}^{m},\rho) is bi-Lipschitz equivalent to 𝒯m\mathcal{T}^{m} (equipped with the Euclidean metric).

Lemma 10.7.

For each n0n\geq 0, the metric space (Jn,ρ)(J_{n},\rho) isometrically embeds into 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

Proof.

First, for each integer n0n\geq 0, JnJ_{n} is closed (since it’s compact by Lemma 10.4) and connected by Lemma 10.5; hence, as a subset of 𝒯m\mathcal{T}^{m}, it is itself a geodesic metric tree [BT21, Lemma 3.3].

Let γ:I𝕋m,𝐚\gamma:I\rightarrow\mathbb{T}^{m,\operatorname{\bf{a}}} be the unique arc connecting [1()][1^{(\infty)}] with [2()][2^{(\infty)}], where II denotes the unit interval. Define for each integer n0n\geq 0,

𝒥n:=wAnϕw(γ(I))𝕋m,𝐚.\mathcal{J}_{n}:=\bigcup_{w\in A^{n}}\phi_{w}(\gamma(I))\subset\mathbb{T}^{m,\operatorname{\bf{a}}}.

We claim that for all integers n0n\geq 0 there exists an isometric map fn:Jn𝒥nf_{n}:J_{n}\to\mathcal{J}_{n} satisfying fn(12)=[1()]f_{n}(-\frac{1}{2})=[1^{(\infty)}] and fn(12)=[2()]f_{n}(\frac{1}{2})=[2^{(\infty)}]. We prove the claim by induction on nn.

The case n=0n=0 is trivial, since (γ(I),ρ𝐚)(\gamma(I),\rho_{\operatorname{\bf{a}}}) and (J0,ρ)(J_{0},\rho) are both isometric to II, due to 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} being geodesic.

Assume now that for some integer n0n\geq 0 there exists an isometric homeomorphism fn:Jn𝒥nf_{n}:J_{n}\to\mathcal{J}_{n} with fn(12)=[1()]f_{n}(-\frac{1}{2})=[1^{(\infty)}] and fn(12)=[2()]f_{n}(\frac{1}{2})=[2^{(\infty)}]. For each jAj\in A, set Jn,j=ψj(Jn)J_{n,j}=\psi_{j}(J_{n}) and 𝒥n,j=ϕj(𝒥n)\mathcal{J}_{n,j}=\phi_{j}(\mathcal{J}_{n}). Since ψj\psi_{j} has the same scaling factor as ϕj\phi_{j}, we have that the map

gn,j=ϕjfn(ψj|Jn,j)1:Jn,j𝒥n,jg_{n,j}=\phi_{j}\circ f_{n}\circ(\psi_{j}|J_{n,j})^{-1}:J_{n,j}\to\mathcal{J}_{n,j}

is an isometric homeomorphism. Note that 12Jn,1-\frac{1}{2}\in J_{n,1} and 12Jn,2\frac{1}{2}\in J_{n,2} since 12=ψ1(ψ1(n)(12))-\frac{1}{2}=\psi_{1}(\psi_{1^{(n)}}(-\frac{1}{2})) and 12=ψ2(ψ2(n)(12))\frac{1}{2}=\psi_{2}(\psi_{2^{(n)}}(\frac{1}{2})). Hence,

gn,1(12)=gn,1(ψ1(12))=ϕ1(fn(12))=ϕ1([1()])=[1()].g_{n,1}(-\tfrac{1}{2})=g_{n,1}(\psi_{1}(-\tfrac{1}{2}))=\phi_{1}(f_{n}(-\tfrac{1}{2}))=\phi_{1}([1^{(\infty)}])=[1^{(\infty)}].

Similarly, we get gn,2(12)=[2()]g_{n,2}(\frac{1}{2})=[2^{(\infty)}]. Note that by 12,12Jn-\frac{1}{2},\frac{1}{2}\in J_{n} for all nn, we also have 0Jn,j0\in J_{n,j} for all jAj\in A. As a result, for distinct j,jAj,j^{\prime}\in A we have

(10.1) {0}Jn,jJn,j=ψj(Jn)ψj(Jn)ψj(𝒯m)ψj(𝒯m){0}.\{0\}\subset J_{n,j}\cap J_{n,j^{\prime}}=\psi_{j}(J_{n})\cap\psi_{j^{\prime}}(J_{n})\subset\psi_{j}(\mathcal{T}^{m})\cap\psi_{j^{\prime}}(\mathcal{T}^{m})\subset\{0\}.

Hence the sets Jn,jJ_{n,j} have only 0 as common point. Similarly, the sets 𝒥n,j\mathcal{J}_{n,j} have only [12()][12^{(\infty)}] as common point (see Lemma 6.2). Observe that

Jn+1=wAn+1ψw(J0)=jAwAnψj(ψw(J0))=jAJn,j.J_{n+1}=\bigcup_{w\in A^{n+1}}\psi_{w}(J_{0})=\bigcup_{j\in A}\bigcup_{w\in A^{n}}\psi_{j}(\psi_{w}(J_{0}))=\bigcup_{j\in A}J_{n,j}.

Similarly, 𝒥n+1=wAn+1ϕw(γ(I))=j=1m𝒥n,j\mathcal{J}_{n+1}=\bigcup_{w\in A^{n+1}}\phi_{w}(\gamma(I))=\bigcup_{j=1}^{m}\mathcal{J}_{n,j}.

Define fn+1:(Jn+1,ρ)(𝒥n+1,ρ𝐚)f_{n+1}:(J_{n+1},\rho)\to(\mathcal{J}_{n+1},\rho_{\operatorname{\bf{a}}}) by fn+1|Jn,j=gn,jf_{n+1}|J_{n,j}=g_{n,j}. By (10.1) the map fn+1f_{n+1} is well defined, since fn+1(0)=gn,j(0)=[12()]f_{n+1}(0)=g_{n,j}(0)=[12^{(\infty)}] for all j{1,,m}j\in\{1,\dots,m\}, fn+1(12)=gn,1(12)=[1()]f_{n+1}(-\frac{1}{2})=g_{n,1}(-\frac{1}{2})=[1^{(\infty)}], and fn+1(12)=gn,2(12)=[2()]f_{n+1}(\frac{1}{2})=g_{n,2}(\frac{1}{2})=[2^{(\infty)}].

To finish the inductive step, we show that fn+1f_{n+1} is an isometry. To this end, fix x,yJn+1x,y\in J_{n+1}. The case x,yJn,jx,y\in J_{n,j} for some jAj\in A follows by the fact that each gn,jg_{n,j} is an isometry. Assume now that xJn,jx\in J_{n,j} and yJn,jy\in J_{n,j^{\prime}} for two distinct j,jAj,j^{\prime}\in A. By the geodesicity of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} and (Jn+1,ρ)(J_{n+1},\rho),

d𝐚(fn+1(x),fn+1(y))\displaystyle d_{\mathbf{a}}(f_{n+1}(x),f_{n+1}(y)) =d𝐚(gn,j(x),[12()])+d𝐚([12()],gn,j(y))\displaystyle=d_{\mathbf{a}}(g_{n,j}(x),[12^{(\infty)}])+d_{\mathbf{a}}([12^{(\infty)}],g_{n,j^{\prime}}(y))
=ρ(x,0)+ρ(0,y)\displaystyle=\rho(x,0)+\rho(0,y)
=ρ(x,y).\displaystyle=\rho(x,y).\qed
Proof of Proposition 10.1.

Recall the definitions of sets (Jn)n0(J_{n})_{n\in\mathbb{N}_{0}}, (𝒥n)n0(\mathcal{J}_{n})_{n\in\mathbb{N}_{0}}, and isometric homeomorphisms fn:(Jn,ρ)𝒥nf_{n}:(J_{n},\rho)\to\mathcal{J}_{n} from the proof of Lemma 10.7.

By Lemma 10.4, each geodesic tree (Jn,ρ)(J_{n},\rho) is a subset of (𝒯m,ρ)(\mathcal{T}^{m},\rho). By Lemma 10.7 we have

distH(𝕋m,𝐚,𝒥n)=sup[v]𝕋m,𝐚d𝐚([v],𝒥n)\displaystyle\operatorname{dist}_{H}(\mathbb{T}^{m,\operatorname{\bf{a}}},\mathcal{J}_{n})=\sup_{[v]\in\mathbb{T}^{m,\operatorname{\bf{a}}}}d_{\operatorname{\bf{a}}}([v],\mathcal{J}_{n}) sup[v]𝕋m,𝐚infwAnd𝐚([v],[w12()])\displaystyle\leq\sup_{[v]\in\mathbb{T}^{m,\operatorname{\bf{a}}}}\inf_{w\in A^{n}}d_{\operatorname{\bf{a}}}([v],[w12^{(\infty)}])
supwAndiam𝕋wm,𝐚\displaystyle\leq\sup_{w\in A^{n}}\operatorname{diam}{\mathbb{T}^{m,\operatorname{\bf{a}}}_{w}}
2n\displaystyle\leq 2^{-n}

where in the third inequality we used the fact that {[w12()]:wAn}𝒥n\{[w12^{(\infty)}]:w\in A^{n}\}\subset\mathcal{J}_{n}.

Therefore, we have

(10.2) limndistGH(𝕋m,𝐚,(Jn,ρ))limndistH(𝕋m,𝐚,𝒥n)=0.\lim_{n\to\infty}\operatorname{dist}_{GH}(\mathbb{T}^{m,\operatorname{\bf{a}}},(J_{n},\rho))\leq\lim_{n\to\infty}\operatorname{dist}_{H}(\mathbb{T}^{m,\operatorname{\bf{a}}},\mathcal{J}_{n})=0.

Fix n1n\geq 1 and x𝒯mx\in\mathcal{T}^{m}. Then there exists w(n)Anw(n)\in A^{n} such that x𝒯w(n)mx\in\mathcal{T}^{m}_{w(n)}. By Lemma 10.6, 𝒯m\mathcal{T}^{m} is LL-bi-Lipschitz homeomorphic to (𝒯m,ρ)(\mathcal{T}^{m},\rho) for some L1L\geq 1. Therefore,

ρ(x,ψw(n)(0))diamρ𝒯w(n)mLdiam𝒯w(n)mLΔ𝐚(w(n))L2n,\displaystyle\rho(x,\psi_{w(n)}(0))\leq\operatorname{diam}_{\rho}{\mathcal{T}^{m}_{w(n)}}\leq L\operatorname{diam}{\mathcal{T}^{m}_{w(n)}}\leq L\Delta_{\operatorname{\bf{a}}}(w(n))\leq L2^{-n},

which implies that

(10.3) limndistH((𝒯m,ρ),(Jn,ρ))limnsupx𝒯minfwAnρ(x,ψw(0))=0.\lim_{n\to\infty}\operatorname{dist}_{H}((\mathcal{T}^{m},\rho),(J_{n},\rho))\leq\lim_{n\to\infty}\sup_{x\in\mathcal{T}^{m}}\inf_{w\in A^{n}}\rho(x,\psi_{w}(0))=0.

Hence, by triangle inequality and (10.2), (10.3), the Gromov-Hausdorff distance of (𝒯m,ρ)(\mathcal{T}^{m},\rho) and 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is zero, which means that they are isometric to each other. As a result, 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} is bi-Lipschitz homeomorphic to 𝒯m\mathcal{T}^{m} (with the Euclidean metric). ∎

11. Quasisymmetric embeddings

In this section, we show the following proposition.

Proposition 11.1.

If TT is a quasiconformal tree with Val(T)=n\operatorname{Val}(T)=n and uniform branch separation, then TT quasisymmetrically embeds into a uniformly nn-branching quasiconformal tree.

Towards this end, we need to define a notion of “gluing” for geodesic metric spaces, and establish certain properties.

Definition 11.2.

Let (X,dX,𝐱I)(X,d_{X},\mathbf{x}_{I}), (Yi,dYi,yi)(Y_{i},d_{Y_{i}},y_{i}) be pointed geodesic metric spaces, where II is a countable subset of \mathbb{N}, xiXx_{i}\in X, yiYiy_{i}\in Y_{i} for all iIi\in I, and 𝐱I:=(x1,x2,)\mathbf{x}_{I}:=(x_{1},x_{2},\dots). In what follows, we identify X×{0}X\times\{0\} with XX, and for given iIi\in I we identify Yi×{i}Y_{i}\times\{i\} with YiY_{i}.

Define a pseudo-metric on (X×{0})iI(Yi×{i})=XiIYi(X\times\{0\})\cup\bigcup_{i\in I}(Y_{i}\times\{i\})=X\cup\bigcup_{i\in I}Y_{i} by

ρ(z,w)={dX(z,w)if z,wXdYi(z,w)if z,wYi for some iIdX(z,xi)+dYi(yi,w)if zXwYi for some iI,dX(w,xi)+dYi(yi,z)if zYiwX for some iI,dYi(z,yi)+dX(xi,xj)+dYj(yj,w)if zYiwYj for distinct i,jI.\rho(z,w)=\begin{cases}d_{X}(z,w)&\text{if $z,w\in X$}\\ d_{Y_{i}}(z,w)&\text{if $z,w\in Y_{i}$ for some $i\in I$}\\ d_{X}(z,x_{i})+d_{Y_{i}}(y_{i},w)&\text{if $z\in X$, $w\in Y_{i}$ for some $i\in I$},\\ d_{X}(w,x_{i})+d_{Y_{i}}(y_{i},z)&\text{if $z\in Y_{i}$, $w\in X$ for some $i\in I$},\\ d_{Y_{i}}(z,y_{i})+d_{X}(x_{i},x_{j})+d_{Y_{j}}(y_{j},w)&\text{if $z\in Y_{i}$, $w\in Y_{j}$ for distinct $i,j\in I$}.\\ \end{cases}

The pseudometric ρ\rho gives rise to a metric dd on the quotient space (XiIYi)/(X\cup\bigcup_{i\in I}Y_{i})/\sim where w1w2w_{1}\sim w_{2} if ρ(w1,w2)=0\rho(w_{1},w_{2})=0. Define now the geodesic gluing of XX and YiY_{i} (at 𝐱I\mathbf{x}_{I} and yiy_{i})

(X,dX,𝐱I)iI(Yi,dYi,yi):=((XiIYi)/,d).(X,d_{X},\mathbf{x}_{I})\bigvee_{i\in I}(Y_{i},d_{Y_{i}},y_{i}):=((X\cup\bigcup_{i\in I}Y_{i})/\sim,d).

