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

The topology of SLEκSLE_{\kappa} is random for κ>4\kappa>4

Stephen Yearwood
Abstract

We study the topology of SLESLE curves for κ>4\kappa>4. More precisely, we show that, a.s., there is no homeomorphism Φ:¯¯\Phi:\overline{{\mathbb{H}}}\rightarrow\overline{{\mathbb{H}}}, taking the range of one independent SLESLE curve to another for κ(4,8)\kappa\in(4,8). Furthermore, we extend the result to κ8\kappa\geq 8 by showing that there is no homeomorphism taking one SLESLE curve to another, when viewed as curves modulo parametrization.

1 Introduction

1.1 Initial Overview

The Schramm-Loewner Evolutions (SLEκSLE_{\kappa}), introduced by Oded Schramm [14], describes a family of probability distributions, parameterized by κ>0\kappa>0 on non-self traversing curves connecting two boundary points in a planar, simply connected domain. They are characterized by a conformal invariance condition and a domain Markov property. They were initially observed as possible candidates for the scaling limits of various discrete lattice models in statistical physics; we now know that some of these convergences do hold, and so SLESLE exhibits a universality in its definition.

SLEκSLE_{\kappa} is the random growth of a set KtK_{t}, as described through a conformal map gt(z)g_{t}(z) on the complement of KtK_{t}. This map gt(z)g_{t}(z) is the solution of the Loewner differential equation driven by a Brownian motion, whose ‘speed’ is determined by a single parameter κ\kappa. Rhodes and Schramm in [13] showed that for κ8\kappa\neq 8, a.s. there is a (unique) continuous path η:[0,)¯\eta:[0,\infty)\to\overline{{\mathbb{H}}} such that for each t>0t>0 the set KtK_{t} is the union of η[0,t]\eta[0,t] and the bounded connected components of η[0,t]{\mathbb{H}}\setminus\eta[0,t]. This has also been shown for κ=8\kappa=8, but was dealt with separately. We call the path η\eta the SLE trace or SLESLE curve. It was shown as well in [13] that limt|η(t)|=\underset{t\to\infty}{\text{lim}}|\eta(t)|=\infty a.s. We will need the following facts about the curve which are proven in [13]:

  • If κ4\kappa\leq 4, then η\eta is simple with η(0,)\eta(0,\infty)\subset{\mathbb{H}}.

  • If 4<κ<84<\kappa<8, then η(0,)\eta(0,\infty) has double points and intersects {\mathbb{R}}.

  • If κ8\kappa\geq 8, the curve is space-filling.

There are three (well studied) variants of SLESLE : chordal SLESLE, which connects two boundary points (prime ends) in a given domain, radial SLESLE, which connects a boundary point to an interior point, and whole-plane SLESLE, which connects two points on the Riemann sphere. In this paper, we will focus on chordal SLESLE, but we expect that our results can be generalized to other cases. One can also see [7, 2, 15] for some expository work on SLESLE which go into details beyond the scope of this paper.

Most works on SLESLE have focused on its geometric and probabilistic properties, e.g., Hausdorff dimensions of various subsets of the curves, formulas for the probabilities of various events, and connections to other random objects. In this work, we will address a very basic question about the topology of SLESLE: namely, is the topology of the curve deterministic? Said differently, if we have two independent chordal SLEκ curves η1\eta^{1} and η2\eta^{2} (viewed as curves modulo time parametrization), does there a.s. exist a homeomorphism ¯¯\overline{\mathbb{H}}\rightarrow\overline{\mathbb{H}} taking η1\eta^{1} to η2\eta^{2}?

Since SLEκSLE_{\kappa} is a simple curve for κ4\kappa\leq 4, the answer to the above question is clearly affirmative in this case. For κ>4\kappa>4, however, the answer is less obvious. On the one hand, many events for SLEκ occur with probability strictly between 0 and 1 (see Section 2 of [11]) so there are many opportunities for one of η1\eta^{1} or η2\eta^{2} to do something that the other does not. On the other hand, it is common for seemingly very different fractal sets to be homeomorphic. For example, if K1K_{1} and K2K_{2} are compact, non-empty, totally disconnected subsets of {\mathbb{C}} without isolated points (e.g., Cantor-type sets), then there is a homeomorphism from {\mathbb{C}} to {\mathbb{C}} which takes K1K_{1} to K2K_{2} [12]. The main results of this paper show that the topology of SLEκ is random for κ>4\kappa>4. For κ(4,8)\kappa\in(4,8), we prove the stronger statement that the topology of the range is random. The results of this paper are in a similar vein to those of [10], which shows that an SLEκSLE_{\kappa} curve for κ(4,8)\kappa\in(4,8) is not determined by its range. Both this paper and [10] answer a seemingly simple question about SLESLE whose answer is much less obvious than one might initially expect.

1.2 Summary of results

Theorem 1.

The topology of chordal SLEκSLE_{\kappa} is not deterministic in the following sense: Fix κ(4,8),\kappa\in(4,8), and consider two instances of SLEκSLE_{\kappa}, η1\eta^{1} and η2\eta^{2} in {\mathbb{H}}. Then a.s. there is no homeomorphism on ¯\overline{{\mathbb{H}}} taking the range of η1\eta^{1} to the range of η2\eta^{2}.

