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

Decomposing multitwists

Alastair N. Fletcher [email protected] Department of Mathematical Sciences, Northern Illinois University, Dekalb, IL 60115, USA  and  Vyron Vellis [email protected] Department of Mathematics, The University of Tennessee, Knoxville, TN 37966, USA
Abstract.

The Decomposition Problem in the class LIP(𝕊2)LIP(\mathbb{S}^{2}) is to decompose any bi-Lipschitz map f:𝕊2𝕊2f:\mathbb{S}^{2}\to\mathbb{S}^{2} as a composition of finitely many maps of arbitrarily small isometric distortion. In this paper, we construct a decomposition for certain bi-Lipschitz maps which spiral around every point of a Cantor set XX of Assouad dimension strictly smaller than one. These maps are constructed by considering a collection of Dehn twists on the Riemann surface 𝕊2X\mathbb{S}^{2}\setminus X. The decomposition is then obtained via a bi-Lipschitz path which simultaneously unwinds these Dehn twists. As part of our construction, we also show that X𝕊2X\subset\mathbb{S}^{2} is uniformly disconnected if and only if the Riemann surface 𝕊2X\mathbb{S}^{2}\setminus X has a pants decomposition whose cuffs have hyperbolic length uniformly bounded above, which may be of independent interest.

Key words and phrases:
uniformly disconnected set, bi-Lipschitz decomposition, hyperbolic surface of infinite type, bi-Lipschitz path
2020 Mathematics Subject Classification:
Primary 30L05; Secondary 30F45, 37C10
A.F. was supported by a grant from the Simons Foundation, #352034. V.V. was partially supported by the NSF DMS grant 1952510.

1. Introduction

A bi-Lipschitz homeomorphism f:XYf:X\to Y between metric spaces is a homeomorphism that roughly preserves absolute distances; specifically, there exists L1L\geq 1 such that

L1dX(x,y)dY(f(x),f(y))LdX(x,y)L^{-1}d_{X}(x,y)\leq d_{Y}(f(x),f(y))\leq Ld_{X}(x,y)

for all x,yXx,y\in X. We then say that ff is an LL-bi-Lipschitz map. The smallest such constant LL is called the isometric distortion of ff. Letting 𝕊n\mathbb{S}^{n} be the sphere of dimension nn, we denote by LIP(𝕊n)LIP(\mathbb{S}^{n}) the class of orientation preserving homeomorphisms of 𝕊n\mathbb{S}^{n}.

A central problem in bi-Lipschitz geometry is whether a bi-Lipschitz map can be decomposed into bi-Lipschitz mappings of arbitrarily small isometric distortion.

Conjecture 1.1 (Decomposition Problem).

Let n1n\geq 1 and let fLIP(𝕊n)f\in LIP(\mathbb{S}^{n}). Then for every ϵ>0\epsilon>0 we can find homeomorphisms fkLIP(𝕊n)f_{k}\in LIP(\mathbb{S}^{n}), for k=1,,mk=1,\ldots,m, such that ff can be written as a composition f=fmf1f=f_{m}\circ\ldots\circ f_{1}, where each fkf_{k} has isometric distortion at most 1+ϵ1+\epsilon.

The case n=1n=1 is elementary: suppose I,JI,J are intervals in \mathbb{R} and f:IJf:I\to J is an LL-bi-Lipschitz map. Then ff can be written as f=f2f1f=f_{2}\circ f_{1}, where

f1(x)=x0x|f(t)|λ𝑑t,f_{1}(x)=\int_{x_{0}}^{x}|f^{\prime}(t)|^{\lambda}\>dt,

x0x_{0} is fixed, λ=logLα\lambda=\log_{L}\alpha, f1f_{1} is α\alpha-bi-Lipschitz and f2=ff11f_{2}=f\circ f_{1}^{-1} is L/αL/\alpha-bi-Lipschitz.

However, for n2n\geq 2, the Decomposition Problem has been so far elusive. It is clear that affine bi-Lipschitz mappings can be factored into affine mappings of small isometric distortion, but beyond this, only certain specific examples have been considered. Freedman and He [FH88] studied the logarithmic spiral map sk(z)=zeiklog|z|s_{k}(z)=ze^{ik\log|z|}, which is an LL-bi-Lipschitz map with |k|=L1/L|k|=L-1/L. Gutlyanskii and Martio [GM01] studied a related class of mappings in dimension 22, and generalized this to a class of volume preserving bi-Lipschitz automorphisms of the unit ball 𝔹3\mathbb{B}^{3} in three dimensions.

Although in this paper we focus on LIP(𝕊2)LIP(\mathbb{S}^{2}), the Decompsition Problem can also be asked for the class of quasiconformal homeomorphisms of 𝕊n\mathbb{S}^{n}. In dimension 22, the fact that every quasiconformal map arises as a solution of the Beltrami equation can be leveraged to show that the Decomposition Problem has a positive solution here; see [Leh87, Theorem 4.7]. Since every orientation preserving bi-Lipschitz map is also quasiconformal, in dimension 22 we are able to find a decomposition of bi-Lipschitz maps, but only into quasiconformal maps of small conformal distortion. Observe, however, that quasiconformal maps need not be bi-Lipschitz.

A similar problem was studied by the first named author and Markovic in [FM12]. There it was shown that C1C^{1} diffeomorphisms of 𝕊n\mathbb{S}^{n}, for n2n\geq 2, can be decomposed into bi-Lipschitz maps of arbitrarily small isometric distortion. This solves the Decomposition Problem for C1C^{1} bi-Lipschitz maps, but of course, bi-Lipschitz maps are only guaranteed to be differentiable almost everywhere.

In this paper, we study the Decomposition Problem for a class of maps in LIP(𝕊2)LIP(\mathbb{S}^{2}) which spiral around every point of a Cantor set, with small Assouad dimension; see below for definitions. Necessarily these maps are not differentiable at any point of the Cantor set in question. This can be viewed as a generalization of the result of Freedman and He, although they were motivated to give estimates on the number of maps required in the decomposition. Our constructions will be involved enough that we will not address this question here, and be content to just find a decomposition.

Maps which spiral around every point of a Cantor set simultaneously are not new. Such mappings were constructed by Astala et al in [AIPS15] in order to give sharp examples of the multifractal spectrum; see in particular the proof of Theorem 5.1 and Figure 7 in [AIPS15].

1.1. Uniformly disconnected sets and hyperbolic geometry

We identify the topological sphere 𝕊2\mathbb{S}^{2} with the one point compactification 2{}\mathbb{R}^{2}\cup\{\infty\}, and equip it with the chordal metric. If X𝕊2X\subset\mathbb{S}^{2} is a Cantor set, then by applying a chordal isometry we may assume that X2X\subset\mathbb{R}^{2}. Having done this, we may then view S:=𝕊2XS:=\mathbb{S}^{2}\setminus X as a Riemann surface of infinite type.

The bi-Lipschitz maps that we will decompose arise from a collection of Dehn twists on the surface SS. For the mappings we define to be bi-Lipschitz, we need some control on the ring domains on which the Dehn twists are defined. Informally, these ring domains cannot be too thin, and their boundaries cannot be too wiggly.

To address the first of these points, we recall some hyperbolic geometry. The surface SS has a pants decomposition, that is, S=i=1PiS=\bigcup_{i=1}^{\infty}P_{i}, where each PiP_{i} is a topological sphere with three disks removed. The collection of boundary curves of the pairs of pants, called the cuffs of the decomposition, may be enumerated by (αj)j=1(\alpha_{j})_{j=1}^{\infty}. Each αj\alpha_{j} is a simple closed curve on SS and generates a class [αj][\alpha_{j}] of simple closed curves that are freely homotopic to αj\alpha_{j}.

We denote by S(αj)\ell_{S}(\alpha_{j}) the hyperbolic length of αj\alpha_{j} and by S[αj]\ell_{S}[\alpha_{j}] the infimum of hyperbolic lengths of closed curves in SS homotopic to αj\alpha_{j}. We suppress the subscript SS if the context is clear. It is well-known a Cantor set X2X\subset\mathbb{R}^{2} is uniformly perfect if and only if for any pants decomposition of 𝕊2X\mathbb{S}^{2}\setminus X, the associated cuffs (αj)j=1(\alpha_{j})_{j=1}^{\infty} satisfy infjS[αj]>0\inf_{j}\ell_{S}[\alpha_{j}]>0; see [Pom79]. Recall that a non-degenerate metric space XX is uniformly perfect if there exists a constant C>1C>1 such that for any xXx\in X and every positive r<diamXr<\operatorname{diam}{X}, we have that B(x,r)B(x,r/C)B(x,r)\setminus B(x,r/C)\neq\emptyset. Informally, this means that any ring domain separating XX cannot be too thick.

Uniform disconnectedness is, in a sense, the opposite of uniform perfectness; a metric space XX is uniformly disconnected if there exists a constant C>1C>1 such that for any xXx\in X and every positive r<diamXr<\operatorname{diam}{X}, there exists Xx,rXX_{x,r}\subset X that contains xx such that diamXx,rr\operatorname{diam}{X_{x,r}}\leq r and

dist(Xx,r,XXx,r)r/C.\operatorname{dist}(X_{x,r},X\setminus X_{x,r})\geq r/C.

It is natural to ask whether XX being uniformly disconnected implies analogous geometric properties of the surface SS. Our first result gives such a characterization.

Theorem 1.2.

A Cantor set X2X\subset\mathbb{R}^{2} is uniformly disconnected if and only if there exists a pants decomposition for S=𝕊2XS=\mathbb{S}^{2}\setminus X such that the associated cuffs (αj)j=1(\alpha_{j})_{j=1}^{\infty} satisfy supj[αj]<\sup_{j}\ell[\alpha_{j}]<\infty.

By a uniformization theorem of David and Semmes [DS97], a set X2X\subset\mathbb{R}^{2} is quasisymmetrically homeomorphic to the standard ternary Cantor set 𝒞\mathcal{C} if and only if it is compact, uniformly perfect, and uniformly disconnected. Therefore, by Theorem 1.2 and [Pom79], it follows that a Cantor set X2X\subset\mathbb{R}^{2} is quasisymmetrically homeomorphic to 𝒞\mathcal{C} if and only if there exists a constant C>1C>1 and a pants decomposition for S=𝕊2XS=\mathbb{S}^{2}\setminus X such that the associated cuffs (αj)j=1(\alpha_{j})_{j=1}^{\infty} satisfy

C1[αj]C,for all j.C^{-1}\leq\ell[\alpha_{j}]\leq C,\qquad\text{for all $j$}.

1.2. Dehn multi-twists

Here we outline how our bi-Lipschitz mappings are constructed. Full definitions and discussion will follow in the sequel. The first step is the following proposition which is a corollary of Theorem 1.2.

Proposition 1.3.

Given c1c\geq 1, there exists L>1L>1, kk\in\mathbb{N}, and a finite set

{gi:B¯(0,1)B(0,11L)2}i=1k\{g_{i}:\overline{B}(0,1)\setminus B(0,1-\tfrac{1}{L})\to\mathbb{R}^{2}\}_{i=1}^{k}

of LL-bi-Lipschitz conformal maps with the following property. Let X2X\subset\mathbb{R}^{2} be a cc-uniformly disconnected Cantor set and let (αj)j=1(\alpha_{j})_{j=1}^{\infty} be the cuffs from Theorem 1.2. There exist mutually disjoint closed ring domains Rj2XR_{j}\subset\mathbb{R}^{2}\setminus X homotopic to αj\alpha_{j}, and similarities (ϕj)j=1(\phi_{j})_{j=1}^{\infty} of 2\mathbb{R}^{2} such that for each jj\in\mathbb{N} there exists i(j){1,,k}i(j)\in\{1,\dots,k\} with Rj:=ϕjgi(j)(B¯(0,1)B(0,11L))R_{j}:=\phi_{j}\circ g_{i(j)}(\overline{B}(0,1)\setminus B(0,1-\frac{1}{L})). Moreover, for each jj\in\mathbb{N}, the bounded component of ϕj1(Rj)\phi_{j}^{-1}(R_{j}) has diameter equal to 1.

This proposition says that given a pants decomposition of 𝕊2X\mathbb{S}^{2}\setminus X, we can find a collection of rings on which our map ff will be supported with the property that, up to similarity, the rings are chosen from a finite set. This finiteness will lead to a certain uniformity in the Dehn twists that define ff.

More precisely, fix a Cantor set X2X\subset\mathbb{R}^{2} for which the Assouad dimension satisfies dimAX<1\dim_{A}X<1. It then follows from [Luu98] that XX is uniformly disconnected. Let RjR_{j} and fj:=ϕjgi(j)f_{j}:=\phi_{j}\circ g_{i(j)} be the ring domains and conformal maps, respectively, from Proposition 1.3. Then, a Dehn twist can be defined on each Rj¯\overline{R_{j}} by

f|Rj¯:=fj𝔇fj1f|\overline{R_{j}}:=f_{j}\circ\mathfrak{D}\circ f_{j}^{-1}

where 𝔇:B¯(0,1)B(0,11L)B¯(0,1)B(0,11L)\mathfrak{D}:\overline{B}(0,1)\setminus B(0,1-\frac{1}{L})\to\overline{B}(0,1)\setminus B(0,1-\frac{1}{L}) is the Dehn twist

𝔇(r,θ)=(r,θ+2πL(1r)).\mathfrak{D}(r,\theta)=(r,\theta+2\pi L(1-r)).

Let f:22f:\mathbb{R}^{2}\to\mathbb{R}^{2} be given by the Dehn twist in each Rj¯\overline{R_{j}} as above, and the identity elsewhere. The uniform bi-Lipschitz constant of maps gjg_{j} guarantees that ff is a bi-Lipschitz map; see Lemma 6.2. The main theorem of this paper reads as follows:

Theorem 1.4.

If X2X\subset\mathbb{R}^{2} is a Cantor set with dimA(X)<1\dim_{A}(X)<1 and if ff is the bi-Lipschitz map defined above, then given ϵ>0\epsilon>0, there exists NN\in\mathbb{N} such that f=fNf1f=f_{N}\circ\ldots\circ f_{1}, where each fjf_{j}, for j=1,,Nj=1,\ldots,N, is (1+ϵ)(1+\epsilon)-bi-Lipschitz.

It is worth pointing out that if the rings RjR_{j} can initially be chosen to be round rings, such as those constructed in [AIPS15], then the assumption dimA(X)<1\dim_{A}(X)<1 can be replaced by uniform disconnectedness, and we can decompose ff directly in this case. In fact, the assumption dimAX<1\dim_{A}{X}<1 can be dropped, see Section 8 for an example), and we conjecture that it can be replaced by uniform disconnectedness.

1.3. Strategy of the proof

The crux of the proof is to construct a bi-Lipschitz path from the identity to ff. Bi-Lipschitz paths were introduced in [FM12] to provide a way to deform one bi-Lipschitz mapping to another in a controlled way. Partitioning the path into small subintervals yields the required decomposition.

Consider first the special case where each of the rings RjR_{j} are round, see Figure 1. Writing VjV_{j} for the bounded component of the complement of RjR_{j}, we can unwind the Dehn twist supported in RjR_{j} in the obvious way, and extend this unwinding via the identity in the unbounded component of the complement of RjR_{j} and via a path of rotations in VjV_{j}. This unwinding can happen in each ring RjR_{j} and the corresponding domain VjV_{j} simultaneously for all jj. The point is that on a given RjR_{j}, the unwinding will act via finitely many rotations (one for each ring RkR_{k} such that RjVkR_{j}\subset V_{k}) and then via the unwinding on RjR_{j}.

Refer to caption
Figure 1. Round rings and Dehn twists

This idealized case is, however, not the most general case. Complications arise once RjR_{j} are not round rings. In particular, it may certainly be the case that the two rings Rk1,Rk2R_{k_{1}},R_{k_{2}} contained in VjV_{j}, cannot be flowed isometrically around DjD_{j}, see Figure 2

Refer to caption
Figure 2. Rings that are not round

Our resolution to this issue is to use the hypothesis that dimAX<1\dim_{A}X<1 to show that the intersection XVjX\cap V_{j} may be covered by small islands which can be flowed into a relatively small ball contained in VjV_{j}. The point is that while the next level of rings down from RjR_{j} may not be flowed around VjV_{j}, we can pass through finitely many levels, say NN, to obtain a collection of rings which can be flowed around VjV_{j}.

Consequently, to unwind the Dehn twist in RjR_{j}, we concatenate three bi-Lipschitz paths in VjV_{j}: one to move the rings NN levels down into a given disk contained in VjV_{j}, one to act as a conjugate of rotations in VjV_{j}, and then the third to undo the first path. It follows that we may apply this construction simultaneously in the collection of levels that differ by NN to yield a bi-Lipschitz path. Applying this construction NN times, we may concatenate the resulting bi-Lipschitz paths to obtain one path from the identity to ff itself.

1.4. Outline of the paper

In Section 2, we recall the basic definitions and properties of the objects we will use. In Section 3, we prove Theorem 1.2. In Section 4, we prove some technical results on bi-Lipschitz paths. In Section 5, we study how to collapse sets of Assouad dimension less than 11 into small disks. Finally in Section 6, we prove Proposition 1.3, and in Section 7 we prove how the map ff in Theorem 1.4 can be decomposed into bi-Lipschitz mappings of small isometric distortion. Finally, in Section 8 we construct a multitwist map with a singular set of Assouad dimension close to 2 that can be decomposed using the techniques of the paper.

1.5. Acknowledgements

The authors are indebted to Vladimir Markovic for suggesting this problem and many discussions thereupon.

2. Preliminaries

2.1. Modulus of ring domains

Given a family Γ\Gamma of curves in n\mathbb{R}^{n}, define the conformal modulus

Mod(Γ)=infρnρ(x)n𝑑x\operatorname{Mod}(\Gamma)=\inf_{\rho}\int_{\mathbb{R}^{n}}\rho(x)^{n}\,dx

where the infimum is taken over all Borel ρ:n[0,)\rho:\mathbb{R}^{n}\to[0,\infty) such that γρ𝑑s1\int_{\gamma}\rho\,ds\geq 1 for all locally rectifiable γΓ\gamma\in\Gamma.

Here and for the rest, given a ring domain RR in 2\mathbb{R}^{2} with boundary components γ1\gamma_{1} and γ2\gamma_{2}, we denote by M(R)M(R) the modulus of the family of curves in RR that join γ1\gamma_{1} with γ2\gamma_{2}. Observe that the larger M(R)M(R) is, the thinner the ring domain RR is. It is well known [Loe59] that there exists a decreasing function ψ:(0,)(0,)\psi:(0,\infty)\to(0,\infty) such that, if RR is a ring domain with outer boundary component γ1\gamma_{1} and inner boundary component γ2\gamma_{2}, then

(2.1) M(R)ψ(dist(γ1,γ2)diamγ2).M(R)\geq\psi\left(\frac{\operatorname{dist}(\gamma_{1},\gamma_{2})}{\operatorname{diam}{\gamma_{2}}}\right).

2.2. Assouad dimension

A set XNX\subset\mathbb{R}^{N} is ss-homogeneous for some s0s\geq 0 if there exists C>0C>0 such that for every bounded set AXA\subset X, any ϵ(0,diamA)\epsilon\in(0,\operatorname{diam}{A}), and any ϵ\epsilon-separated set VAV\subset A,

cardVC(ϵ1diamA)s.\operatorname{card}{V}\leq C(\epsilon^{-1}\operatorname{diam}{A})^{s}.

Recall that a set VAV\subset A is ϵ\epsilon-separated if for any distinct x,yVx,y\in V we have |xy|ϵ|x-y|\geq\epsilon.

If we want to emphasize on the constant CC, we say that XX is (C,s)(C,s)-homogeneous. Note that every subset of N\mathbb{R}^{N} is NN-homogeneous. Moreover, if 0s1s20\leq s_{1}\leq s_{2} and XX is s1s_{1}-homogeneous, then it is also s2s_{2}-homogeneous. The Assouad dimension of a set XNX\subset\mathbb{R}^{N} is defined as

dimA(X)=inf{s0:X is s-homogeneous}.\dim_{A}(X)=\inf\{s\geq 0:X\text{ is $s$-homogeneous}\}.

2.3. Hyperbolic geometry

Suppose X𝕊2X\subset\mathbb{S}^{2} is a Cantor set and S=𝕊2XS=\mathbb{S}^{2}\setminus X is a hyperbolic Riemann surface with a pants decomposition. Here, we recall how the cuffs (αj)(\alpha_{j}) of the decomposition can be related to the thickness of ring domains embedded in the surface, see Figure 3.

Refer to caption
Figure 3. On the left, we have a pants decomposition for 𝕊2X\mathbb{S}^{2}\setminus X with a particular pair of pants shaded. On the right, we have a topological model of a pair of pants, again shaded, with the arrowed curve a geodesic cuff αj\alpha_{j} and the ring domain in black an example of RjR_{j}^{\prime}.
Proposition 2.1.

For each jj there exists a ring domain Rj𝕊2XR_{j}^{\prime}\subset\mathbb{S}^{2}\setminus X that contains αj\alpha_{j} such that domains RjR_{j}^{\prime} are mutually disjoint and

M(Rj)=(αj)2arcsin(e(αj)).M(R_{j}^{\prime})=\frac{\ell(\alpha_{j})}{2\arcsin(e^{-\ell(\alpha_{j})})}.

This result is assuredly standard. Maskit [Mas85] proves this for finite type surfaces, but since we will be applying this to infinite type surfaces, we give a proof for the convenience of the reader. We need the following Collar Lemma.

Lemma 2.2 ([ALP+11, Lemma 2.2]).

There exist pairwise disjoint collars (Cj)j(C_{j})_{j} of cuffs (αj)j(\alpha_{j})_{j} given by

Cj={zS:dS(z,γ)B((αj))},C_{j}=\{z\in S:d_{S}(z,\gamma)\leq B(\ell(\alpha_{j}))\},

where dSd_{S} denotes the hyperbolic metric on SS and

B(t)=12log(1+2et1).B(t)=\frac{1}{2}\log\left(1+\frac{2}{e^{t}-1}\right).
Proof of Proposition 2.1.

Let CjC_{j} be the collars from Lemma 2.2. These collars are necessarily ring domains.

Since SS is a hyperbolic Riemann surface, we can consider its lift to the strip model of the hyperbolic plane. More precisely, let Σ={z:|Im(z)|<π/2}\Sigma=\{z\in\mathbb{C}:|\operatorname{Im}(z)|<\pi/2\}. Then the hyperbolic metric density on Σ\Sigma is given by λΣ(z)=sec(Im(z))\lambda_{\Sigma}(z)=\sec(\operatorname{Im}(z)) (see for example [BM07, Example 7.9]). Since we can identify SS with Σ/G\Sigma/G, where GG is a covering group of deck transformations, we can lift αj\alpha_{j} so that its lift is contained in the real axis in Σ\Sigma. Moreover, CjC_{j} can be lifted to a rectangle in Σ\Sigma whose closure is given by R=[r,r]×[s,s]R=[-r,r]\times[-s,s].

Here, we have dΣ(r,r)=d_{\Sigma}(-r,r)=\ell and dΣ(is,is)=2B((αj))d_{\Sigma}(-is,is)=2B(\ell(\alpha_{j})). Since the hyperbolic metric and the Euclidean metric coincide on the real axis in Σ\Sigma, we have r=(αj)/2r=\ell(\alpha_{j})/2. Next,

2B((αj))=dΣ(is,is)=sssectdt=2ln(secs+tans).2B(\ell(\alpha_{j}))=d_{\Sigma}(-is,is)=\int_{-s}^{s}\sec t\>dt=2\ln(\sec s+\tan s).

Solving this for ss, we see that

s=arcsin(tanh(B((αj))))s=\arcsin(\tanh(B(\ell(\alpha_{j}))))

and hence

s=arcsin(e(αj)).s=\arcsin(e^{-\ell(\alpha_{j})}).

Finally, M(Cj)M(C_{j}) is equal to the modulus of the path family joining the rr-sides to the ss-sides of the rectangle RR. Thus

M(Cj)=rs=S(αj)2arcsin(eS(αj)).M(C_{j})=\frac{r}{s}=\frac{\ell_{S}(\alpha_{j})}{2\arcsin(e^{-\ell_{S}(\alpha_{j})})}.\qed

We will also need the following result of Wolpert.

Lemma 2.3 (Wolpert [Wol81]).