It is easy to see that the geodesic gluing of geodesic spaces is itself a geodesic space. Intuitively, we are simply identifying xix_{i} with yiy_{i} for all iIi\in I. It is clear that XX and YiY_{i} isometrically embed into the above geodesic gluing. We make the convention that points of XX and YiY_{i} are treated as points of a “subspace” in the geodesic gluing metric space, in order to ease the notation.

Lemma 11.3.

Let XX, YiY_{i} be geodesic metric trees, for all iIi\in I. Assume that, either cardI<\operatorname{card}I<\infty, or cardI=\operatorname{card}I=\infty and diamYi0\operatorname{diam}Y_{i}\to 0 as ii\to\infty. Then, the geodesic gluing (X,dX,𝐱I)iI(Yi,dYi,yi)(X,d_{X},\mathbf{x}_{I})\bigvee_{i\in I}(Y_{i},d_{Y_{i}},y_{i}) is a geodesic metric tree.

Proof.

Denote by (Z,d)(Z,d) the geodesic gluing of XX and YiY_{i} at 𝐱I\mathbf{x}_{I} and yiy_{i}. The fact that (Z,d)(Z,d) is a geodesic metric space follows by Definition 11.2. We prove the case where II is infinite, and the case cardI<\operatorname{card}I<\infty is similar. We first show that ZZ is compact. Let (zn)n(z_{n})_{n\in\mathbb{N}} be a sequence in ZZ. If znXz_{n}\in X for infinitely many nn, then there is an accumulation point in XX, and, hence, a convergent subsequence in XZX\subset Z. Similarly, if znYiz_{n}\in Y_{i} for some iIi\in I and infinitely many nn, then there is an accumulation point in YiY_{i}, and a convergent subsequence in YiY_{i}. For the last case, suppose that znYnz_{n}\in Y_{n} for all nn\in\mathbb{N}, by passing to a subsequence if necessary. Consider the sequence (xn)n(x_{n})_{n\in\mathbb{N}} in ZZ, where xnx_{n} is the term of 𝐱I\mathbf{x}_{I} at which we glue the metric space YnY_{n}. Since XX is compact, there exists a convergent subsequence (xnj)j(x_{n_{j}})_{j\in\mathbb{N}} that converges to a point zXz\in X. We have

d(znj,z)d(znj,xnj)+d(xnj,z)=d(znj,ynj)+d(ynj,z)diamYnj+d(xnj,z).d(z_{n_{j}},z)\leq d(z_{n_{j}},x_{n_{j}})+d(x_{n_{j}},z)=d(z_{n_{j}},y_{n_{j}})+d(y_{n_{j}},z)\leq\operatorname{diam}Y_{n_{j}}+d(x_{n_{j}},z).

By letting jj\to\infty, it follows that znjzZz_{n_{j}}\to z\in Z. Therefore, the arbitrary sequence (zn)n(z_{n})_{n\in\mathbb{N}} in ZZ always has a convergent subsequence, proving that ZZ is compact.

We show that there exists a unique arc between any two distinct points in ZZ; hence, ZZ is connected. Let x,yXx,y\in X be distinct. There exists a unique arc [x,y][x,y] in XX, which is also an arc in ZZ. Assume towards contradiction that there exists another arc α[x,y]\alpha\neq[x,y] in ZZ that connects xx and yy. There is a fixed iIi\in I for which α(Yi{yi})\alpha\cap(Y_{i}\setminus\{y_{i}\})\neq\emptyset, because otherwise αX\alpha\subset X, which would contradict the uniqueness of [x,y][x,y] in XX. Let zα(Yi{yi})z\in\alpha\cap(Y_{i}\setminus\{y_{i}\}). We can write XYi={yi}X\cap Y_{i}=\{y_{i}\}, due to xiyix_{i}\sim y_{i} and an abuse of notation that we follow henceforth, identifying xix_{i} with yiy_{i} as points in ZZ. This implies that

α=[x,z](z,y]=[x,yi](yi,z](z,yi](yi,y],\alpha=[x,z]\cup(z,y]=[x,y_{i}]\cup(y_{i},z]\cup(z,y_{i}]\cup(y_{i},y],

which shows that α\alpha is not an arc. If x,yYix,y\in Y_{i} for some iIi\in I, the proof is similar. Suppose xXx\in X and yYiy\in Y_{i} for some iIi\in I. The arc β=[x,yi](yi,y]\beta=[x,y_{i}]\cup(y_{i},y] connects xx and yy, but due to XYi={yi}X\cap Y_{i}=\{y_{i}\}, any other possible arc connecting these two points must intersect yiy_{i}. Similarly to the previous case, it can be shown that [x,xi]X[x,x_{i}]\subset X and [yi,y]Y[y_{i},y]\subset Y are unique arcs in ZZ. Hence, by the identification [x,xi]=[x,yi][x,x_{i}]=[x,y_{i}] in ZZ, we have that β\beta is the unique arc in ZZ that connects xx and yy. Suppose that xYix\in Y_{i} and yYjy\in Y_{j} for iji\neq j. It can be shown similarly that the arc [x,yi](yi,yj](yj,y][x,y_{i}]\cup(y_{i},y_{j}]\cup(y_{j},y], which connects xx and yy in ZZ, is unique.

The geodesicity of ZZ implies that ZZ is locally connected. Therefore, ZZ is a geodesic metric tree. ∎

Recall the notion of height HT(p)H_{T}(p) from (1.1). For most of this section, we need to replace it by a new notion. Given a metric tree TT, a branch point pp in TT, and a branch BB of TT at pp, define hT(p,B)h_{T}(p,B) to be the maximal distance from pp within BB, i.e.,

(11.1) hT(p,B):=maxxBd(x,p).h_{T}(p,B):=\max_{x\in B}d(x,p).

We need the following relation between branches of distinct branch points, in order to establish certain properties involving the above maximal distance.

Lemma 11.4.

Let pTp\in T, BTi(p)B_{T}^{i}(p) be a branch of TT at pp, for some i{1,,Val(p)}i\in\{1,\dots,\operatorname{Val}(p)\}, and qBTi(p)q\in B_{T}^{i}(p) that is not a leaf. If BTj(q)B_{T}^{j}(q) is a branch of qq such that pBTj(q)p\in B_{T}^{j}(q), j{1,,Val(q)}j\in\{1,\dots,\operatorname{Val}(q)\}, then BTk(q)BTi(p)B_{T}^{k}(q)\subset B_{T}^{i}(p) for all kjk\neq j.

Proof.

Note that if BB is a connected component of T{p}T\setminus\{p\}, then B={p}\partial B=\{p\} by [BT21, Lemma 3.2(ii)]. If pp is a leaf, then the result holds due to BTi(p)=T{p}B_{T}^{i}(p)=T\setminus\{p\} and pBTk(q)p\notin B_{T}^{k}(q) for all kjk\neq j. Suppose that pp is not a leaf. Let zBTk(q)z\in B_{T}^{k}(q) for kjk\neq j. Since pBTj(q)p\in B_{T}^{j}(q), it follows that q[z,p]q\in[z,p] by [BT21, Lemma 3.2(iii)]. If zBT(p)z\in B_{T}^{\ell}(p) for i\ell\neq i, then the arc [z,p)[z,p) lies in BT(p)B_{T}^{\ell}(p). But [z,p]=[z,q](q,p][z,p]=[z,q]\cup(q,p], and due to qBTi(p)q\in B_{T}^{i}(p) we have that [z,p]BTi(p)[z,p]\cap B_{T}^{i}(p)\neq\emptyset. This contradicts the fact that BTi(p)B_{T}^{i}(p) and BT(p)B_{T}^{\ell}(p) are distinct connected components of T{p}T\setminus\{p\}. Hence, zBTi(p)z\in B_{T}^{i}(p). ∎

If TT is a geodesic tree, then the maximum in (11.1) occurs at leaves of TT, i.e.,

(11.2) hT(p,B)=max{d(x,p):xleaf on B}h_{T}(p,B)=\max\{d(x,p):x\,\,\text{leaf on }\,B\}

Indeed, assume towards contradiction that TT is a geodesic tree, and that the maximum in (11.1) occurs at a point xBx\in B that is not a leaf. This implies that there exists a branch BT(x)B_{T}(x) of TT at xx that does not contain pp. By Lemma 11.4, we have BT(x)BB_{T}(x)\subset B. Let qBT(x)q\in B_{T}(x). Since qq and pp lie on different branches at xx, we have x[p,q]x\in[p,q] by [BT21, Lemma 3.2(iii)]. Therefore, by geodecisity we have

d(p,q)=d(p,x)+d(x,q).d(p,q)=d(p,x)+d(x,q).

This implies that d(p,q)>d(p,x)d(p,q)>d(p,x) for qBT(p)q\in B_{T}(p), which contradicts the assumption. Hence, the maximum cannot be achieved at xx, unless xx is a leaf.

Let pTp\in T be a branch point, and let BT1(p),BT2(p),B_{T}^{1}(p),B_{T}^{2}(p),\dots be an enumeration of the branches of TT at pp with hT(p,BT1(p))hT(p,BT2(p)).h_{T}(p,B_{T}^{1}(p))\geq h_{T}(p,B_{T}^{2}(p))\geq\dots. We define the new height of pp to be

(11.3) hT(p)=hT(p,BT3(p)).h_{T}(p)=h_{T}(p,B_{T}^{3}(p)).

To simplify the notation henceforth, we also denote hT(p,BTj(p))h_{T}(p,B_{T}^{j}(p)) by hT(p,Bj)h_{T}(p,B_{j}), for all jj.

The two notions of height are related in a way that allows us to interchange them in certain cases. We say that a metric tree TT is uniformly nn-branching with respect to the new height hTh_{T}, for some integer n3n\geq 3, if every branch point of TT has valence nn, and TT has uniform branch growth, uniformly relatively separated branch points, and uniformly relative dense branch points with respect to hTh_{T}, i.e., Definitions 1.7, 1.2, and 1.6 are true for TT with HT(p)H_{T}(p) and diamBTi(p)\operatorname{diam}B_{T}^{i}(p) replaced by hT(p)h_{T}(p) and hT(p,Bi(p))h_{T}(p,B_{i}(p)), respectively. The following lemma shows that the new height and the one defined at (1.1) are comparable at every branch point.

Lemma 11.5.

Given a branch point pTp\in T,

(11.4) hT(p)HT(p),h_{T}(p)\simeq H_{T}(p),

where HT(p)H_{T}(p) is defined at (1.1), and the constant C()C(\simeq) is independent of the metric tree TT and the branch point pp.

Moreover, TT is uniformly nn-branching with respect to hTh_{T}, if, and only if, TT is uniformly nn-branching with respect to HTH_{T}.

Proof.

Fix a branch point pTp\in T. We have two different enumerations of the branches of TT at pp. We denote by B1H,B2H,B_{1}^{H},B_{2}^{H},\dots the enumeration of the branches at pp in non-increasing order with respect to their diameters, and by B1h,B2h,B_{1}^{h},B_{2}^{h},\dots the enumeration of the branches of pp in non-increasing order with respect to the maximal distances defined in (11.2). We need to show that diamB3HhT(p,B3h)\operatorname{diam}B_{3}^{H}\simeq h_{T}(p,B_{3}^{h}). Fix a branch BkhB_{k}^{h} at pp. If x,yBkhx,y\in B_{k}^{h}, then d(x,y)d(x,p)+d(p,y)2hT(p,Bkh)d(x,y)\leq d(x,p)+d(p,y)\leq 2h_{T}(p,B_{k}^{h}), which implies that diamBkh2hT(p,Bkh)\operatorname{diam}B_{k}^{h}\leq 2h_{T}(p,B_{k}^{h}). On the other hand, if xBkhx\in B_{k}^{h}, then d(x,p)diamBkh¯=diamBkhd(x,p)\leq\operatorname{diam}\overline{B_{k}^{h}}=\operatorname{diam}B_{k}^{h}. Therefore, we have shown that

(11.5) diamBkhhT(p,Bkh).\operatorname{diam}B_{k}^{h}\simeq h_{T}(p,B_{k}^{h}).

Set B3h=BiHB_{3}^{h}=B_{i}^{H} and B3H=BjhB_{3}^{H}=B_{j}^{h} for some i,j{1,,Val(p)}i,j\in\{1,\dots,\operatorname{Val}(p)\}. The rest of the proof is a case study on i,ji,j.

If i=3i=3, then the heights are comparable by (11.5).

If i{1,2}i\in\{1,2\}, we have

diamB3HdiamBiH=diamB3hhT(p).\operatorname{diam}B_{3}^{H}\leq\operatorname{diam}B_{i}^{H}=\operatorname{diam}B_{3}^{h}\simeq h_{T}(p).