We consider the left and right boundaries of an SLESLE curve η\eta (which are boundary-touching SLE16/κ(ρ¯)SLE_{16/\kappa}(\bar{\rho}) curves). These curves form ‘bubbles’ in {\mathbb{H}} (which we characterize explicitly in a later section) which we use as the primary observable to prove Theorem 1.

The result also holds for κ8\kappa\geq 8. The proof is similar, though a bit more work is needed in the setup.

Theorem 2.

The topology of chordal SLEκSLE_{\kappa} is not deterministic for κ8\kappa\geq 8 in the following sense: Consider two SLEκSLE_{\kappa} curves, η1\eta^{1} and η2\eta^{2} in {\mathbb{H}}. Then a.s. there is no homeomorphism Φ:¯¯\Phi:\overline{{\mathbb{H}}}\rightarrow\overline{{\mathbb{H}}} such that Φ(η1)=η2\Phi(\eta^{1})=\eta^{2} viewed as curves modulo time parametrization.

Remark 1.

Notice that in Theorem 2, we care about parametrized curves, and so preservation of ranges in this setting makes less sense. Recall SLESLE in this instance is plane filling.

As a natural extension, one can think about the behavior of these curves for varying κ\kappa.

Conjecture 1.1.

Let κ1,κ2>4\kappa_{1},\kappa_{2}>4 be distinct. Let η1\eta_{1} (resp. η2\eta_{2}) be a chordal SLEκ1{}_{\kappa_{1}} (resp. SLEκ2{}_{\kappa_{2}}) in \mathbb{H}. Almost surely, there is no homeomorphism Φ:¯¯\Phi:\overline{\mathbb{H}}\rightarrow\overline{\mathbb{H}} such that Φ(η1)=η2\Phi(\eta^{1})=\eta^{2} viewed as curves modulo time parametrization. If one of κ1\kappa_{1} or κ2\kappa_{2} is in (4,8)(4,8), a.s., there is no such homeomorphism which takes the range of η1\eta_{1} to the range of η2\eta_{2}.

We expect that this conjecture can be proved using similar ideas to the ones in this paper, but one would have to explicitly compute some of the quantities involved to show that they are κ\kappa-dependent.

2 Preliminaries

Here we discuss a few SLESLE basics as well as how one defines the more general SLEκ(ρ¯)SLE_{\kappa}(\bar{\rho}) processes. Recall that we write :={z:𝔪(z)>0}.{\mathbb{H}}:=\{z\in{\mathbb{C}}:\mathfrak{Im}(z)>0\}. If KK is a bounded closed subset of {\mathbb{H}} such that K{\mathbb{H}}\setminus K is simply connected, then we call KK a hull in {\mathbb{H}} w.r.t. \infty. For such KK, there is a unique gKg_{K} that maps K{\mathbb{H}}\setminus K conformally onto {\mathbb{H}} such that gK(z)=z+az+O(1z2)g_{K}(z)=z+\dfrac{a}{z}+O\left(\dfrac{1}{z^{2}}\right) as zz\to\infty. The quantity aa is known as the half plane capacity of KK, and is denoted by hcapK\text{hcap}{K}. It can be shown that a0a\geq 0. The map gKg_{K} is said to satisfy the hydrodynamic normalization at infinity. For a real interval II, let 𝒞(I)\mathcal{C}(I) denote the real-valued continuous functions on I. Suppose U𝒞([0,T])U\in\mathcal{C}([0,T]) for some T(0,]T\in(0,\infty]. The chordal Loewner equation driven by UU is as follows:

g˙t(z)=2gt(z)Utg0(z)=z\dot{g}_{t}(z)=\dfrac{2}{g_{t}(z)-U_{t}}\qquad g_{0}(z)=z (1)

For 0t<T0\leq t<T, let KtK_{t} and gtg_{t} be the chordal Loewner hulls and maps, respectively, driven by UtU_{t}. Suppose that for every t[0,T)t\in[0,T),

ηt:=limz,zUtgt1(z)\eta_{t}:=\underset{z\in{\mathbb{H}},z\to U_{t}}{\lim}g_{t}^{-1}(z)\in{\mathbb{H}}\cup{\mathbb{R}}

exists, and ηt,0t<T\eta_{t},0\leq t<T, is a continuous curve. Then for every t[0,T),Ktt\in[0,T),K_{t} is the complement of the unbounded component of η((0,t]){\mathbb{H}}\setminus\eta((0,t]). We call η\eta the chordal Loewner trace driven by UtU_{t}. In general, however, such a curve may not exist depending on the choice of driving function.

An SLEκSLE_{\kappa} in {\mathbb{H}} from 0 to \infty is defined by the random family of conformal maps gtg_{t} obtained by solving the Loewner ODE driven by Brownian motion. In particular, we let Ut=κBtU_{t}=\sqrt{\kappa}B_{t}, where BtB_{t} is a standard Brownian motion. An SLEκSLE_{\kappa} connecting boundary points xx and yy of an arbitrary simply connected Jordan domain can be constructed as the image of an SLEκSLE_{\kappa} on {\mathbb{H}} under a conformal transformation Ψ:D\Psi\colon{\mathbb{H}}\to D sending 0 to xx and \infty to yy. SLESLE curves are characterized by scale invariance and the domain Markov property, and are viewed modulo reparametrization. The almost sure continuity of the curves of these processes has also been shown in [13].