Let f:SSf:S\to S^{\prime} be a KK-quasiconformal homeomorphism between hyperbolic Riemann surfaces S,SS,S^{\prime}. Let α\alpha be a closed geodesic in SS, and let α\alpha^{\prime} be the unique closed geodesic in SS^{\prime} that is homotopic to f(α)f(\alpha). Then

K1S[α]S[α]KS[α].K^{-1}\ell_{S}[\alpha]\leq\ell_{S^{\prime}}[\alpha^{\prime}]\leq K\ell_{S}[\alpha].

2.4. Square thickenings

We recall some terminology and notation from [Mac99]. Given α>0\alpha>0 define

𝒢α:={αn+[0,α]2:n2}and𝒢α1:={e:e is an edge of some S𝒢α}.\displaystyle\mathscr{G}_{\alpha}:=\{\alpha\textbf{n}+[0,\alpha]^{2}:\textbf{n}\in\mathbb{Z}^{2}\}\quad\text{and}\quad\mathscr{G}_{\alpha}^{1}:=\{e:\text{$e$ is an edge of some $S\in\mathscr{G}_{\alpha}$}\}.

Given a set W2W\subset\mathbb{R}^{2} define WαW^{\alpha} to be the collection of all squares in 𝒢α\mathscr{G}_{\alpha} that intersect with WW. For δ>0\delta>0, define the δ\delta-square thickening

𝒯δ(W)=(W4δ)δ,\mathcal{T}_{\delta}(W)=(W^{4\delta})^{\delta},

see Figure 4.

Refer to caption
Figure 4. The shaded region is WW. The black curve is the boundary of W4δW^{4\delta} and the blue curve is the boundary of 𝒯δ(W)\mathcal{T}_{\delta}(W).
Lemma 2.4 ([Mac99, Lemma 2.1]).

If WW is a bounded subset of the plane and δ>0\delta>0, then the boundary of 𝒯δ(W)\mathcal{T}_{\delta}(W) is a finite union of mutually disjoint polygonal Jordan curves made of edges in 𝒢δ1\mathscr{G}^{1}_{\delta} and

(2.2) δdist(x,W)8δfor all x𝒯δ.\delta\leq\operatorname{dist}(x,W)\leq 8\delta\qquad\text{for all $x\in\partial\mathcal{T}_{\delta}$}.

2.5. Symbolic notation

At several junctures in this paper, it will be convenient to use symbolic notation to describe our constructions.

Given an integer k0k\geq 0, we denote by {1,2}k\{1,2\}^{k} the set of words formed from the alphabet {1,2}\{1,2\} that have length exactly kk. Conventionally, we set {1,2}0={ε}\{1,2\}^{0}=\{\varepsilon\} where ε\varepsilon is the empty word. We also denote by {1,2}=k0{1,2}k\{1,2\}^{*}=\bigcup_{k\geq 0}\{1,2\}^{k} the set of all finite words formed from {1,2}\{1,2\}. Given a word w{1,2}w\in\{1,2\}^{*}, we denote by |w||w| the length of ww with the convention |ε|=0|\varepsilon|=0.

3. Uniformly disconnected Cantor sets and hyperbolic geometry

In this section, we prove Theorem 1.2. One direction of the theorem is given in Section 3.1 and the other direction is given in Section 3.2.

3.1. Assuming uniformly disconnected

Here we prove the necessary direction of Theorem 1.2.

Proposition 3.1.

Let X𝕊2X\subset\mathbb{S}^{2} be a cc-uniformly disconnected Cantor set. There exists M>0M>0 depending only on cc, and there exists a pants decomposition for the Riemann surface S=𝕊2XS=\mathbb{S}^{2}\setminus X such that the associated cuffs (αj)(\alpha_{j}) satisfy supj[αj]<M\sup_{j}\ell[\alpha_{j}]<M.

Denote by 𝒞\mathcal{C} the standard one-third Cantor set and by S0S_{0} the Riemann surface S0=𝕊2𝒞S_{0}=\mathbb{S}^{2}\setminus\mathcal{C}.

Lemma 3.2 ([Vel21, Corollary A]).

If X𝕊2X\subset\mathbb{S}^{2} is a uniformly disconnected set, then there exists a quasiconformal map f:𝕊2𝕊2f:\mathbb{S}^{2}\to\mathbb{S}^{2} such that f(X)𝒞f(X)\subset\mathcal{C}.

As observed in [Shi18, p.5], the pairs of pants in the pants decomposition of S0S_{0} can be chosen to be conformally equivalent to one another. It follows that each such pair of pants has the same cuff lengths. To see this, suppose PP and PP^{\prime} are two pairs of pants in this decomposition with a conformal map h:PPh:P\to P^{\prime}. Let RR and RR^{\prime} be the respective doubles of PP and PP^{\prime}, that is, RR and RR^{\prime} are genus two surfaces. Then hh extends via reflection to a conformal map h~:RR\widetilde{h}:R\to R^{\prime} and hence h~\widetilde{h} is a hyperbolic isometry. Restricting h~\widetilde{h} to the cuffs of PP, we see that PP and PP^{\prime} have the same cuff lengths.

In particular, we conclude that there exist a constant q>0q>0 and a pair of pants decomposition of S0S_{0} with cuffs (Cj)(C_{j}) such that

(3.1) supjS0[Cj]=q.\sup_{j}\ell_{S_{0}}[C_{j}]=q.
Proof of Proposition 3.1.

Let X2X\subset\mathbb{R}^{2} be a uniformly disconnected Cantor set, and let ff be the quasiconformal map from Lemma 3.2 such that f(X)𝒞f(X)\subset\mathcal{C}. We will use the pants decomposition with cuffs (Cj)(C_{j}) for S0S_{0}. Since f(S)S0f(S)\supset S_{0}, we may use a subset of the (Cj)(C_{j}) to generate a pair of pants decomposition for f(S)f(S). This subset can be labelled as (Cjk)(C_{j_{k}}) and we, for brevity, will denote it by (βk)(\beta_{k}).

Suppose (ηj)(\eta_{j}) are the cuffs of a pants decomposition of SS. Then each ηj\eta_{j} is homotopic to f1(βk)f^{-1}(\beta_{k}) for some kk and vice versa. Hence, if we assume for a contradiction that S[ηjm]\ell_{S}[\eta_{j_{m}}]\to\infty, it follows via Lemma 2.3 that f(S)[βkm]\ell_{f(S)}[\beta_{k_{m}}]\to\infty.

Since f(S)S0f(S)\supset S_{0}, the subordination principle for the hyperbolic metric implies that if γ\gamma is any path in S0S_{0}, then f(S)(γ)S0(γ)\ell_{f(S)}(\gamma)\leq\ell_{S_{0}}(\gamma). In particular, we conclude that S0[βkm]\ell_{S_{0}}[\beta_{k_{m}}]\to\infty. This contradicts (3.1). ∎

3.2. Towards uniformly disconnected

Here we prove the sufficient direction of Theorem 1.2.

Proposition 3.3.

Let X2X\subset\mathbb{R}^{2} be a Cantor set and suppose that the Riemann surface S=𝕊2XS=\mathbb{S}^{2}\setminus X has a pants decomposition (Pj)(P_{j}) where the cuffs (αj)(\alpha_{j}) satisfy supj[αj]<L<\sup_{j}\ell[\alpha_{j}]<L<\infty. Then XX is cc-uniformly disconnected for some cc depending only on LL.

Recall the symbolic notation from Section 2.5.

Lemma 3.4.

A totally bounded metric space XX is uniformly disconnected if and only if there exists a set 𝒲{1,2}\mathcal{W}\subset\{1,2\}^{*}, a constant δ>0\delta>0 and a collection of subsets {Xw:w𝒲}\{X_{w}:w\in\mathcal{W}\} with the following properties.

  1. (i)

    The empty word ε𝒲\varepsilon\in\mathcal{W} and Xε=XX_{\varepsilon}=X.

  2. (ii)

    If wi𝒲wi\in\mathcal{W} for some i{1,2}i\in\{1,2\} and w{1,2}w\in\{1,2\}^{*}, then w𝒲w\in\mathcal{W} and XwiXwX_{wi}\subset X_{w}.

  3. (iii)

    If XwX_{w} is a point for some w𝒲w\in\mathcal{W}, then w1𝒲w1\in\mathcal{W}, w2𝒲w2\not\in\mathcal{W}, and Xw1=XwX_{w1}=X_{w}.

  4. (iv)

    If XwX_{w} has at least two points for some w𝒲w\in\mathcal{W}, then w1,w2𝒲w1,w2\in\mathcal{W}, Xw=Xw1Xw2X_{w}=X_{w1}\cup X_{w2}, and

    (3.2) dist(Xw1,Xw2)δmax{diamXw1,diamXw2}.\operatorname{dist}(X_{w1},X_{w2})\geq\delta\max\{\operatorname{diam}{X_{w1}},\operatorname{diam}{X_{w2}}\}.

The constant of uniform disconnectedness and δ\delta are quantitatively related.

Proof.

Assume first that XX is cc-uniformly disconnected. Set Xε=XX_{\varepsilon}=X. Assume now that for some w{1,2}w\in\{1,2\}^{*}, we have defined a nonempty set XwXX_{w}\subset X. If XwX_{w} is a single point, then set Xw1=XwX_{w1}=X_{w}. Assume now that XwX_{w} contains at least two points and fix xXwx\in X_{w}. By the uniform disconnectedness of XwX_{w}, there exists EXwE\subset X_{w} such that xEx\in E, diamE12diamXw\operatorname{diam}{E}\leq\frac{1}{2}\operatorname{diam}{X_{w}} and dist(E,XwE)(2c)1diamXw\operatorname{dist}(E,X_{w}\setminus E)\geq(2c)^{-1}\operatorname{diam}{X_{w}}. Set Xw1=EX_{w1}=E and Xw2=XwEX_{w2}=X_{w}\setminus E. Note that

dist(Xw1,Xw2)(2c)1diamXw(2c)1max{diamXw1,diamXw2}.\operatorname{dist}(X_{w1},X_{w2})\geq(2c)^{-1}\operatorname{diam}{X_{w}}\geq(2c)^{-1}\max\{\operatorname{diam}{X_{w1}},\operatorname{diam}{X_{w2}}\}.

Setting 𝒲\mathcal{W} to be the set of all words w{1,2}w\in\{1,2\}^{*} for which XwX_{w} has been defined, it is easy to see that {Xw:w𝒲}\{X_{w}:w\in\mathcal{W}\} satisfies (i)–(iv) with δ=(2c)1\delta=(2c)^{-1}.

Suppose now that there exists 𝒲{1,2}\mathcal{W}\subset\{1,2\}^{*}, δ>0\delta>0 and a collection {Xw:w𝒲}\{X_{w}:w\in\mathcal{W}\} satisfying (i)–(iv). We first show that if (in)(i_{n}) is a sequence in {1,2}\{1,2\} such that i1in𝒲i_{1}\cdots i_{n}\in\mathcal{W} for all nn\in\mathbb{N}, then limndiamXi1in=0\lim_{n\to\infty}\operatorname{diam}{X_{i_{1}\cdots i_{n}}}=0. Assume for a contradiction that there exists d>0d>0 and a sequence (in)(i_{n}) in {1,2}\{1,2\} such that i1in𝒲i_{1}\cdots i_{n}\in\mathcal{W} and diamXi1in>d\operatorname{diam}{X_{i_{1}\cdots i_{n}}}>d for all nn\in\mathbb{N}. Fix x1XεXi1x_{1}\in X_{\varepsilon}\setminus X_{i_{1}} and for each nn\in\mathbb{N} fix xnXi1in1Xi1inx_{n}\in X_{i_{1}\cdots i_{n-1}}\setminus X_{i_{1}\cdots i_{n}}. By (3.2), for any distinct i,ji,j\in\mathbb{N}, |xixj|δd|x_{i}-x_{j}|\geq\delta d. Then, the set {xn:n}\{x_{n}:n\in\mathbb{N}\} is not totally bounded and we reach a contradiction.

We prove now that XX is uniformly disconnected. If XX contains a single point, then the claim is trivial. Assume now that diamX>0\operatorname{diam}{X}>0 and let xXx\in X and r(0,diamX)r\in(0,\operatorname{diam}{X}). Let w𝒲w\in\mathcal{W} be the maximal word (in word-length) such that xXwx\in X_{w} and diamXwr\operatorname{diam}{X_{w}}\geq r. Write w=i1inw=i_{1}\cdots i_{n} and assume that xXwin+1x\in X_{wi_{n+1}} where in+1{1,2}i_{n+1}\in\{1,2\}. Setting E=Xwin+1E=X_{wi_{n+1}}, we have that diamE<r\operatorname{diam}{E}<r while

dist(E,XE)\displaystyle\operatorname{dist}(E,X\setminus E) =minj=1,,n+1dist(Xi1ij,Xi1ij1Xi1ij)\displaystyle=\min_{j=1,\dots,n+1}\operatorname{dist}(X_{i_{1}\cdots i_{j}},X_{i_{1}\cdots i_{j-1}}\setminus X_{i_{1}\cdots i_{j}})
min{dist(Xw1,Xw2),δminj=1,,ndiamXi1ij}\displaystyle\geq\min\left\{\operatorname{dist}(X_{w1},X_{w2}),\delta\min_{j=1,\dots,n}\operatorname{diam}{X_{i_{1}\cdots i_{j}}}\right\}
=min{dist(Xw1,Xw2),δdiamXw}.\displaystyle=\min\left\{\operatorname{dist}(X_{w1},X_{w2}),\delta\operatorname{diam}{X_{w}}\right\}.

By the triangle inequality and (3.2),

diamXwdiamXw1+dist(Xw1,Xw2)+diamXw2(1+2δ1)dist(Xw1,Xw2).\operatorname{diam}{X_{w}}\leq\operatorname{diam}{X_{w1}}+\operatorname{dist}(X_{w1},X_{w2})+\operatorname{diam}{X_{w2}}\leq(1+2\delta^{-1})\operatorname{dist}(X_{w1},X_{w2}).

Therefore,

dist(E,XE)min{δ,(1+2δ1)1}diamXw(1+2δ1)1r\operatorname{dist}(E,X\setminus E)\geq\min\left\{\delta,(1+2\delta^{-1})^{-1}\right\}\operatorname{diam}{X_{w}}\geq(1+2\delta^{-1})^{-1}r

and XX is cc-uniformly disconnected with c=(1+2δ1)1c=(1+2\delta^{-1})^{-1}. ∎

We now show Proposition 3.3 and thus complete the proof of Theorem 1.2.

Proof of Proposition 3.3.

We assume that each αj\alpha_{j} has been chosen to minimize S(αj)\ell_{S}(\alpha_{j}) in its homotopy class. By Proposition 2.1 there exist disjoint ring domains RjR_{j}^{\prime} in 2X\mathbb{R}^{2}\setminus X containing the cuffs αj\alpha_{j} such that supjM(Rj)m\sup_{j}M(R_{j}^{\prime})\leq m for some mm depending only on LL. For each jj let VjV_{j}^{\prime} and UjU_{j}^{\prime} be the bounded and unbounded, respectively, components of 2Rj\mathbb{R}^{2}\setminus R_{j}^{\prime}. We relabel the ring domains RjR_{j}^{\prime} in the following way.

Firstly, we remark that there exist three indices j1,j2,j3j_{1},j_{2},j_{3} such that Vj1,Vj2,Vj3V_{j_{1}}^{\prime},V_{j_{2}}^{\prime},V_{j_{3}}^{\prime} are mutually disjoint and XX is contained in Vj1Vj2Vj3V_{j_{1}}^{\prime}\cup V_{j_{2}}^{\prime}\cup V_{j_{3}}^{\prime}. For l=1,2,3l=1,2,3, we denote Rl,ε:=RjlR_{l,\varepsilon}^{\prime}:=R_{j_{l}}^{\prime}. Inductively, assume that for some l{1,2,3}l\in\{1,2,3\} and some w{1,2}w\in\{1,2\}^{*} we have defined Rl,w=RjR_{l,w}^{\prime}=R_{j}^{\prime} for an index jj. There exist two indices i1,i2i_{1},i_{2} such that

  1. (i)

    Vi1,Vi2VjV_{i_{1}}^{\prime},V_{i_{2}}^{\prime}\subset V_{j}^{\prime} and Vi1Vi2=V_{i_{1}}^{\prime}\cap V_{i_{2}}^{\prime}=\emptyset;

  2. (ii)

    if ViVjV_{i}^{\prime}\subset V_{j}^{\prime} for some ii, then ViVi1V_{i}^{\prime}\subset V_{i_{1}}^{\prime} or ViVi2V_{i}^{\prime}\subset V_{i_{2}}^{\prime}.

Set now Rl,w1:=Ri1R_{l,w1}^{\prime}:=R_{i_{1}}^{\prime} and Rl,w2:=Ri2R_{l,w2}^{\prime}:=R_{i_{2}}^{\prime}.

If for some jj, l{1,2,3}l\in\{1,2,3\} and w{1,2}w\in\{1,2\}^{*} we have defined Rl,w=RjR_{l,w}^{\prime}=R_{j}^{\prime}, then define Vl,w:=VjV_{l,w}^{\prime}:=V_{j}^{\prime} and Ul,w:=UjU_{l,w}^{\prime}:=U_{j}^{\prime}. Set also Xl,w=XVl,wX_{l,w}=X\cap V_{l,w}^{\prime}. It is easy to see that for each l=1,2,3l=1,2,3, the collection {Xl,w:w{1,2}}\{X_{l,w}:w\in\{1,2\}^{*}\} satisfies (i)–(iii) of Lemma 3.4 for Xl,εX_{l,\varepsilon}.

Fix now l{1,2,3}l\in\{1,2,3\}. By (2.1) we have that there exists d>0d>0 depending only on mm such that for all w{1,2}w\in\{1,2\}^{*}

dist(Ul,w,Vl,w)ddiamVl,w.\operatorname{dist}(U_{l,w}^{\prime},V_{l,w}^{\prime})\geq d\operatorname{diam}{V_{l,w}^{\prime}}.

Therefore, for each w{1,2}w\in\{1,2\}^{*}

dist(Xl,w1,Xl,w2)maxi=1,2dist(Xl,wi,Vl,wi)\displaystyle\operatorname{dist}(X_{l,w1},X_{l,w2})\geq\max_{i=1,2}\operatorname{dist}(X_{l,wi},\partial V_{l,wi}^{\prime}) maxi=1,2dist(Vl,wi,Ul,wi)\displaystyle\geq\max_{i=1,2}\operatorname{dist}(V_{l,wi}^{\prime},U_{l,wi}^{\prime})
dmaxi=1,2diamVl,wi\displaystyle\geq d\max_{i=1,2}\operatorname{diam}{V_{l,wi}^{\prime}}
dmaxi=1,2diamXl,wi.\displaystyle\geq d\max_{i=1,2}\operatorname{diam}{X_{l,wi}}.

Working as above, we can deduce that for all distinct l,l{1,2,3}l,l^{\prime}\in\{1,2,3\}, we have

dist(Xl,ε,Xl,ε)dmax{diamXl,ε,diamXl,ε}.\operatorname{dist}(X_{l,\varepsilon},X_{l^{\prime},\varepsilon})\geq d\max\{\operatorname{diam}{X_{l,\varepsilon}},\operatorname{diam}{X_{l^{\prime},\varepsilon}}\}.

Since XX is compact, by Lemma 3.4, XX is CC-uniformly disconnected with CC depending only on dd, hence only on mm, hence only on LL. ∎

4. Bi-Lipschitz paths

Our strategy to proving Theorem 1.4 is to use bi-Lipschitz paths to yield the required decomposition. We recall the following definition from [FM12].

Definition 4.1.

Let (X,dX)(X,d_{X}) be a metric space. A path H:[0,1]LIP(X)H:[0,1]\rightarrow LIP(X) is called a bi-Lipschitz path if for every ϵ>0\epsilon>0, there exists δ>0\delta>0 such that if s,t[0,1]s,t\in[0,1] with |st|<δ\arrowvert s-t\arrowvert<\delta, the following two conditions hold:

  1. (i)

    for all xXx\in X, dX(HsHt1(x),x)<ϵd_{X}(H_{s}\circ H_{t}^{-1}(x),x)<\epsilon;

  2. (ii)

    we have that HsHt1H_{s}\circ H_{t}^{-1} is (1+ϵ)(1+\epsilon)-bi-Lipschitz with respect to dXd_{X}.

In this paper bi-Lipschitz paths are denoted by capital letters F,G,H,F,G,H,\cdots. Given two bi-Lipschitz maps f,g:XXf,g:X\to X, two bi-Lipschitz paths F,G:[0,1]LIP(X)F,G:[0,1]\to LIP(X), and a subset EXE\subset X, we define

  1. (i)

    the concatenation of FF with GG to be the biLipschitz path H:[0,1]LIP(X)H:[0,1]\to LIP(X) with Ht=F2tH_{t}=F_{2t} for t[0,1/2]t\in[0,1/2] and Ht=G2t1H_{t}=G_{2t-1} for t[1/2,1]t\in[1/2,1], and we may then concatenate finitely many bi-Lipschitz paths in the obvious way;

  2. (ii)

    the restriction F|E:[0,1]LIP(E)F|E:[0,1]\to LIP(E) by (F|E)t=Ft|E(F|E)_{t}=F_{t}|E;

  3. (iii)

    the composition FGF\circ G by (FG)t=FtGt(F\circ G)_{t}=F_{t}\circ G_{t} for all t[0,1]t\in[0,1];

  4. (iv)

    the composition fFgf\circ F\circ g by (fFg)t=fFtg(f\circ F\circ g)_{t}=f\circ F_{t}\circ g for all t[0,1]t\in[0,1].

We emphasize that in (iii) and (iv) here, the compositions need not be bi-Lipschitz paths. Much of our work will involve showing that our constructions are made carefully enough that when we do need to compose or conjugate, we do still have a bi-Lipschitz path. For illustrative purposes, we include examples where (iii) and (iv) fail to give a bi-Lipschitz path.

Example 4.2.

Let L>1L>1, B=B(0,1/3)2B=B(0,1/3)\subset\mathbb{R}^{2} and let f:22f:\mathbb{R}^{2}\to\mathbb{R}^{2} be an LL-bi-Lipschitz map which is the identity on 2B\mathbb{R}^{2}\setminus B. Then define

f~(z)={f(zn)+n,zB(n,1/3)n{0,1,2,}z,otherwise.\widetilde{f}(z)=\begin{cases}f(z-n)+n,&\text{$z\in B(n,1/3)$, $n\in\{0,1,2,\ldots\}$}\\ z,&\text{otherwise.}\end{cases}

Clearly f~\widetilde{f} is also an LL-bi-Lipschitz map.

Set Ft=eiπtzF_{t}=e^{i\pi t}z for t[0,1]t\in[0,1] and H=IdFf~H=\text{Id}\circ F\circ\tilde{f} where Id:22\text{Id}:\mathbb{R}^{2}\to\mathbb{R}^{2} is the identity map. Then,

Ht1Hs(z)=f~1(ei(st)πf~(z)).H_{t}^{-1}\circ H_{s}(z)=\widetilde{f}^{-1}(e^{i(s-t)\pi}\widetilde{f}(z)).

Suppose ϵ<L1\epsilon<L-1, δ>0\delta>0 and |st|=δ|s-t|=\delta. Then there exists NN\in\mathbb{N} large enough that B(Nei(st)π,1/3)B(N,1/3)=B(Ne^{i(s-t)\pi},1/3)\cap B(N,1/3)=\emptyset. Hence on B(N,1/3)B(N,1/3) we have that Ht1HsH_{t}^{-1}\circ H_{s} agrees with a composition of a rotation and f~\widetilde{f}. This means that Ht1HsH_{t}^{-1}\circ H_{s} is not (1+ϵ)(1+\epsilon)-bi-Lipschitz and hence HH is not a bi-Lipschitz path. We conclude that IdFf~\text{Id}\circ F\circ\tilde{f} is not a bi-Lipschitz path.

Using the same example and setting G:[0,1]LIP(2)G:[0,1]\to LIP(\mathbb{R}^{2}) being the constant path f~\tilde{f} we see that FGF\circ G is not a bi-Lipschitz path. Hence, compositions of bi-Lipschitz paths are not always bi-Lipschitz paths.

It is worth pointing out that in a bi-Lipschitz path, the elements are bi-Lipschitz with uniform constant.

Lemma 4.3.

Suppose H:[0,1]LIP(X)H:[0,1]\to LIP(X) is a bi-Lipschitz path. Then there exists L>1L>1 such that HtH_{t} is an LL-bi-Lipschitz map for each t[0,1]t\in[0,1].

Proof.