For the other side of (11.4), note that there is integer k3k\geq 3 such that either B1h=BkHB_{1}^{h}=B_{k}^{H}, or B2h=BkHB_{2}^{h}=B_{k}^{H}. Suppose that B1h=BkHB_{1}^{h}=B_{k}^{H}, and the proof in the other case is similar. We then have

diamB3HdiamBkH=diamB1hhT(p,B1h)hT(p),\operatorname{diam}B_{3}^{H}\geq\operatorname{diam}B_{k}^{H}=\operatorname{diam}B_{1}^{h}\simeq h_{T}(p,B_{1}^{h})\geq h_{T}(p),

and (11.4) follows in this case.

Suppose that i>3i>3. We have that

diamB3HdiamBiH=diamB3hhT(p).\operatorname{diam}B_{3}^{H}\geq\operatorname{diam}B_{i}^{H}=\operatorname{diam}B_{3}^{h}\simeq h_{T}(p).

For the other side of the inequality, if j3j\geq 3 then

diamB3H=diamBjhhT(p,Bjh)hT(p).\operatorname{diam}B_{3}^{H}=\operatorname{diam}B_{j}^{h}\simeq h_{T}(p,B_{j}^{h})\leq h_{T}(p).

If j{1,2}j\in\{1,2\}, then either B1H=BhB_{1}^{H}=B_{\ell}^{h}, or B2H=BhB_{2}^{H}=B_{\ell}^{h}, for some integer 3\ell\geq 3. Suppose that B1H=B1hB_{1}^{H}=B_{\ell_{1}}^{h}, and the proof in the other case is similar. It follows that

diamB3HdiamB1H=diamB1hhT(p,B1h)hT(p).\operatorname{diam}B_{3}^{H}\leq\operatorname{diam}B_{1}^{H}=\operatorname{diam}B_{\ell_{1}}^{h}\simeq h_{T}(p,B_{\ell_{1}}^{h})\leq h_{T}(p).

Therefore (11.4) holds.

It remains to show that the uniform nn-branching property with respect to hTh_{T} is equivalent to that with respect to HTH_{T}. We prove one direction, namely that the uniform nn-branching property with respect to hTh_{T} implies that with respect to HTH_{T}, and the other direction is similar. Suppose that TT is uniformly nn-branching with respect to hTh_{T}. The uniform branch separation and uniform density of branch points with respect to HTH_{T} follow immediately by (11.4). It remains to show the uniform branch growth of TT with respect to HTH_{T}. We assume that Val(T)=n4\operatorname{Val}(T)=n\geq 4, since the case Val(T)=3\operatorname{Val}(T)=3 is trivial. Fix a branch point pTp\in T with Val(T,p)=m{4,,n}\operatorname{Val}(T,p)=m\in\{4,\dots,n\}. We need to show that diamBmHdiamB3H\operatorname{diam}B_{m}^{H}\gtrsim\operatorname{diam}B_{3}^{H}. Let BmH=BihB_{m}^{H}=B_{i}^{h} for some i{1,,m}i\in\{1,\dots,m\}. If i{1,2}i\in\{1,2\}, then

diamBmH=diamBihhT(p,Bih)hT(p,B3h)=hT(p)HT(p)=diamB3H.\operatorname{diam}B_{m}^{H}=\operatorname{diam}B_{i}^{h}\geq h_{T}(p,B_{i}^{h})\geq h_{T}(p,B_{3}^{h})=h_{T}(p)\simeq H_{T}(p)=\operatorname{diam}B_{3}^{H}.

If i3i\geq 3, then

diamBmH=diamBihhT(p,Bih)hT(p,B3h)=hT(p)HT(p)=diamB3H,\operatorname{diam}B_{m}^{H}=\operatorname{diam}B_{i}^{h}\geq h_{T}(p,B_{i}^{h})\simeq h_{T}(p,B_{3}^{h})=h_{T}(p)\simeq H_{T}(p)=\operatorname{diam}B_{3}^{H},

by the uniform branch growth of TT with respect to hTh_{T}. Therefore, TT has uniform branch growth with respect to HTH_{T}. ∎

For the rest of this section, we employ the new height hTh_{T}, and the enumeration of branches at a given branch point is with respect to the non-increasing order induced by the maximal distance within each branch defined in (11.2) (and not the one induced by the diameter of each branch). In addition, we often refer to the uniform nn-branching property without specifying the employed height notion, wherever that is clear from the context.

11.1. Overview of proof of Proposition 11.1

The proof of Proposition 11.1 splits into three steps. At each step, we construct a quasisymmetric embedding of a given quasiconformal tree TT into a tree TT^{\prime} with one extra property. In particular, given a tree T0𝒬𝒞𝒯(n)T_{0}\in\mathscr{QCT}^{*}(n),

  • in §11.2 we show that T0T_{0} quasisymmetrically embeds into a tree T1𝒬𝒞𝒯(n)T_{1}\in\mathscr{QCT}^{*}(n) that has uniform branch growth,

  • in §11.3 we show that T1T_{1} quasisymmetrically embeds into an nn-valent tree T2𝒬𝒞𝒯(n)T_{2}\in\mathscr{QCT}^{*}(n) that has uniform branch growth,

  • in §11.4 we show that T2T_{2} quasisymmetrically embeds into an nn-valent tree T3𝒬𝒞𝒯(n)T_{3}\in\mathscr{QCT}^{*}(n) that has uniform branch growth and uniform branch density (hence T3T_{3} is uniformly nn-branching).

In each case, applying a uniformization result of Bonk and Meyer [BM20], we may assume that all trees are geodesic and our embeddings will be in fact isometric. The isometric embedding discussed in Definition 11.2 allows us to view each TiT_{i} (i=0,1,2i=0,1,2) as a subspace of the geodesic gluing Ti+1T_{i+1}. Henceforth, we do not address the aforementioned isometric embeddings directly, and we focus on simply constructing Ti+1T_{i+1} “on top” of TiT_{i}.

11.2. Step 1: uniform branch growth

The first step in the proof of Proposition 11.1 is the following proposition.

Proposition 11.6.

Let TT be a quasiconformal tree with Val(T)=m\operatorname{Val}(T)=m, and uniformly separated branch points. Then TT quasisymmetrically embeds into a doubling geodesic tree TT^{\prime} with Val(T)=m\operatorname{Val}(T^{\prime})=m, uniformly separated branch points, and uniform branch growth.

Proof.

Any quasiconformal tree is quasisymmetrically equivalent to a geodesic metric tree by [BM20, Theorem 1.2]. Hence, we may assume that the tree (T,d)(T,d) is a geodesic doubling metric tree with uniformly separated branch points and Val(T)=m\operatorname{Val}(T)=m (since these properties are quasisymmetrically invariant).

Let S={pT:pis a branch point ofTwithVal(T,p)4}S=\{p\in T:p\ \text{is a branch point of}\ T\ \text{with}\operatorname{Val}(T,p)\geq 4\}. If S=S=\emptyset, then TT trivially has uniform branch growth. Suppose that SS\neq\emptyset. For all pp in SS, let ip{3,,Val(T,p)}i_{p}\in\{3,\dots,\operatorname{Val}(T,p)\} be the maximal index such that hT(p)=hT(p,Bip)h_{T}(p)=h_{T}(p,B_{i_{p}}). The branches at branch points where ip=Val(T,p)i_{p}=\operatorname{Val}(T,p) satisfy the uniform branch growth property. Consider the set S={pS:ip{3,,Val(T,p)1}}S^{\prime}=\{p\in S:i_{p}\in\{3,\dots,\operatorname{Val}(T,p)-1\}\}. Similarly, we may assume that SS^{\prime}\neq\emptyset. Fix a branch point pp in SS^{\prime}. By (11.2), for every j{ip+1,,Val(T,p)}j\in\{i_{p}+1,\dots,\operatorname{Val}(T,p)\} there exists a leaf qpjBTj(p)q_{p}^{j}\in B_{T}^{j}(p) with hT(p,Bj)=d(p,qpj)h_{T}(p,B_{j})=d(p,q_{p}^{j}). Note that for a given jj, there could be uncountably many qpjq_{p}^{j} on the branch BTj(p)B_{T}^{j}(p). Hence, for every j{ip+1,,Val(T,p)}j\in\{i_{p}+1,\dots,\operatorname{Val}(T,p)\} we fix some qpjBTj(p)q_{p}^{j}\in B_{T}^{j}(p) with hT(p,Bj)=d(p,qpj)h_{T}(p,B_{j})=d(p,q_{p}^{j}), and we define

p={qpjT:hT(p,Bj)=d(p,qp)for allj{ip+1,,Val(T,p)}}.\mathcal{L}_{p}=\{q_{p}^{j}\in T:\ h_{T}(p,B_{j})=d(p,q_{p})\ \text{for all}\ j\in\{i_{p}+1,\dots,\operatorname{Val}(T,p)\}\}.

Note that cardpVal(T,p)3\operatorname{card}\mathcal{L}_{p}\leq\operatorname{Val}(T,p)-3, and suppose there exists pTp\in T branch point with p\mathcal{L}_{p}\neq\emptyset, because otherwise TT has uniform branch growth. Set =pSp\mathcal{L}^{*}=\bigcup_{p\in S^{\prime}}\mathcal{L}_{p}. By [Why63, Chapter V, (1.3)(iv)] the branch points of any metric tree are countable. Therefore, \mathcal{L}^{*} is also a countable set, i.e. ={x1,x2,}\mathcal{L}^{*}=\{x_{1},x_{2},\dots\}. Let II be an index set with card=cardI\operatorname{card}\mathcal{L}^{*}=\operatorname{card}I. Suppose that cardI=\operatorname{card}I=\infty, set 𝐱I=(x1,)\mathbf{x}_{I}=(x_{1},\dots), and the proof is similar in the case cardI<\operatorname{card}I<\infty. We construct the tree TT^{\prime} into which we isometrically embed TT (seen as a subspace of TT^{\prime}), by gluing a Euclidean line segment of suitable length on TT at each qpjq_{p}^{j}, for every pSp\in S^{\prime}. This ensures that TT^{\prime} has uniform branch growth, even for branches at pp, where the property does not necessarily hold for TT.

For every xx\in\mathcal{L}^{*} there exists a pSp\in S^{\prime} such that x=qpjx=q_{p}^{j}, for some j{ip+1,,Val(T,p)}j\in\{i_{p}+1,\dots,\operatorname{Val}(T,p)\}. Let x\ell_{x} be a Euclidean line segment with diamx=hT(p)hT(p,Bj)\operatorname{diam}\ell_{x}=h_{T}(p)-h_{T}(p,B_{j}), and fix yxy_{x} to be one of the endpoints of x\ell_{x}. We first need to make sure that for every xx\in\mathcal{L}^{*}, the line segment x\ell_{x} is well-defined, i.e., we have to show that there is a unique branch point pSp\in S^{\prime} such that x=qpjx=q_{p}^{j}. It is enough to verify that pp=\mathcal{L}_{p}\cap\mathcal{L}_{p^{\prime}}=\emptyset for all p,pSp,p^{\prime}\in S^{\prime} distinct.

To this end, suppose that pBTi(p)p\in B_{T}^{i}(p^{\prime}) for i{1,,Val(T,p)}i\in\{1,\dots,\operatorname{Val}(T,p^{\prime})\} and that pBTj(p)p^{\prime}\in B_{T}^{j}(p) for j{1,,Val(T,p)}j\in\{1,\dots,\operatorname{Val}(T,p)\}. By Lemma 11.4,

(11.6) BTm(p)BTj(p)for all mi,\displaystyle B_{T}^{m}(p^{\prime})\subset B_{T}^{j}(p)\quad\text{for all }m\neq i,
(11.7) BTn(p)BTi(p)for all nj.\displaystyle B_{T}^{n}(p)\subset B_{T}^{i}(p^{\prime})\quad\text{for all }n\neq j.

Assume towards contradiction that there is a leaf qppq\in\mathcal{L}_{p}\cap\mathcal{L}_{p^{\prime}}. There are three possible cases to consider, depending on the intersection of the arcs [p,q][p,q] and [p,q][p^{\prime},q]. Note that {q}[p,q][p,q]\{q\}\subsetneq[p,q]\cap[p^{\prime},q] by the unique arc property of TT.

Case 1. Suppose [p,q][p,q]=[p,q][p,q]\cap[p^{\prime},q]=[p,q]. Since [p,q][p,q][p,q]\subset[p^{\prime},q], by (11.6), (11.7) we have qBTi(p)q\in B_{T}^{i}(p^{\prime}). Due to qpq\in\mathcal{L}_{p}, there is jq{4,,Val(T,p)}j_{q}\in\{4,\dots,\operatorname{Val}(T,p)\} with hT(p,Bjq)=d(p,q)h_{T}(p,B_{j_{q}})=d(p,q), and there are at least three branches BTj1(p)B_{T}^{j_{1}}(p), BTj2(p)B_{T}^{j_{2}}(p), BTj3(p)B_{T}^{j_{3}}(p), j1,j2,j3{1,Val(T,p)}j_{1},j_{2},j_{3}\in\{1,\dots\operatorname{Val}(T,p)\} , with

min{hT(p,Bj1),hT(p,Bj2),hT(p,Bj3)}>hT(p,Bjq),\min\{h_{T}(p,B_{j_{1}}),h_{T}(p,B_{j_{2}}),h_{T}(p,B_{j_{3}})\}>h_{T}(p,B_{j_{q}}),

by definition of p\mathcal{L}_{p}. Hence, there exists j{1,Val(T,p)}j^{\prime}\in\{1,\dots\operatorname{Val}(T,p)\} with hT(p,Bj)>hT(p,Bjq)=d(p,q)h_{T}(p,B_{j^{\prime}})>h_{T}(p,B_{j_{q}})=d(p,q) and BTj(p)BTj(p)B_{T}^{j^{\prime}}(p)\neq B_{T}^{j}(p). By (11.7) we have BTj(p)BTi(p)B_{T}^{j^{\prime}}(p)\subset B_{T}^{i}(p^{\prime}), which due to qBTi(p)q\in B_{T}^{i}(p^{\prime}) and qpq\in\mathcal{L}_{p^{\prime}} implies

hT(p,Bi)=d(p,q)=d(p,p)+d(p,q)<d(p,p)+hT(p,Bj)=d(p,p)+d(p,qj),h_{T}(p^{\prime},B_{i})=d(p^{\prime},q)=d(p^{\prime},p)+d(p,q)<d(p,p^{\prime})+h_{T}(p,B_{j^{\prime}})=d(p,p^{\prime})+d(p,q_{j^{\prime}}),

for some leaf qjBTj(p)BTi(p)q_{j^{\prime}}\in B_{T}^{j^{\prime}}(p)\subset B_{T}^{i}(p^{\prime}). The right hand side in the above relation is at most d(p,qj)hT(p,Bi)d(p^{\prime},q_{j^{\prime}})\leq h_{T}(p^{\prime},B_{i}), which leads to a contradiction.