SLE(κ;ρ¯)SLE(\kappa;\bar{\rho}), which is often written as SLEκ(ρ¯L;ρ¯R)SLE_{\kappa}(\bar{\rho}_{L};\bar{\rho}_{R}), is the stochastic process one obtains by solving (1) with a modification on the driving process UtU_{t}, which we now discuss. It is a natural generalization of SLEκSLE_{\kappa} in which one keeps track of additional marked points which are called force points. Fix xL¯=(xl,L<<x1,L0)\bar{x_{L}}=(x_{l,L}<\dots<x_{1,L}\leq 0) and x¯R=(0x1,R<<xr,R)\bar{x}_{R}=(0\leq x_{1,R}<\dots<x_{r,R}). We associate with each xi,qx_{i,q} for qL,Rq\in{L,R} a weight ρi,q\rho_{i,q}\in{\mathbb{R}}. An SLEκ(ρ¯L;ρ¯R)SLE_{\kappa}(\bar{\rho}_{L};\bar{\rho}_{R}) process with force points (x¯L;x¯R)(\bar{x}_{L};\bar{x}_{R}) is the measure on continuously growing compact hulls KtK_{t} generated by the Loewner chain with UtU_{t} replaced by the solution to the system of SDEs given by

dUt=i=1lρi,LUtVi,Ldt+i=1rρi,RUtVi,Rdt+κdBtdU_{t}=\sum\limits_{i=1}^{l}\dfrac{\rho_{i,L}}{U_{t}-V^{i,L}}dt+\sum\limits_{i=1}^{r}\dfrac{\rho_{i,R}}{U_{t}-V^{i,R}}dt+\sqrt{\kappa}\,dB_{t} (2)
dVti,q=2Vti,qdt;V0i,q=xi,q,i,q{L,R}dV_{t}^{i,q}=\dfrac{2}{V_{t}^{i,q}}dt;\quad V_{0}^{i,q}=x_{i,q},\quad i\in{\mathbb{N}},\quad q\in\{L,R\} (3)

The existence and uniqueness of solutions to (1) is discussed in [6]. In particular, it is shown that there is a unique solution to (1) until the first time tt that Ut=Vtj,qU_{t}=V_{t}^{j,q} where i=1jρi,q2\sum\limits_{i=1}^{j}\rho^{i,q}\leq-2 for q{L,R}q\in\{L,R\} (we call this time the continuation threshold). In particular,if i=1jρi,q>2\sum\limits_{i=1}^{j}\rho^{i,q}>-2 for all 1j|ρ¯q|1\leq j\leq|\bar{\rho}^{q}| for q{L,R}q\in\{L,R\}, then (1) has a unique solution for all times tt. This even holds when one or both of the x1,qx^{1,q} are zero. Note that in [8] it is proved that SLEκ(ρ¯)SLE_{\kappa}(\bar{\rho}) is generated by a curve and is transient. In this paper, we need only consider the case of two force points, though the main ingredients for the proofs are stated in more generality.

For κ>4\kappa>4, there is also significant interest in the hulls that are generated by the SLEκSLE_{\kappa} curves. Duplantier conjectured in [4, 5] the duality between SLEκSLE_{\kappa} and SLE16/κSLE_{16/\kappa}, which says that the boundary of an SLEκSLE_{\kappa} hull behaves like an SLE16/κSLE_{16/\kappa} curve, for κ>4\kappa>4. Many versions of this duality have also been shown in [16, 17, 3, 8, 9].

Lemma 4.9 in [8] asserts that, for κ>4\kappa>4, the outer boundary η\eta^{\prime} of an SLEκSLE_{\kappa} curve is an SLEκ(ρ¯)SLE_{\kappa}(\bar{\rho}) process. The SLESLE curves are realized as flow lines of the Gaussian free field (i.e SLEκ(ρ¯)SLE_{\kappa}(\bar{\rho}) curves coupled with the Gaussian free field in {\mathbb{H}}), with the outer boundaries described as counterflow lines (in which the coupling is done with the negation of the Gaussian field). Though we do not need this machinery as presented in [8] and [11], it serves as an excellent framework for proving some general properties of SLEκ(ρ¯)SLE_{\kappa}(\bar{\rho}), some of which we rely on to prove the main results. We state one such fact as follows:

Lemma 2.1.