Clearly H0H_{0} is an L0L_{0}-bi-Lipschitz map for some L0>1L_{0}>1. Set ϵ=1\epsilon=1 and the corresponding δ>0\delta>0 so that condition (ii) holds. In particular, for every t1(0,δ)t_{1}\in(0,\delta), by condition (ii) applied to (Ht1H01)H0(H_{t_{1}}\circ H_{0}^{-1})\circ H_{0}, the map HtH_{t} is 2L02L_{0}-bi-Lipschitz. Next, for every t2[δ,2δ)t_{2}\in[\delta,2\delta), there exists t1(0,δ)t_{1}\in(0,\delta) with |t2t1|<δ|t_{2}-t_{1}|<\delta. Applying condition (ii) to (Ht2Ht11)(Ht1H01)H0(H_{t_{2}}\circ H_{t_{1}}^{-1})\circ(H_{t_{1}}\circ H_{0}^{-1})\circ H_{0}, we see that Ht2H_{t_{2}} is 22L02^{2}L_{0}-bi-Lipschitz.

Continuing inductively, we see that for any t[0,1]t\in[0,1], HtH_{t} is 21/δ+1L02^{\lfloor 1/\delta\rfloor+1}L_{0}-bi-Lipschitz. ∎

Next we show that if the restrictions of HH on three sets whose union is 2\mathbb{R}^{2} are bi-Lipschitz paths, then HH is a bi-Lipschitz path quantitatively.

Lemma 4.4.

Let A,B,C2A,B,C\subset\mathbb{R}^{2} be closed sets such that 2=ABC\mathbb{R}^{2}=A\cup B\cup C. Suppose that for any ϵ>0\epsilon>0, there exists δ>0\delta>0 such that if s,t[0,1]s,t\in[0,1] with |st|<δ|s-t|<\delta, then the two conditions in Definition 4.1 hold simultaneously for H|AH|A, H|BH|B, H|CH|C. Then the two conditions in Definition 4.1 also hold for HH with the same δ\delta.

Proof.

Fix ϵ>0\epsilon>0 and let δ>0\delta>0 such that the two conditions in Definition 4.1 hold simultaneously for H|AH|A, H|BH|B, H|CH|C. Let s,t[0,1]s,t\in[0,1] such that |st|<δ|s-t|<\delta. If x2x\in\mathbb{R}^{2}, then without loss of generality we may assume that xAx\in A and we have

|HsHt1(x)x|=|(H|A)s(H|A)t1(x)x|<ϵ|H_{s}\circ H_{t}^{-1}(x)-x|=|(H|A)_{s}\circ(H|A)_{t}^{-1}(x)-x|<\epsilon

and HH satisfies (i).

For (ii), it suffices to show that HsHt1H_{s}\circ H_{t}^{-1} is (1+ϵ)(1+\epsilon)-Lipschitz. Let x,y2x,y\in\mathbb{R}^{2}. If both xx and yy belong to the same set from A,B,CA,B,C, then (ii) follows immediately. Assume without loss of generality that xAx\in A and yBAy\in B\setminus A. Let z[x,y]Az\in[x,y]\cap A such that A([y,z]{z})=A\cap([y,z]\setminus\{z\})=\emptyset. Since yBAy\in B\setminus A, we have that zyz\neq y.

There are now two cases. First, if zBz\in B then

|HsHt1(x)HsHt1(y)|\displaystyle|H_{s}\circ H_{t}^{-1}(x)-H_{s}\circ H_{t}^{-1}(y)| |HsHt1(x)HsHt1(z)|\displaystyle\leq|H_{s}\circ H_{t}^{-1}(x)-H_{s}\circ H_{t}^{-1}(z)|
+|HsHt1(z)HsHt1(y)|\displaystyle\quad+|H_{s}\circ H_{t}^{-1}(z)-H_{s}\circ H_{t}^{-1}(y)|
|(H|A)s(H|A)t1(x)(H|A)s(H|A)t1(z)|\displaystyle\leq|(H|A)_{s}\circ(H|A)_{t}^{-1}(x)-(H|A)_{s}\circ(H|A)_{t}^{-1}(z)|
+|(H|B)s(H|B)t1(z)(H|B)s(H|B)t1(y)|\displaystyle\quad+|(H|B)_{s}\circ(H|B)_{t}^{-1}(z)-(H|B)_{s}\circ(H|B)_{t}^{-1}(y)|
(1+ϵ)(|xz|+|zy|)\displaystyle\leq(1+\epsilon)(|x-z|+|z-y|)
=(1+ϵ)|xy|.\displaystyle=(1+\epsilon)|x-y|.

Second, if zCz\in C then we have two sub-cases. If also yCy\in C, we have the same argument as above, with the role of BB played by CC. If yCy\notin C, then let wC[z,y]w\in C\cap[z,y] be such that C([w,y]{w})=C\cap([w,y]\setminus\{w\})=\emptyset. We have wyw\neq y by construction. Since zACz\in A\cap C and wBCw\in B\cap C, we have

|HsHt1(x)HsHt1(y)|\displaystyle|H_{s}\circ H_{t}^{-1}(x)-H_{s}\circ H_{t}^{-1}(y)| |HsHt1(x)HsHt1(z)|\displaystyle\leq|H_{s}\circ H_{t}^{-1}(x)-H_{s}\circ H_{t}^{-1}(z)|
+|HsHt1(z)HsHt1(w)|\displaystyle\quad+|H_{s}\circ H_{t}^{-1}(z)-H_{s}\circ H_{t}^{-1}(w)|
+|HsHt1(w)HsHt1(y)|\displaystyle\quad+|H_{s}\circ H_{t}^{-1}(w)-H_{s}\circ H_{t}^{-1}(y)|
|(H|A)s(H|A)t1(x)(H|A)s(H|A)t1(z)|\displaystyle\leq|(H|A)_{s}\circ(H|A)_{t}^{-1}(x)-(H|A)_{s}\circ(H|A)_{t}^{-1}(z)|
+|(H|C)s(H|C)t1(z)(H|C)s(H|C)t1(w)|\displaystyle\quad+|(H|C)_{s}\circ(H|C)_{t}^{-1}(z)-(H|C)_{s}\circ(H|C)_{t}^{-1}(w)|
+|(H|B)s(H|B)t1(w)(H|B)s(H|B)t1(y)|\displaystyle\quad+|(H|B)_{s}\circ(H|B)_{t}^{-1}(w)-(H|B)_{s}\circ(H|B)_{t}^{-1}(y)|
(1+ϵ)(|xz|+|zw|+|wy|)\displaystyle\leq(1+\epsilon)(|x-z|+|z-w|+|w-y|)
=(1+ϵ)|xy|.\displaystyle=(1+\epsilon)|x-y|.\qed

Our next result involves the removability of a Cantor set for a bi-Lipschitz path.

Proposition 4.5.

Let X2X\subset\mathbb{R}^{2} be a Cantor set. For each 0t10\leq t\leq 1, suppose Ft:22F_{t}:\mathbb{R}^{2}\to\mathbb{R}^{2} is a continuous mapping such that F|2XF|\mathbb{R}^{2}\setminus X is a bi-Lipschitz path. Then FF extends to a bi-Lipschitz path on 2\mathbb{R}^{2}.

Proof.

First, it is well-known that a bi-Lipschitz map on UU can be extended to a bi-Lipschitz map on the metric closure of UU. Hence the hypothesis that FF is a bi-Lipschitz path on 2X\mathbb{R}^{2}\setminus X, and Lemma 4.3 imply that there exists L1L\geq 1 so that each FtF_{t} is LL-bi-Lipschitz on 2\mathbb{R}^{2}.

Next, we show that property (i) holds in the definition of a bi-Lipschitz path. Suppose ϵ>0\epsilon>0 is given and find δ>0\delta>0 so that if |st|<δ|s-t|<\delta then for any z2Xz\in\mathbb{R}^{2}\setminus X, |Ft(z)Fs(z)|<ϵ/2|F_{t}(z)-F_{s}(z)|<\epsilon/2. If xXx\in X, find a sequence (xn)(x_{n}) in 2X\mathbb{R}^{2}\setminus X with xnx0x_{n}\to x_{0}. Then if we choose NN\in\mathbb{N} so that |xxn|<ϵ/4L|x-x_{n}|<\epsilon/4L for nNn\geq N, we have

|Ft(x)Fs(x)|\displaystyle|F_{t}(x)-F_{s}(x)| |Ft(x)Ft(xn)|+|Ft(xn)Fs(xn)|+|Fs(xn)Fs(x)|\displaystyle\leq|F_{t}(x)-F_{t}(x_{n})|+|F_{t}(x_{n})-F_{s}(x_{n})|+|F_{s}(x_{n})-F_{s}(x)|
2L|xxn|+ϵ/2\displaystyle\leq 2L|x-x_{n}|+\epsilon/2
<ϵ.\displaystyle<\epsilon.

Hence condition (i) is satisfied.

Turning now to property (ii), let xXx\in X and y2y\in\mathbb{R}^{2} and find sequences (xn),(yn)(x_{n}),(y_{n}) in 2X\mathbb{R}^{2}\setminus X with xnxx_{n}\to x and ynyy_{n}\to y. Given ϵ>0\epsilon>0, find δ>0\delta>0 so that if |st|<δ|s-t|<\delta and z,w2Xz,w\in\mathbb{R}^{2}\setminus X then

|Ft(z)Ft(w)|<(1+ϵ/3)|Fs(z)Fs(w)|.|F_{t}(z)-F_{t}(w)|<(1+\epsilon/3)|F_{s}(z)-F_{s}(w)|.

Next, choose NN\in\mathbb{N} large enough that if nNn\geq N then |xxn|<L2ϵ|xy|/3|x-x_{n}|<L^{-2}\epsilon|x-y|/3 and |yyn|<ϵL2|xy|/3|y-y_{n}|<\epsilon L^{-2}|x-y|/3. Hence

|Ft(x)Ft(xn)|L|xxn|<ϵ(3L)1|xy|ϵ|Fs(x)Fs(y)|/3|F_{t}(x)-F_{t}(x_{n})|\leq L|x-x_{n}|<\epsilon(3L)^{-1}|x-y|\leq\epsilon|F_{s}(x)-F_{s}(y)|/3

and

|Ft(yn)Ft(y)|L|yny|<ϵ(3L)1|xy|ϵ|Fs(x)Fs(y)|/3.|F_{t}(y_{n})-F_{t}(y)|\leq L|y_{n}-y|<\epsilon(3L)^{-1}|x-y|\leq\epsilon|F_{s}(x)-F_{s}(y)|/3.

It follows that

|Ft(x)Ft(y)|\displaystyle|F_{t}(x)-F_{t}(y)| |Ft(x)Ft(xn)|+|Ft(xn)Ft(yn)|+|Ft(yn)Ft(y)|\displaystyle\leq|F_{t}(x)-F_{t}(x_{n})|+|F_{t}(x_{n})-F_{t}(y_{n})|+|F_{t}(y_{n})-F_{t}(y)|
<(1+ϵ)|Fs(x)Fs(y)|.\displaystyle<(1+\epsilon)|F_{s}(x)-F_{s}(y)|.

We conclude that property (ii) is satisfied and hence FF is a bi-Lipschitz path on 2\mathbb{R}^{2}. ∎

4.1. Uniform families of bi-Lipschitz paths

For the construction in the proof of Theorem 1.4, it will be useful to consider collections of bi-Lipschitz paths with uniform control. To that end we make the following definition.

Definition 4.6.

A collection \mathcal{H} of bi-Lipschitz paths H:[0,1]LIP(X)H:[0,1]\rightarrow LIP(X) on a common metric space XX is a uniform family of bi-Lipschitz paths if

  1. (i)

    there exists L1L\geq 1 so that H0H_{0} has isometric distortion bounded above by LL for all HH\in\mathcal{H},

  2. (ii)

    given ϵ>0\epsilon>0, there exists δ>0\delta>0 so that if s,t[0,1]s,t\in[0,1] with |st|<δ|s-t|<\delta then the two conditions in Definition 4.1 hold simultaneously for all HH\in\mathcal{H}.

It is clear from Definition 4.6 and Lemma 4.3 that there is a uniform bound on the isometric distortion of any map from any path in a uniform family of bi-Lipschitz paths. We have the following composition result.

Lemma 4.7.

Let \mathcal{H} and 𝒢\mathcal{G} be two uniform families of bi-Lipschitz paths so that for each G𝒢G\in\mathcal{G} and each t[0,1]t\in[0,1], GtG_{t} is an isometry. Then the family ={GH:G𝒢,H}\mathcal{F}=\{G\circ H:G\in\mathcal{G},H\in\mathcal{H}\} is a uniform family of bi-Lipschitz paths.

Proof.

Given ϵ>0\epsilon>0, find δ>0\delta>0 so that both conditions in Definition 4.1 and Definition 4.6 hold for |st|<δ|s-t|<\delta, all HH\in\mathcal{H} and all G𝒢G\in\mathcal{G}. Fixing G𝒢G\in\mathcal{G} and HH\in\mathcal{H}, let F=GHF=G\circ H. Then using the fact that GtG_{t} is an isometry,

|Ft(z)Fs(z)|\displaystyle|F_{t}(z)-F_{s}(z)| =|Gt(Ht(z))Gs(Hs(z))|\displaystyle=|G_{t}(H_{t}(z))-G_{s}(H_{s}(z))|
=|Gt(Ht(z))Gt(Hs(z))+Gt(Hs(z))Gs(Hs(z))|\displaystyle=|G_{t}(H_{t}(z))-G_{t}(H_{s}(z))+G_{t}(H_{s}(z))-G_{s}(H_{s}(z))|
|Ht(z)Hs(z)|+|Gt(Hs(z))Gs(Hs(z))|\displaystyle\leq|H_{t}(z)-H_{s}(z)|+|G_{t}(H_{s}(z))-G_{s}(H_{s}(z))|
2ϵ,\displaystyle\leq 2\epsilon,

which verifies that condition (i) of Definition 4.1 holds uniformly for all paths in \mathcal{F}. For condition (ii), we have

|Ft(z)Ft(w)|\displaystyle|F_{t}(z)-F_{t}(w)| =|Gt(Ht(z))Gt(Ht(w))|\displaystyle=|G_{t}(H_{t}(z))-G_{t}(H_{t}(w))|
=|Ht(z)Ht(w)|\displaystyle=|H_{t}(z)-H_{t}(w)|
(1+ϵ)|Hs(z)Hs(w)|\displaystyle\leq(1+\epsilon)|H_{s}(z)-H_{s}(w)|
=(1+ϵ)|Gs(Hs(z))Gs(Hs(w))|\displaystyle=(1+\epsilon)|G_{s}(H_{s}(z))-G_{s}(H_{s}(w))|
=(1+ϵ)|Fs(z)Fs(w)|,\displaystyle=(1+\epsilon)|F_{s}(z)-F_{s}(w)|,

which verifies that condition (ii) of Definition 4.1 holds uniformly for all paths in \mathcal{F}. Finally, since G0G_{0} is a isometry, and the isometric distortion of H0H_{0} is uniformly bounded above, it follows that the same is true for any FF\in\mathcal{F}. ∎

For our next result, we see that a family of conjugates of a bi-Lipschitz path by controlled dilations is uniform.

Lemma 4.8.

Let F:[0,1]LIP(n)F:[0,1]\to LIP(\mathbb{R}^{n}) be a bi-Lipschitz path and let c1c\geq 1. The family

={ϕFϕ1:ϕ is a similarity of n with scaling factor at most c}\mathcal{F}=\{\phi\circ F\circ\phi^{-1}:\text{$\phi$ is a similarity of $\mathbb{R}^{n}$ with scaling factor at most $c$}\}

is a uniform family of bi-Lipschitz paths.

Proof.

Fix ϕ:nn\phi:\mathbb{R}^{n}\to\mathbb{R}^{n} to be a similarity of scaling factor λc\lambda\leq c.

First, suppose F0F_{0} has isometric distortion L0L_{0}. Then we have

|ϕF0ϕ1(z)ϕF0ϕ1(w)|=λ|F0ϕ1(z)F0ϕ1(w)||\phi\circ F_{0}\circ\phi^{-1}(z)-\phi\circ F_{0}\circ\phi^{-1}(w)|=\lambda|F_{0}\circ\phi^{-1}(z)-F_{0}\circ\phi^{-1}(w)|

from which it easily follows that ϕF0ϕ1\phi\circ F_{0}\circ\phi^{-1} is L0L_{0}-bi-Lipschitz.

Next, given ϵ>0\epsilon>0, find δ>0\delta>0 so that the two conditions in Definition 4.1 hold for FtF_{t}. If s,t[0,1]s,t\in[0,1] with |st|<δ|s-t|<\delta, then

|ϕFtϕ1(z)ϕFsϕ1(z))|=λ|Ftϕ1(z)Fsϕ1(z)|<λϵcϵ|\phi\circ F_{t}\circ\phi^{-1}(z)-\phi\circ F_{s}\circ\phi^{-1}(z))|=\lambda|F_{t}\circ\phi^{-1}(z)-F_{s}\circ\phi^{-1}(z)|<\lambda\epsilon\leq c\epsilon

and we conclude property (i) of Definition 4.1 holds uniformly in \mathcal{F}. Finally,

|ϕFtϕ1(z)ϕFtϕ1(w)|\displaystyle|\phi\circ F_{t}\circ\phi^{-1}(z)-\phi\circ F_{t}\circ\phi^{-1}(w)| =λ|Ftϕ1(z)Ftϕ1(w)|\displaystyle=\lambda|F_{t}\circ\phi^{-1}(z)-\circ F_{t}\circ\phi^{-1}(w)|
λ(1+ϵ)|Ftϕ1(z)Ftϕ1(w)|\displaystyle\leq\lambda(1+\epsilon)|F_{t}\circ\phi^{-1}(z)-\circ F_{t}\circ\phi^{-1}(w)|
=(1+ϵ)|ϕFsϕ1(z)ϕFsϕ1(w)|,\displaystyle=(1+\epsilon)|\phi\circ F_{s}\circ\phi^{-1}(z)-\phi\circ F_{s}\circ\phi^{-1}(w)|,

from which we conclude condition (ii) in Definition 4.1 holds uniformly in \mathcal{F}. ∎

4.2. Bi-Lipschitz paths on triangles

As part of our construction, we will be using specific bi-Lipschitz paths which deform triangles in 2\mathbb{R}^{2}. Let TT be a triangle in \mathbb{C}. If the vertices are w1,w2,w3w_{1},w_{2},w_{3}, taken in counterclockwise order, then we may also denote this triangle by T(w1,w2,w3)T(w_{1},w_{2},w_{3}). In our construction, there will be two triangles T1T_{1} and T2T_{2} which share a vertex, and a bi-Lipschitz path G:[0,1]LIP(2)G:[0,1]\to LIP(\mathbb{R}^{2}) such that G0G_{0} is the identity in T1T_{1} and G1G_{1} is an affine map from T1T_{1} onto T2T_{2}. We will focus on constructing GtG_{t} inside T1T_{1}.

After conjugating by an affine map, we may assume that T1T_{1} and T2T_{2} share 0 as a vertex, that T1=T(0,1,a)T_{1}=T(0,1,a) and T2=T(0,c,b)T_{2}=T(0,c,b) for 0<arga<π0<\arg a<\pi, 0argc<argbπ0\leq\arg c<\arg b\leq\pi and argbargc<π\arg b-\arg c<\pi. These restrictions ensure that neither T1T_{1} nor T2T_{2} degenerate to line segments.

Proposition 4.9.

There exists a bi-Lipschitz path G:[0,1]LIP(T1)G:[0,1]\to LIP(T_{1}) such that G0G_{0} is the identity and G1G_{1} is the map given by

G1(z)=(ba¯caa¯)z+(acbaa¯)z¯.G_{1}(z)=\left(\frac{b-\overline{a}c}{a-\overline{a}}\right)z+\left(\frac{ac-b}{a-\overline{a}}\right)\overline{z}.
Proof.

First, every real-linear map in \mathbb{C} is of the form Az+Bz¯Az+B\overline{z} for A,BA,B\in\mathbb{C}, and since we require our maps to be orientation-preserving, we have |A|>|B||A|>|B|. Given T1=T(0,1,a)T_{1}=T(0,1,a) and T2=(0,c,b)T_{2}=(0,c,b), it is elementary to check that

g1(z)=(ba¯caa¯)z+(acbaa¯)z¯g_{1}(z)=\left(\frac{b-\overline{a}c}{a-\overline{a}}\right)z+\left(\frac{ac-b}{a-\overline{a}}\right)\overline{z}

fixes 0, maps 11 to cc and maps aa to bb.

Set γ1(t)=ct+(1t)\gamma_{1}(t)=ct+(1-t) for 0t10\leq t\leq 1 and γ2(t)=bt+(1t)a\gamma_{2}(t)=bt+(1-t)a for 0t10\leq t\leq 1. Then define

Gt(z)=(γ2(t)a¯γ1(t)aa¯)z+(aγ1(t)γ2(t)aa¯)z¯.G_{t}(z)=\left(\frac{\gamma_{2}(t)-\overline{a}\gamma_{1}(t)}{a-\overline{a}}\right)z+\left(\frac{a\gamma_{1}(t)-\gamma_{2}(t)}{a-\overline{a}}\right)\overline{z}.

For any t[0,1]t\in[0,1], GtG_{t} maps T1T_{1} onto the triangle with vertices 0,γ1(t)0,\gamma_{1}(t) and γ2(t)\gamma_{2}(t).

For ease of notation, we define the map h:=GtGs1h:=G_{t}\circ G_{s}^{-1}, which maps T(0,γ1(s),γ2(s))T(0,\gamma_{1}(s),\gamma_{2}(s)) onto T(0,γ1(t),γ2(t))T(0,\gamma_{1}(t),\gamma_{2}(t)). Since h(z)=γ1(t)α(z/γ1(s))h(z)=\gamma_{1}(t)\alpha(z/\gamma_{1}(s)), where α\alpha maps T(0,1,γ2(s)/γ1(s))T(0,1,\gamma_{2}(s)/\gamma_{1}(s)) onto T(0,1,γ2(t)/γ1(t))T(0,1,\gamma_{2}(t)/\gamma_{1}(t)), we can compute that

(4.1) h(z)\displaystyle h(z) =(γ2(t)/γ1(t)γ2(s)/γ1(s)¯γ2(s)/γ1(s)γ2(s)/γ1(s)¯)(γ1(t)zγ1(s))\displaystyle=\left(\frac{\gamma_{2}(t)/\gamma_{1}(t)-\overline{\gamma_{2}(s)/\gamma_{1}(s)}}{\gamma_{2}(s)/\gamma_{1}(s)-\overline{\gamma_{2}(s)/\gamma_{1}(s)}}\right)\left(\frac{\gamma_{1}(t)z}{\gamma_{1}(s)}\right)
+(γ2(s)/γ1(s)γ2(t)/γ1(t)γ2(s)/γ1(s)γ2(s)/γ1(s)¯)(γ1(t)z¯γ1(s)¯).\displaystyle\qquad+\left(\frac{\gamma_{2}(s)/\gamma_{1}(s)-\gamma_{2}(t)/\gamma_{1}(t)}{\gamma_{2}(s)/\gamma_{1}(s)-\overline{\gamma_{2}(s)/\gamma_{1}(s)}}\right)\left(\frac{\gamma_{1}(t)\overline{z}}{\overline{\gamma_{1}(s)}}\right).

We collect some estimates that we will need. First we may suppose there exists R>0R>0 so that

(4.2) 1R|γi(t)|R,\frac{1}{R}\leq|\gamma_{i}(t)|\leq R,

for all t[0,1]t\in[0,1] and i{1,2}i\in\{1,2\}. Note also that |z|R|z|\leq R for zT(0,γ1(s),γ2(s))z\in T(0,\gamma_{1}(s),\gamma_{2}(s)). Next, since the interior of every open triangle Gt(T(0,1,a))G_{t}(T(0,1,a)) is contained in the upper half-plane and the triangles do not degenerate, there exists r>0r>0 so that

(4.3) supt[0,1]|γ2(t)γ1(t)(γ2(t)γ1(t))¯|>r.\sup_{t\in[0,1]}\ \left|\frac{\gamma_{2}(t)}{\gamma_{1}(t)}-\overline{\left(\frac{\gamma_{2}(t)}{\gamma_{1}(t)}\right)}\right|>r.