Case 2. Suppose [p,q][p,q]=[p,q][p,q]\cap[p^{\prime},q]=[p^{\prime},q]. The proof is similar to that of Case 1.

Case 3. Suppose [p,q][p,q]=[r,q][p,q]\cap[p^{\prime},q]=[r,q], for some rT{p,p}r\in T\setminus\{p,p^{\prime}\}. This implies by [BT21, Lemma 3.2(iii)] that p,q,rp^{\prime},q,r lie on the same branch of TT at pp. Hence, qBTj(p)q\in B_{T}^{j}(p), and by qpq\in\mathcal{L}_{p} we have

(11.8) hT(p,Bj)=d(p,q)d(p,),h_{T}(p,B_{j})=d(p,q)\geq d(p,\ell),

for any leaf BTj(p)\ell\in B_{T}^{j}(p). By pBTj(p)p^{\prime}\in B_{T}^{j}(p) and (11.6), for any m{1,,Val(T,p)}m\in\{1,\dots,\operatorname{Val}(T,p^{\prime})\} with mim\neq i we have that

(11.9) d(p,)=d(p,r)+d(r,p)+d(p,),d(p,\ell)=d(p,r)+d(r,p^{\prime})+d(p^{\prime},\ell),

for any leaf BTm(p)\ell\in B_{T}^{m}(p^{\prime}). Fix some BTm(p)B_{T}^{m}(p^{\prime}) for mim\neq i, and a leaf BTm(p)BTj(p)\ell\in B_{T}^{m}(p^{\prime})\subset B_{T}^{j}(p). By (11.8), (11.9), and d(p,r)+d(r,q)=d(p,q)d(p,r)+d(r,q)=d(p,q), we have

d(p,r)+d(r,q)d(p,r)+d(r,p)+d(p,).d(p,r)+d(r,q)\geq d(p,r)+d(r,p^{\prime})+d(p^{\prime},\ell).

Since rpr\neq p^{\prime}, the above implies that d(r,q)>d(p,)d(r,q)>d(p^{\prime},\ell). But [r,q][p,q][r,q]\subset[p^{\prime},q], so

d(p,)<d(r,q)d(p,q).d(p^{\prime},\ell)<d(r,q)\leq d(p^{\prime},q).

The branch BTm(p)B_{T}^{m}(p^{\prime}) and the leaf BTm(p)\ell\in B_{T}^{m}(p^{\prime}) are arbitrary, which means that the above inequality shows that hT(p,B)<d(p,q)h_{T}(p^{\prime},B)<d(p^{\prime},q), for all branches BBTi(p)B\neq B_{T}^{i}(p^{\prime}) of TT at pp^{\prime}. This contradicts the fact that qpq\in\mathcal{L}_{p^{\prime}}, since there is no need to glue a line segment at qq to increase a branch of TT at pp^{\prime}.

Recall that diamx>0\operatorname{diam}\ell_{x}>0 for all xx\in\mathcal{L}^{*}, since xx is a leaf in p\mathcal{L}_{p} for some pSp\in S^{\prime}, and by definition of SS^{\prime}. We define the geodesic gluing (T,d):=(T,d,𝐱I)iI(xi,di,yxi)(T^{\prime},d^{\prime}):=(T,d,\mathbf{x}_{I})\bigvee_{i\in I}(\ell_{x_{i}},d_{i},y_{x_{i}}), where did_{i} is the Euclidean metric on xi\ell_{x_{i}}. By Lemma 11.5, [BT21, Lemma 3.9], and diamxihT(pi)\operatorname{diam}\ell_{x_{i}}\leq h_{T}(p_{i}) for all iIi\in I, we have that diamxi0\operatorname{diam}\ell_{x_{i}}\to 0 as ii\to\infty. By Lemma 11.3 this implies that (T,d)(T^{\prime},d^{\prime}) is a geodesic metric tree. Note that the gluing is between TT at leaves and line segments at endpoints, which ensures that the branch points of TT and of TT^{\prime} are in one-to-one correspondence. Therefore, Val(T)=m\operatorname{Val}(T^{\prime})=m.

Addressing uniform branch separation and uniform branch growth requires to show that the height of branch points in TT is the same as in TT^{\prime}. Fix a branch point pp of TT and let BTk(p)B_{T}^{k}(p) be a branch of TT at pp for k{1,,Val(T,p)}k\in\{1,\dots,\operatorname{Val}(T,p)\}. The definition of TT^{\prime}, the convention TTT\subset T^{\prime}, and the fact that TT^{\prime} is a tree imply that there exists a unique n~k{1,,Val(T,p)}\tilde{n}_{k}\in\{1,\dots,\operatorname{Val}(T^{\prime},p)\} such that BTk(p)BTn~k(p)B_{T}^{k}(p)\subset B_{T^{\prime}}^{\tilde{n}_{k}}(p). Recall that ip{3,,Val(T,p)1}i_{p}\in\{3,\dots,\operatorname{Val}(T,p)-1\}, and hT(p,B3)=hT(p,Bip)h_{T}(p,B_{3})=h_{T}(p,B_{i_{p}}). We show that hT(p)=hT(p,Bj)h_{T}(p)=h_{T^{\prime}}(p,B_{j}) for all j{3,,Val(T,p)}j\in\{3,\dots,\operatorname{Val}(T^{\prime},p)\}, by considering three cases.

Case (i). Suppose that k{ip+1,,Val(T,p)}k\in\{i_{p}+1,\dots,\operatorname{Val}(T,p)\}. Let qkBTk(p)pq_{k}\in B_{T}^{k}(p)\cap\mathcal{L}_{p} such that hT(p,Bk)=d(p,qk)h_{T}(p,B_{k})=d(p,q_{k}). Let xBTn~k(p)x\in B_{T^{\prime}}^{\tilde{n}_{k}}(p). If xTx\in T, then

d(p,x)=d(p,x)d(p,qk)=hT(p,Bk)<hT(p).d^{\prime}(p,x)=d(p,x)\leq d(p,q_{k})=h_{T}(p,B_{k})<h_{T}(p).

If xTTx\in T^{\prime}\setminus T, then xrx\in\ell_{r} for some leaf rBTk(p)r\in B_{T}^{k}(p) that lies in \mathcal{L}^{*}. If r=qkpr=q_{k}\in\mathcal{L}_{p}, then

d(p,x)=d(p,qk)+d(qk,x)hT(p,Bk)+diamqk=hT(p).d^{\prime}(p,x)=d^{\prime}(p,q_{k})+d^{\prime}(q_{k},x)\leq{h_{T}(p,B_{k})+\operatorname{diam}\ell_{q_{k}}=h_{T}(p)}.

If rqkr\neq q_{k}, there is pTp^{\prime}\in T branch point with hT(p,Bj)=d(p,r)h_{T}(p^{\prime},B_{j})=d(p^{\prime},r), for some j{ip+1,,Val(T,p)}j\in\{i_{p^{\prime}}+1,\dots,\operatorname{Val}(T,p^{\prime})\}. Note that pBTk(p)p^{\prime}\in B_{T}^{k}(p). Suppose otherwise that pBTk(p)p^{\prime}\in B_{T}^{k^{\prime}}(p) for some kkk^{\prime}\neq k, say k=1k^{\prime}=1. Then, by Lemma 11.4, BT2(p)BTj(p)B_{T}^{2}(p)\subset B_{T}^{j}(p^{\prime}). Furthermore, let q2BT2(p)q_{2}\in B_{T}^{2}(p) be such that d(p,q2)=hT(p,B2)d^{\prime}(p,q_{2})=h_{T}(p,B_{2}). By k2k\neq 2 and [BT21, Lemma 3.2(iii)] p[pq2][p,r]p\in[p^{\prime}q_{2}]\cap[p^{\prime},r]. Hence, by geodesicity of TT

d(p,q2)=d(p,p)+d(p,q2)=d(p,p)+hT(p,B2)>d(p,p)+d(p,r)=d(p,r)d^{\prime}(p^{\prime},q_{2})=d^{\prime}(p^{\prime},p)+d^{\prime}(p,q_{2})=d^{\prime}(p^{\prime},p)+h_{T}(p,B_{2})>d^{\prime}(p^{\prime},p)+d^{\prime}(p,r)=d^{\prime}(p^{\prime},r)

which is contradiction since q2BTk(p)q_{2}\in B_{T}^{k}(p) and rr\in\mathcal{L}^{*}. Repeating Lemma 11.4 we have BTi(p)BTk(p)B_{T}^{i}(p^{\prime})\subset B_{T}^{k}(p) for all ini\neq n, where BTn(p)B_{T}^{n}(p^{\prime}) is the branch of TT at pp^{\prime} that contains pp. By k1k\neq 1 and Lemma 11.4 we have BT1(p)BTn(p)B_{T}^{1}(p)\subset B_{T}^{n}(p^{\prime}). This implies n=1n=1 due to kip+1k\geq i_{p}+1, since otherwise h(p,Bk)h(p,B1)h(p,B_{k})\geq h(p,B_{1}), leading to contradiction. Thus, x,px,p lie in different components of pp^{\prime}, implying p[p,x]p^{\prime}\in[p,x] by [BT21, Lemma 3.2(iii)]. Similarly, p[r,p]p^{\prime}\in[r^{\prime},p], where rBT2(p)BTk(p)r^{\prime}\in B_{T}^{2}(p^{\prime})\subset B_{T}^{k}(p) such that hT(p,B2)=d(p,r)h_{T}(p^{\prime},B_{2})=d(p^{\prime},r^{\prime}). It follows

d(p,x)\displaystyle d^{\prime}(p,x) =d(p,p)+d(p,x)\displaystyle=d^{\prime}(p,p^{\prime})+d^{\prime}(p^{\prime},x)
=d(p,p)+d(p,r)+d(r,x)\displaystyle=d^{\prime}(p,p^{\prime})+d^{\prime}(p^{\prime},r)+d^{\prime}(r,x)
d(p,p)+hT(p,Bj)+diamr\displaystyle\leq d^{\prime}(p,p^{\prime})+h_{T}(p^{\prime},B_{j})+\operatorname{diam}\ell_{r}
d(p,p)+d(p,r)\displaystyle\leq d^{\prime}(p,p^{\prime})+d^{\prime}(p^{\prime},r^{\prime})
=d(p,r)\displaystyle=d^{\prime}(p,r^{\prime})
hT(p,Bk)\displaystyle\leq h_{T}(p,B_{k})
hT(p).\displaystyle\leq h_{T}(p).

Therefore, since xBTn~k(p)x\in B_{T^{\prime}}^{\tilde{n}_{k}}(p) is arbitrary, we have shown that hT(p,Bn~k)hT(p)h_{T^{\prime}}(p,B_{\tilde{n}_{k}})\leq h_{T}(p). For the other side of the inequality note that if yBTk(p)y\in B_{T}^{k}(p) is an endpoint of qk\ell_{q_{k}} distinct from qkq_{k}, then

d(p,y)=d(p,qk)+d(qk,y)=hT(p,Bk)+diamqk=hT(p).d^{\prime}(p,y)=d^{\prime}(p,q_{k})+d^{\prime}(q_{k},y)=h_{T}(p,B_{k})+\operatorname{diam}\ell_{q_{k}}=h_{T}(p).