Fix κ>0\kappa>0. Suppose that η\eta is an SLEκ(ρ¯L;ρ¯R)SLE_{\kappa}(\bar{\rho}_{L};\bar{\rho}_{R}) process in {\mathbb{H}} from 0 to \infty with force points located at (x¯L;x¯R)(\bar{x}_{L};\bar{x}_{R}) with x1,L=0x_{1,L}=0^{-} and x1,R=0+x_{1,R}=0^{+} (possibly by taking ρ1,q=0\rho_{1,q}=0 for q{L,R}q\in\{L,R\}). Assume that ρ1,L,ρ1,R>2\rho_{1,L},\,\rho_{1,R}>-2. Fix kk\in{\mathbb{N}} such that ρ=j=1kρj,R(κ24,κ22)\rho=\sum\limits_{j=1}^{k}\rho_{j,R}\in(\frac{\kappa}{2}-4,\frac{\kappa}{2}-2) and ϵ>0\epsilon>0. There exists p1>0p_{1}>0 depending only on κ,maxi,q|ρi,q|,ρ\kappa,\,\emph{max}_{i,q}|\rho_{i,q}|,\,\rho, and ϵ\epsilon such that if |x2,q|ϵ|x_{2,q}|\geq\epsilon for q{L,R},xk+1,Rxk,Rϵq\in\{L,R\},x_{k+1,R}\,-\,x_{k,R}\geq\epsilon, and xk,Rϵ1x_{k,R}\leq\epsilon^{-1} then the following is true. Suppose that γ\gamma is a simple curve starting from 0, terminating in [xk,R,xk+1,R][x_{k,R},x_{k+1,R}], and otherwise does not hit \partial{\mathcal{H}}. Let A(ϵ)A(\epsilon) be the ϵ\epsilon- neighborhood of γ([0,T])\gamma([0,T]) and let

σ1=inf{t0:η(t)(xk,R,xk+1,R)}andσ2=inf{t0:η(t)A(ϵ)}\sigma_{1}=\emph{inf}\{t\geq 0:\eta(t)\in(x_{k,R},x_{k+1,R})\}\quad\emph{and}\quad\sigma_{2}=\emph{inf}\{t\geq 0:\eta(t)\notin A(\epsilon)\}

. Then [σ1<σ2]p1{\mathbb{P}}[\sigma_{1}<\sigma_{2}]\geq p_{1}.

Intuitively, Lemma 2.1 tells us that an SLEκ(ρ¯L;ρ¯R)SLE_{\kappa}(\bar{\rho}_{L};\bar{\rho}_{R}) process has a positive chance to stay close to any fixed deterministic curve for a positive amount of time.

Proof.

This is lemma 2.5 in [11]. ∎

3 Proof of Theorem 1

Consider the left and right boundaries of the SLESLE curve η\eta, which are boundary-touching SLE16κ(ρ)SLE_{\frac{16}{\kappa}}(\rho) curves, with force points starting at 0. In fact, the left boundary of SLEκSLE_{\kappa} turns out to be SLE16/κ(16κ4;8κ2)SLE_{16/\kappa}(\frac{16}{\kappa}-4;\frac{8}{\kappa}-2) and by symmetry, the right boundary is SLE16/κ(8κ2;16κ4)SLE_{16/\kappa}(\frac{8}{\kappa}-2;\frac{16}{\kappa}-4). This can be deduced from Proposition 7.31 in [8]. These curves are shown in the figure below. The open region between the left and right boundaries has countably many connected components, which are separated by the intersection points of the left and right boundaries, i.e., the cut points of η\eta. These connected components have a total ordering, and come in four types:

  • Type 0: Neither the left nor the right boundary of the component intersects the real line.

  • Type 1: Only the right boundary intersects the real line.

  • Type 2: Only the left boundary intersects the real line.

  • Type 3: The left and right boundaries both intersect the real line.

Refer to caption
Figure 1: We view the complement of the SLESLE curve as the union of two boundary-touching SLEκ(ρ¯)SLE_{\kappa}(\bar{\rho}) processes. We observe ‘bubbles’ of four types, which we use in constructing the observable invariant.

Note that η\eta is a continuous curve that travels between the positive and negative real axes between any two consecutive components of type 3. This shows that the components of type 3 form a discrete set, to which we may assign a labeling by the integers - written as

(U1,U0,U1,U2,)(\dots U_{-1},U_{0},U_{1},U_{2},\dots)

uniquely, modulo index shift. For concreteness, we choose the indexing for the sequence so that U0U_{0} is the first type 3 bubble which has Euclidean diameter at least 1. We remark here that our construction relies on a few tail triviality arguments, and so we require the following:

Lemma 3.1.

Suppose t>0t>0 and let ata_{t} (resp. btb_{t}) be the last time before tt at which η\eta hits the left (resp. right) boundary. Then η|[0,t]\eta|_{[0,t]} determines the set of bubbles (i.e. connected components of the region between the left and right boundaries) which are formed before time min{at,bt}\min\{a_{t},b_{t}\} as well as their types.

Proof.

This follows trivially from the fact that η\eta cannot cross itself and η([min{at,bt},t])\eta([\min\{a_{t},b_{t}\},t]) disconnects all of the bubbles formed before time min{at,bt}\min\{a_{t},b_{t}\} from η(t).\eta(t).

Between pairs of consecutive type 3 bubbles, UiU_{i} and Ui+1U_{i+1}, we may either observe a type 1 or 2 bubble, or we may not. Let EiE_{i} be the event that there is a type 1 or type 2 bubble between UiU_{i} and Ui+1U_{i+1}, and define

X:=(𝟙E1,𝟙E0,𝟙E1,𝟙E2,)X:=(\dots\mathbbm{1}_{E_{-1}},\mathbbm{1}_{E_{0}},\mathbbm{1}_{E_{1}},\mathbbm{1}_{E_{2}},\dots)

the bi-infinite sequence of 0’s and 1’s consisting of the indicators of the EiE_{i}’s.