We also observe that

(4.4) |γ1(t)γ1(s)|=|c1||ts|,|γ2(t)γ2(s)|=|ba||ts|,|\gamma_{1}(t)-\gamma_{1}(s)|=|c-1|\cdot|t-s|,\quad|\gamma_{2}(t)-\gamma_{2}(s)|=|b-a|\cdot|t-s|,

and hence that

(4.5) |γ1(t)γ1(s)|1+|γ1(t)γ1(s)||γ1(s)|1+R|c1||ts|.\left|\frac{\gamma_{1}(t)}{\gamma_{1}(s)}\right|\leq 1+\frac{|\gamma_{1}(t)-\gamma_{1}(s)|}{|\gamma_{1}(s)|}\leq 1+R|c-1||t-s|.

Finally, via an elementary calculation we have

γ2(s)γ1(s)γ2(t)γ1(t)=(bac)(st)(1+t(c1))(1+s(c1))\frac{\gamma_{2}(s)}{\gamma_{1}(s)}-\frac{\gamma_{2}(t)}{\gamma_{1}(t)}=\frac{(b-ac)(s-t)}{(1+t(c-1))(1+s(c-1))}

and hence by (4.2) we have

(4.6) |γ2(s)γ1(s)γ2(t)γ1(t)|R2|bac||st|.\left|\frac{\gamma_{2}(s)}{\gamma_{1}(s)}-\frac{\gamma_{2}(t)}{\gamma_{1}(t)}\right|\leq R^{2}|b-ac||s-t|.

We can now prove property (i) for showing GG is a bi-Lipschitz path. Given ϵ>0\epsilon>0, we choose δ<ϵ/ξ\delta<\epsilon/\xi, with ξ\xi chosen below, so that if |ts|<δ|t-s|<\delta with s,t[0,1]s,t\in[0,1], then by (4.1),

|h(z)z|\displaystyle|h(z)-z| =|z(γ2(t)γ2(s)γ1(s)+(γ2(s)¯|γ1(s)|2)(γ1(s)γ1(t)))(γ2(s)γ1(s)γ2(s)γ1(s)¯)1\displaystyle=\left|z\left(\frac{\gamma_{2}(t)-\gamma_{2}(s)}{\gamma_{1}(s)}+\left(\frac{\overline{\gamma_{2}(s)}}{|\gamma_{1}(s)|^{2}}\right)\cdot\left(\gamma_{1}(s)-\gamma_{1}(t)\right)\right)\left(\frac{\gamma_{2}(s)}{\gamma_{1}(s)}-\overline{\frac{\gamma_{2}(s)}{\gamma_{1}(s)}}\right)^{-1}\right.
+z¯(γ2(s)/γ1(s)γ2(t)/γ1(t)γ2(s)/γ1(s)γ2(s)/γ1(s)¯)(γ1(t)γ1(s)¯)|.\displaystyle\qquad+\left.\overline{z}\left(\frac{\gamma_{2}(s)/\gamma_{1}(s)-\gamma_{2}(t)/\gamma_{1}(t)}{\gamma_{2}(s)/\gamma_{1}(s)-\overline{\gamma_{2}(s)/\gamma_{1}(s)}}\right)\left(\frac{\gamma_{1}(t)}{\overline{\gamma_{1}(s)}}\right)\right|.

Using (4.2), (4.3) , (4.4) and (4.6), we obtain

|h(z)z|Rr(R|ba||ts|+R3|c1||ts|)+R5|bac|r|ts|.|h(z)-z|\leq\frac{R}{r}\left(R|b-a||t-s|+R^{3}|c-1||t-s|\right)+\frac{R^{5}|b-ac|}{r}|t-s|.

By choosing

ξ=R2|ba|r+R4|c1|r+R5|bac|r,\xi=\frac{R^{2}|b-a|}{r}+\frac{R^{4}|c-1|}{r}+\frac{R^{5}|b-ac|}{r},

we obtain |h(z)z|<ϵ|h(z)-z|<\epsilon for zT(0,γ1(s),γ2(s))z\in T(0,\gamma_{1}(s),\gamma_{2}(s)) as required.

Next, we prove property (ii). If h(z)=Az+Bz¯h(z)=Az+B\overline{z} is orientation preserving, then |A|>|B||A|>|B| and hh is bi-Lipschitz with isometric distortion given by

(4.7) max{|A|+|B|,1|A||B|}.\max\left\{|A|+|B|,\frac{1}{|A|-|B|}\right\}.

In our setting,

|A|=|γ1(t)γ1(s)||γ2(t)γ1(t)γ2(s)γ1(s)¯||γ2(s)γ1(s)γ2(s)γ1(s)¯|1.|A|=\left|\frac{\gamma_{1}(t)}{\gamma_{1}(s)}\right|\cdot\left|\frac{\gamma_{2}(t)}{\gamma_{1}(t)}-\overline{\frac{\gamma_{2}(s)}{\gamma_{1}(s)}}\right|\cdot\left|\frac{\gamma_{2}(s)}{\gamma_{1}(s)}-\overline{\frac{\gamma_{2}(s)}{\gamma_{1}(s)}}\right|^{-1}.

We can compute that

γ2(t)/γ1(t)γ2(s)/γ1(s)¯γ2(s)/γ1(s)γ2(s)/γ1(s)¯=1+γ2(t)/γ1(t)γ2(s)/γ1(s)γ2(s)/γ1(s)γ2(s)/γ1(s)¯\frac{\gamma_{2}(t)/\gamma_{1}(t)-\overline{\gamma_{2}(s)/\gamma_{1}(s)}}{\gamma_{2}(s)/\gamma_{1}(s)-\overline{\gamma_{2}(s)/\gamma_{1}(s)}}=1+\frac{\gamma_{2}(t)/\gamma_{1}(t)-\gamma_{2}(s)/\gamma_{1}(s)}{\gamma_{2}(s)/\gamma_{1}(s)-\overline{\gamma_{2}(s)/\gamma_{1}(s)}}

and hence by (4.5) and (4.6) we obtain

(4.8) |A|(1+R|c1||ts|)(1+R2|bac|r|ts|).|A|\leq\left(1+R|c-1||t-s|\right)\left(1+\frac{R^{2}|b-ac|}{r}|t-s|\right).

We also have that

|B|=|γ1(t)γ1(s)¯||γ2(t)γ1(t)γ2(s)γ1(s)||γ2(s)γ1(s)γ2(s)γ1(s)¯|1|B|=\left|\frac{\gamma_{1}(t)}{\overline{\gamma_{1}(s)}}\right|\cdot\left|\frac{\gamma_{2}(t)}{\gamma_{1}(t)}-\frac{\gamma_{2}(s)}{\gamma_{1}(s)}\right|\cdot\left|\frac{\gamma_{2}(s)}{\gamma_{1}(s)}-\overline{\frac{\gamma_{2}(s)}{\gamma_{1}(s)}}\right|^{-1}

and it follows from the comment after (4.2), (4.3) and (4.6) that

(4.9) |B|R4|bac|r|st|.|B|\leq\frac{R^{4}|b-ac|}{r}|s-t|.

It then follows easily from (4.7), (4.8) and (4.9) that given ϵ>0\epsilon>0 we can choose δ>0\delta>0 so that if s,t[0,1]s,t\in[0,1] with |st|<δ|s-t|<\delta, then GtGs1G_{t}\circ G_{s}^{-1} is (1+ϵ)(1+\epsilon)-bi-Lipschitz, completing the proof that GtG_{t} is a bi-Lipschitz path. ∎

4.3. Dehn twists and conjugates

Suppose R2R\subset\mathbb{R}^{2} is a ring domain. Then we can define a Dehn twist 𝔇\mathfrak{D} in RR as follows. There exists η(0,1)\eta\in(0,1) and a conformal map bijection g:SRg:S\to R, where S=B(0,1)B¯(0,1η)S=B(0,1)\setminus\overline{B}(0,1-\eta) is a round ring. By the conformal invariance of the modulus of ring domains, η\eta is uniquely defined. The Dehn twist in SS is given in polar coordinates by

𝔇(r,θ)=(r,θ+2π1η(1r)),\mathfrak{D}(r,\theta)=\left(r,\theta+2\pi\tfrac{1}{\eta}(1-r)\right),

and then the Dehn twist in RR is given by g𝔇g1g\circ\mathfrak{D}\circ g^{-1}.

Lemma 4.10.

Let η>0\eta>0, and consider SS and 𝔇\mathfrak{D} as above. For 0t10\leq t\leq 1 and reiθSre^{i\theta}\in S, set

Dt(r,θ)=(r,θ+2πt1η(1r)).D_{t}(r,\theta)=\left(r,\theta+2\pi t\tfrac{1}{\eta}(1-r)\right).

Then DD is a bi-Lipschitz path in SS connecting the identity to 𝔇\mathfrak{D}.

Proof.

For convenience, for 1ηr11-\eta\leq r\leq 1, set h(r)=1η(1r)h(r)=\frac{1}{\eta}(1-r). Let z=reiθSz=re^{i\theta}\in S. If s,t[0,1]s,t\in[0,1], we have

|Dt(z)Ds(z)|\displaystyle|D_{t}(z)-D_{s}(z)| =|ze2πith(|z|)ze2πish(|z|)|=|z||1e2πi(st)h(|z|)|.\displaystyle=|ze^{2\pi ith(|z|)}-ze^{2\pi ish(|z|)}|=|z|\left|1-e^{2\pi i(s-t)h(|z|)}\right|.

Since |z|1|z|\leq 1 and h(|z|)[0,1]h(|z|)\in[0,1], it is clear that the first condition in the definition of a bi-Lipschitz path is satisfied.

Next, it is clear that

Dt1(reiθ)=rei(θ2πth(r)),D_{t}^{-1}(re^{i\theta})=re^{i(\theta-2\pi th(r))},

and hence for z,wSz,w\in S and s,t[0,1]s,t\in[0,1] we have

|DsDt1(z)DsDt1(w)|\displaystyle|D_{s}\circ D_{t}^{-1}(z)-D_{s}\circ D_{t}^{-1}(w)| =|ze2πih(|z|)(st)we2πih(|w|)(st)|\displaystyle=|ze^{2\pi ih(|z|)(s-t)}-we^{2\pi ih(|w|)(s-t)}|
=|zwe2πi(st)(h(|w|)h(|z|))|.\displaystyle=|z-we^{2\pi i(s-t)(h(|w|)-h(|z|))}|.

Since hh is linear, it follows that

|h(|w|)h(|z|)|=1η||w||z||1η|wz|.|h(|w|)-h(|z|)|=\tfrac{1}{\eta}\left||w|-|z|\right|\leq\tfrac{1}{\eta}|w-z|.

We conclude that there exists C>0C>0 independent of z,wz,w such that

|DsDt1(z)DsDt1(w)|\displaystyle|D_{s}\circ D_{t}^{-1}(z)-D_{s}\circ D_{t}^{-1}(w)| =|zwe2πi(st)(h(|w|)h(|z|))|\displaystyle=|z-we^{2\pi i(s-t)(h(|w|)-h(|z|))}|
|zw|+|w||1e2πi(st)(h(|w|)h(|z|))|\displaystyle\leq|z-w|+|w|\left|1-e^{2\pi i(s-t)(h(|w|)-h(|z|))}\right|
|zw|(1+C|st|),\displaystyle\leq|z-w|(1+C|s-t|),

from which the second condition in the definition of a bi-Lipschitz path is satisfied. ∎

We will need to know that conformal conjugates of DtD_{t} are also bi-Lipschitz paths. As was observed in [FM12, Remark 2.6], the conjugate of a bi-Lipschitz path on a closed manifold by a conformal map is a bi-Lipschitz path and this cannot be weakened to conjugation by a diffeomorphism. However, here we have closed ring domains, and so this remark does not immediately apply.

Proposition 4.11.

Suppose SS is a round annulus, RR is a ring domain with smooth boundary components, g:SRg:S\to R is a conformal map and F:[0,1]LIP(S)F:[0,1]\to LIP(S) is a bi-Lipschitz path such that Ft(S)=SF_{t}(S)=S for all t[0,1]t\in[0,1]. If H=gFg1H=g\circ F\circ g^{-1}, then HH is a bi-Lipschitz path with Ht(R)=RH_{t}(R)=R for each t[0,1]t\in[0,1].

Proof.

We start with condition (i) from Definition 4.1. Since the boundary components of RR are assumed smooth, the Riemann map gg and all its derivatives extend continuously to S\partial S, see [BK87, p.24]. In particular, there exists an upper bound MM for both |g||g^{\prime}| and |(g1)||(g^{-1})^{\prime}| on S¯\overline{S} and R¯\overline{R}, respectively. Hence gg and g1g^{-1} are MM-bi-Lipschitz maps.

Given ϵ>0\epsilon>0, find δ>0\delta>0 so that if s,t[0,1]s,t\in[0,1] satisfy |st|<δ|s-t|<\delta then

|FsFt1(z)z|<ϵ/M|F_{s}\circ F_{t}^{-1}(z)-z|<\epsilon/M

for all zRz\in R. Then

|HsHt1(z)z|\displaystyle|H_{s}\circ H_{t}^{-1}(z)-z| =|gFsFt1g1(z)g(g1(z))|\displaystyle=|g\circ F_{s}\circ F_{t}^{-1}\circ g^{-1}(z)-g(g^{-1}(z))|
M|FsFt1g1(z)g1(z)|\displaystyle\leq M|F_{s}\circ F_{t}^{-1}\circ g^{-1}(z)-g^{-1}(z)|
ϵ.\displaystyle\leq\epsilon.

Hence condition (i) holds.

Next, for condition (ii), given ϵ>0\epsilon>0, find δ>0\delta>0 so that if s,t[0,1]s,t\in[0,1] with

(4.10) |st|<δ, then FsFt1 is (1+ϵ)-bi-Lipschitz.|s-t|<\delta,\quad\text{ then }F_{s}\circ F_{t}^{-1}\text{ is $(1+\epsilon)$-bi-Lipschitz.}

Consider the functions p:S¯×S¯p:\overline{S}\times\overline{S}\to\mathbb{C} and q:R¯×R¯q:\overline{R}\times\overline{R}\to\mathbb{C} defined by

(4.11) p(z,w)={g(z)g(w)zwg(w) if zw0 if z=wp(z,w)=\begin{cases}\frac{g(z)-g(w)}{z-w}-g^{\prime}(w)&\text{ if $z\neq w$}\\ 0&\text{ if $z=w$}\end{cases}

and

(4.12) q(z,w)={g1(z)g1(w)zw(g1)(w) if zw0 if z=w.q(z,w)=\begin{cases}\frac{g^{-1}(z)-g^{-1}(w)}{z-w}-(g^{-1})^{\prime}(w)&\text{ if $z\neq w$}\\ 0&\text{ if $z=w$.}\end{cases}

By differentiability of gg and g1g^{-1}, both pp and qq are continuous functions on compact sets in 2\mathbb{C}^{2} and hence bounded. That is, there exists C>0C>0 so that |p(z,w)|C|p(z,w)|\leq C for all (z,w)S¯×S¯(z,w)\in\overline{S}\times\overline{S}, and |q(z,w)|C|q(z,w)|\leq C, for all (z,w)R¯×R¯(z,w)\in\overline{R}\times\overline{R}. Hence given ϵ>0\epsilon^{\prime}>0, there exists r>0r>0 so that if z,wS¯z,w\in\overline{S} with |zw|<r|z-w|<r, then

(4.13) |p(z,w)|<ϵ.|p(z,w)|<\epsilon^{\prime}.

By reducing rr if necessary, by the same reasoning we can also assume that if z,wR¯z,w\in\overline{R} with |zw|<r|z-w|<r then

(4.14) |q(z,w)|<ϵ.|q(z,w)|<\epsilon^{\prime}.

Now, let z,wRz,w\in R with |zw|<r/[M(1+ϵ)]|z-w|<r/[M(1+\epsilon)]. Set u=FsFt1g1(z)u=F_{s}\circ F_{t}^{-1}\circ g^{-1}(z) and v=FsFt1g1(w)v=F_{s}\circ F_{t}^{-1}\circ g^{-1}(w). Then since FsFt1F_{s}\circ F_{t}^{-1} is (1+ϵ)(1+\epsilon)-bi-Lipschitz and g1g^{-1} is MM-bi-Lipschitz, we have |uv|<r|u-v|<r. Hence by (4.11) we have

(4.15) |HsHt1(z)HsHt1(w)|=|g(u)g(v)|=|g(v)+p(u,v)||uv|.|H_{s}\circ H_{t}^{-1}(z)-H_{s}\circ H_{t}^{-1}(w)|=|g(u)-g(v)|=|g^{\prime}(v)+p(u,v)|\cdot|u-v|.

Next, again using the fact that FsFt1F_{s}\circ F_{t}^{-1} is (1+ϵ)(1+\epsilon)-bi-Lipschitz, we obtain

(4.16) |uv|(1+ϵ)|g1(z)g1(w)|.|u-v|\leq(1+\epsilon)|g^{-1}(z)-g^{-1}(w)|.

Using (4.12), we have

(4.17) |g1(z)g1(w)|=|(g1)(w)+q(z,w)||zw|.|g^{-1}(z)-g^{-1}(w)|=|(g^{-1})^{\prime}(w)+q(z,w)|\cdot|z-w|.

Combining (4.15), (4.16) and (4.17), we obtain

(4.18) |HsHt1(z)HsHt1(w)||g(v)+p(u,v)|(1+ϵ)|(g1)(w)+q(z,w)||zw|.|H_{s}\circ H_{t}^{-1}(z)-H_{s}\circ H_{t}^{-1}(w)|\leq|g^{\prime}(v)+p(u,v)|\cdot(1+\epsilon)\cdot|(g^{-1})^{\prime}(w)+q(z,w)|\cdot|z-w|.

Next, we have

|g(v)\displaystyle|g^{\prime}(v) (g1)(w)|\displaystyle(g^{-1})^{\prime}(w)|
=|g(g1(w))(g1)(w)+[g(Fs(Ft1(g1(w))))g(g1(w))](g1)(w)|.\displaystyle=|g^{\prime}(g^{-1}(w))(g^{-1})^{\prime}(w)+\left[g^{\prime}(F_{s}(F_{t}^{-1}(g^{-1}(w))))-g^{\prime}(g^{-1}(w))\right](g^{-1})^{\prime}(w)|.

Using condition (i) of FF being a bi-Lipschitz path, the fact that |(g1)||(g^{-1})^{\prime}| is bounded and the fact that gg^{\prime} is uniformly continuous on S¯\overline{S}, by shrinking δ\delta if necessary, we may conclude that

(4.19) |g(v)(g1)(w)|1+ϵ.|g^{\prime}(v)(g^{-1})^{\prime}(w)|\leq 1+\epsilon.

Combining (4.10), (4.13), (4.14), (4.18), (4.19),and the bounds for the derivatives of gg, g1g^{-1} we obtain

|HsHt1(z)HsHt1(w)|\displaystyle|H_{s}\circ H_{t}^{-1}(z)-H_{s}\circ H_{t}^{-1}(w)| (1+ϵ)(|g(v)(g1)(w)|+|p(u,v)|(g1))(w)|\displaystyle\leq(1+\epsilon)\left(|g^{\prime}(v)(g^{-1})^{\prime}(w)|+|p(u,v)|(g^{-1})^{\prime})(w)|\right.
+|q(z,w)g(v)|+|p(u,v)q(z,w)|)|zw|\displaystyle\left.\quad+\>|q(z,w)g^{\prime}(v)|+|p(u,v)q(z,w)|\right)|z-w|
(1+ϵ)((1+ϵ)+2Mϵ+(ϵ)2)|zw|.\displaystyle\leq(1+\epsilon)\left((1+\epsilon)+2M\epsilon^{\prime}+(\epsilon^{\prime})^{2}\right)|z-w|.

In particular, given η>0\eta>0 we can find δ>0\delta>0 and r>0r>0 so that if s,t[0,1]s,t\in[0,1] with |st|<δ|s-t|<\delta, then for any z,wR¯z,w\in\overline{R} with |zw|<r|z-w|<r we have

(4.20) |HsHt1(z)HsHt1(w)|(1+η)|zw|.|H_{s}\circ H_{t}^{-1}(z)-H_{s}\circ H_{t}^{-1}(w)|\leq(1+\eta)|z-w|.

To show that condition (ii) holds, suppose for a contradiction that it does not. Then we can find η>0\eta>0 and sequences sn,tns_{n},t_{n} in [0,1][0,1] with |sntn|0|s_{n}-t_{n}|\to 0 and sequences zn,wnz_{n},w_{n} in R¯\overline{R} for which

(4.21) |HsnHtn1(zn)HsnHtn1(wn)znwn|>1+η\left|\frac{H_{s_{n}}\circ H_{t_{n}}^{-1}(z_{n})-H_{s_{n}}\circ H_{t_{n}}^{-1}(w_{n})}{z_{n}-w_{n}}\right|>1+\eta

for all nn. By passing to subsequences, we may assume that znz0z_{n}\to z_{0} and wnw0w_{n}\to w_{0}. If z0=w0z_{0}=w_{0} then we obtain a contradiction to (4.20). Otherwise, suppose |z0w0|=ξ|z_{0}-w_{0}|=\xi and find NNN\in N so that if nNn\geq N then |znwn|>ξ/2|z_{n}-w_{n}|>\xi/2. By condition (i), we have

|HsnHtn1(zn)\displaystyle|H_{s_{n}}\circ H_{t_{n}}^{-1}(z_{n}) HsnHtn1(wn)|\displaystyle-H_{s_{n}}\circ H_{t_{n}}^{-1}(w_{n})|
|HsnHtn1(zn)zn|+|znwn|+|HsnHtn1(wn)wn|\displaystyle\leq|H_{s_{n}}\circ H_{t_{n}}^{-1}(z_{n})-z_{n}|+|z_{n}-w_{n}|+|H_{s_{n}}\circ H_{t_{n}}^{-1}(w_{n})-w_{n}|
|znwn|+2ϵ\displaystyle\leq|z_{n}-w_{n}|+2\epsilon
(1+4ϵξ)|znwn|.\displaystyle\leq(1+\frac{4\epsilon}{\xi})|z_{n}-w_{n}|.

Since |sntn|0|s_{n}-t_{n}|\to 0, we can choose nn large enough so that 4ϵ/ξ<η4\epsilon/\xi<\eta and hence contradict (4.21). We conclude that condition (ii) holds and hence HtH_{t} is a bi-Lipschitz path. ∎

4.4. Interpolation in an annulus

In this subsection, we will prove the following interpolation result.

Proposition 4.12.

Suppose T>1T>1, and let R={z:1|z|T}R=\{z\in\mathbb{C}:1\leq|z|\leq T\} with boundary components S1={z:|z|=1}S_{1}=\{z:|z|=1\} and ST={z:|z|=T}S_{T}=\{z:|z|=T\}. Let P:[0,1]LIP(S1)P:[0,1]\to LIP(S_{1}) and Q:[0,1]LIP(ST)Q:[0,1]\to LIP(S_{T}) be bi-Lipschitz paths such that P0=P1P_{0}=P_{1} is the identity on S1S_{1}, Q0=Q1Q_{0}=Q_{1} is the identity on STS_{T}, and argPt,argQt\arg P_{t},\arg Q_{t} are strictly increasing in tt. Then there exists a bi-Lipschitz path F:[0,1]LIP(R)F:[0,1]\to LIP(R), with F0=F1F_{0}=F_{1} the identity on RR and FtF_{t} extends PtP_{t} and QtQ_{t} for each tt.

We start with the following fairly elementary estimate.

Lemma 4.13.

Suppose that c1,c2>0c_{1},c_{2}>0, c3[3,3]c_{3}\in[-3,3] and aa\in\mathbb{R}. For any ϵ>0\epsilon>0 and any δ1,δ2(ϵ,ϵ)\delta_{1},\delta_{2}\in(-\epsilon,\epsilon), δ3(2ϵ,2ϵ)\delta_{3}\in(-2\epsilon,2\epsilon), we have

|a+i(c1+c2+c3x)+i(δ1c1+δ2c2+δ3a)|(1+8ϵ)|a+i(c1+c2+c3a)|.|a+i(c_{1}+c_{2}+c_{3}x)+i(\delta_{1}c_{1}+\delta_{2}c_{2}+\delta_{3}a)|\leq(1+8\epsilon)|a+i(c_{1}+c_{2}+c_{3}a)|.
Proof.

We consider three cases.