Case (ii). Suppose that k{3,,ip}k\in\{3,\dots,i_{p}\} and that hT(p,B2)>hT(p,B3)h_{T}(p,B_{2})>h_{T}(p,B_{3}). Let qkBTk(p)q_{k}\in B_{T}^{k}(p) with hT(p)=d(p,qk)h_{T}(p)=d(p,q_{k}). We claim that no line segment is glued on the leaf qkTq_{k}\in T. Then it follows with similar arguments as in Case (i) that hT(p,Bn~k)=hT(p)h_{T^{\prime}}(p,B_{\tilde{n}_{k}})=h_{T}(p). Assume towards contradiction that qk\ell_{q_{k}} is a line segment glued on the leaf qkq_{k}. Hence, there exists n{1,,Val(T,p)}n\in\{1,\dots,\operatorname{Val}(T,p)\} and p′′BTn(p)Sp^{\prime\prime}\in B_{T}^{n}(p)\cap S^{\prime} branch point such that hT(p′′,Bm)=d(p′′,qk)h_{T}(p^{\prime\prime},B_{m})=d(p^{\prime\prime},q_{k}) for some m{ip′′+1,,Val(T,p′′)}m\in\{i_{p^{\prime\prime}}+1,\dots,\operatorname{Val}(T,p^{\prime\prime})\}. If n=kn=k, then p′′,qkBTk(p)p^{\prime\prime},q_{k}\in B_{T}^{k}(p). By Lemma 11.4 and similar arguments as in Case (i), we have pBT1(p′′)p\in B_{T}^{1}(p^{\prime\prime}), and the rest of the branches of TT at p′′p^{\prime\prime} are subsets of BTk(p)B_{T}^{k}(p). Let rBT2(p′′)r\in B_{T}^{2}(p^{\prime\prime}) such that hT(p′′,B2)=d(p′′,r)h_{T}(p^{\prime\prime},B_{2})=d(p^{\prime\prime},r). Similarly to Case (i), it can be shown that p′′[p,r][p,qk]p^{\prime\prime}\in[p,r]\cap[p,q_{k}]. Hence, d(p′′,r)>d(p′′,qk)d^{\prime}(p^{\prime\prime},r)>d^{\prime}(p^{\prime\prime},q_{k}) due to p′′Sp^{\prime\prime}\in S^{\prime}. Adding d(p,p′′)d^{\prime}(p,p^{\prime\prime}) to each side results in d(p,r)>hT(p,Bk)d^{\prime}(p,r)>h_{T}(p,B_{k}), which is a contradiction by rBTk(p)r\in B_{T}^{k}(p). The case where n{2,3,,ip}{k}n\in\{2,3,\dots,i_{p}\}\setminus\{k\} follows by Lemma 11.4 and (11.2). Lastly, if n=1n=1, then p[p′′,qk][p′′,q2]p\in[p^{\prime\prime},q_{k}]\cap[p^{\prime\prime},q_{2}], since they lie on different branches of pp in TT. Therefore, d(p,q2)>d(p,qk)d^{\prime}(p,q_{2})>d^{\prime}(p,q_{k}) due to hT(p,B2)>hT(p,B3)h_{T}(p,B_{2})>h_{T}(p,B_{3}). This implies d(p′′,q2)>d(p′′,qk)=hT(p′′,Bk)d^{\prime}(p^{\prime\prime},q_{2})>d^{\prime}(p^{\prime\prime},q_{k})=h_{T}(p^{\prime\prime},B_{k}) which is contradiction.

Case (iii). Suppose that k{3,,ip}k\in\{3,\dots,i_{p}\}, qkBTk(p)q_{k}\in B_{T}^{k}(p) with hT(p)=d(p,qk)h_{T}(p)=d(p,q_{k}), and that hT(p,B2)=hT(p,B3)h_{T}(p,B_{2})=h_{T}(p,B_{3}). This case is almost identical to Case (ii), with the exception of the case n=1n=1, where the fact that hT(p,B2)>hT(p,B3)h_{T}(p,B_{2})>h_{T}(p,B_{3}) was crucial. We may assume that qk\ell_{q_{k}} is a line segment glued on the leaf qkq_{k} (since otherwise we are in Case (ii)), and p′′BT1(p)Sp^{\prime\prime}\in B_{T}^{1}(p)\cap S^{\prime} is a branch point with hT(p′′,Bm)=d(p′′,qk)h_{T}(p^{\prime\prime},B_{m})=d(p^{\prime\prime},q_{k}) for some m{ip′′+1,,Val(T,p′′)}m\in\{i_{p^{\prime\prime}}+1,\dots,\operatorname{Val}(T,p^{\prime\prime})\}. We claim that p′′{qp2,,qpip}={qk}\mathcal{L}_{p^{\prime\prime}}\cap\{q_{p}^{2},\dots,q_{p}^{i_{p}}\}=\{q_{k}\}, where qpaBTa(p)q_{p}^{a}\in B_{T}^{a}(p) with hT(p,Ba)=d(p,qka)h_{T}(p,B_{a})=d(p,q_{k}^{a}) for all a{2,,ip}a\in\{2,\dots,i_{p}\}. Without loss of generality assume that ip=3i_{p}=3. Suppose there are qs1,qs2q_{s_{1}},q_{s_{2}}\in\mathcal{L}^{*} with s1,s2Ss_{1},s_{2}\in S^{\prime} and qs1=qp2,qs2=qp3q_{s_{1}}=q_{p}^{2},q_{s_{2}}=q_{p}^{3}. Similarly to Case (ii), we may assume that s1,s2BT1(p)s_{1},s_{2}\in B_{T}^{1}(p). By Lemma 11.4 we have qs1,qs2BTis1+1(s1)BTis2+1(s2)q_{s_{1}},q_{s_{2}}\in B_{T}^{i_{s_{1}}+1}(s_{1})\cap B_{T}^{i_{s_{2}}+1}(s_{2}) where is1,is23i_{s_{1}},i_{s_{2}}\geq 3. Without loss of generality we may assume that is1=is2=3i_{s_{1}}=i_{s_{2}}=3. If s2[s1,p]s_{2}\in[s_{1},p], by Lemma 11.4 the branches BT2(s2),BT4(s2)B_{T}^{2}(s_{2}),B_{T}^{4}(s_{2}) are contained in BT4(s1)B_{T}^{4}(s_{1}). Let zBT2(s2)z\in B_{T}^{2}(s_{2}) with hT(s2,B2)=d(s2,z)h_{T}(s_{2},B_{2})=d^{\prime}(s_{2},z). Then, d(s2,z)>d(s2,qs2)d^{\prime}(s_{2},z)>d^{\prime}(s_{2},q_{s_{2}}), which by hT(p,B2)=hT(p,B3)h_{T}(p,B_{2})=h_{T}(p,B_{3}) and d(s1,qs1)=hT(s1,B4)d^{\prime}(s_{1},q_{s_{1}})=h_{T}(s_{1},B_{4}) implies

d(s1,z)>d(s1,qs2)=d(s1,p)+d(p,qs2)=d(s1,p)+d(p,qs1)=hT(s1,B4),d^{\prime}(s_{1},z)>d^{\prime}(s_{1},q_{s_{2}})=d^{\prime}(s_{1},p)+d^{\prime}(p,q_{s_{2}})=d^{\prime}(s_{1},p)+d^{\prime}(p,q_{s_{1}})=h_{T}(s_{1},B_{4}),

which is a contradiction. The case s1[s2,p]s_{1}\in[s_{2},p] is similar. If s1[s2,p]s_{1}\notin[s_{2},p] and s2[s1,p]s_{2}\notin[s_{1},p], then s2BT4(s1)s_{2}\in B_{T}^{4}(s_{1}) and s1BT4(s2)s_{1}\in B_{T}^{4}(s_{2}) by [BT21, Lemma 3.2(iii)], since pBT4(s1)BT4(s2)p\in B_{T}^{4}(s_{1})\cap B_{T}^{4}(s_{2}). Then BT1(s2)BT4(s1)B_{T}^{1}(s_{2})\subset B_{T}^{4}(s_{1}) and BT1(s1)BT4(s2)B_{T}^{1}(s_{1})\subset B_{T}^{4}(s_{2}) by Lemma 11.4. Hence, we have

hT(s2,B1)hT(s1,B4)<hT(s1,B1)hT(s2,B4),h_{T}(s_{2},B_{1})\leq h_{T}(s_{1},B_{4})<h_{T}(s_{1},B_{1})\leq h_{T}(s_{2},B_{4}),

which is a contradiction by s1,s2Ss_{1},s_{2}\in S^{\prime}. Hence, there is no line segment glued on qp2BT2(p)q_{p}^{2}\in B_{T}^{2}(p), for which it’s true that d(p,qp2)=hT(p,B2)=hT(p,B3)d(p,q_{p}^{2})=h_{T}(p,B_{2})=h_{T}(p,B_{3}). It follows with similar arguments as in Case (i) that hT(p,Bn~k)=hT(p)h_{T^{\prime}}(p,B_{\tilde{n}_{k}})=h_{T}(p).

Therefore, for any branch point pTp\in T^{\prime}, we have shown hT(p,BTj(p))=hT(p)h_{T^{\prime}}(p,B_{T^{\prime}}^{j}(p))=h_{T}(p), for any branch BTj(p)B_{T^{\prime}}^{j}(p) of TT^{\prime} at pp. The uniform branch growth follows trivially, and the uniform branch separation follows by the uniform separation of TT (Lemma 11.5). Lastly, by Lemma 2.5 we have that TT^{\prime} is doubling, completing the proof. ∎

11.3. Step 2: uniform valence

The second step in the proof of Proposition 11.1 is the next proposition.

Proposition 11.7.

Let (T,d)(T,d) be a quasiconformal tree with Val(T)=m\operatorname{Val}(T)=m, uniformly separated branch points, and uniform branch growth. Then, TT quasisymmetrically embeds into (T,d)(T^{\prime},d^{\prime}), where TT^{\prime} is an mm-valent geodesic doubling tree with uniform branch separation and uniform branch growth.

Note that given the pointed metric spaces (Yi,dYi,yi)(Y_{i},d_{Y_{i}},y_{i}), for iIi\in I, we can define the geodesic gluing iI(Yi,dYi,yi)\bigvee_{i\in I}(Y_{i},d_{Y_{i}},y_{i}) by identifying yiy_{i} and yjy_{j} for all i,jIi,j\in I. More specifically, denote by X0={0}X_{0}=\{0\} the trivial singleton metric space, and define iI(Yi,dYi,yi):=(X0,d0,𝐱I)iI(Yi,dYi,yi)\bigvee_{i\in I}(Y_{i},d_{Y_{i}},y_{i}):=(X_{0},d_{0},\mathbf{x}_{I})\bigvee_{i\in I}(Y_{i},d_{Y_{i}},y_{i}) as in Definition 11.2, where xi=0x_{i}=0 for all iIi\in I. We make use of this specific type of gluing in what follows.

Proof of Proposition 11.7.

Similarly to Step 1, by applying [BM20, Theorem 1.2], we may assume that the tree (T,d)(T,d) is a geodesic doubling metric tree with uniformly separated branch points, uniform branch growth and Val(T)=m\operatorname{Val}(T)=m (since these properties are quasisymmetrically invariant).

Let S={pT:Val(T,p){3,,m1}}S=\{p\in T:\operatorname{Val}(T,p)\in\{3,\dots,m-1\}\}. Unlike Step 1, we glue Euclidean line segments on branch points of TT, and, specifically, on every pSp\in S.

Let {p1,p2,}\{p_{1},p_{2},\dots\} be a fixed enumeration of SS. Set nk=Val(T,pk)n_{k}=\operatorname{Val}(T,p_{k}), and let IpkI_{p_{k}} be an index set with cardIpk=mnk\operatorname{card}I_{p_{k}}=m-n_{k}, for all kk. For pkSp_{k}\in S, let ki\ell_{k}^{i} be Euclidean line segments with diamki=hT(pk,Bnk)\operatorname{diam}\ell_{k}^{i}=h_{T}(p_{k},B_{n_{k}}), for all iIpki\in I_{p_{k}}. Fix ykiy_{k}^{i} to be one of the endpoints of ki\ell_{k}^{i}. For each kk\in\mathbb{N}, define Yk=iIpk(ki,di,yki)Y_{k}=\bigvee_{i\in I_{p_{k}}}(\ell_{k}^{i},d_{i},y_{k}^{i}), where did_{i} is the Euclidean metric on ki\ell_{k}^{i}.

Set (T,d):=(T,d,𝐩I)kI(Yk,dYk,zk)(T^{\prime},d^{\prime}):=(T,d,\mathbf{p}_{I})\bigvee_{k\in I}(Y_{k},d_{Y_{k}},z_{k}), where II is an index set with cardI=cardS\operatorname{card}I=\operatorname{card}S, 𝐩I=(p1,p2,)\mathbf{p}_{I}=(p_{1},p_{2},\dots), and zkykiz_{k}\sim y_{k}^{i} for all iIpki\in I_{p_{k}}, i.e., for each kIk\in I the point zkz_{k} is the “identification” of all points ykikiy_{k}^{i}\in\ell_{k}^{i}. Suppose cardI=\operatorname{card}I=\infty, and the proof is similar in the case cardI<\operatorname{card}I<\infty. By [BT21, Lemma 3.9], we have diamYk0\operatorname{diam}Y_{k}\to 0 as kk\to\infty. Hence, (T,d)(T^{\prime},d^{\prime}) is a geodesic metric tree by Lemma 11.3.

It remains to show that TT^{\prime} is mm-valent, with uniform branch separation and uniform branch growth. Note that the gluing is between TT at branch points and line segments at endpoints, which implies that the branch points of TT and those of TT^{\prime} are in one-to-one correspondence. Therefore, by the choice of nkn_{k} and YkY_{k}, we have Val(T,p)=m\operatorname{Val}(T^{\prime},p)=m for all branch points of TT^{\prime}.

Let pp be a branch point of TT^{\prime}. Then pp is a branch point of TT, and for every j{1,,Val(T,p)}j\in\{1,\dots,\operatorname{Val}(T,p)\} there is a unique n~j\tilde{n}_{j} such that BTj(p)BTn~j(p)B_{T}^{j}(p)\subset B_{T^{\prime}}^{\tilde{n}_{j}}(p). Let xBTn~j(p)x\in B_{T^{\prime}}^{\tilde{n}_{j}}(p). If xBTj(p)x\in B_{T}^{j}(p), then

d(p,x)=d(p,x)hT(p,Bj).d^{\prime}(p,x)=d(p,x)\leq h_{T}(p,B_{j}).