Lemma 3.2.

For any fixed deterministic bi-infinite sequence of 0’s and 11’s xx, we have [X=x]=0{\mathbb{P}}[X=x]=0.

Proof.

Consider a left-infinite sequence y=(y2,y1,y0)y=(\dots y_{-2},y_{-1},y_{0}). For kk\in\mathbb{N}, let AkA_{k} be the event that {Xk1,Xk=y}\{\dots X_{-k-1},X_{-k}=y\}. We wish to show that [A0]=0{\mathbb{P}}[A_{0}]=0. We will argue this by contradiction, but we first require a bit of setup. For r>0r\in{\mathbb{R}}_{>0}, nn\in{\mathbb{N}}, let Kr(n)K_{r}^{(n)} be the nnth smallest kk such that the Euclidean diameter of UkU_{k} is at least rr. Now, we claim that [AK1(n)]=0{\mathbb{P}}[A_{K_{1}^{(n)}}]=0 for all nn. We argue to the contrary, and so we assume that there exists some nn such that [AK1(n)]>0{\mathbb{P}}\left[A_{K_{1}^{(n)}}\right]>0. Note that by scale invariance, [AKr(n)]{\mathbb{P}}\left[A_{K_{r}^{(n)}}\right] is independent of rr, and so depends only on nn. Consider the event i=0miAK1m(n)\bigcap_{i=0}^{\infty}\bigcup_{m\geq i}A_{K_{\frac{1}{m}}^{(n)}}, which is a tail event for the Brownian motion that drives the SLESLE, for every choice of nn. To see this, note that Lemma 3.1 implies that for each tt, t\mathcal{F}_{t} determines AKr(n)A_{K_{r}^{(n)}} for each rr which is small enough so that the bubble UKr(n)U_{K_{r}^{(n)}} is formed before time min{at,bt}\min\{a_{t},b_{t}\}. Thus, by continuity from above, we note that

[i=0miAK1m(n)][AK1(n)]>0{\mathbb{P}}\left[\bigcap_{i=0}^{\infty}\bigcup_{m\geq i}A_{K_{\frac{1}{m}}^{(n)}}\right]\geq{\mathbb{P}}\left[A_{K_{1}^{(n)}}\right]>0

and so the Blumenthal 010-1 law implies that, a.s., there exists a sequence {rj}0\{r_{j}\}\rightarrow 0 such that the events AKrj(n)A_{K_{r_{j}}^{(n)}} occur for all jj. This implies that there exist infinitely many kk such that AkA_{k} occurs. Thus, it follows that a.s., \exists infinitely many kk such that

(Xk1,Xk)=y(\dots X_{-k-1},X_{-k})=y

forcing the sequence yy to be periodic. We claim that this implies that the sequence {Xk}\{X_{k}\} is periodic. Indeed, Let mm be the period of yy. Since there are arbitrarily large kk for which (Xk1,Xk)=y(...X_{-k-1},X_{-k})=y and yy is periodic, it follows that with probability tending to 11 as r0r\rightarrow 0, the sequence
(XKr(n)1,XKr(n))(...X_{-K_{r}^{(n)}-1},X_{-K_{r}^{(n)}}) is equal to (yj1,yj)(...y_{-j-1},y_{-j}) for some j=1,,mj=1,...,m. By scale invariance, the probability that this is the case for all values of rr is equal to 1. Thus, as rr\rightarrow\infty, we see that the entire sequence {Xk}\{X_{k}\} is equal to yy, shifted by some j=1,,mj=1,...,m. This means that if we observe (Xk1,Xk)(...X_{-k-1},X_{-k}) for some kk, we can determine the rest of the sequence {Xk}\{X_{k}\}, forcing this sequence to be itself periodic.

For t>0t>0, we have that by Lemma 3.1 t\mathcal{F}_{t} determines the sequence (Xl1,Xl)(\dots X_{-l-1},X_{-l}) for some ll, which by periodicity is enough to determine the sequence {Xk}\{X_{k}\}. Thus, by Lemma 3.1, t\mathcal{F}_{t} determines {Xk}\{X_{k}\} modulo an index shift for each t>0t>0, and hence the sequence {Xk}\{X_{k}\} is deterministic modulo an index shift. The goal now is to recursively apply Lemma 2.1 to arrive at a contradiction.

Proposition 3.3.

Let 𝒵\mathcal{Z} be a finite sequence of 0’s and 11’s which does not appear in yy, with |𝒵|=m|\mathcal{Z}|=m. Then it must hold that

[{X1,X2,,Xm}={𝒵1,𝒵2,,𝒵m}]>0.{\mathbb{P}}\left[\{X_{1},X_{2},\dots,X_{m}\}=\{\mathcal{Z}_{1},\mathcal{Z}_{2},\dots,\mathcal{Z}_{m}\}\right]>0.