Case 1: ac30ac_{3}\geq 0. Then,

|a+i(c1+c2+c3a)\displaystyle|a+i(c_{1}+c_{2}+c_{3}a) +i(δ1c1+δ2c2+δ3a)|\displaystyle+i(\delta_{1}c_{1}+\delta_{2}c_{2}+\delta_{3}a)|
|a+i(c1+c2+c3a)|+|δ1|c1+|δ2|c2+|δ3||a|\displaystyle\leq|a+i(c_{1}+c_{2}+c_{3}a)|+|\delta_{1}|c_{1}+|\delta_{2}|c_{2}+|\delta_{3}||a|
|a+i(c1+c2+c3a)|+(2ϵ)|i(c1+c2+c3a)|+2ϵ|a|\displaystyle\leq|a+i(c_{1}+c_{2}+c_{3}a)|+(2\epsilon)|i(c_{1}+c_{2}+c_{3}a)|+2\epsilon|a|
(1+4ϵ)|a+i(c1+c2+c3a)|\displaystyle\leq(1+4\epsilon)|a+i(c_{1}+c_{2}+c_{3}a)|

Case 2: ac3<0ac_{3}<0 and c1+c2+c3a12(c1+c2)c_{1}+c_{2}+c_{3}a\leq\frac{1}{2}(c_{1}+c_{2}). We have that

|a|>12|c3|(c1+c2)16(c1+c2).|a|>\frac{1}{2|c_{3}|}(c_{1}+c_{2})\geq\frac{1}{6}(c_{1}+c_{2}).

Therefore,

|a+i(c1+c2+c3a)\displaystyle|a+i(c_{1}+c_{2}+c_{3}a) +i(δ1c1+δ2c2+δ3a)|\displaystyle+i(\delta_{1}c_{1}+\delta_{2}c_{2}+\delta_{3}a)|
|a+i(c1+c2+c3a)|+|δ1|c1+|δ2|c2+|δ3||a|\displaystyle\leq|a+i(c_{1}+c_{2}+c_{3}a)|+|\delta_{1}|c_{1}+|\delta_{2}|c_{2}+|\delta_{3}||a|
|a+i(c1+c2+c3a)|+ϵ(c1+c2)+2ϵ|a|\displaystyle\leq|a+i(c_{1}+c_{2}+c_{3}a)|+\epsilon(c_{1}+c_{2})+2\epsilon|a|
|a+i(c1+c2+c3a)|+8ϵ|a|\displaystyle\leq|a+i(c_{1}+c_{2}+c_{3}a)|+8\epsilon|a|
(1+8ϵ)|a+i(c1+c2+c3a)|.\displaystyle\leq(1+8\epsilon)|a+i(c_{1}+c_{2}+c_{3}a)|.

Case 3: ac3<0ac_{3}<0 and c1+c2+c3a>12(c1+c2)c_{1}+c_{2}+c_{3}a>\frac{1}{2}(c_{1}+c_{2}). We have that

|a+i(c1+c2+c3a)\displaystyle|a+i(c_{1}+c_{2}+c_{3}a) +i(δ1c1+δ2c2+δ3a)|\displaystyle+i(\delta_{1}c_{1}+\delta_{2}c_{2}+\delta_{3}a)|
|a+i(c1+c2+c3a)|+|δ1|c1+|δ2|c2+|δ3||a|\displaystyle\leq|a+i(c_{1}+c_{2}+c_{3}a)|+|\delta_{1}|c_{1}+|\delta_{2}|c_{2}+|\delta_{3}||a|
|a+i(c1+c2+c3a)|+ϵ(c1+c2)+2ϵ|a|\displaystyle\leq|a+i(c_{1}+c_{2}+c_{3}a)|+\epsilon(c_{1}+c_{2})+2\epsilon|a|
|a+i(c1+c2+c3a)|+2ϵ|i(c1+c2+c3a)|+2ϵ|a|\displaystyle\leq|a+i(c_{1}+c_{2}+c_{3}a)|+2\epsilon|i(c_{1}+c_{2}+c_{3}a)|+2\epsilon|a|
(1+4ϵ)|a+i(c1+c2+c3a)|.\displaystyle\leq(1+4\epsilon)|a+i(c_{1}+c_{2}+c_{3}a)|.\qed

Next, we prove an interpolation result on strips.

Lemma 4.14.

Suppose that F,G:[0,1]LIP()F,G:[0,1]\to LIP(\mathbb{R}) be bi-Lipschitz paths with F0(x)=G0(x)=xF_{0}(x)=G_{0}(x)=x for all xx\in\mathbb{R}, F1(x)=G1(x)=x+2πF_{1}(x)=G_{1}(x)=x+2\pi for all xx\in\mathbb{R}, Ft,GtF_{t},G_{t} are 2π2\pi-periodic for all t[0,1]t\in[0,1] and Ft(x),Gt(x)F_{t}(x),G_{t}(x) are both strictly increasing in tt for a fixed xx. Let M>0M>0 and let SS be the strip S={z:0<Re(z)<M}S=\{z\in\mathbb{C}:0<\operatorname{Re}(z)<M\}. Then there exists a bi-Lipschitz path H:[0,1]LIP(S¯)H:[0,1]\to LIP(\overline{S}) which extends to S\partial S with Ht(iy)=iFt(y)H_{t}(iy)=iF_{t}(y) and Ht(M+iy)=M+iGt(y)H_{t}(M+iy)=M+iG_{t}(y) for 0t10\leq t\leq 1 and yy\in\mathbb{R}. Moreover, H0(z)=zH_{0}(z)=z and H1(z)=z+2πH_{1}(z)=z+2\pi for all zSz\in S.

Proof.

We define HtH_{t} via the obvious convex interpolation in SS. That is, we set

Ht(x+iy)=x+i(Gt(y)+(1x/M)(Ft(y)Gt(y)))H_{t}(x+iy)=x+i\left(G_{t}(y)+(1-x/M)(F_{t}(y)-G_{t}(y))\right)

for 0t10\leq t\leq 1, 0xM0\leq x\leq M and yy\in\mathbb{R}. Clearly H0H_{0} is the identity and H1H_{1} is a translation by 2πi2\pi i. We need to show that HtH_{t} is a bi-Lipschitz path.

We start by showing that each HtH_{t} is a bi-Lipschitz map. Using Lemma 4.3, suppose that FtF_{t} is LL-bi-Lipschitz and GtG_{t} is λ\lambda-bi-Lipschitz for all t[0,1]t\in[0,1]. Setting z=x+iyz=x+iy and w=x+iyw=x^{\prime}+iy^{\prime}, we have

|Ht(z)Ht(w)|\displaystyle|H_{t}(z)-H_{t}(w)|
|xx|+|xMGt(y)xMGt(y)+(1xM)Ft(y)(1xMFt(y)|\displaystyle\leq|x-x^{\prime}|+\left|\tfrac{x}{M}G_{t}(y)-\tfrac{x^{\prime}}{M}G_{t}(y^{\prime})+(1-\tfrac{x}{M})F_{t}(y)-(1-\tfrac{x^{\prime}}{M}F_{t}(y^{\prime})\right|
|xx|+(1xM)|Ft(y)Ft(y)|+xM|Gt(y)Gt(y)|+|xx|M|Gt(y)Ft(y)|\displaystyle\leq|x-x^{\prime}|+(1-\tfrac{x}{M})|F_{t}(y)-F_{t}(y^{\prime})|+\tfrac{x}{M}|G_{t}(y)-G_{t}(y^{\prime})|+\frac{|x-x^{\prime}|}{M}|G_{t}(y^{\prime})-F_{t}(y^{\prime})|
(1+2π)|xx|+max{L,λ}|yy|\displaystyle\leq(1+2\pi)|x-x^{\prime}|+\max\{L,\lambda\}|y-y^{\prime}|
max{L,λ,1+2π}(|xx|+|yy|)\displaystyle\leq\max\{L,\lambda,1+2\pi\}(|x-x^{\prime}|+|y-y^{\prime}|)
2max{L,λ,1+2π}|zw|.\displaystyle\leq 2\max\{L,\lambda,1+2\pi\}|z-w|.

For the lower bound, we consider two cases. First, set C=min{M8πL,M8πλ,12}.\displaystyle C=\min\left\{\frac{M}{8\pi L},\frac{M}{8\pi\lambda},\frac{1}{2}\right\}.

Case 1. Suppose that |xx|C|yy||x-x^{\prime}|\geq C|y-y^{\prime}|. It follows that

|Ht(z)Ht(w)|\displaystyle|H_{t}(z)-H_{t}(w)| |xx|C2(|xx|+|yy|)C2|zw|.\displaystyle\geq|x-x^{\prime}|\geq\frac{C}{2}(|x-x^{\prime}|+|y-y^{\prime}|)\geq\frac{C}{2}|z-w|.

Case 2. Suppose that |xx|<C|yy||x-x^{\prime}|<C|y-y^{\prime}|. Without loss of generality, assume that yyy^{\prime}\leq y. Then,

|Ht\displaystyle|H_{t} (z)Ht(w)|\displaystyle(z)-H_{t}(w)|
|(1xM)(Ft(y)Ft(y))+xM(Gt(y)Gt(y))+(Gt(y)Ft(y))(xx)/M|\displaystyle\geq|(1-\tfrac{x}{M})(F_{t}(y)-F_{t}(y^{\prime}))+\tfrac{x}{M}(G_{t}(y)-G_{t}(y^{\prime}))+(G_{t}(y^{\prime})-F_{t}(y^{\prime}))(x-x^{\prime})/M|
(1xM)(Ft(y)Ft(y))+xM(Gt(y)Gt(y))|Gt(y)Ft(y)||xx|/M\displaystyle\geq(1-\tfrac{x}{M})(F_{t}(y)-F_{t}(y^{\prime}))+\tfrac{x}{M}(G_{t}(y)-G_{t}(y^{\prime}))-|G_{t}(y^{\prime})-F_{t}(y^{\prime})||x-x^{\prime}|/M
min{L1,λ1}|yy|2π|xx|/M\displaystyle\geq\min\{L^{-1},\lambda^{-1}\}|y-y^{\prime}|-2\pi|x-x^{\prime}|/M
(min{L1,λ1}2πCM)|yy|\displaystyle\geq\left(\min\{L^{-1},\lambda^{-1}\}-\frac{2\pi C}{M}\right)|y-y^{\prime}|
2πCM|yy|\displaystyle\geq\frac{2\pi C}{M}|y-y^{\prime}|
πCM(|xx|+|yy|)\displaystyle\geq\frac{\pi C}{M}(|x-x^{\prime}|+|y-y^{\prime}|)
πCM|zw|.\displaystyle\geq\frac{\pi C}{M}|z-w|.

Next, we show that hth_{t} satisfies condition (i) of Definition 4.1. From Definition 4.1 (i), by setting u=Ft1(x)u=F_{t}^{-1}(x), it follows that given ϵ>0\epsilon>0, we may find δ>0\delta>0 so that if |st|<δ|s-t|<\delta, then |Fs(u)Ft(u)|<ϵ|F_{s}(u)-F_{t}(u)|<\epsilon for all uu\in\mathbb{R}. The same holds true for GtG_{t}. Now suppose zSz\in S and ht1(z)=x+iyh_{t}^{-1}(z)=x+iy. Then we have

z=x+i(Gt(y)+(1xM)(Ft(y)Gt(y)))z=x+i\left(G_{t}(y)+(1-\tfrac{x}{M})(F_{t}(y)-G_{t}(y))\right)

and

HsHt1(z)=x+i(Gs(y)+(1xM)(Fs(y)Gs(y))).H_{s}\circ H_{t}^{-1}(z)=x+i\left(G_{s}(y)+(1-\tfrac{x}{M})(F_{s}(y)-G_{s}(y))\right).

Therefore, we obtain

|HsHt1(z)z|=|x(Gt(y)Gs(y))+(1xM)(Fs(y)Ft(y))|<ϵ.|H_{s}\circ H_{t}^{-1}(z)-z|=\left|x(G_{t}(y)-G_{s}(y))+(1-\tfrac{x}{M})(F_{s}(y)-F_{t}(y))\right|<\epsilon.

Hence condition (i) of Definition 4.1 is satisfied.

Finally, we show that hth_{t} satisfies condition (ii) of Definition 4.1. Note that

Hs(z)Hs(w)\displaystyle H_{s}(z)-H_{s}(w)
=(xx)+i[(1xM)Fs(y)(1xM)Fs(y)]+i[xGs(y)xGs(y)]/M\displaystyle=(x-x^{\prime})+i\left[(1-\tfrac{x}{M})F_{s}(y)-(1-\tfrac{x^{\prime}}{M})F_{s}(y^{\prime})\right]+i\left[xG_{s}(y)-x^{\prime}G_{s}(y^{\prime})\right]/M
=(xx)+i[(1xM)(Fs(y)Fs(y))+xM(Gs(y)Gs(y))\displaystyle=(x-x^{\prime})+i[(1-\tfrac{x}{M})(F_{s}(y)-F_{s}(y^{\prime}))+\tfrac{x}{M}(G_{s}(y)-G_{s}(y^{\prime}))
+(Gs(y)Fs(y))(xx)/M].\displaystyle\quad+(G_{s}(y^{\prime})-F_{s}(y^{\prime}))(x-x^{\prime})/M].

Fix ϵ>0\epsilon>0. We know that there exists δ>0\delta>0 such that if |ts|<δ|t-s|<\delta, then

|Fs(y)Fs(y)|\displaystyle|F_{s}(y)-F_{s}(y^{\prime})| (1+ϵ)|Ft(y)Ft(y)|,\displaystyle\leq(1+\epsilon)|F_{t}(y)-F_{t}(y^{\prime})|,
|Gs(y)Gs(y)|\displaystyle|G_{s}(y)-G_{s}(y^{\prime})| (1+ϵ)|Gt(y)Gt(y)|\displaystyle\leq(1+\epsilon)|G_{t}(y)-G_{t}(y^{\prime})|

and

|Fs(y)Ft(y)|ϵ,|Gs(y)Gt(y)|ϵ.\displaystyle|F_{s}(y)-F_{t}(y)|\leq\epsilon,\quad|G_{s}(y)-G_{t}(y)|\leq\epsilon.

Therefore,

Hs(z)\displaystyle H_{s}(z) Hs(w)\displaystyle-H_{s}(w)
=(xx)+i[(1xM)(Ft(y)Ft(y))(1+δ1)+xM(Gt(y)Gt(y))(1+δ2)\displaystyle=(x-x^{\prime})+i[(1-\tfrac{x}{M})(F_{t}(y)-F_{t}(y^{\prime}))(1+\delta_{1})+\tfrac{x}{M}(G_{t}(y)-G_{t}(y^{\prime}))(1+\delta_{2})
+(Gt(y)Ft(y)+δ3)(xx)/M]\displaystyle\quad+(G_{t}(y^{\prime})-F_{t}(y^{\prime})+\delta_{3})(x-x^{\prime})/M]
=a+i(c1+c2+c3a+δ1c1+δ2c2+δ3a)\displaystyle=a+i(c_{1}+c_{2}+c_{3}a+\delta_{1}c_{1}+\delta_{2}c_{2}+\delta_{3}a)
=a+i(c1+c2+c3a)+(δ1c1+δ2c2+δ3a)i\displaystyle=a+i(c_{1}+c_{2}+c_{3}a)+(\delta_{1}c_{1}+\delta_{2}c_{2}+\delta_{3}a)i
=Ht(z)Ht(w)+(δ1c1+δ2c2+δ3a)i\displaystyle=H_{t}(z)-H_{t}(w)+(\delta_{1}c_{1}+\delta_{2}c_{2}+\delta_{3}a)i

where

a\displaystyle a =xx\displaystyle=x-x^{\prime}
c1\displaystyle c_{1} =(1xM)(Ft(y)Ft(y))\displaystyle=(1-\tfrac{x}{M})(F_{t}(y)-F_{t}(y^{\prime}))
c2\displaystyle c_{2} =x(Gt(y)Gt(y))/M\displaystyle=x(G_{t}(y)-G_{t}(y^{\prime}))/M
c3\displaystyle c_{3} =(Gt(y)Ft(y))/M\displaystyle=(G_{t}(y^{\prime})-F_{t}(y^{\prime}))/M

and δ1,δ2,δ3\delta_{1},\delta_{2},\delta_{3} are functions of x,y,x,y,s,tx,y,x^{\prime},y^{\prime},s,t satisfying

|δ1|<ϵ,|δ2|<ϵ,|δ3|<2ϵ.\displaystyle|\delta_{1}|<\epsilon,\quad|\delta_{2}|<\epsilon,\quad|\delta_{3}|<2\epsilon.

Now, it follows from Lemma 4.13 that

|Hs(z)Hs(w)|(1+8ϵ)|Ht(z)Ht(w)||H_{s}(z)-H_{s}(w)|\leq(1+8\epsilon)|H_{t}(z)-H_{t}(w)|

and condition (ii) of Definition 4.1 is satisfied. ∎

We are now in a position to prove Proposition 4.12.

Proof of Proposition 4.12.

The idea is to lift via the exponential function and then use Lemma 4.14. To that end, define P~\widetilde{P} and Q~\widetilde{Q} via the functional equations Pexp=expP~P\circ\exp=\exp\circ\widetilde{P} and Qexp=expQ~Q\circ\exp=\exp\circ\widetilde{Q}.

Since the exponential function is conformal and has uniformly bounded derivative on the strip S={z:0Re(z)lnT}S=\{z:0\leq\operatorname{Re}(z)\leq\ln T\}, we conclude via the same argument as in Proposition 4.11 that P~\widetilde{P} and Q~\widetilde{Q} are bi-Lipschitz paths in the lines {z:Re(z)=0}\{z:\operatorname{Re}(z)=0\} and {z:Re(z)=lnT}\{z:\operatorname{Re}(z)=\ln T\} respectively.

Applying Lemma 4.14 to the strip S={z:0Re(z)lnT}S=\{z:0\leq\operatorname{Re}(z)\leq\ln T\} with boundary bi-Lipschitz path P~\widetilde{P} and Q~\widetilde{Q}, we obtain a bi-Lipschitz path F~\widetilde{F} which extends the boundary bi-Lipschitz paths.

Since Ft~\widetilde{F_{t}} is 2π2\pi-periodic by construction, we obtain the required bi-Lipschitz path FF via Fexp=expF~F\circ\exp=\exp\circ\widetilde{F}, again using the fact that the exponential function has uniformly bounded derivative in SS. ∎

5. Bi-Lipschitz collapsing for sets of small Assouad dimension

The goal in this section is to show that for a Cantor set X2X\subset\mathbb{R}^{2} with dimAX<1\dim_{A}X<1, we can cover it by small topological disks that can then by collapsed via a bi-Lipschitz path into a small disk. This is the content of Proposition 5.1 below, see Figure 5 for a schematic.

Refer to caption
Figure 5. The larger domain is Ω\Omega, the shaded ball is BB, the PL curves give the boundaries of the components of 𝒯δ(X)\mathcal{T}_{\delta}(X) and the arrows indicate that the bi-Lipschitz path HtH_{t} constructed in Proposition 5.1 moves these components into BB in an isometric way.
Proposition 5.1.

Let C>0C>0, c1c\geq 1, s[0,1)s\in[0,1), η(0,1)\eta\in(0,1) and let Ω2\Omega\subset\mathbb{R}^{2} be a domain with diamΩ=1\operatorname{diam}\Omega=1 such that for any x,yΩx,y\in\Omega there exists a path γx,y:[0,1]Ω\gamma_{x,y}:[0,1]\to\Omega such that γx,y(0)=x\gamma_{x,y}(0)=x, γx,y(1)=y\gamma_{x,y}(1)=y and

(5.1) dist(γx,y,U)(2c)1min{dist(x,Ω),dist(y,Ω)}.\operatorname{dist}(\gamma_{x,y},\partial U)\geq(2c)^{-1}\min\{\operatorname{dist}(x,\partial\Omega),\operatorname{dist}(y,\partial\Omega)\}.

Let XΩX\subset\Omega be (C,s)(C,s)-homogeneous with dist(X,Ω)>η\operatorname{dist}(X,\partial\Omega)>\eta. There exists ϵ>0\epsilon>0 so that if z,wΩz,w\in\Omega have distance at least 2ϵ2\epsilon from Ω\partial\Omega, then the disk B(z,ϵ)B(z,\epsilon) can be deformed continuously and isometrically to B(w,ϵ)B(w,\epsilon) in Ω\Omega. There exists δ>0\delta>0 so that if B=B(z,ϵ)ΩB=B(z,\epsilon)\subset\Omega is a disk of radius ϵ\epsilon with center zz satisfying dist(z,Ω)2ϵ\operatorname{dist}(z,\partial\Omega)\geq 2\epsilon, there exists a bi-Lipschitz path H:[0,1]LIP(Ω)H:[0,1]\to LIP(\Omega) such that

  1. (i)

    H1H_{1} maps the closed neighborhood 𝒯δ(X)\mathcal{T}_{\delta}(X) of XX into BB;

  2. (ii)

    for each t[0,1]t\in[0,1] and each component DD of 𝒯δ(X)\mathcal{T}_{\delta}(X), the map Ht|DH_{t}|D is an isometry.

A couple of remarks are in order.

Remark 5.2.

First, condition (5.1) on Ω\Omega is inspired by, but slightly weaker than, the well-known cc-John property. Second, for the rest of this section, we call curves γx,y\gamma_{x,y} cc-cigar curves. Finally, if c>cc^{\prime}>c, then there exists a piecewise linear (abbv. PL) cc^{\prime}-cigar curve σ\sigma joining xx with yy in Ω\Omega. In light of this observation, we will assume from now on that all cigar curves are PL.

5.1. Convex sets

Given a set ENE\subset\mathbb{R}^{N}, we denote by Hull(E)\operatorname{Hull}(E) the closed convex hull of EE, that is, the intersection of all closed convex sets that contain EE. Such a set is itself convex and diam(Hull(E))=diam(E)\operatorname{diam}(\operatorname{Hull}(E))=\operatorname{diam}(E).

Lemma 5.3.

Let ENE\subset\mathbb{R}^{N} be a bounded set. If x,yHull(E)x,y\in\operatorname{Hull}(E) and |xy|=diam(Hull(E))|x-y|=\operatorname{diam}(\operatorname{Hull}(E)), then x,yE¯x,y\in\overline{E}.

Proof.

For a contradiction, assume that xx is not in E¯\overline{E}. That is, r:=dist(x,E¯)>0r:=\operatorname{dist}(x,\overline{E})>0. Let PNP\subset\mathbb{R}^{N} be the (N1)(N-1)-plane that contains xx and is orthogonal to the line segment [x,y][x,y]. Then, since |xy|=diam(Hull(E))|x-y|=\operatorname{diam}(\operatorname{Hull}(E)), it follows that Hull(E)\operatorname{Hull}(E) lies on H¯\overline{H} where HH is one of the two components of NP\mathbb{R}^{N}\setminus P. Therefore,

E(HB(y,|xy|))B(x,r).E\subset(H\cap B(y,|x-y|))\setminus B(x,r).

Then, setting δ=dist(B(x,r)B(y,|xy|),P)\delta=\operatorname{dist}(\partial B(x,r)\cap\partial B(y,|x-y|),P) we have that the set

{zH:dist(z,P)δ}Hull(E)\{z\in H:\operatorname{dist}(z,P)\geq\delta\}\cap\text{Hull}(E)

is a convex set which contains EE and is a proper subset of Hull(E)\operatorname{Hull}(E), which is a contradiction. ∎

Lemma 5.4.

Let E1,,EnE_{1},\dots,E_{n} be sets in N\mathbb{R}^{N}. There exists l{1,,n}l\in\{1,\dots,n\} and there exist mutually disjoint convex closed sets Δ1,,Δl\Delta_{1},\dots,\Delta_{l} in N\mathbb{R}^{N} such that each EiE_{i} is contained in some Δj\Delta_{j} and

j=1ldiamΔji=1ndiamEi.\sum_{j=1}^{l}\operatorname{diam}{\Delta_{j}}\leq\sum_{i=1}^{n}\operatorname{diam}{E_{i}}.
Proof.

If one of the sets EiE_{i} is unbounded, then set l=1l=1, Δ1=N\Delta_{1}=\mathbb{R}^{N} and the claim is trivial.

Assume now that all sets EiE_{i} are bounded. In this case, the construction of the convex sets Δj\Delta_{j} is in an inductive fashion.

Step 1. For each i{1,,n}i\in\{1,\dots,n\}, let Δi(1)=Hull(Ei)\Delta_{i}^{(1)}=\operatorname{Hull}(E_{i}). If the sets Δi(1)\Delta_{i}^{(1)} are mutually disjoint, then set Δi=Δi(1)\Delta_{i}=\Delta_{i}^{(1)} and the procedure terminates; if some intersect, proceed to the next step.