Suppose that xBTj(p)x\notin B_{T}^{j}(p), which implies that xkix\in\ell_{k}^{i} for some pkSBTj(p)p_{k}\in S\cap B_{T}^{j}(p) and iIpki\in I_{p_{k}}. Then, by diamki=hT(pk,Bnk)\operatorname{diam}\ell_{k}^{i}=h_{T}(p_{k},B_{n_{k}}) we have

d(p,x)\displaystyle d^{\prime}(p,x) =d(p,pk)+di(pk,x)\displaystyle=d(p,p_{k})+d_{i}(p_{k},x)
d(p,pk)+diamki\displaystyle\leq d(p,p_{k})+\operatorname{diam}\ell_{k}^{i}
=d(p,pk)+hT(pk,Bnk)\displaystyle=d(p,p_{k})+h_{T}(p_{k},B_{n_{k}})
=d(p,pk)+d(pk,qk),\displaystyle=d(p,p_{k})+d(p_{k},q_{k}),

for some leaf qkBTnk(pk)q_{k}\in B_{T}^{n_{k}}(p_{k}). Note that BTnk(pk)BTj(p)B_{T}^{n_{k}}(p_{k})\subset B_{T}^{j}(p), by pkBTj(p)p_{k}\in B_{T}^{j}(p) and Lemma 11.4. Hence, the above implies d(p,x)hT(p,Bj)d^{\prime}(p,x)\leq h_{T}(p,B_{j}). Therefore, since xBTn~j(p)x\in B_{T^{\prime}}^{\tilde{n}_{j}}(p) was arbitrary, we showed that hT(p,Bn~j)hT(p,Bj)h_{T^{\prime}}(p,B_{\tilde{n}_{j}})\leq h_{T}(p,B_{j}), and the other inequality follows by BTj(p)BTn~j(p)B_{T}^{j}(p)\subset B_{T^{\prime}}^{\tilde{n}_{j}}(p). As a result, the maximal distances of branches of TT are the same as those of branches of TT^{\prime} at every branch point. The uniform growth and separation for TT^{\prime} follow by those of TT, and the fact that TT^{\prime} is doubling follows by Lemma 2.5. ∎

11.4. Step 3: uniform density of branch points

The final step in the proof of Proposition 11.1 is the following proposition.

Proposition 11.8.

Let (T,d)(T,d) be an mm-valent quasiconformal tree with uniform branch separation and uniform branch growth with m3m\geq 3. Then, (T,d)(T,d) quasisymmetrically embeds into a geodesic uniformly mm-branching quasiconformal tree (T,d)(T^{\prime},d^{\prime}).

For the rest of this section fix an integer m3m\geq 3, an alphabet A={1,,m}A=\{1,\dots,m\}, and a weight 𝐚\operatorname{\bf{a}}. For some fixed constant c(0,1)c\in(0,1) denote by c𝕋m,𝐚c\mathbb{T}^{m,\operatorname{\bf{a}}} the metric space (A/,cdA,𝐚)(A^{\mathbb{N}}/\sim,cd_{A,\operatorname{\bf{a}}}), i.e., a rescaled copy of the tree 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} of diameter cc.

As with the previous steps, we construct TT^{\prime} “on top” of TT with appropriate geodesic gluing. By rescaling the metric, we may assume that diamT=1\operatorname{diam}T=1.

For this step, we use the decomposition of the tree TT as in [BM20, Section 5]. Namely, there exist finite sets VnT\textbf{V}^{n}\subset T, for all nn\in\mathbb{N}, consisting of double points of TT with V1V2\textbf{V}^{1}\subset\textbf{V}^{2}\subset\dots. Each point vVnv\in\textbf{V}^{n} is called an nn-vertex. The closure of a component of TVnT\setminus\textbf{V}^{n} is called an nn-tile, and the set of all nn-tiles is denoted by Xn\textbf{X}^{n} of level nn. It is also convenient to denote V0=\textbf{V}^{0}=\emptyset, X0={T}\textbf{X}^{0}=\{T\} with TT being the only 0-tile. This specific collection V1,V2,\textbf{V}^{1},\textbf{V}^{2},\dots satisfies certain properties for some fixed small enough constants β1\beta\geq 1, γ>0\gamma>0, and δ(0,1/3β)\delta\in(0,1/3\beta).

Lemma 11.9.
  1. (i)

    For each vVnv\in\textbf{V}^{n}, hT(v,B2)cβδnh_{T}(v,B_{2})\geq c\beta\delta^{n}, where cc is a constant that solely depends on the comparability constant from (11.4).

  2. (ii)

    For each vVnv\in\textbf{V}^{n} and branch point pTp\in T, d(v,p)cγmin{hT(p),cδn}d(v,p)\geq c\gamma\min\{h_{T}(p),c\delta^{n}\}.

  3. (iii)

    For u,vVnu,v\in\textbf{V}^{n} distinct, d(u,v)δnd(u,v)\geq\delta^{n}.

  4. (iv)

    For each nn-tile XX, diam(X)δn\operatorname{diam}(X)\simeq\delta^{n}. In particular, δndiam(X)3βδn\delta^{n}\leq\operatorname{diam}(X)\leq 3\beta\delta^{n}.

  5. (v)

    Each nn-tile XX is equal to the union of all (n+1)(n+1)-tiles X0X_{0} with X0XX_{0}\subset X.

  6. (vi)

    Each nn-tile XX contains at least three (n+1)(n+1)-tiles.

  7. (vii)

    Each nn-tile XX is a subtree of TT with X𝐕n\partial X\subset\mathbf{V}_{n}.

  8. (viii)

    If vv is an nn-vertex and XX an nn-tile with vXv\in X, then vXv\in\partial X. Moreover, there exists precisely one (n+1)(n+1)-tile X0XX_{0}\subset X with vX0v\in X_{0}.

  9. (ix)

    There is a constant KK\in\mathbb{N} such that for each n0n\in\mathbb{N}_{0} and each nn-tile XX, the number of (n+1)(n+1)-tiles contained in XX are at most KK.

Proof.

Claims (i), (ii) are immediate by [BM20, Equations (5.5), (5.6)], Lemma 11.5 and the fact that 0<c10<c\leq 1.

Claims (iii), (iv) are proved in [BM20, Equations (5.2), (5.3)]. Claim (v) is proved in [BM20, Lemma 5.1 (viii)]. Claim (vi) is proved in [BM20, Lemma 5.4 (i)]. Claim (vii) is proved in [BM20, Lemma 5.1 (i)]. Claim (viii) is proved in [BM20, Lemma 5.1 (ix)]. Claim (ix) is proved in [BM20, Lemma 5.7 (ii)]. ∎

We also need to determine the new height of branch points of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}.

Lemma 11.10.

Let wAw\in A^{*} and p=[w12()]p=[w12^{(\infty)}] be a branch point of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}. Then h𝕋m,𝐚(p)=𝐚(3)Δ(w)h_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p)=\operatorname{\bf{a}}(3)\Delta(w), and [1()][1^{(\infty)}] lies either on B𝕋m,𝐚1(p)B_{\mathbb{T}^{m,\operatorname{\bf{a}}}}^{1}(p), or on B𝕋m,𝐚2(p)B_{\mathbb{T}^{m,\operatorname{\bf{a}}}}^{2}(p).

Proof.

Let wAw\in A^{*}, then p=[w12()]p=[w12^{(\infty)}] is a branch point of 𝕋wm,𝐚\mathbb{T}_{w}^{m,\operatorname{\bf{a}}} with mm distinct branches 𝕋w1m,𝐚,,𝕋wmm,𝐚\mathbb{T}_{w1}^{m,\operatorname{\bf{a}}},\dots,\mathbb{T}_{wm}^{m,\operatorname{\bf{a}}} in 𝕋wm,𝐚\mathbb{T}_{w}^{m,\operatorname{\bf{a}}}.

Observe that if i1i\neq 1, then by Remark 4.7

h𝕋m,𝐚(p,𝕋wim,𝐚)dA,𝐚([wi1()],[wi2()])=𝐚(i)Δ(w),h_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p,\mathbb{T}_{wi}^{m,\operatorname{\bf{a}}})\geq d_{A,\operatorname{\bf{a}}}([wi1^{(\infty)}],[wi2^{(\infty)}])=\operatorname{\bf{a}}(i)\Delta(w),

and similarly h𝕋m,𝐚(p,𝕋w1m,𝐚)𝐚(1)Δ(w)h_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p,\mathbb{T}_{w1}^{m,\operatorname{\bf{a}}})\geq\operatorname{\bf{a}}(1)\Delta(w). On the other hand, for each iAi\in A

h𝕋m,𝐚(p,𝕋wim,𝐚)diam𝕋wim,𝐚=𝐚(i)Δ(w).h_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p,\mathbb{T}_{wi}^{m,\operatorname{\bf{a}}})\leq\operatorname{diam}\mathbb{T}_{wi}^{m,\operatorname{\bf{a}}}=\operatorname{\bf{a}}(i)\Delta(w).

Hence, h𝕋m,𝐚(p,𝕋wim,𝐚)=𝐚(i)Δ(w)h_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p,\mathbb{T}_{wi}^{m,\operatorname{\bf{a}}})=\operatorname{\bf{a}}(i)\Delta(w) for all i{1,,m}i\in\{1,\dots,m\}. By Lemma 6.7 it follows that h𝕋m,𝐚(p)=𝐚(3)Δ(w)h_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p)=\operatorname{\bf{a}}(3)\Delta(w) (note that Lemma 6.7 is stated for an arbitrary order of the branches at pp). For the second part of the statement, suppose that [1()]B𝕋m,𝐚i[1^{(\infty)}]\in B_{\mathbb{T}^{m,\operatorname{\bf{a}}}}^{i}. Let n0n\geq 0 be a maximal integer such that w=1(n)vw=1^{(n)}v, vA|w|nv\in A^{|w|-n}, v(1)1v(1)\neq 1. We have shown that

h𝕋m,𝐚(p,Bj)=diam𝕋wjm,𝐚=Δ(wj)=𝐚(j)Δ(w)2|w|1,h_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p,B_{j})=\operatorname{diam}\mathbb{T}_{wj}^{m,\operatorname{\bf{a}}}=\Delta(wj)=\operatorname{\bf{a}}(j)\Delta(w)\leq 2^{-|w|-1},

for all j{3,,m}j\in\{3,\dots,m\}. On the other hand,

h𝕋m,𝐚(p,Bi)\displaystyle h_{\mathbb{T}^{m,\operatorname{\bf{a}}}}(p,B_{i}) ρ([1()],[1(n)v12()])\displaystyle\geq\rho([1^{(\infty)}],[1^{(n)}v12^{(\infty)}])
=2nρ([1()],[v12()])\displaystyle=2^{-n}\rho([1^{(\infty)}],[v12^{(\infty)}])
=2n(ρ([1()],[12()])+ρ([12()],[v12()]))\displaystyle=2^{-n}\left(\rho([1^{(\infty)}],[12^{(\infty)}])+\rho([12^{(\infty)}],[v12^{(\infty)}])\right)
2n1,\displaystyle\geq 2^{-n-1},

where we used the geodesicity of 𝕋m,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}} and the fact that [12()][12^{(\infty)}] is on the arc that connects [1()],[v12()][1^{(\infty)}],[v12^{(\infty)}]. Due to n|w|n\leq|w|, it follows that i{1,2}i\in\{1,2\}. ∎

Proof of Proposition 11.8.

Similarly to the previous steps, applying [BM20, Theorem 1.2], we may assume that the tree (T,d)(T,d) is a geodesic doubling mm-valent metric tree with uniformly separated branch points and uniform branch growth (since these properties are quasisymmetrically invariant).

If TT has uniformly dense branch points, the statement is trivially true. Suppose TT does not have uniformly dense branch points. We use the decomposition V1V2\textbf{V}^{1}\subset\textbf{V}^{2}\subset\dots of TT as defined in [BM20, Section 5], which provides a countable set of double points dense in TT, where we plan on gluing branches of sufficient (new) height, so that the resulting space is a geodesic, doubling, uniformly mm-branching metric tree.

Let {v1,v2,}\{v_{1},v_{2},\dots\} be a fixed enumeration of V:=n=1Vn\textbf{V}^{*}:=\bigcup_{n=1}^{\infty}\textbf{V}^{n}. By Lemma 11.9 (vi) and (vii), each tile XXnX\in\textbf{X}^{n} contains at least 3 tiles of the next level Xn+1\textbf{X}^{n+1} and XVn\partial X\subset\textbf{V}^{n}. Hence, the sequence V1V2\textbf{V}^{1}\subset\textbf{V}^{2}\subset\dots is a strictly increasing sequence. Therefore, we can write V=n=1Wn\textbf{V}^{*}=\bigcup_{n=1}^{\infty}W_{n}, where the sets Wn:=𝐕n𝐕n1W_{n}:=\mathbf{V}^{n}\setminus\mathbf{V}^{n-1} are disjoint, have cardinality Mn:=cardWnM_{n}:=\operatorname{card}W_{n}\in\mathbb{N}, and assume that the aforementioned enumeration of 𝐕\mathbf{V}^{*} is the union of consecutive enumerations of each WnW_{n}, i.e., 𝐕={v11,,vM11,v12,,vM22,}\mathbf{V}^{*}=\{v_{1}^{1},\dots,v^{1}_{M_{1}},v_{1}^{2},\dots,v_{M_{2}}^{2},\dots\}. Thus, if xVx\in\textbf{V}^{*}, then there exist unique nn\in\mathbb{N} and i{1,,Mn}i\in\{1,\dots,M_{n}\} for which x=vinWnx=v_{i}^{n}\in W_{n}.