Note that the existence of such a 𝒵\mathcal{Z} follows from the periodicity of yy. With this result, we can conclude that the sequence {Xk}\{X_{k}\} can contain any finite sequence of 0’s and 11’s with positive probability, and hence cannot be periodic and deterministic modulo index shift. We delay the proof of the proposition to state the following key lemma, which uses the fact that the outer boundaries of the curve are SLEκ(ρL;ρR)SLE_{\kappa}(\rho_{L};\rho_{R}) processes, and more specifically the right boundary, ηR\eta^{R}, conditioned on the left boundary, ηL\eta^{L}, has distribution of SLE16κ(8κ;16κ4)SLE_{\frac{16}{\kappa}}(-\frac{8}{\kappa};\frac{16}{\kappa}-4) (see Lemma 7.1 in [8]):

Refer to caption
Figure 2: We condition on the left boundary (pictured as the orange curve) and run the right boundary until we first form a type 3 bubble of diameter at least 1 (blue). At this time (denoted ηR(τ)\eta^{R}(\tau)), we have two options: either the right boundary hits [0,)[0,\infty) before hitting the left boundary again (green), thus forming a type 3 bubble, or it hits the left boundary first (red), forming a type 1 bubble before forming the next type 3 bubble. These events each occur with positive probability.
Lemma 3.4.

Let τ\tau be a stopping time for ηR\eta^{R} given ηL\eta^{L}, at which ηR\eta^{R} forms a type 3 bubble denoted UkτU_{k_{\tau}}. Let EkτE_{k_{\tau}} be the event that there is a type 1 or type 2 bubble between UkτU_{k_{\tau}} and Ukτ+1U_{k_{\tau}+1}, as defined previously. Then,

0<[Ekτ|ηL,η|[0,τ]R]<1.0<{\mathbb{P}}\left[E_{k_{\tau}}\,\Big{|}\,\eta^{L},\eta^{R}_{|_{[0,\tau]}}\right]<1.
Proof.

With some setup, this is a straightforward application of Lemma 2.1. Indeed, let zτ:=ηR(τ)z_{\tau}:=\eta^{R}(\tau) and define CzτC_{z_{\tau}} to be the connected component of ηL\eta^{L}\setminus{\mathbb{R}} containing zτz_{\tau}. Set

s1:=inf{t>τ:ηR[0,)},s2:=inf{t>τ:ηRηL(Czτ(,0]).}s^{1}:=\inf\{t>\tau:\eta^{R}\cap[0,\infty)\neq\emptyset\},\qquad s^{2}:=\inf\{t>\tau:\eta^{R}\cap\eta^{L}\setminus(C_{z_{\tau}}\cup(-\infty,0])\neq\emptyset.\}

By Lemma 2.1, we have that

[s2>s1|ηL,η|[0,τ]R]>0;[s2s1|ηL,η|[0,τ]R]>0{\mathbb{P}}\left[s^{2}>s^{1}\Big{|}\,\eta^{L},\eta^{R}_{|_{[0,\tau]}}\right]>0;\qquad{\mathbb{P}}\left[s^{2}\leq s^{1}\Big{|}\,\eta^{L},\eta^{R}_{|_{[0,\tau]}}\right]>0

where the second inequality follows from symmetry considerations. Indeed, we can simply apply Lemma 2.1 to the curve ηR\eta^{R}, under the conditional law given ηL\eta^{L}. In this case, an interval on the left boundary corresponds to a segment of ηL\eta^{L}. Note that these probabilities are strictly less than 11 as they are both positive and complementary. With this, and appealing to the setting of Fig. 2, we have that ηR[τ,)\eta^{R}[\tau,\infty), conditioned on ηL,η|[0,τ]R\eta^{L},\eta^{R}_{|_{[0,\tau]}}, will either first intersect the left boundary and form a type 1 bubble before forming another type 3 bubble, or it will intersect [0,)[0,\infty) before hitting the left boundary again, forming another type 3 bubble. In particular, the event that a type 1 bubble is formed after UkτU_{k_{\tau}} occurs with probability strictly between 0 and 11 as desired. ∎

Proof of Proposition 3.3.

We define a sequence of stopping times as follows: For a given bubble UiU_{i}, let τi\tau_{i} be the corresponding time at which UiU_{i} is formed. By our choice of indexing of the type 3 bubbles, we have that

τ0:=1st time we form a type 3 bubble of Euclidean diameter at least 1 \tau_{0}:=1\textsuperscript{st}\text{ time we form a type 3 bubble of Euclidean diameter at least 1 }
τ1:=1st time after τ0 we form a type 3 bubble\tau_{1}:=1\textsuperscript{st}\text{ time after }\tau_{0}\text{ we form a type 3 bubble}
\vdots
τm:=1st time after τm1 we form a type 3 bubble.\tau_{m}:=1\textsuperscript{st}\text{ time after }\tau_{m-1}\text{ we form a type 3 bubble.}

Note that EkτiE_{k_{\tau_{i}}}is measurable with respect to ηL\eta^{L} and ηR|[0,τi+1]\eta^{R}|_{[0,\tau_{i+1}]}, and for each i{1,2,,m}i\in\{1,2,\dots,m\}, we have that by Lemma 3.4,

0<[Ekτi|ηL,η|[0,τi]R]<1.0<{\mathbb{P}}\left[E_{k_{\tau_{i}}}\,\Big{|}\,\eta^{L},\eta^{R}_{|_{[0,\tau_{i}]}}\right]<1.