Inductive Step. Suppose that for some k{1,,n1}k\in\{1,\dots,n-1\} we have defined closed convex sets Δ1(k),,Δnk+1(k)\Delta_{1}^{(k)},\dots,\Delta_{n-k+1}^{(k)} such that at least two of them intersect. In particular, let 1i0<j0n11\leq i_{0}<j_{0}\leq n-1 be such that Δi0(k)Δj0(k)\Delta_{i_{0}}^{(k)}\cap\Delta_{j_{0}}^{(k)}\neq\emptyset. We now define Δi(k+1)\Delta^{(k+1)}_{i} for i{1,,nk}i\in\{1,\dots,n-k\} as follows:

  • if i<i0i<i_{0} or if i0<i<j0i_{0}<i<j_{0}, then set Δi(k+1)=Δi(k)\Delta^{(k+1)}_{i}=\Delta^{(k)}_{i};

  • if i=i0i=i_{0}, then set Δi0(k+1)=Hull(Δi0(k)Δj0(k))\Delta^{(k+1)}_{i_{0}}=\operatorname{Hull}(\Delta^{(k)}_{i_{0}}\cup\Delta^{(k)}_{j_{0}});

  • if j0inkj_{0}\leq i\leq n-k, then set Δi(k+1)=Δi+1(k)\Delta^{(k+1)}_{i}=\Delta^{(k)}_{i+1}.

Note that

diamΔi0(k+1)=diam(Δi0(k)Δj0(k))diamΔi0(k)+diamΔj0(k).\operatorname{diam}{\Delta^{(k+1)}_{i_{0}}}=\operatorname{diam}(\Delta^{(k)}_{i_{0}}\cup\Delta^{(k)}_{j_{0}})\leq\operatorname{diam}{\Delta^{(k)}_{i_{0}}}+\operatorname{diam}{\Delta^{(k)}_{j_{0}}}.

If the sets Δi(k+1)\Delta_{i}^{(k+1)} are mutually disjoint, then set Δi=Δi(k+1)\Delta_{i}=\Delta_{i}^{(k+1)} and the procedure terminates; if some intersect, proceed to the next step.

It is clear that the procedure above will terminate in mm steps for some m{1,,n}m\in\{1,\dots,n\}. The sets Δ1,,Δnm+1\Delta_{1},\dots,\Delta_{n-m+1} produced are convex, mutually disjoint, and each EiE_{i} is contained in some Δj\Delta_{j}. It remains to show that

(5.2) i=1nm+1diamΔii=1ndiamEi.\sum_{i=1}^{n-m+1}\operatorname{diam}{\Delta_{i}}\leq\sum_{i=1}^{n}\operatorname{diam}{E_{i}}.

To prove (5.2), first note that for all i{1,,n}i\in\{1,\dots,n\}, diamEi=diamΔi(1)\operatorname{diam}{E_{i}}=\operatorname{diam}{\Delta_{i}^{(1)}}. Therefore, if m=1m=1, then (5.2) follows.

Suppose now that m2m\geq 2. Fix k{1,,m1}k\in\{1,\dots,m-1\} and let i0,j0{1,,nk+1}i_{0},j_{0}\in\{1,\dots,n-k+1\} be as in the construction of domains Δi(k+1)\Delta_{i}^{(k+1)}. Then,

i=1nk+1diamΔi(k)\displaystyle\sum_{i=1}^{n-k+1}\operatorname{diam}{\Delta_{i}^{(k)}} =i{1,,nk+1}{i0,j0}diamΔi(k)+diamΔi0(k)+diamΔj0(k)\displaystyle=\sum_{i\in\{1,\dots,n-k+1\}\setminus\{i_{0},j_{0}\}}\operatorname{diam}{\Delta_{i}^{(k)}}+\operatorname{diam}{\Delta_{i_{0}}^{(k)}}+\operatorname{diam}{\Delta_{j_{0}}^{(k)}}
i{1,,nk}{i0}diamΔi(k+1)+diamΔi0(k+1)\displaystyle\geq\sum_{i\in\{1,\dots,n-k\}\setminus\{i_{0}\}}\operatorname{diam}{\Delta_{i}^{(k+1)}}+\operatorname{diam}{\Delta_{i_{0}}^{(k+1)}}
=i=1nkdiamΔi(k+1).\displaystyle=\sum_{i=1}^{n-k}\operatorname{diam}{\Delta_{i}^{(k+1)}}.

Now by induction, (5.2) follows. ∎

Lemma 5.5.

Let Ω2\Omega\subset\mathbb{R}^{2} be a domain with non-empty boundary, and let ΔΩ\Delta\subset\Omega be a compact convex set with PL boundary. Let δ(0,1)\delta\in(0,1), let

0<r<(1δ)dist(Δ,Ω),0<r<(1-\delta)\operatorname{dist}(\Delta,\partial\Omega),

and let γ:[0,1]Ω\gamma:[0,1]\to\Omega be a PL curve in Ω\Omega with γ(0)Δ\gamma(0)\in\Delta and |γ(t)γ(0)|<r|\gamma(t)-\gamma(0)|<r for all t[0,1]t\in[0,1]. Then there exists a bi-Lipschitz path H:[0,1]LIP(Ω)H:[0,1]\to LIP(\Omega) such that

  1. (i)

    for each t[0,1]t\in[0,1], Ht|ΩH_{t}|\partial\Omega is the identity;

  2. (ii)

    for each t[0,1]t\in[0,1], Ht|ΔH_{t}|\Delta is a translation mapping with Ht(γ(0))=γ(t)H_{t}(\gamma(0))=\gamma(t).

Proof.

Without loss of generality, we may assume that γ\gamma is a straight line segment; in the general case of PL curves γ\gamma, concatenate the bi-Lipschitz paths from the various segments of γ\gamma and re-parameterize if necessary. Assume then, that γ:[0,1]Ω\gamma:[0,1]\to\Omega with γ(t)=γ(0)+tv\gamma(t)=\gamma(0)+tv for some vv\in\mathbb{C} with |v|<r|v|<r.

By the hypotheses, Δ\Delta is a convex polygon with vertices v1,,vnv_{1},\ldots,v_{n}. Fix z0Δz_{0}\in\Delta and for i{1,,n}i\in\{1,\ldots,n\} let wiw_{i} be the point on the ray from z0z_{0} through viv_{i} that is distance (1δ/2)dist(Δ,Ω)(1-\delta/2)\operatorname{dist}(\Delta,\partial\Omega) away from viv_{i} (and outside Δ\Delta). Let YY be the convex hull of w1,,wnw_{1},\ldots,w_{n} and set d=dist(Δ,Y)>0d=\operatorname{dist}(\Delta,\partial Y)>0.

Triangulate the PL ring domain YΔ¯\overline{Y\setminus\Delta} via triangles T1,,TmT_{1},\ldots,T_{m} which have, alternately, one or two vertices contained in Y\partial Y.

Given a direction eiθe^{i\theta}, we will construct a bi-Lipschitz path which moves Δ\Delta onto Δ1={z:z=z+deiθ/2,zΔ}\Delta_{1}=\{z:z=z^{\prime}+de^{i\theta}/2,z^{\prime}\in\Delta\}. For zΔz\in\Delta we just define

Ht(z)=(z+deiθ/2)t+(1t)z.H_{t}(z)=(z+de^{i\theta}/2)t+(1-t)z.

If TiT_{i} has two vertices on Y\partial Y and third vertex ξ1Δ\xi_{1}\in\partial\Delta, then we apply the bi-Lipschitz path from Proposition 4.9 (conjugated by a suitable similarity) which fixes the two vertices in Y\partial Y and moves ξ1\xi_{1} to ξ1+deiθ/2Y\xi_{1}+de^{i\theta}/2\in Y.

If TiT_{i} has one vertex on Y\partial Y and two vertices ξ1,ξ2\xi_{1},\xi_{2} in X\partial X, then we apply the bi-Lipschitz path from Proposition 4.9 (again conjugated by a suitable similarity) which fixes the vertex in Y\partial Y and moves ξj\xi_{j} to ξj+deiθ/2\xi_{j}+de^{i\theta}/2 for j=1,2j=1,2.

This piecewise construction yields a bi-Lipschitz path which moves Δ\Delta to Δ1\Delta_{1} and fixes every point of Y\partial Y and hence can be extended to fix every point of ΩY\Omega\setminus Y. By concatenating a finite number of bi-Lipschitz paths, we may move XX along any PL path in Ω\Omega, as long as we avoid Ω\partial\Omega, such that the path acts as a translation on XX. ∎

For the rest of the paper, given a bounded set X2X\subset\mathbb{R}^{2}, a number r>0r>0 and a curve γ:[0,a]2\gamma:[0,a]\to\mathbb{R}^{2} with γ(0)X\gamma(0)\in X, we denote

𝒩(X,γ,r):=t[0,a](γ(t)γ(0)+N(X,r)).\mathcal{N}(X,\gamma,r):=\bigcup_{t\in[0,a]}\left(\gamma(t)-\gamma(0)+N(X,r)\right).

5.2. Proof of Proposition 5.1

The first claim about the existence of such an ϵ\epsilon follows by following a cc-cigar curve from zz to ww. Henceforth, fix B=B(z0,ϵ)B=B(z_{0},\epsilon).

Suppose first that diamX=0\operatorname{diam}{X}=0, that is X={x0}X=\{x_{0}\} for some x0Ωx_{0}\in\Omega. Let γ\gamma be a PL cc-cigar path that joins x0x_{0} with x0x_{0} in Ω\Omega. Let Δ\Delta be a compact convex set with PL boundary contained in B(x0,r)B(x_{0},r) with r<min{ϵ,13dist(x0,Ω)}r<\min\{\epsilon,\tfrac{1}{3}\operatorname{dist}(x_{0},\partial\Omega)\}. We then apply Lemma 5.5 to find the required bi-Lipschitz path H:[0,1]LIP(Ω)H:[0,1]\to LIP(\Omega) such that for any t[0,1]t\in[0,1], Ht(x0)=γ(t)H_{t}(x_{0})=\gamma(t).

Suppose now and for the rest of the proof of Proposition 5.1 that diamX>0\operatorname{diam}{X}>0. Set

(5.3) δ=(min{η,ϵ}216cC)11s.\delta=\left(\frac{\min\{\eta,\epsilon\}}{216cC}\right)^{\frac{1}{1-s}}.

We may assume that C>1C>1, hence δ\delta is less than 11. Then let VV be a δ\delta-net of XX and let D1,,DnD_{1},\dots,D_{n} be the components of 𝒯δ(X)\mathcal{T}_{\delta}(X).

Since δ<η/20\delta<\eta/20, we have that

dist(𝒯δ(X),Ω)dist(X,Ω)distH(𝒯δ(X),X)η8δη/2\operatorname{dist}(\mathcal{T}_{\delta}(X),\partial\Omega)\geq\operatorname{dist}(X,\partial\Omega)-\operatorname{dist}_{H}(\mathcal{T}_{\delta}(X),X)\geq\eta-8\delta\geq\eta/2

where distH\operatorname{dist}_{H} denotes the Hausdorff distance.

Let i{1,,n}i\in\{1,\dots,n\}. For each xDix\in\partial D_{i} there exists zXz\in X such that |xz|8δ|x-z|\leq 8\delta and there exists vVv\in V such that |zv|δ|z-v|\leq\delta. Therefore, for every xDix\in\partial D_{i}, dist(x,V)9δ\operatorname{dist}(x,V)\leq 9\delta and it follows that

(5.4) diamDi18δcard(VDi).\operatorname{diam}{D_{i}}\leq 18\delta\operatorname{card}(V\cap D_{i}).

Therefore,

(5.5) i=1ndiamDi18δcard(V)18Cδ1s=(12c)1min{η,ϵ}.\sum_{i=1}^{n}\operatorname{diam}{D_{i}}\leq 18\delta\operatorname{card}(V)\leq 18C\delta^{1-s}=(12c)^{-1}\min\{\eta,\epsilon\}.

The construction of the bi-Lipschitz path HH consists of two parts. In the first part we construct at most n1n-1 many bi-Lipschitz paths that "gather the sets DiD_{i} together" and in the second part we construct a bi-Lipschitz path that leads the cluster of gathered sets DiD_{i} into the disk BB.

5.2.1. Part 1

The construction in this part is in an inductive manner.

Step 0. Apply Lemma 5.4 for the sets D1,,DnD_{1},\dots,D_{n} and obtain closed mutually disjoint convex sets Δ1(1),,Δk0(0)\Delta_{1}^{(1)},\dots,\Delta_{k_{0}}^{(0)} for some positive integer k0{1,,n}k_{0}\in\{1,\dots,n\}. Note that

i=1k0diamΔi(0)i=1ndiamDi(12c)1min{η,ϵ}.\displaystyle\sum_{i=1}^{k_{0}}\operatorname{diam}{\Delta^{(0)}_{i}}\leq\sum_{i=1}^{n}\operatorname{diam}{D_{i}}\leq(12c)^{-1}\min\{\eta,\epsilon\}.

Moreover, the sets Δ1(0),,Δk0(0)\Delta_{1}^{(0)},\dots,\Delta_{k_{0}}^{(0)} are contained in Ω\Omega and for each i{1,,k0}i\in\{1,\dots,k_{0}\}

dist(Δi(0),Ω)dist(𝒯δ(X),Ω)diamΔi(0)η/2(12c)1η>13η.\displaystyle\operatorname{dist}(\Delta_{i}^{(0)},\partial\Omega)\geq\operatorname{dist}(\mathcal{T}_{\delta}(X),\partial\Omega)-\operatorname{diam}{\Delta_{i}^{(0)}}\geq\eta/2-(12c)^{-1}\eta>\tfrac{1}{3}\eta.

If k0=1k_{0}=1, then the procedure terminates and we proceed to Part 2; otherwise proceed to the next step.

Inductive step. Suppose that for some positive integer m{0,,n2}m\in\{0,\dots,n-2\} we have defined disjoint closed convex sets Δ1(m1),,Δkm1(m1)Ω\Delta^{(m-1)}_{1},\dots,\Delta^{(m-1)}_{k_{m-1}}\subset\Omega such that 2km1nm+12\leq k_{m-1}\leq n-m+1 and the following three properties hold.

  1. (P1)

    For each i{1,,km1}i\in\{1,\dots,k_{m-1}\} there exists j{1,,n}j\in\{1,\dots,n\} with DjΔi(m1)D_{j}\subset\Delta^{(m-1)}_{i};

  2. (P2)

    We have

    i=1km1diamΔi(m1)<(6c)1min{η,ϵ};\sum_{i=1}^{k_{m-1}}\operatorname{diam}{\Delta^{(m-1)}_{i}}<(6c)^{-1}\min\{\eta,\epsilon\};
  3. (P3)

    For each i{1,,km1}i\in\{1,\dots,k_{m-1}\}, dist(Δi(m1),Ω)>η/3\operatorname{dist}(\Delta_{i}^{(m-1)},\partial\Omega)>\eta/3.

Let γm:[0,1]Ω\gamma_{m}:[0,1]\to\Omega be a PL cc-cigar curve with γm(0)XΔ1(m1)\gamma_{m}(0)\in X\cap\Delta_{1}^{(m-1)} and γm(1)XΔ2(m1)\gamma_{m}(1)\in X\cap\Delta^{(m-1)}_{2}. By (5.1) and inductive assumption (P3), we have that for all t[0,1]t\in[0,1],

(5.6) dist(γm(t),Ω)(2c)1min{dist(γm(0),Ω),dist(γm(1),Ω)}(2c)1η.\operatorname{dist}(\gamma_{m}(t),\partial\Omega)\geq(2c)^{-1}\min\{\operatorname{dist}(\gamma_{m}(0),\partial\Omega),\operatorname{dist}(\gamma_{m}(1),\partial\Omega)\}\geq(2c)^{-1}\eta.

Using inductive assumption (P2), we can find a number rm>0r_{m}>0 such that

  1. (i)

    rm<13dist(Δ1(m1),Δi(m1))r_{m}<\tfrac{1}{3}\operatorname{dist}(\Delta^{(m-1)}_{1},\Delta^{(m-1)}_{i}) for all i{2,,km1}i\in\{2,\dots,k_{m-1}\}

  2. (ii)

    rm<(6c)1min{η,ϵ}i=1km1diamΔi(m1)r_{m}<(6c)^{-1}\min\{\eta,\epsilon\}-\sum_{i=1}^{k_{m-1}}\operatorname{diam}{\Delta^{(m-1)}_{i}}.

The second property of rmr_{m} implies that

rm<(2c)1dist(Δ1(m1),Ω)diamΔ1(m1)r_{m}<(2c)^{-1}\operatorname{dist}(\Delta^{(m-1)}_{1},\partial\Omega)-\operatorname{diam}{\Delta^{(m-1)}_{1}}

which, along with (5.6), implies that 𝒩(Δ1(m1),γm,2rm)Ω\mathcal{N}(\Delta_{1}^{(m-1)},\gamma_{m},2r_{m})\subset\Omega. Let

Tm=sup{t[0,1]:𝒩(Δ1(m1),γm|[0,t],rm)j=2km1Δj(m1)=}.T_{m}=\sup\left\{t\in[0,1]:\mathcal{N}(\Delta_{1}^{(m-1)},\gamma_{m}|_{[0,t]},r_{m})\cap\bigcup_{j=2}^{k_{m}-1}\Delta_{j}^{(m-1)}=\emptyset\right\}.

Since dist(Δ1(m1),Δi(m1))3rm\operatorname{dist}(\Delta_{1}^{(m-1)},\Delta_{i}^{(m-1)})\geq 3r_{m} for all i1i\neq 1, we have that Tm>0T_{m}>0. Let i0{2,,km1}i_{0}\in\{2,\dots,k_{m-1}\} be such that

𝒩(Δ1(m1),γm|[0,Tm],rm)Δi0(m1).\mathcal{N}(\Delta_{1}^{(m-1)},\gamma_{m}|_{[0,T_{m}]},r_{m})\cap\Delta_{i_{0}}^{(m-1)}\neq\emptyset.

For simplicity, we may assume that i0=2i_{0}=2. Denote by H(m)H^{(m)} the bi-Lipschitz path given from Lemma 5.5 for the curve γ=γm|[0,Tm]\gamma=\gamma_{m}|_{[0,T_{m}]}. Consider now the disjoint closed sets

E1=H1(m)(Δ1(m1))Δ2(m1),E2=Δ3(m1),,Ekm1=Δkm1(m1)E_{1}=H_{1}^{(m)}(\Delta^{(m-1)}_{1})\cup\Delta^{(m-1)}_{2},\quad E_{2}=\Delta^{(m-1)}_{3},\quad\dots,\quad E_{k_{m}-1}=\Delta^{(m-1)}_{k_{m-1}}

and apply Lemma 5.4 to the sets EjE_{j} to obtain mutually disjoint closed convex sets Δ1(m),,Δkm(m)\Delta^{(m)}_{1},\dots,\Delta^{(m)}_{k_{m}} with kmkm11k_{m}\leq k_{m-1}-1. We note that

  1. (i)

    for each i{1,,km}i\in\{1,\dots,k_{m}\} there exists j{1,,n}j\in\{1,\dots,n\} with DjΔi(m)D_{j}\subset\Delta^{(m)}_{i};

  2. (ii)
    i=1kmdiamΔi(m)i=1km1diamΔi(m1)+rm<(6c)1min{η,ϵ}.\sum_{i=1}^{k_{m}}\operatorname{diam}{\Delta^{(m)}_{i}}\leq\sum_{i=1}^{k_{m-1}}\operatorname{diam}{\Delta^{(m-1)}_{i}}+r_{m}<(6c)^{-1}\min\{\eta,\epsilon\}.

It follows that Δ1(m),,Δkm(m)\Delta_{1}^{(m)},\dots,\Delta_{k_{m}}^{(m)} are contained in Ω\Omega and in fact, for each i{1,,km}i\in\{1,\dots,k_{m}\}

dist(Δi(2),Ω)dist(𝒯δ(X),Ω)diamΔi(m)\displaystyle\operatorname{dist}(\Delta_{i}^{(2)},\partial\Omega)\geq\operatorname{dist}(\mathcal{T}_{\delta}(X),\partial\Omega)-\operatorname{diam}{\Delta_{i}^{(m)}} η/2(6c)1>η/3.\displaystyle\geq\eta/2-(6c)^{-1}>\eta/3.

Therefore, we have verified that inductive assumptions (P1)–(P3) hold for mm. If km=1k_{m}=1 the procedure terminates and we proceed to Part 2; otherwise proceed to the next step.

After pp steps, for some p{0,,n1}p\in\{0,\dots,n-1\}, we have kp=1k_{p}=1. By the choice of δ\delta and numbers r1,,rpr_{1},\dots,r_{p}, the final convex set Δ1(p)\Delta^{(p)}_{1} satisfies properties (P1)–(P3); precisely, we have

  1. (i)

    diamΔ1(p)<(6c)1min{η,ϵ}\operatorname{diam}{\Delta^{(p)}_{1}}<(6c)^{-1}\min\{\eta,\epsilon\};

  2. (ii)

    there exists i{1,,n}i\in\{1,\dots,n\} such that DiΔ1(p)D_{i}\subset\Delta^{(p)}_{1},

  3. (iii)

    Δ1(p)Ω\Delta^{(p)}_{1}\subset\Omega and dist(Δ1(p),Ω)>η/3\operatorname{dist}(\Delta^{(p)}_{1},\partial\Omega)>\eta/3.

5.2.2. Part 2

Let z0Ωz_{0}\in\Omega be the center of BB and let γp+1:[0,1]Ω\gamma_{p+1}:[0,1]\to\Omega be a PL cc-cigar curve in Ω\Omega with γp+1(0)XΔ1(p)\gamma_{p+1}(0)\in X\cap\Delta^{(p)}_{1} and γp+1(1)=z0\gamma_{p+1}(1)=z_{0}. If z0XΔ1(p)z_{0}\in X\cap\Delta^{(p)}_{1}, then we can choose γp+1\gamma_{p+1} to be constant. By (5.1), we have that for all t[0,1]t\in[0,1],

(5.7) dist(γp+1(t),Ω)\displaystyle\operatorname{dist}(\gamma_{p+1}(t),\partial\Omega) (2c)1min{dist(γp+1(0),Ω),dist(γp+1(1),Ω)}\displaystyle\geq(2c)^{-1}\min\{\operatorname{dist}(\gamma_{p+1}(0),\partial\Omega),\operatorname{dist}(\gamma_{p+1}(1),\partial\Omega)\}
(2c)1min{ϵ,η}.\displaystyle\geq(2c)^{-1}\min\{\epsilon,\eta\}.

Let rm+1r_{m+1} be a positive number with rm+1<(6c)1min{η,ϵ}r_{m+1}<(6c)^{-1}\min\{\eta,\epsilon\}. Then (5.7) implies that

𝒩(Δ1(p),γp+1,rp+1)Ω.\mathcal{N}(\Delta_{1}^{(p)},\gamma_{p+1},r_{p+1})\subset\Omega.

Let now H(p+1)H^{(p+1)} be the bi-Lipschitz path given from Lemma 5.5 for γ=γp+1\gamma=\gamma_{p+1}. If p=0p=0, then we define H:[0,1]LIP(N)H:[0,1]\to LIP(\mathbb{R}^{N}) with H=H(p+1)H=H^{(p+1)}. If p1p\geq 1, we concatenate the bi-Lipschitz paths H(1),,H(p+1)H^{(1)},\dots,H^{(p+1)} and we obtain the desired bi-Lipschitz path HH.

6. A multitwist bi-Lipschitz map

In §6.1 we prove Proposition 1.3 while in §6.2 we show that the multitwist map in Theorem 7.1 is bi-Lipschitz.

6.1. Proof of Proposition 1.3

In this subsection we prove Proposition 1.3. To that end, we require the following “egg-yolk principle" lemma which is a simple application of Koebe’s Distortion Theorem.

Lemma 6.1.

Given δ>0\delta>0, there exists L0>1L_{0}>1 with the following property. If UU is a domain in 2\mathbb{R}^{2}, KUK\subset U is a compact connected set with dist(K,U)δdiamK\operatorname{dist}(K,\partial U)\geq\delta\operatorname{diam}{K}, x0Kx_{0}\in K is a point, and f:U2f:U\to\mathbb{R}^{2} is an injective conformal map, then for all x,yKx,y\in K,

L01|f(x0)||xy||f(x)f(y)|L0|f(x0)||xy|.L_{0}^{-1}|f^{\prime}(x_{0})||x-y|\leq|f(x)-f(y)|\leq L_{0}|f^{\prime}(x_{0})||x-y|.
Proof.