We plan on gluing m2m-2 isometric copies of cδn𝕋m,𝐚c\delta^{n}\mathbb{T}^{m,\operatorname{\bf{a}}} at [1()][1^{(\infty)}] and each double point xWnx\in W_{n}. Here, cc is a constant that solely depends on the comparability constant from (11.4). Thus, for every nn\in\mathbb{N}, and every i{1,,Mn}i\in\{1,\dots,M_{n}\}, set

S0=0,Sn:=M1++Mn,S_{0}=0,\ S_{n}:=M_{1}+\dots+M_{n},
Ωni={(Sn1+i1)(m2)+1,,(Sn1+i)(m2)},\Omega_{n}^{i}=\{(S_{n-1}+i-1)(m-2)+1,\dots,(S_{n-1}+i)(m-2)\},

and Ωn:=iΩni={Sn1(m2)+1,,Sn(m2)}\Omega_{n}:=\bigcup_{i}\Omega_{n}^{i}=\{S_{n-1}(m-2)+1,\dots,S_{n}(m-2)\}. We define

(Ykn,dk,n):=cδn𝕋m,𝐚×{k},(Y_{k}^{n},d_{k,n}):=c\delta^{n}\mathbb{T}^{m,\operatorname{\bf{a}}}\times\{k\},

for all kΩnk\in\Omega_{n}. Set ykn:=([1()],k)Ykny_{k}^{n}:=([1^{(\infty)}],k)\in Y_{k}^{n}, which we also denote by [1()]kn[1^{(\infty)}]_{k}^{n}, and write [2()]kn:=([2()],k)Ykn[2^{(\infty)}]_{k}^{n}:=([2^{(\infty)}],k)\in Y_{k}^{n}. Note that to define YknY_{k}^{n}, we briefly used the more detailed notation from Definition 11.2, in order to point out how at every double point xWnx\in W_{n} we need isometric, but distinct copies of the scaled metric tree cδn𝕋m,𝐚c\delta^{n}\mathbb{T}^{m,\operatorname{\bf{a}}}. Henceforth, we return to the convention where we identify points in YknY_{k}^{n} with points in cδn𝕋m,𝐚c\delta^{n}\mathbb{T}^{m,\operatorname{\bf{a}}}. In addition, it is understood that if nn is fixed, then kk is used to denote integers in Ωn\Omega_{n}, and we write Yk=YknY_{k}=Y_{k}^{n}, dk=dk,nd_{k}=d_{k,n}, yk=[1()]k=ykny_{k}=[1^{(\infty)}]_{k}=y_{k}^{n}, and [2()]k=[2()]kn[2^{(\infty)}]_{k}=[2^{(\infty)}]_{k}^{n}. On the other hand, if kk is an arbitrary positive integer, then the corresponding n=nkn=n_{k} is the unique integer such that kΩnk\in\Omega_{n}. For every kk\in\mathbb{N}, set uk=vin𝐕u_{k}=v_{i}^{n}\in\mathbf{V}^{*}, for all i=iki=i_{k} and the unique n=nkn=n_{k} with kΩnik\in\Omega_{n}^{i}. Set 𝐮:=(u1,u2,)\mathbf{u}_{\mathbb{N}}:=(u_{1},u_{2},\dots), where the terms are not necessarily distinct, due to the choice of uku_{k}. We claim that

(T,d):=(T,d,𝐮)k(Ykn,dk,n,ykn)(T^{\prime},d^{\prime}):=(T,d,\mathbf{u}_{\mathbb{N}})\bigvee_{k\in\mathbb{N}}(Y_{k}^{n},d_{k,n},y_{k}^{n})

is the space with the desired properties, into which TT isometrically embeds.

By Lemma 11.5, [BT21, Lemma 3.9], and Lemma 11.3, it can be shown that (T,d)(T^{\prime},d^{\prime}) is a geodesic metric tree (similarly to the proofs of Propositions 11.6, 11.7). By Lemma 11.5, we can assume that TT has uniform branch separation and uniform branch growth with respect to hTh_{T}. In order to address uniform branch growth and uniform branch separation for TT^{\prime}, we need to determine the height hTh_{T^{\prime}} of branch points of TT^{\prime}. To this end, let pTp\in T^{\prime} branch point. If p=vinWnp=v_{i}^{n}\in W_{n} for some nn\in\mathbb{N}, i{1,,Mn}i\in\{1,\dots,M_{n}\}, then, by Lemma 11.9 (i) and β1\beta\geq 1, the height of pp is determined by the distances within the corresponding Yk=YknY_{k}=Y_{k}^{n}, kΩnik\in\Omega_{n}^{i}, which are glued at pp. In particular, since all the glued trees YkY_{k} are isometric, we have for any kΩnik\in\Omega_{n}^{i} and j=jk=k(Sn1+i1)(m2)+2{3,,m}j=j_{k}=k-(S_{n-1}+i-1)(m-2)+2\in\{3,\dots,m\} that

hT(p,Bj)=hT([1()]k,Yk)dk([1()]k,[2()]k)=cδn.h_{T^{\prime}}(p,B_{j})=h_{T^{\prime}}([1^{(\infty)}]_{k},Y_{k})\geq d_{k}([1^{(\infty)}]_{k},[2^{(\infty)}]_{k})=c\delta^{n}.

On the other hand, hT([1()]k,Yk)cδndiam𝕋m,𝐚h_{T^{\prime}}([1^{(\infty)}]_{k},Y_{k})\leq c\delta^{n}\operatorname{diam}\mathbb{T}^{m,\operatorname{\bf{a}}}, which with the above implies hT(p,Bj)=cδnh_{T^{\prime}}(p,B_{j})=c\delta^{n}. If pp is a branch point of TT, then again by Lemma 11.9 (i), and similar arguments as in the proofs of Propositions 11.6, 11.7, we have that hT(p)=hT(p)h_{T^{\prime}}(p)=h_{T}(p). Lastly, if pp is a branch point of YknY_{k}^{n} for some nn\in\mathbb{N} and kΩnk\in\Omega_{n}, i.e., p=[w12()]kp=[w12^{(\infty)}]_{k} for some wAw\in A^{*}, then, by Lemma 11.10,

hT(p)=cδn𝐚(3)Δ(w)cδn.h_{T^{\prime}}(p)=c\delta^{n}\operatorname{\bf{a}}(3)\Delta(w)\leq c\delta^{n}.

Note that the only branch points of TT^{\prime} that are not branch points on TT or on some YknY_{k}^{n} are the points vin𝐕v_{i}^{n}\in\mathbf{V}^{*}. Since TT and YknY_{k}^{n} have uniform branch growth, and hT(vin,Bj)=cδnh_{T^{\prime}}(v_{i}^{n},B_{j})=c\delta^{n} for all j{3,,m}j\in\{3,\dots,m\}, it follows that TT^{\prime} also has uniform branch growth (note that by Remark 6.10, the uniform branch growth constant is independent of diam𝕋m,𝐚\operatorname{diam}\mathbb{T}^{m,\operatorname{\bf{a}}} and, hence, independent of diamYkn\operatorname{diam}Y_{k}^{n}).

Let p1,p2p_{1},p_{2} be two distinct branch points of TT^{\prime}, and let c1,c2c_{1},c_{2} be the uniform branch separation constants of TT and YknY_{k}^{n}, respectively (note that by Remark 6.10, c2c_{2} is independent of diamYkn\operatorname{diam}Y_{k}^{n} and, thus, independent of k,nk,n). Suppose that p1Tp_{1}\in T. If p2T𝐕p_{2}\in T\setminus\mathbf{V}^{*}, then

d(p1,p2)=d(p1,p2)c1min{hT(p1),hT(p2)}=c1min{hT(p1),hT(p2)}.d^{\prime}(p_{1},p_{2})=d(p_{1},p_{2})\geq c_{1}\min\{h_{T}(p_{1}),h_{T}(p_{2})\}=c_{1}\min\{h_{T^{\prime}}(p_{1}),h_{T^{\prime}}(p_{2})\}.

If p2Wnp_{2}\in W_{n} for some nn\in\mathbb{N}, then, by Lemma 11.9(ii),

d(p1,p2)cγmin{hT(p1),cδn}=cγmin{hT(p1),hT(p2)}.d^{\prime}(p_{1},p_{2})\geq c\gamma\min\{h_{T^{\prime}}(p_{1}),c\delta^{n}\}=c\gamma\min\{h_{T^{\prime}}(p_{1}),h_{T^{\prime}}(p_{2})\}.

Similarly, if p2Yknp_{2}\in Y_{k}^{n} for some n,kn,k\in\mathbb{N}, and v=ukWnv=u_{k}\in W_{n} is the nn-vertex at which we glue YknY_{k}^{n}, it follows that

d(p1,p2)d(p1,v)cγmin{hT(p1),cδn}cγmin{hT(p1),hT(p2)}.d^{\prime}(p_{1},p_{2})\geq d^{\prime}(p_{1},v)\geq c\gamma\min\{h_{T^{\prime}}(p_{1}),c\delta^{n}\}\geq c\gamma\min\{h_{T^{\prime}}(p_{1}),h_{T^{\prime}}(p_{2})\}.

Suppose that p1Wnp_{1}\in W_{n} for some nn\in\mathbb{N}. If p2Wnp_{2}\in W_{n^{\prime}} for some nn^{\prime}\in\mathbb{N}, then, without loss of generality, assume that nnn\geq n^{\prime}. By Lemma 11.9 (v), p2p_{2} lies in some nn-tile XX. Hence, by Lemma 11.9(iii),

d(p1,p2)δnhT(p1)=min{hT(p1),hT(p2)}.d^{\prime}(p_{1},p_{2})\geq\delta^{n}\geq h_{T^{\prime}}(p_{1})=\min\{h_{T^{\prime}}(p_{1}),h_{T^{\prime}}(p_{2})\}.

If p2Yknp_{2}\in Y_{k^{\prime}}^{n^{\prime}} for some n,kn^{\prime},k^{\prime}\in\mathbb{N}, then let v=uk𝐕v^{\prime}=u_{k^{\prime}}\in\mathbf{V}^{*} be the unique nn^{\prime}-vertex to which we glue YknY_{k^{\prime}}^{n^{\prime}}. If vp1v^{\prime}\neq p_{1}, then, by geodesicity and similar arguments to the previous case, it follows that

d(p1,p2)d(p1,v)δnmin{hT(p1),hT(v)}min{hT(p1),hT(p2)}.d^{\prime}(p_{1},p_{2})\geq d^{\prime}(p_{1},v^{\prime})\geq\delta^{n}\geq\min\{h_{T^{\prime}}(p_{1}),h_{T^{\prime}}(v^{\prime})\}\geq\min\{h_{T^{\prime}}(p_{1}),h_{T^{\prime}}(p_{2})\}.

If v=p1v^{\prime}=p_{1}, by similar arguments to those in the proof of Lemma 11.10,

d(p1,p2)=dk,n([1()]k,p2)hYkn(p2)=hT(p2)=min{hT(p1),hT(p2)}.d^{\prime}(p_{1},p_{2})=d_{k^{\prime},n^{\prime}}([1^{(\infty)}]_{k^{\prime}},p_{2})\geq h_{Y_{k^{\prime}}^{n^{\prime}}}(p_{2})=h_{T^{\prime}}(p_{2})=\min\{h_{T^{\prime}}(p_{1}),h_{T^{\prime}}(p_{2})\}.

If p1Yknp_{1}\in Y_{k}^{n} and p2Yknp_{2}\in Y_{k^{\prime}}^{n^{\prime}}, for n,n,k,kn,n^{\prime},k,k^{\prime}\in\mathbb{N}, then

d(p1,p2)c2min{hT(p1),hT(p2)},d^{\prime}(p_{1},p_{2})\geq c_{2}\min\{h_{T^{\prime}}(p_{1}),h_{T^{\prime}}(p_{2})\},

which follows similarly to previous cases, so we omit the details. Therefore, TT^{\prime} has uniformly separated branch points. By Lemma 2.5, TT^{\prime} is doubling.

It remains to show that TT^{\prime} has uniformly dense branch points. Let x,yTx,y\in T^{\prime} with xyx\neq y, and let c3c_{3} be the uniform density constant of YknY_{k}^{n} for all n,kn,k\in\mathbb{N} (note that by Remark 6.10, c3c_{3} is independent of k,nk,n). If x,yTx,y\in T, then let n0n\geq 0 be the maximal integer such that x,yXx,y\in X, where XX is an nn-tile. Hence, by Lemma 11.9 (v), we have xX1x\in X_{1} and yX2y\in X_{2}, where X1,X2XX_{1},X_{2}\subset X are (n+1)(n+1)-tiles with X1X2X_{1}\neq X_{2}. Since x,yx,y cannot both lie on the boundary of the same (n+1)(n+1)-tile, we may assume that xX1x\notin\partial X_{1}. Note that by Lemma 11.9 (vii), XX is a subtree of TT^{\prime}, which implies [x,y]X[x,y]\subset X. We claim that the appropriate branch point of TT^{\prime} in [x,y][x,y] for the uniform density condition is the unique point pp in X1[x,y]\partial X_{1}\cap[x,y]. Since X1X_{1} is an (n+1)(n+1)-tile, we have that pVn+1p\in\textbf{V}^{n+1}, by Lemma 11.9 (vii). If pVnp\in\textbf{V}^{n}, then pXp\in\partial X, by Lemma 11.9 (viii). Therefore, [p,y](TX)[p,y]\cap(T^{\prime}\setminus X)\neq\emptyset, which is a contradiction, since [p,y][x,y]X[p,y]\subset[x,y]\subset X. Hence, pVn+1Vn=Wn+1p\in\textbf{V}^{n+1}\setminus\textbf{V}^{n}=W_{n+1}, which implies that hT(p)=cδn+1h_{T^{\prime}}(p)=c\delta^{n+1}. Also, by geodecisity, the diameter of the tile XX is equal to the sum of diameters of all (n+1)(n+1)-tiles contained in XX. By Lemma 11.9 (iv) and (ix), this implies that

d(x,y)K3βδn+1=3Kc1βhT(p),d^{\prime}(x,y)\leq K3\beta\delta^{n+1}=3Kc^{-1}\beta h_{T^{\prime}}(p),

for some constant KK depending only on the doubling constant of the tree TT.