Thus, it follows that

[Xi=𝒵i|X1=𝒵1,,Xi1=𝒵i1]>0.{\mathbb{P}}[X_{i}=\mathcal{Z}_{i}|X_{1}=\mathcal{Z}_{1},...,X_{i-1}=\mathcal{Z}_{i-1}]>0.

To finish the proof, we note that since {Xj=Zj}\{X_{j}=Z_{j}\} is determined by ηL\eta^{L} and ηR|[0,τi]\eta^{R}|_{[0,\tau_{i}]} for i<ji<j, so

[X1=𝒵1,,Xi=𝒵i]=𝔼[[Xi=𝒵i|ηL,ηR|[0,τi]]𝟙X1=𝒵1,,Xi1=𝒵i1].{\mathbb{P}}[X_{1}=\mathcal{Z}_{1},...,X_{i}=\mathcal{Z}_{i}]=\mathbb{E}\left[{\mathbb{P}}[X_{i}=\mathcal{Z}_{i}\,|\,\eta^{L},\eta^{R}|_{[0,\tau_{i}]}]\mathbbm{1}_{{X_{1}=\mathcal{Z}_{1},...,X_{i-1}=\mathcal{Z}_{i-1}}}\right].

The probability within the expectation on the right hand side is always positive, and so inducting on ii (and setting i=mi=m as a final step) yields the desired result. ∎

By Proposition 3.3, we see that {Xk}\{X_{k}\} can contain any finite sequence of 0’s and 11’s not contained in yy, implying that {Xk}\{X_{k}\} cannot be deterministic modulo index shift. This is a contradiction. Thus, [AK1(n)]=0{\mathbb{P}}[A_{K_{1}^{(n)}}]=0 for every nn.

Thus, by scale invariance we see that [AKr(n)]=0{\mathbb{P}}[A_{K_{r}^{(n)}}]=0 for every rr and nn. Note that every kk is equal to Kr(n)K_{r}^{(n)} for some rational rr and some nn. Indeed, every kkth bubble has some positive diameter, and there are at most finitely many bubbles before it of larger diameter. Thus, we can set nn to be the number of bubbles before the kkth bubble with diameter exceeding that of the kkth bubble, and simply let rr be any rational number slightly smaller than this diameter. From this, it follows that

[k such that Ak occurs ][nr>0AKr(n)]=0.{\mathbb{P}}[\exists\,k\text{ such that }A_{k}\text{ occurs }]\leq{\mathbb{P}}\left[\bigcup_{n\in{\mathbb{N}}}\bigcup_{r\in{\mathbb{Q}}_{>0}}A_{K_{r}}^{(n)}\right]=0.

In particular, we have that [A0]=0{\mathbb{P}}[A_{0}]=0.

Proof of Theorem.

Now let η1\eta^{1} and η2\eta^{2} be two independent SLESLE’s. In order for η1\eta^{1}\cup{\mathbb{R}} and η2\eta^{2}\cup{\mathbb{R}} to be homeomorphic via a homeomorphism that takes {\mathbb{R}} to {\mathbb{R}}, it must be the case that the corresponding bi-infinite sequences X1X^{1} and X2X^{2} differ by at most an index shift. Indeed, any homeomorphism has to preserve the bi-infinite sequence of connected components lying between the left and right boundaries of the curve, as well as the types of these components. Thus, by the above argument, the probability that X1X^{1} is equal to any of the countably many possible index shifted versions of X2X^{2} is zero. Hence the probability that η1\eta^{1}\cup{\mathbb{R}} and η2\eta^{2}\cup{\mathbb{R}} are homeomorphic, via a homeomorphism that takes {\mathbb{R}} to {\mathbb{R}}, is 0. ∎

4 Proof of Theorem 2

Here, we require a more subtle argument that relies on a less obvious statistic of observation. In this section, we fix κ8\kappa\geq 8 and recall SLEκSLE_{\kappa}, in this instance, is plane filling. Let η\eta be an instance of SLEκSLE_{\kappa} in ¯\overline{{\mathbb{H}}}. We are interested in the successive crossing times (about the origin) of the curve η\eta, i.e., the times at which η\eta hits the real line again, just after having hit it on the opposite side of the origin. We look at one such crossing time, and consider the part of the SLESLE within this, observing the times it goes back and forth between the boundaries of the crossing excursion. As pictured below in Fig. 3, these left and right crossings (within the curve) define a sequence of marked points {Xk}\{X_{k}\} along the boundary, which accumulate only at the tip of the curve. Via the corresponding Loewner map gtηg_{t}^{\eta}, we may conformally map this configuration as shown in Fig. 3, so that the tip goes to 0, and we abtain a sequence of marked points along the left boundary. Notice these marked points are determined by the past, so we can condition on their locations, and the future will still be an SLESLE by the Markov property.