If KK is a single point, the claim is trivial. Assume for the rest that diamK=d>0\operatorname{diam}{K}=d>0. Let VV be a maximal (δd/4)(\delta d/4)-separated subset of KK containing x0x_{0}. By the doubling property of 2\mathbb{R}^{2}, there exists NN\in\mathbb{N} depending only on δ\delta such that cardVN\operatorname{card}{V}\leq N.

By the Koebe Distortion Theorem (see for example [GM05, Theorem I.4.5] and [Pom92, Theorem 1.3]), there exists a universal A>1A>1 such that for any zKz\in K and for any w,w1,w2B(z,12δd)w,w_{1},w_{2}\in B(z,\frac{1}{2}\delta d) we have

(6.1) A1|f(w)||w1w2||f(w1)f(w2)|A|f(w)||w1z2|A^{-1}|f^{\prime}(w)||w_{1}-w_{2}|\leq|f(w_{1})-f(w_{2})|\leq A|f^{\prime}(w)||w_{1}-z_{2}|
(6.2) A1|f(z)||f(w)|A|f(z)|A^{-1}|f^{\prime}(z)|\leq|f^{\prime}(w)|\leq A|f^{\prime}(z)|
(6.3) dist(f(z),f(U))A1δd|f(z)|.\operatorname{dist}(f(z),\partial f(U))\geq A^{-1}\delta d|f^{\prime}(z)|.

By (6.2), we have that for all xKx\in K,

(6.4) AN|f(x0)||f(x)|AN|f(x0)|.A^{-N}|f^{\prime}(x_{0})|\leq|f^{\prime}(x)|\leq A^{N}|f^{\prime}(x_{0})|.

We show that f|Kf|K is (L1|f(x0)|)(L_{1}|f^{\prime}(x_{0})|)-Lipschitz for some L1>0L_{1}>0 depending only on δ\delta. Fix x,yKx,y\in K and consider two cases. If |xy|<δd/2|x-y|<\delta d/2, then by (6.1) and (6.4)

AN|f(x0)||xy||f(x)f(y)|AN|f(x0)||xy|.A^{-N}|f^{\prime}(x_{0})||x-y|\leq|f(x)-f(y)|\leq A^{N}|f^{\prime}(x_{0})||x-y|.

Suppose now that |xy|δd/2|x-y|\geq\delta d/2. Then, there exist z,zVz,z^{\prime}\in V such that xB(z,δd/4)x\in B(z,\delta d/4) and yB(z,δd/4)y\in B(z^{\prime},\delta d/4), and by connectedness of KK, there exist distinct z1,,zlVz_{1},\dots,z_{l}\in V such that z1=zz_{1}=z, zl=zz_{l}=z^{\prime}, and for all j{1,,l1}j\in\{1,\dots,l-1\}, |zjzj+1|<δd/2|z_{j}-z_{j+1}|<\delta d/2. Therefore,

|f(x)f(y)|\displaystyle|f(x)-f(y)| |f(x)f(z)|+i=1l1|f(zi+1)f(zi)|+|f(x)f(z)|\displaystyle\leq|f(x)-f(z)|+\sum_{i=1}^{l-1}|f(z_{i+1})-f(z_{i})|+|f(x^{\prime})-f(z^{\prime})|
|f(x0)|(N+1)AN(δd/2)\displaystyle\leq|f^{\prime}(x_{0})|(N+1)A^{N}(\delta d/2)
|f(x0)|(N+1)AN|xy|.\displaystyle\leq|f^{\prime}(x_{0})|(N+1)A^{N}|x-y|.

By (6.3) we have that dist(w,f(U))ANδd|f(x0)|\operatorname{dist}(w,\partial f(U))\geq A^{-N}\delta d|f^{\prime}(x_{0})| for all wf(K)w\in f(K). On the other hand, since f|Kf|K is L1L_{1}-Lipschitz, we have that diamf(K)L1|f(x0)|d\operatorname{diam}{f(K)}\leq L_{1}|f^{\prime}(x_{0})|d. Therefore,

dist(f(K),f(U))(L1AN)1δdiamf(K).\operatorname{dist}(f(K),\partial f(U))\geq(L_{1}A^{N})^{-1}\delta\operatorname{diam}{f(K)}.

Then, working as above, we can find L2>0L_{2}>0 depending only on L1L_{1} and NN (hence only on δ\delta) such that f1|f(K)f^{-1}|f(K) is (L2|(f1)(f(x0))|)(L_{2}|(f^{-1})^{\prime}(f(x_{0}))|)-Lipschitz. Therefore, for all x,yKx,y\in K

|xy|L2|f(x0)||f(x)f(y)||x-y|\leq\frac{L_{2}}{|f^{\prime}(x_{0})|}|f(x)-f(y)|

and the proof is complete. ∎

We can now prove Proposition 1.3.

Proof of Proposition 1.3.

Let X2X\subset\mathbb{R}^{2} be a cc-uniformly disconnected set. By Theorem 1.2 we know that there is a geodesic pants decomposition of the hyperbolic Riemann surface S:=𝕊2XS:=\mathbb{S}^{2}\setminus X so that the cuffs (αj)(\alpha_{j}) have uniformly bounded hyperbolic length. By Proposition 2.1, there exist mutually disjoint ring domains (Rj)(R_{j}^{\prime}) which are thickenings of (αj)(\alpha_{j}) with a uniform upper bound M0M_{0} on their moduli.

For each jj, denote by VjV_{j}^{\prime} and UjU_{j}^{\prime} the bounded and unbounded, respectively, components of 2Rj\mathbb{R}^{2}\setminus R_{j}^{\prime}. Let ζj\zeta_{j} be a similarity of 2\mathbb{R}^{2} such that diamζj1(Vj)=1\operatorname{diam}{\zeta_{j}^{-1}(V_{j}^{\prime})}=1 and 0ζj1(Vj)0\in\zeta_{j}^{-1}(V_{j}^{\prime}). By (2.1), there exists ϵ0\epsilon_{0} depending only on M0M_{0} (hence only on cc) such that dist(ζj1(Uj),ζj1(Vj))ϵ0\operatorname{dist}(\partial\zeta_{j}^{-1}(U_{j}^{\prime}),\partial\zeta_{j}^{-1}(V_{j}^{\prime}))\geq\epsilon_{0}. By Lemma 2.4, there exists a polygonal Jordan curve γj\gamma_{j} with edges in 𝒢ϵ0/161\mathscr{G}_{\epsilon_{0}/16}^{1} which encloses ζj1(Vj)\zeta_{j}^{-1}(V_{j}^{\prime}) and satisfies

ϵ0/16dist(x,ζj1(Vj))ϵ0/2,for all xγj.\epsilon_{0}/16\leq\operatorname{dist}(x,\zeta_{j}^{-1}(V_{j}^{\prime}))\leq\epsilon_{0}/2,\qquad\text{for all $x\in\gamma_{j}$}.

Applying Lemma 2.4, there exists a polygonal Jordan curve Γj\Gamma_{j} with edges in 𝒢ϵ0/321\mathscr{G}_{\epsilon_{0}/32}^{1} which encloses γj\gamma_{j} and satisfies

ϵ0/32dist(x,γj)ϵ0/4,for all xΓj.\epsilon_{0}/32\leq\operatorname{dist}(x,\gamma_{j})\leq\epsilon_{0}/4,\qquad\text{for all $x\in\Gamma_{j}$}.

The ring domain Rj′′R_{j}^{\prime\prime} bounded by γj\gamma_{j} and Γj\Gamma_{j} satisfies

  1. (i)

    dist(γj,Γj)ϵ0/32\operatorname{dist}(\gamma_{j},\Gamma_{j})\geq\epsilon_{0}/32,

  2. (ii)

    1diamRj′′1+32ϵ01\leq\operatorname{diam}{R_{j}^{\prime\prime}}\leq 1+\frac{3}{2}\epsilon_{0} and

  3. (iii)

    dist(x,ζj(Rj))ϵ0/16\operatorname{dist}(x,\partial\zeta_{j}(R_{j}^{\prime}))\geq\epsilon_{0}/16, for all xRj′′x\in R_{j}^{\prime\prime}.

It follows that Rj′′[132ϵ0,1+32ϵ0]2R_{j}^{\prime\prime}\subset[-1-\frac{3}{2}\epsilon_{0},1+\frac{3}{2}\epsilon_{0}]^{2} and since the boundary curves of Rj′′R_{j}^{\prime\prime} are made of edges in 𝒢ϵ0/321\mathscr{G}_{\epsilon_{0}/32}^{1}, there are at most kk many different domains Rj′′R_{j}^{\prime\prime}, with kk depending only on ϵ0\epsilon_{0}, hence only on cc.

There exists δ0(0,1)\delta_{0}\in(0,1) depending only on ϵ0\epsilon_{0} (hence only on cc) and for each jj there exists δj(0,1δ0)\delta_{j}\in(0,1-\delta_{0}), and there exists a conformal map

ψj:B(0,1)B¯(0,δj)Rj′′.\psi_{j}:B(0,1)\setminus\overline{B}(0,\delta_{j})\to R_{j}^{\prime\prime}.

Setting

K:={134δ0|x|114δ0}U:=B(0,1)B¯(0,δj),K:=\{1-\tfrac{3}{4}\delta_{0}\leq|x|\leq 1-\tfrac{1}{4}\delta_{0}\}\subset U:=B(0,1)\setminus\overline{B}(0,\delta_{j}),

we have dist(K,U)δ0/4\operatorname{dist}(K,\partial U)\geq\delta_{0}/4 and diamK=2δ0/2\operatorname{diam}{K}=2-\delta_{0}/2. Hence by Lemma 6.1, we have that |ψj(134δ0)|1ψj|\psi_{j}^{\prime}(1-\tfrac{3}{4}\delta_{0})|^{-1}\psi_{j} restricted on KK is a L0L_{0}-bi-Lipschitz, where L0L_{0} depends only on δ0\delta_{0} (hence only on cc). Moreover,

1L0(2δ0/2)diamψj(K)L0diamK|ψj(134δ0)|L0diamψj(K)diamKL02(2+3ϵ0)2δ0/2.\frac{1}{L_{0}(2-\delta_{0}/2)}\leq\frac{\operatorname{diam}{\psi_{j}(K)}}{L_{0}\operatorname{diam}{K}}\leq|\psi_{j}^{\prime}(1-\tfrac{3}{4}\delta_{0})|\leq L_{0}\frac{\operatorname{diam}{\psi_{j}(K)}}{\operatorname{diam}{K}}\leq L_{0}\frac{\sqrt{2}(2+3\epsilon_{0})}{2-\delta_{0}/2}.

For each jj\in\mathbb{N}, let λj=diamψj(B(0,134δ0))[1,1+3ϵ0/2]\lambda_{j}=\operatorname{diam}{\psi_{j}(\partial B(0,1-\frac{3}{4}\delta_{0}))}\in[1,1+3\epsilon_{0}/2]. It follows that the map

(λj)1ψj|K(\lambda_{j})^{-1}\psi_{j}|K

is L1L_{1}-bi-Lipschitz for some L1L_{1} depending on L0,ϵ0,δ0L_{0},\epsilon_{0},\delta_{0}, hence only on cc.

To complete the proof set

L=max{4δ02δ0,4L14δ0},L=\max\left\{\frac{4-\delta_{0}}{2\delta_{0}},\frac{4L_{1}}{4-\delta_{0}}\right\},

define conformal maps

gj:B(0,1)¯B(0,11/L)2withgj(x)=(λj)1ψj|K((1δ0/4)x),g_{j}:\overline{B(0,1)}\setminus B(0,1-1/L)\to\mathbb{R}^{2}\qquad\text{with}\qquad g_{j}(x)=(\lambda_{j})^{-1}\psi_{j}|K((1-\delta_{0}/4)x),

and define similarities

ϕj:22withϕj(x)=(λj)1ζj1(x).\phi_{j}:\mathbb{R}^{2}\to\mathbb{R}^{2}\qquad\text{with}\qquad\phi_{j}(x)=(\lambda_{j})^{-1}\zeta_{j}^{-1}(x).

Since L4δ02δ0L\geq\frac{4-\delta_{0}}{2\delta_{0}}, we have that (1δ0/4)xK(1-\delta_{0}/4)x\in K for all xB(0,1)¯B(0,11/L)x\in\overline{B(0,1)}\setminus B(0,1-1/L). Moreover, since L4L14δ0L\geq\frac{4L_{1}}{4-\delta_{0}}, we have that gjg_{j} is LL-bi-Lipschitz. Since there are at most kk many domains Rj′′R_{j}^{\prime\prime}, there are at most kk many conformal maps gjg_{j}. ∎

Setting

fj=ϕjgi(j)andRj=fj(B(0,1)¯B(0,11/L))f_{j}=\phi_{j}\circ g_{i(j)}\qquad\text{and}\qquad R_{j}=f_{j}(\overline{B(0,1)}\setminus B(0,1-1/L))

where ϕj\phi_{j} and gi(j)g_{i(j)} are as in the statement of Proposition 1.3, and applying Lemma 6.1 to the ring

K=B(0,1ϵ0/8)¯B(0,17ϵ0/8),K^{\prime}=\overline{B(0,1-\epsilon_{0}/8)}\setminus B(0,1-7\epsilon_{0}/8),

we see that there exists ξ>0\xi>0 so that

(6.5) dist(Rj,Rj)dist(Kj,Rj)ξdiamRj\operatorname{dist}(\partial R_{j}^{\prime},R_{j})\geq\operatorname{dist}(\partial K_{j},R_{j})\geq\xi\operatorname{diam}{R_{j}}

for all jj.

6.2. A multitwist bi-Lipschitz map

For the rest of this section we fix a cc-uniformly disconnected Cantor set X2X\subset\mathbb{R}^{2}. By Proposition 1.3, we obtain kk\in\mathbb{N}, L>1L>1, a finite set {g1,,gk}\{g_{1},\dots,g_{k}\} of LL-bi-Lipschitz conformal maps defined on B¯(0,1)B(0,11L)\overline{B}(0,1)\setminus B(0,1-\tfrac{1}{L}), similarities (ϕj)j(\phi_{j})_{j\in\mathbb{N}} and ring domains RjR_{j} such that for each jj\in\mathbb{N} there exists i(j){1,,k}i(j)\in\{1,\dots,k\}

(6.6) Rj=fj(B¯(0,1)B(0,11L))withfj=ϕjgi(j).R_{j}=f_{j}(\overline{B}(0,1)\setminus B(0,1-\tfrac{1}{L}))\qquad\text{with}\qquad f_{j}=\phi_{j}\circ g_{i(j)}.

Let f:22f:\mathbb{R}^{2}\to\mathbb{R}^{2} be a map such that ff is the identity outside of the union of Rj¯\overline{R_{j}}, while for each jj\in\mathbb{N}, f|Rj=fj𝔇fj1f|R_{j}=f_{j}\circ\mathfrak{D}\circ f_{j}^{-1} with

𝔇(r,θ)=(r,θ+2πL(1r)).\mathfrak{D}(r,\theta)=(r,\theta+2\pi L(1-r)).
Lemma 6.2.

The map ff is L0L_{0}-bi-Lipschitz with L0L_{0} depending only on cc.

Proof.

It is fairly elementary to see that 𝔇\mathfrak{D} is L1L_{1}-bi-Lipschitz for some L1>1L_{1}>1 depending only on LL (hence only on cc). It follows that for each jj\in\mathbb{N}, f|Rjf|R_{j} is L2L1L^{2}L_{1}-bi-Lipschitz. Since ff is the identity outside of the union of Rj¯\overline{R_{j}} (and hence bi-Lipschitz), we get that ff is an L2L_{2}-bounded length distortion map for some L2>1L_{2}>1 depending only on LL. That is,

L21(γ)(f(γ))L2(γ)L_{2}^{-1}\ell(\gamma)\leq\ell(f(\gamma))\leq L_{2}\ell(\gamma)

for any rectifiable curve γ\gamma, with \ell denoting length. The proof is completed by recalling that every bounded length distortion homeomorphism of 2\mathbb{R}^{2} (or any quasiconvex space) is bi-Lipschitz quantitatively. ∎

7. Decomposion and proof of Theorem 7.1

In this section we will prove the following result, which immediately implies Theorem 1.4.

Theorem 7.1.

Suppose the Assouad dimension of XX is less than 11 and ff is the bi-Lipschitz map from §6.2. Then there exists a bi-Lipschitz path H:[0,1]LIP(2)H:[0,1]\to LIP(\mathbb{R}^{2}) such that H0=fH_{0}=f and H1H_{1} is the identity.

The proof comprises of 4 steps. In the first step we relabel the ring domains RjR_{j} obtained from Proposition 1.3. In the second step we use Proposition 5.1 to unwind the Dehn twists in each RjR_{j} without changing small neighborhoods of XX. In the third step we compose the bi-Lipschitz paths from the second step to perform unwindings arbitrarily close to XX. Finally, in the fourth step, we use the uniformity of our maps to take a limit in the sequence of bi-Lipschitz paths obtained from the third step and recover the desired bi-Lipschitz path.

For the rest, we denote by (Rj)j(R_{j})_{j\in\mathbb{N}}, (ϕj)j(\phi_{j})_{j\in\mathbb{N}}, {g1,,gk}\{g_{1},\dots,g_{k}\}, and

(fj)j=(ϕjgi(j))j(f_{j})_{j\in\mathbb{N}}=(\phi_{j}\circ g_{i(j)})_{j\in\mathbb{N}}

the ring domains, similarities, and conformal maps, respectively, from Proposition 1.3.

7.1. Step 1: Relabelling the ring domains RjR_{j}.

This step is similar to the proof of Proposition 3.3.

For each jj\in\mathbb{N} let VjV_{j} and UjU_{j} be the bounded and unbounded, respectively, components of 2Rj¯\mathbb{R}^{2}\setminus\overline{R_{j}}.

Let ε\varepsilon be the empty word. There exist three distinct l1,l2,l3l_{1},l_{2},l_{3}\in\mathbb{N} such that

  1. (i)

    for all jj\in\mathbb{N}, there exists i{1,2,3}i\in\{1,2,3\} such that RjVliR_{j}\subset V_{l_{i}} and

  2. (ii)

    for all jj\in\mathbb{N} and all i{1,2,3}i\in\{1,2,3\}, RliVj=R_{l_{i}}\cap V_{j}=\emptyset.

For each l{1,2,3}l\in\{1,2,3\}, we denote Rl,ε=RilR_{l,\varepsilon}=R_{i_{l}} where ε\varepsilon denotes the empty word.

Inductively, suppose that for some l{1,2,3}l\in\{1,2,3\} and for some finite word w{1,2}w\in\{1,2\}^{*} we have labelled Rl,w=Rj0R_{l,w}=R_{j_{0}} where j0j_{0}\in\mathbb{N}. Then there exist exactly two distinct j1,j2j_{1},j_{2}\in\mathbb{N} such that

  1. (i)

    Rj1,Rj2Vj0R_{j_{1}},R_{j_{2}}\subset V_{j_{0}} and

  2. (ii)

    for all j{j1,j2}j\in\mathbb{N}\setminus\{j_{1},j_{2}\} with RjVj0R_{j}\subset V_{j_{0}}, either RjVj1R_{j}\subset V_{j_{1}}, or RjVj2R_{j}\subset V_{j_{2}}.

We denote Rl,w1=Rj1R_{l,w1}=R_{j_{1}} and Rl,w2=Rj2R_{l,w2}=R_{j_{2}}.

Thus, we have that {Rj:j}={Rl,w:l{1,2,3},w{1,2}}\{R_{j}:j\in\mathbb{N}\}=\{R_{l,w}:l\in\{1,2,3\},w\in\{1,2\}^{*}\}. Given l{1,2,3}l\in\{1,2,3\} and w{1,2}w\in\{1,2\}^{*} we denote by Vl,wV_{l,w} and Ul,wU_{l,w} the bounded and unbounded, respectively, components of 2Rl,w¯\mathbb{R}^{2}\setminus\overline{R_{l,w}}. Further, denote by Xl,wX_{l,w} the intersection Xl,w=XVl,wX_{l,w}=X\cap V_{l,w}.

Moreover, if Rj=Rl,wR_{j}=R_{l,w} we set ϕl,w=ϕj\phi_{l,w}=\phi_{j} and fl,w=fjf_{l,w}=f_{j}. In particular, fl,w=ϕl,wgi(l,w)f_{l,w}=\phi_{l,w}\circ g_{i(l,w)}.

By Proposition 1.3 we have that for all l{1,2,3}l\in\{1,2,3\}, w{1,2}w\in\{1,2\}^{*} and i{1,2}i\in\{1,2\}

(7.1) diamRl,widiamRl,wdiamVl,wdiamRl,wdiamRl,w2dist(Vl,w,Ul,w)diamRl,w.\frac{\operatorname{diam}{R_{l,wi}}}{\operatorname{diam}{R_{l,w}}}\leq\frac{\operatorname{diam}{V_{l,w}}}{\operatorname{diam}{R_{l,w}}}\leq\frac{\operatorname{diam}{R_{l,w}}-2\operatorname{dist}(V_{l,w},U_{l,w})}{\operatorname{diam}{R_{l,w}}}.

Suppose dist(Vl,w,Ul,w)\operatorname{dist}(V_{l,w},U_{l,w}) is realized by |xy||x-y|. Then since x,yRl,wx,y\in\partial R_{l,w} and ff is the identity there, we have by (6.6) that for some jj\in\mathbb{N},

(7.2) dist(Vl,w,Ul,w)\displaystyle\operatorname{dist}(V_{l,w},U_{l,w}) =|f(x)f(y)|diamRl,wL|(fl,w1)(x)(fl,w1)(y)|diamRl,wL2.\displaystyle=|f(x)-f(y)|\geq\frac{\operatorname{diam}{R_{l,w}}}{L}|(f_{l,w}^{-1})^{\prime}(x)-(f_{l,w}^{-1})^{\prime}(y)|\geq\frac{\operatorname{diam}{R_{l,w}}}{L^{2}}.

We conclude via (7.1) that

(7.3) diamRl,widiamRl,w11L2.\frac{\operatorname{diam}{R_{l,wi}}}{\operatorname{diam}{R_{l,w}}}\leq 1-\frac{1}{L^{2}}.

7.2. Step 2: Unwinding the Dehn twist in Rl,wR_{l,w} while acting as isometries on neighbourhoods of Xl,wX_{l,w}.

For each l{1,2,3}l\in\{1,2,3\} and w{1,2}w\in\{1,2\}^{*} we define a bi-Lipschitz path Hl,w:[0,1]LIP(2)H_{l,w}:[0,1]\to LIP(\mathbb{R}^{2}) as follows.

First, set Hl,w|Ul,wH_{l,w}|U_{l,w} to be the identity. Second, define Hl,w|Rl,wH_{l,w}|R_{l,w} so that for each t[0,1]t\in[0,1]

(Hl,w|Rl,w)t=fl,wD1t(fl,w)1(H_{l,w}|R_{l,w})_{t}=f_{l,w}\circ D_{1-t}\circ(f_{l,w})^{-1}

recalling DtD_{t} from Lemma 4.10.

Lemma 7.2.

The family of bi-Lipschitz paths

:={Hl,w|Rl,w:l{1,2,3},w{1,2}},\mathcal{F}:=\left\{H_{l,w}|R_{l,w}:l\in\{1,2,3\},w\in\{1,2\}^{*}\right\},

which unwinds the Dehn twist in each Rl,wR_{l,w}, is a uniform family of bi-Lipschitz paths.

Proof.

For each i{1,,k}i\in\{1,\dots,k\} and each t[0,1]t\in[0,1] set Hti=giD1tgi1H^{i}_{t}=g_{i}\circ D_{1-t}\circ g_{i}^{-1}. By Proposition 4.11, each HiH^{i} is a bi-Lipschitz path. Now for each i{1,,k}i\in\{1,\dots,k\} let

𝒢i={ϕl,wHi(ϕl,w)1:l{1,2,3},w{1,2}}.\mathcal{G}^{i}=\left\{\phi_{l,w}\circ H^{i}\circ(\phi_{l,w})^{-1}:l\in\{1,2,3\},w\in\{1,2\}^{*}\right\}.

Since XX is bounded, there exists cc depending on the diameter of XX such that each ϕl,w\phi_{l,w} has a scaling factor at most cc. Therefore, by Lemma 4.8, 𝒢i\mathcal{G}^{i} is a uniform family of bi-Lipschitz paths. Note that i=1k𝒢i\mathcal{F}\subset\bigcup_{i=1}^{k}\mathcal{G}^{i} so \mathcal{F} is a uniform family of bi-Lipschitz paths as a finite union of uniform families of bi-Lipschitz paths. ∎

Before defining Hl,w|Vl,wH_{l,w}|V_{l,w} we make some remarks.