Suppose that xTx\in T and yYkny\in Y_{k}^{n} for some n,kn,k\in\mathbb{N}. Let v=ukWnv=u_{k}\in W_{n} be the vertex to which we glue YknY_{k}^{n}. Suppose first that xvx\neq v. Let n0n^{\prime}\geq 0 be the maximal integer such that x,vZx,v\in Z, where ZZ is a nn^{\prime}-tile. If n<nn^{\prime}<n, by geodesicity, and by applying the previous case to x,vx,v, we have

d(x,y)=d(x,v)+d(v,y)K3βδn+1+diamYkn(3Kc1β+1)hT(p),d^{\prime}(x,y)=d^{\prime}(x,v)+d^{\prime}(v,y)\leq K3\beta\delta^{n^{\prime}+1}+\operatorname{diam}Y_{k}^{n}\leq(3Kc^{-1}\beta+1)h_{T^{\prime}}(p),

where pp is the branch point lying on the intersection of [x,v][x,v] and the boundary of the (n+1)(n^{\prime}+1)-tile that contains xx. On the other hand, if nnn^{\prime}\geq n, we claim that vv is the desired branch point of TT^{\prime}. Indeed, by hT(v)=cδnh_{T^{\prime}}(v)=c\delta^{n} and Lemma 11.9 (iv),

d(x,y)=d(x,v)+d(v,y)3βδn+diamYkn(3c1β+1)hT(v).d^{\prime}(x,y)=d^{\prime}(x,v)+d^{\prime}(v,y)\leq 3\beta\delta^{n^{\prime}}+\operatorname{diam}Y_{k}^{n}\leq(3c^{-1}\beta+1)h_{T^{\prime}}(v).

Suppose that x=vx=v. Then d(x,y)=d(v,y)diamYkn=cδn=hT(v)d^{\prime}(x,y)=d^{\prime}(v,y)\leq\operatorname{diam}Y_{k}^{n}=c\delta^{n}=h_{T^{\prime}}(v).

Lastly, if xYknx\in Y_{k}^{n} and yYkny\in Y_{k^{\prime}}^{n^{\prime}}, for n,k,n,kn,k,n^{\prime},k^{\prime}\in\mathbb{N}, it can be shown that

d(x,y)=d(x,v)+d(v,y)max{c3,(6Kc1β+1)}hT(p),d^{\prime}(x,y)=d^{\prime}(x,v)+d^{\prime}(v,y)\leq\max\{c_{3},(6Kc^{-1}\beta+1)\}h_{T^{\prime}}(p),

for some branch point pTp\in T^{\prime}, with similar case studies as in the previous cases.

It follows by all the above cases that TT^{\prime} has uniformly dense branch points. Therefore, TT^{\prime} is uniformly mm-branching. ∎

12. Proofs of Theorem 1.3 and Theorem 1.9

We are finally able to give the proof of Theorem 1.3 and Theorem 1.9. Fix for the rest of this section a number ϵ>0\epsilon>0. Fix also a weight 𝐚\operatorname{\bf{a}} such that n=1𝐚(n)1+ϵ<1\sum_{n=1}^{\infty}\operatorname{\bf{a}}(n)^{1+\epsilon}<1.

Recall from §5.1 our convention that 𝕋m,𝐚𝕋n,𝐚𝕋,𝐚\mathbb{T}^{m,\operatorname{\bf{a}}}\subset\mathbb{T}^{n,\operatorname{\bf{a}}}\subset\mathbb{T}^{\infty,\operatorname{\bf{a}}} if 2nm2\leq n\leq m are integers. Set 𝕋n=𝕋n,𝐚\mathbb{T}^{n}=\mathbb{T}^{n,\operatorname{\bf{a}}} for n{2,3,}n\in\{2,3,\dots\} and 𝕋=𝕋,𝐚\mathbb{T}^{\infty}=\mathbb{T}^{\infty,\operatorname{\bf{a}}}.

By Lemma 4.9, and the argument following the lemma, each 𝕋n\mathbb{T}^{n} is self-similar in the sense of §2.2. By Lemma 4.8, each 𝕋n\mathbb{T}^{n} and 𝕋\mathbb{T}^{\infty} are metric trees, and by Proposition 5.1 each 𝕋n\mathbb{T}^{n} and 𝕋\mathbb{T}^{\infty} are geodesic trees. By Proposition 8.1, each 𝕋n\mathbb{T}^{n} is Ahlfors regular and the Hausdorff dimension of 𝕋\mathbb{T}^{\infty} is at most 1+ϵ1+\epsilon.

By Theorem 9.1, a metric space TT is an uniformly nn-branching quasiconformal tree if and only if TT is quasisymmetric to 𝕋n\mathbb{T}^{n}. This proves Theorem 1.9.

By Proposition 10.1 each 𝕋n\mathbb{T}^{n} bi-Lipschitz embeds into 2\mathbb{R}^{2} with the embedded image being quasiconvex. By Proposition 5.1 and Proposition 6.1, 𝕋2\mathbb{T}^{2} is isometric to the Euclidean unit interval [0,1][0,1], while the CSST is quasisymmetric to 𝕋3\mathbb{T}^{3} by Theorem 1.9 and the fact that the CSST is uniformly 3-branching [BM22]. Lastly, by Proposition 7.1, the Vicsek fractal is uniformly 4-branching, so it is quasisymmetric equivalent to 𝕋4\mathbb{T}^{4}. This proves Theorem 1.3(i).

By Lemma 5.10, spaces 𝕋n\mathbb{T}^{n} converge to 𝕋\mathbb{T}^{\infty} as nn\to\infty in the Gromov-Hausdorff sense. This proves Theorem 1.3(ii).

Lastly, let TT be a quasiconformal metric tree in 𝒬𝒞𝒯(n)\mathscr{QCT}^{*}(n). By Proposition 11.1, TT quasisymmetrically embeds into a uniformly nn-branching quasiconformal tree TT^{\prime}. By Theorem 1.9 TT^{\prime} is quasisymetrically equivalent to 𝕋n\mathbb{T}^{n}. Hence, TT quasisymmetrically embeds into 𝕋n\mathbb{T}^{n}. This proves Theorem 1.3(iii).

References

  • [Ald91a] David Aldous. The continuum random tree. I. Ann. Probab., 19(1):1–28, 1991.
  • [Ald91b] David Aldous. The continuum random tree. II. An overview. In Stochastic analysis (Durham, 1990), volume 167 of London Math. Soc. Lecture Note Ser., pages 23–70. Cambridge Univ. Press, Cambridge, 1991.
  • [Azz15] Jonas Azzam. Hausdorff dimension of wiggly metric spaces. Arkiv för Matematik, 53(1):1 – 36, 2015.
  • [BBI01] Dmitri Burago, Yuri Burago, and Sergei Ivanov. A course in metric geometry, volume 33 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2001.
  • [BC23a] Fabrice Baudoin and Li Chen. Heat kernel gradient estimates for the Vicsek set. arXiv preprint, 2023.
  • [BC23b] Fabrice Baudoin and Li Chen. Sobolev spaces and Poincaré inequalities on the Vicsek fractal. Ann. Fenn. Math., 48(1):3–26, 2023.
  • [Bes84] Mladen Bestvina. Characterizing kk-dimensional universal Menger compacta. Bull. Amer. Math. Soc. (N.S.), 11(2):369–370, 1984.
  • [Bes02] Mladen Bestvina. \mathbb{R}-trees in topology, geometry, and group theory. In Handbook of geometric topology, pages 55–91. North-Holland, Amsterdam, 2002.
  • [Bis14] Christopher J. Bishop. True trees are dense. Inventiones mathematicae, 197(2):433–452, 2014.
  • [BM20] Mario Bonk and Daniel Meyer. Quasiconformal and geodesic trees. Fund. Math., 250(3):253–299, 2020.
  • [BM22] Mario Bonk and Daniel Meyer. Uniformly branching trees. Trans. Amer. Math. Soc., 375(6):3841–3897, 2022.
  • [BT01] Christopher J. Bishop and Jeremy T. Tyson. Conformal dimension of the antenna set. Proc. Amer. Math. Soc., 129(12):3631–3636, 2001.
  • [BT21] Mario Bonk and Huy Tran. The continuum self-similar tree. In Fractal geometry and stochastics VI, volume 76 of Progr. Probab., pages 143–189. Birkhäuser/Springer, Cham, 2021.
  • [CG93] Lennart Carleson and Theodore W. Gamelin. Complex dynamics. Universitext: Tracts in Mathematics. Springer-Verlag, New York, 1993.
  • [Cha80] Janusz J. Charatonik. Open mappings of universal dendrites. Bull. Acad. Polon. Sci. Sér. Sci. Math., 28(9-10):489–494, 1980.
  • [CJY94] Lennart Carleson, Peter W. Jones, and Jean-Christophe Yoccoz. Julia and John. Bol. Soc. Brasil. Mat. (N.S.), 25(1):1–30, 1994.
  • [CSW11] Sarah Constantin, Robert S. Strichartz, and Miles Wheeler. Analysis of the Laplacian and spectral operators on the Vicsek set. Commun. Pure Appl. Anal., 10(1):1–44, 2011.
  • [DEBV23] Guy C. David, Sylvester Eriksson-Bique, and Vyron Vellis. Bi-Lipschitz embeddings of quasiconformal trees. Proc. Amer. Math. Soc., 151(5):2031–2044, 2023.
  • [DV22] Guy C. David and Vyron Vellis. Bi-Lipschitz geometry of quasiconformal trees. Illinois J. Math., 66(2):189–244, 2022.
  • [FG23a] David Freeman and Chris Gartland. Lipschitz functions on quasiconformal trees. Fund. Math., 262(2):153–203, 2023.
  • [FG23b] David Freeman and Chris Gartland. Lipschitz functions on unions and quotients of metric spaces. Studia Math., 273(1):29–61, 2023.
  • [GH12] Frederick W. Gehring and Kari Hag. The ubiquitous quasidisk, volume 184 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2012. With contributions by Ole Jacob Broch.
  • [Gro81] Mikhael Gromov. Groups of polynomial growth and expanding maps. Inst. Hautes Études Sci. Publ. Math., (53):53–73, 1981.
  • [Gro99] Misha Gromov. Metric structures for Riemannian and non-Riemannian spaces, volume 152 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1999. Based on the 1981 French original [MR0682063 (85e:53051)], With appendices by M. Katz, P. Pansu and S. Semmes, Translated from the French by Sean Michael Bates.
  • [GT11] Anupam Gupta and Kunal Talwar. Making doubling metrics geodesic. Algorithmica, 59(1):66–80, 2011.
  • [Hei01] Juha Heinonen. Lectures on analysis on metric spaces. Universitext. Springer-Verlag, New York, 2001.
  • [HM98] B. M. Hambly and V. Metz. The homogenization problem for the Vicsek set. Stochastic Process. Appl., 76(2):167–190, 1998.
  • [Hut81] John E. Hutchinson. Fractals and self-similarity. Indiana Univ. Math. J., 30(5):713–747, 1981.
  • [Kin17] Kyle Kinneberg. Conformal dimension and boundaries of planar domains. Trans. Amer. Math. Soc., 369(9):6511–6536, 2017.
  • [LG06] Jean-Fran¸cois Le Gall. Random real trees. Ann. Fac. Sci. Toulouse Math. (6), 15(1):35–62, 2006.
  • [Li21] Wenbo Li. Quasisymmetric embeddability of weak tangents. Ann. Fenn. Math., 46(2):909–944, 2021.
  • [LNP09] James R. Lee, Assaf Naor, and Yuval Peres. Trees and Markov convexity. Geom. Funct. Anal., 18(5):1609–1659, 2009.
  • [LR18] Peter Lin and Steffen Rohde. Conformal welding of dendrites. preprint, 2018.
  • [McM98] Curtis T. McMullen. Kleinian groups and John domains. Topology, 37(3):485–496, 1998.
  • [Met93] Volker Metz. How many diffusions exist on the Vicsek snowflake? Acta Appl. Math., 32(3):227–241, 1993.
  • [Nad92] Sam B. Nadler, Jr. Continuum theory, volume 158 of Monographs and Textbooks in Pure and Applied Mathematics. Marcel Dekker, Inc., New York, 1992. An introduction.
  • [NV91] Raimo Näkki and Jussi Väisälä. John disks. Exposition. Math., 9(1):3–43, 1991.
  • [Sea07] Mícheál Ó. Searcóid. Metric spaces. Springer Undergraduate Mathematics Series. Springer-Verlag London, Ltd., London, 2007.
  • [Sta71] M. A. Stanko. Solution of Menger’s problem in the class of compacta. Dokl. Akad. Nauk SSSR, 201:1299–1302, 1971.
  • [Str93] Daniel W. Stroock. Probability theory, an analytic view. Cambridge University Press, Cambridge, 1993.
  • [TV80] P. Tukia and J. Väisälä. Quasisymmetric embeddings of metric spaces. Ann. Acad. Sci. Fenn. Ser. A I Math., 5(1):97–114, 1980.
  • [Vic83] T Vicsek. Fractal models for diffusion controlled aggregation. Journal of Physics A: Mathematical and General, 16(17):L647, dec 1983.
  • [Why63] Gordon Thomas Whyburn. Analytic topology, volume Vol. XXVIII of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 1963.
  • [Zho09] Denglin Zhou. Spectral analysis of Laplacians on the Vicsek set. Pacific J. Math., 241(2):369–398, 2009.