First, there exist C>0C>0 and s(0,1)s\in(0,1) such that for any l{1,2,3}l\in\{1,2,3\} and w{1,2}w\in\{1,2\}^{*} the set ϕl,w1(Xl,w)\phi_{l,w}^{-1}(X_{l,w}) is (C,s)(C,s)-homogeneous.

Second, since {ϕl,w1(Vl,w)}l,w\{\phi^{-1}_{l,w}(V_{l,w})\}_{l,w} is a finite collection of Jordan domains with smooth boundary, there exists c>1c>1 such that for all l{1,2,3}l\in\{1,2,3\} and w{1,2}w\in\{1,2\}^{*}, the domain ϕl,w1(Vl,w)\phi^{-1}_{l,w}(V_{l,w}) satisfies (5.1) with constant cc.

Third, by the bi-Lipschitz Schoenflies Theorem [Tuk80, Theorem A], there exists L>1L^{\prime}>1 depending only on LL such that every gig_{i} extends to be an LL^{\prime}-bi-Lipschitz map on B¯(0,1)\overline{B}(0,1). Therefore, for each l{1,2,3}l\in\{1,2,3\} and w{1,2}w\in\{1,2\}^{*} there exists a disk Bl,wϕl,w1(Vl,w)B_{l,w}\subset\phi^{-1}_{l,w}(V_{l,w}) such that

radius(Bl,w)ϵanddist(Bl,w,ϕl,w1(Vl,w))ϵ\text{radius}(B_{l,w})\geq\epsilon\quad\text{and}\quad\operatorname{dist}(B_{l,w},\partial\phi_{l,w}^{-1}(V_{l,w}))\geq\epsilon

with ϵ:=(2L)1(1L1)\epsilon:=(2L^{\prime})^{-1}(1-L^{-1}).

Fourth, by (6.5), there exists η>0\eta>0 such that for all l{1,2,3}l\in\{1,2,3\} and w{1,2}w\in\{1,2\}^{*}

dist(ϕl,w1(Xl,w),ϕl,w1(Vl,w))ηdiamϕl,w1(Vl,w)η.\operatorname{dist}(\phi_{l,w}^{-1}(X_{l,w}),\partial\phi_{l,w}^{-1}(V_{l,w}))\geq\eta\operatorname{diam}{\phi_{l,w}^{-1}(V_{l,w})}\geq\eta.

Let δ\delta be the constant given in (5.3) depending only on C,s,c,η,ϵC,s,c,\eta,\epsilon above. Recall from the proof of Proposition 1.3 that for all l{1,2,3}l\in\{1,2,3\} and w{1,2}w\in\{1,2\}^{*}, ϕl,w1(Vl,w)[132ϵ0,1+32ϵ0]2\phi_{l,w}^{-1}(V_{l,w})\subset[-1-\frac{3}{2}\epsilon_{0},1+\frac{3}{2}\epsilon_{0}]^{2}. Therefore, there exist at most k1k_{1} different configurations for 𝒯δ(ϕl,w1(Xl,w))\mathcal{T}_{\delta}(\phi^{-1}_{l,w}(X_{l,w})) inside ϕl,w1(Vl,w)\phi^{-1}_{l,w}(V_{l,w}). Applying Proposition 5.1 for each of these finitely many cases we obtain bi-Lipschitz paths {H1,,Hk2}\{H_{1},\dots,H_{k_{2}}\} such that for each l{1,2,3}l\in\{1,2,3\} and w{1,2}w\in\{1,2\}^{*}, there exists j(l,w){1,,k2}j(l,w)\in\{1,\dots,k_{2}\} for which

  1. (i)

    Hj(l,w):[0,1]LIP(ϕl,w1(Vl,w))H_{j(l,w)}:[0,1]\to LIP(\phi_{l,w}^{-1}(V_{l,w})),

  2. (ii)

    Hj(l,w)H_{j(l,w)} is an isometry on each component of 𝒯δ(ϕl,w1(Xl,w))\mathcal{T}_{\delta}(\phi^{-1}_{l,w}(X_{l,w})),

  3. (iii)

    (Hj(l,w))1(H_{j(l,w)})_{1} maps 𝒯δ(ϕl,w1(Xl,w))\mathcal{T}_{\delta}(\phi^{-1}_{l,w}(X_{l,w})) onto Bl,wB_{l,w}.

By (7.3) there exists pp\in\mathbb{N}, so that if u{1,2}pu\in\{1,2\}^{p} then

Rl,wuϕl,w(𝒯δ(ϕl,w1(Xl,w))).R_{l,wu}\subset\phi_{l,w}\left(\mathcal{T}_{\delta}(\phi^{-1}_{l,w}(X_{l,w}))\right).

We define Hl,w|Vl,wH_{l,w}|V_{l,w} as follows.

  1. (a)

    For 0t1/30\leq t\leq 1/3, we set

    (Hl,w|Vl,w)t=ϕl,w(Hi(l,w))3tϕl,w1(H_{l,w}|V_{l,w})_{t}=\phi_{l,w}\circ(H_{i(l,w)})_{3t}\circ\phi^{-1}_{l,w}

    to be the path which moves ϕl,w(𝒯δ(ϕl,w1(Xl,w)))\phi_{l,w}\left(\mathcal{T}_{\delta}(\phi^{-1}_{l,w}(X_{l,w}))\right) into the disk ϕl,w(Bl,w)\phi_{l,w}(B_{l,w}).

  2. (b)

    For 2/3t12/3\leq t\leq 1, we set

    (Hl,w|Vl,w)t=(Hwl|Vl,w)1t.(H_{l,w}|V_{l,w})_{t}=(H^{l}_{w}|V_{l,w})_{1-t}.
  3. (c)

    For 1/3t2/31/3\leq t\leq 2/3, we define Hl,w|Vl,wH_{l,w}|V_{l,w} as a path of rotations. Fix l,wl,w and suppose that Bl,w=B(z0,r)B_{l,w}=B(z_{0},r). Find a conformal map

    ψl,w:ϕl,w1(Vl,w)Bl,w¯{z:1<|z|<ρl,w}\psi_{l,w}:\phi^{-1}_{l,w}(V_{l,w})\setminus\overline{B_{l,w}}\to\{z:1<|z|<\rho_{l,w}\}

    for some ρl,w>1\rho_{l,w}>1. Since the boundary ϕl,w1(Vl,w)\phi^{-1}_{l,w}(V_{l,w}) is smooth, ψl,w\psi_{l,w} extends smoothly on ϕl,w1(Vl,w)\partial\phi^{-1}_{l,w}(V_{l,w}). We apply Proposition 4.12 with P,QP,Q given by

    Qt(z)\displaystyle Q_{t}(z) =ψl,wϕl,w1(Hl,w|Vl,w)(z)\displaystyle=\psi_{l,w}\circ\phi^{-1}_{l,w}\circ(H_{l,w}|\partial V_{l,w})(z)
    Pt(z)\displaystyle P_{t}(z) =ψl,w(z0+(zz0)e2π(13t)i)).\displaystyle=\psi_{l,w}(z_{0}+(z-z_{0})e^{2\pi(1-3t)i)}).

    Here Hl,w|Vl,wH_{l,w}|\partial V_{l,w} agrees with Hl,wH_{l,w} on the inner boundary component of Rl,wR_{l,w}, recalling the construction in Lemma 7.2. This yields a bi-Lipschitz path

    Pl,w:[0,1]LIP({z:1|z|ρl,w}).P_{l,w}:[0,1]\to LIP(\{z:1\leq|z|\leq\rho_{l,w}\}).

    By Proposition 4.11, Gl,w:=(ψwl)1PwlψwlG_{l,w}:=(\psi^{l}_{w})^{-1}\circ P^{l}_{w}\circ\psi^{l}_{w} is a bi-Lipschitz path. Since there are finitely many different pairs (ϕl,w1(Vl,w),Bl,w)(\phi^{-1}_{l,w}(V_{l,w}),B_{l,w}), the set

    {Gl,w:l{1,2,3},w{1,2}}\{G_{l,w}:l\in\{1,2,3\},w\in\{1,2\}^{*}\}

    is finite. Set now for 1/3t2/31/3\leq t\leq 2/3,

    (Hl,w|Vl,w)t=ϕl,w(Gl,w)3t1ϕl,w1.(H_{l,w}|V_{l,w})_{t}=\phi_{l,w}\circ(G_{l,w})_{3t-1}\circ\phi^{-1}_{l,w}.

By the finiteness of the family {Gl,w}l,w\{G_{l,w}\}_{l,w}, and working as in Lemma 7.2, we see that {Hl,w|Vl,w:l{1,2,3},w{1,2}}\{H_{l,w}|V_{l,w}:l\in\{1,2,3\},w\in\{1,2\}^{*}\} is a uniform family of bi-Lipschitz paths.

By Lemma 4.4, {Hl,w:l{1,2,3},w{1,2}}\{H_{l,w}:l\in\{1,2,3\},w\in\{1,2\}^{*}\} is a uniform family of bi-Lipschitz paths. The key point in the construction of Hl,wH_{l,w} is that it unwinds the Dehn twist in Rl,wR_{l,w} and acts as an isometry on Rl,wuR_{l,{wu}} for any u{1,2}pu\in\{1,2\}^{p}.

7.3. Step 3: Composing unwindings in a controlled way.

The next step is to combine the paths Hl,wH_{l,w} defined above. Let k{0,1,,p1}k\in\{0,1,\ldots,p-1\}. Define

(F0k)t(z)={(Hl,w)t(z)zRl,wVl,w for |w|=kl{1,2,3},zotherwise.(F^{k}_{0})_{t}(z)=\begin{cases}(H_{l,w})_{t}(z)&\text{$z\in R_{l,w}\cup V_{l,w}$ for $|w|=k$, $l\in\{1,2,3\}$,}\\ z&\text{otherwise.}\end{cases}

This is a bi-Lipschitz path. For example, for k=0k=0, this path unwinds the Dehn twists in the three outermost rings R1,ϵ,R2,ϵ,R3,ϵR_{1,\epsilon},R_{2,\epsilon},R_{3,\epsilon}. Then for jj\in\mathbb{N}, suppose that Fj1kF^{k}_{j-1} has been defined. We then define

(7.4) (Fjk)t(z)={(Fj1k)t(Hl,w)t(z),zRl,wVl,w|w|=k+jpl{1,2,3},(Fj1k)t(z),otherwise.(F^{k}_{j})_{t}(z)=\begin{cases}(F^{k}_{j-1})_{t}\circ(H_{l,w})_{t}(z),&\text{$z\in R_{l,w}\cup V_{l,w}$, $|w|=k+jp$, $l\in\{1,2,3\}$,}\\ (F^{k}_{j-1})_{t}(z),&\text{otherwise.}\end{cases}

If |w|=k+jp|w|=k+jp and l{1,2,3}l\in\{1,2,3\}, then (Fj1k)t(F^{k}_{j-1})_{t} acts as an isometry on Rl,wVl,wR_{l,w}\cup V_{l,w}. Hence Lemma 4.7 implies that the composition in (7.4) gives a bi-Lipschitz path, and we conclude that FjkF^{k}_{j} is a bi-Lipschitz path which unwinds the Dehn twists in Rl,wR_{l,w} for |w|=k,k+p,k+2p,,k+jp|w|=k,k+p,k+2p,\ldots,k+jp and l=1,2,3l=1,2,3.

7.4. Step 4: Taking a limit.

Set FkF^{k} by (Fk)t=limj(Fjk)t(F^{k})_{t}=\lim_{j\to\infty}(F^{k}_{j})_{t} for all t[0,1]t\in[0,1]. We claim that FkF^{k} is a bi-Lipschitz path. To that end, first consider, for nn\in\mathbb{N}, the domain

𝒰n:=|w|=k+npl{1,2,3}Ul,w.\mathcal{U}_{n}:=\bigcup_{\begin{subarray}{c}|w|=k+np\\ l\in\{1,2,3\}\end{subarray}}U_{l,w}.

By construction, on this set we have Fk|𝒰n=Fnk|𝒰nF^{k}|\mathcal{U}_{n}=F^{k}_{n}|\mathcal{U}_{n}, and hence Fk|𝒰nF^{k}|\mathcal{U}_{n} is a bi-Lipschitz path.

Next, note from (7.4) that FjkF^{k}_{j} is obtained from Fj1kF^{k}_{j-1} by modifications from a uniform family of bi-Lipschitz paths (namely, the family {Hl,w}l,w\{H_{l,w}\}_{l,w}) on a region where Fj1kF^{k}_{j-1} acts as a family of isometries in a uniform way. By Lemma 4.7, it follows that the family

{Fjk|n𝒰n:j}\{F^{k}_{j}|\bigcup_{n\in\mathbb{N}}\mathcal{U}_{n}:j\in\mathbb{N}\}

is a uniform family of bi-Lipschitz paths. Hence Fk|n𝒰nF^{k}|\bigcup_{n\in\mathbb{N}}\mathcal{U}_{n} is a bi-Lipschitz path.

Since n𝒰n=2X\bigcup_{n\in\mathbb{N}}\mathcal{U}_{n}=\mathbb{R}^{2}\setminus X, an application of Proposition 4.5 shows that FkF^{k} is in fact a bi-Lipschitz path on all of 2\mathbb{R}^{2} which unwinds the Dehn twists in Rl,wR_{l,w} for |w|k+p|w|\in k+p\mathbb{N}, l=1,2,3l=1,2,3. Hence the concatenation of the finitely many paths F0,F1,,Fp1F^{0},F^{1},\cdots,F^{p-1} yields a bi-Lipschitz path which connects ff to the identity.

8. A decomposable multitwist with singular set of large Assouad dimension

Let DεD_{\varepsilon} be the rectangle [2,2]×[1,1][-\sqrt{2},\sqrt{2}]\times[-1,1], let α(0,1)\alpha\in(0,1) and let

D1\displaystyle D_{1} =[2(112α),212α]×[α1,1α]\displaystyle=[-\sqrt{2}(1-\tfrac{1}{2}\alpha),-\sqrt{2}\tfrac{1}{2}\alpha]\times[\alpha-1,1-\alpha]
D2\displaystyle D_{2} =[212α,2(112α)]×[α1,1α].\displaystyle=[\sqrt{2}\tfrac{1}{2}\alpha,\sqrt{2}(1-\tfrac{1}{2}\alpha)]\times[\alpha-1,1-\alpha].

as in Figure 6. Here ε\varepsilon denotes the empty word. For each i{1,2}i\in\{1,2\} let ϕi\phi_{i} be the similarity of 2\mathbb{R}^{2} mapping DεD_{\varepsilon} onto DiD_{i} with scaling factor 12(1α)\frac{1}{\sqrt{2}}(1-\alpha). Let XX be the Cantor set attractor of the iterated function system {ϕ1,ϕ2}\{\phi_{1},\phi_{2}\}.

Refer to caption
Figure 6. The first two steps in the construction of XX.

By self-similarity, XX is uniformly disconnected and its Assouad dimension is

dimA(X)=log2log2log(1α)\dim_{A}(X)=\frac{\log{2}}{\log{\sqrt{2}}-\log(1-\alpha)}

which is greater than 1 when α\alpha is sufficiently small. Moreover, there exists a multitwist bi-Lipschitz map ff as in Section 6.2, and by self-similarity, the set of maps {gj}\{g_{j}\} in Proposition 1.3 contains one single element.

We claim that the map ff is decomposable. To prove the claim, we follow the arguments in Section 7. We may assume that the domains {Vl,w}\{V_{l,w}\} are exactly the interiors of the rectangles {Dw}\{D_{w}\}. For simplicity, we drop the index ll. The only step in the proof that we need to check (and the only one that requires the assumption on the Assouad dimension) is the existence of bi-Lipschitz paths Hj(w)H_{j(w)}. Since the collection {gj}\{g_{j}\} contains only one element, we only need to construct for each ϵ>0\epsilon>0 a “collapsing” bi-Lipschitz path H:[0,1]DεH:[0,1]\to D_{\varepsilon} which, for some small δ>0\delta>0, is an isometry on each component of 𝒯δ(X)\mathcal{T}_{\delta}(X) and maps 𝒯δ(X)\mathcal{T}_{\delta}(X) into a ball BB in DεD_{\varepsilon} of radius ϵ\epsilon.

We give a rough sketch of the construction of HH and leave the details to the reader. Fix ϵ>0\epsilon>0. Choose β(0,1)\beta\in(0,1) such that (1α)(1+β)<1(1-\alpha)(1+\beta)<1 and choose nn\in\mathbb{N} such that

(1+β)n(1α)n<14ϵ.(1+\beta)^{n}(1-\alpha)^{n}<\tfrac{1}{4}\epsilon.

The bi-Lipschitz path HH is a concatenation of nn bi-Lipschitz paths H1,,HnH_{1},\dots,H_{n}. Let H1H_{1} be the bi-Lipschitz path that is identity outside of w{1,2}n1Dw\bigcup_{w\in\{1,2\}^{n-1}}D_{w} and for each w{1,2}n1w\in\{1,2\}^{n-1}, it moves Dw1D_{w1} towards Dw2D_{w2} so that they both end up in a rectangle DwD_{w}^{\prime} with sides parallel to the axes and side-lengths

4(1+β)(12(1α))n,22(1+β)(12(1α))n.4(1+\beta)(\tfrac{1}{\sqrt{2}}(1-\alpha))^{n},\quad 2\sqrt{2}(1+\beta)(\tfrac{1}{\sqrt{2}}(1-\alpha))^{n}.

The choice of β\beta ensures that DwD_{w}^{\prime} is contained in DwD_{w}. Moreover, H1H_{1} acts as an isometry on DwiD_{wi} for all wi{1,2}nwi\in\{1,2\}^{n}.

Assume now that for some m{1,,n1}m\in\{1,\dots,n-1\} we have defined the paths H1,,HmH_{1},\dots,H_{m} and assume that

  1. (i)

    the concatenation of these paths is the identity outside of w{1,2}nmDw\bigcup_{w\in\{1,2\}^{n-m}}D_{w},

  2. (ii)

    for each w{1,2}nmw\in\{1,2\}^{n-m}, the concatenation has moved XDwX\cap D_{w} inside a rectangle DwDwD_{w}^{\prime}\subset D_{w} with sides parallel to the axes and side-lengths

    22(2)m(1+β)m(12(1α))n,2(2)m(1+β)m(12(1α))n,2\sqrt{2}(\sqrt{2})^{m}(1+\beta)^{m}(\tfrac{1}{\sqrt{2}}(1-\alpha))^{n},\qquad 2(\sqrt{2})^{m}(1+\beta)^{m}(\tfrac{1}{\sqrt{2}}(1-\alpha))^{n},
  3. (iii)

    for each u{1,2}nu\in\{1,2\}^{n} the concatenation of these paths acts as an isometry on DuD_{u}.

Let Hm+1H_{m+1} be the bi-Lipschitz path that is identity outside of w{1,2}nm1Dw\bigcup_{w\in\{1,2\}^{n-m-1}}D_{w} and for each w{1,2}nm1w\in\{1,2\}^{n-m-1}, it moves Dw1D_{w1}^{\prime} towards Dw2D_{w2}^{\prime} so that they both end up in a rectangle DwD_{w}^{\prime} with sides parallel to the axes and side-lengths

22(2)m+1(1+β)m+1(12(1α))n,2(2)m+1(1+β)m+1(12(1α))n.2\sqrt{2}(\sqrt{2})^{m+1}(1+\beta)^{m+1}(\tfrac{1}{\sqrt{2}}(1-\alpha))^{n},\qquad 2(\sqrt{2})^{m+1}(1+\beta)^{m+1}(\tfrac{1}{\sqrt{2}}(1-\alpha))^{n}.

Note that Hm+1H_{m+1} acts as an isometry on DuD_{u} for all u{1,2}nu\in\{1,2\}^{n}.

Finally, the concatenation HH of paths H1,,HnH_{1},\dots,H_{n} is the identity outside of DD, acts as an isometry on DuD_{u} for all u{1,2}nu\in\{1,2\}^{n}, and H(X)H(X) is contained in a rectangle DDD^{\prime}\subset D with side-lengths

22(1+β)n(1α)n,2(1+β)n(1α)n.2\sqrt{2}(1+\beta)^{n}(1-\alpha)^{n},\quad 2(1+\beta)^{n}(1-\alpha)^{n}.

By the choice of nn, the rectangle DD^{\prime} has diameter less than ϵ\epsilon and the proof is complete.

References

  • [AIPS15] K. Astala, T. Iwaniec, I. Prause, and E. Saksman, Bilipschitz and quasiconformal rotation, stretching and multifractal spectra, Pub. Math. IHÉS 121 (2015), 113–154.
  • [ALP+11] Daniele Alessandrini, Lixin Liu, Athanase Papadopoulos, Weixu Su, and Zongliang Sun, On Fenchel-Nielsen coordinates on Teichmüller spaces of surfaces of infinite type, Ann. Acad. Sci. Fenn. Math. 36 (2011), no. 2, 621–659. MR 2865518
  • [BK87] Steven R. Bell and Steven G. Krantz, Smoothness to the boundary of conformal maps, Rocky Mount. J. Math. 17 (1987), no. 1, 23–40.
  • [BM07] A. F. Beardon and D. Minda, The hyperbolic metric and geometric function theory, Quasiconformal mappings and their applications, Narosa, New Delhi, 2007, pp. 9–56. MR 2492498
  • [DS97] Guy David and Stephen Semmes, Fractured fractals and broken dreams, Oxford Lecture Series in Mathematics and its Applications, vol. 7, The Clarendon Press, Oxford University Press, New York, 1997, Self-similar geometry through metric and measure. MR 1616732
  • [FH88] M. Freedman and Z.-X. He, Factoring the logarithmic spiral, Invent. Math. 92 (1988), no. 1, 129–138.
  • [FM12] Alastair Fletcher and Vladimir Markovic, Decomposing diffeomorphisms of the sphere, Bull. Lond. Math. Soc. 44 (2012), no. 3, 599–609. MR 2967005
  • [GM01] V. Gutlyanskii and O. Martio, Rotation estimates and spirals, Conform. Geom. Dyn. 5 (2001), 6–20.
  • [GM05] John B. Garnett and Donald E. Marshall, Harmonic measure, New Mathematical Monographs, vol. 2, Cambridge University Press, Cambridge, 2005. MR 2150803
  • [Leh87] Olli Lehto, Univalent functions and Teichmüller spaces, Graduate Texts in Mathematics, vol. 109, Springer-Verlag, New York, 1987. MR 867407
  • [Loe59] Charles Loewner, On the conformal capacity in space, J. Math. Mech. 8 (1959), 411–414. MR 0104785
  • [Luu98] Jouni Luukkainen, Assouad dimension : antifractal metrization, porous sets, and homogeneous measures, J. Kor. Math. Soc. 35 (1998), no. 1, 23–76.
  • [Mac99] Paul MacManus, Catching sets with quasicircles, Rev. Mat. Iberoamericana 15 (1999), no. 2, 267–277. MR 1715408
  • [Mas85] B. Maskit, Comparison of hyperbolic and extremal lengths, Ann. Acad. Sci. Fenn. 10 (1985), 381–386.
  • [Pom79] C. Pommerenke, Uniformly perfect sets and the poincaré metric, Arch. Math. 32 (1979), 192–199.
  • [Pom92] Ch. Pommerenke, Boundary behaviour of conformal maps, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 299, Springer-Verlag, Berlin, 1992. MR 1217706
  • [Shi18] H. Shiga, On the quasiconformal equivalence of dynamical cantor sets, arXiv:1812.07785 (2018).
  • [Tuk80] Pekka Tukia, The planar Schönflies theorem for Lipschitz maps, Ann. Acad. Sci. Fenn. Ser. A I Math. 5 (1980), no. 1, 49–72. MR 595177
  • [Vel21] Vyron Vellis, Uniformization of Cantor sets with bounded geometry, Conform. Geom. Dyn. 25 (2021), 88–103. MR 4298216
  • [Wol81] Scott Wolpert, The length spectrum as moduli for compact Riemann surfaces, Riemann surfaces and related topics: Proceedings of the 1978 Stony Brook Conference (State Univ. New York, Stony Brook, N.Y., 1978), Ann. of Math. Stud., vol. 97, Princeton Univ. Press, Princeton, N.J., 1981, pp. 515–517. MR 624836