A bit more care is needed in defining these quantities. Let τ(t)\tau(t) be the last time before tt such that η(t)\eta(t)\in{\mathbb{R}}. Define the sets

T:={t:η(τ(t))<0}T+:={t:η(τ(t))>0}T_{-}:=\{t:\eta(\tau(t))<0\}\qquad T_{+}:=\{t:\eta(\tau(t))>0\}

and set 𝒮=T¯T¯+\mathcal{S}=\bar{T}_{-}\cap\bar{T}_{+}. Notice that 𝒮\mathcal{S} is a discrete set since η\eta is continuous, and so it cannot cross back and forth between (,0)(-\infty,0) and (0,)(0,\infty) infinitely many times during any compact time interval contained in (0,)(0,\infty). Thus, we may index the elements of 𝒮\mathcal{S} as a countable sequence of well defined crossing times {τj}\{\tau_{j}\}.

Notice that these are not necessarily stopping times (which poses a problem in applying the strong Markov property), but this can be addressed by adopting some notation from the previous section as follows. Let ηj:=η|[τj1,τj]\eta_{j}:=\eta|_{[\tau_{j-1},\tau_{j}]}, which is the jjth left-right crossing around 0 that we observe. For r>0r>0, let Jr(n)J_{r}^{(n)} be the nnth smallest jj for which the Euclidean diameter of ηj\eta_{j} is at least rr. It is not difficult to see that the set of times {τJr(n)}\{\tau_{J_{r}^{(n)}}\} is indeed a set of stopping times. To see this, let t>0t>0. If one sees η|[0,t]\eta|_{[0,t]}, then one can determine the set {τj:jt}\{\tau_{j}:j\leq t\} (but not necessarily its indexing). This follows from the definition of the times {τj}\{\tau_{j}\} as the intersection points of TT_{-} and T+T_{+}, as shown previously. Hence η|[0,t]\eta|_{[0,t]} determines the set of excursions {ηj:τjt}\{\eta_{j}:\tau_{j}\leq t\}. We have τJr(n)t\tau_{J_{r}^{(n)}}\leq t if and only if this set of excursions includes at least nn elements which have Euclidean diameter at least rr. Hence {τJr(n)t}\{\tau_{J_{r}^{(n)}}\leq t\} is determined by η|[0,t]\eta|_{[0,t]}, which holds for any choice of tt.

Refer to caption
Figure 3: The top picture illustrates a single left-right crossing around 0, with x0=η(τJ)x_{0}=\eta({\tau_{J}}) and the corresponding triangulation in red, determined by the (past) piece of the curve making boundary crossings. We conformally map this down, and we consider intervals [Xk+1,Xk][X_{k+1},X_{k}] in which the tips of future triangles, obtained by left right crossings about 0, may lie. Some intervals may have multiple, while some may have none.

We fix some rr and some nn, and set J:=Jr(n)J:=J_{r}^{(n)}. Within ηJ\eta_{J}, i.e. between the outer boundaries of this crossing, we can keep track of the times at which {ηt;t<τJ}\{\eta_{t};t<\tau_{J}\} sequentially hits these boundaries. More precisely, we let LJL_{J} be the outer boundary of η[0,τJ]\eta[0,\tau_{J}]. We define our sequence of crossing times inductively as follows:

σJ,1:=min{t>τJ1:ηtLJ}\sigma_{J,1}:=\min\{t>\tau_{J-1}:\eta_{t}\cap L_{J}\neq\emptyset\}
σ~J,1:=min{t>σJ,1:ηtLJ1}\tilde{\sigma}_{J,1}:=\min\{t>\sigma_{J,1}:\eta_{t}\cap L_{J-1}\neq\emptyset\}
\vdots
σ~J,k:=min{t>σJ,k:ηtLJ1}\tilde{\sigma}_{J,k}:=\min\{t>\sigma_{J,k}:\eta_{t}\cap L_{J-1}\neq\emptyset\}
σJ,k+1:=min{t>σ~J,k:ηtLJ}\sigma_{J,k+1}:=\min\{t>\tilde{\sigma}_{J,k}:\eta_{t}\cap L_{J}\neq\emptyset\}

and so on. The sequences {σJ,k}k1\{\sigma_{J,k}\}_{k\geq 1} and {σ~J,k}k1\{\tilde{\sigma}_{J,k}\}_{k\geq 1} define two discrete sets of times that our curve successively hits the outer boundaries LJL_{J} and LJ1L_{J-1} respectively. We assume without loss of generality that the JJth excursion goes from left to right. By considering only the outer boundary LJL_{J} (as a priori τJ\tau_{J} is a well-defined stopping time), we can construct a sequence of marked points {XJ,k}k1\{X_{J,k}\}_{k\geq 1} along the negative real axis, via the (shifted) Loewner map which sends η(τJ)\eta(\tau_{J}) to 0. That is to say, XJ,k:=gτJ(η(σJ,k))UτJX_{J,k}:=g_{\tau_{J}}(\eta(\sigma_{J,k}))-U_{\tau_{J}}. As we are considering a fixed JJ, we may write XJ,k:=XkX_{J,k}:=X_{k} for ease.

We let Nk=#{crossing endpoints of the future curve which lie in the interval [Xk+1,Xk]}.N_{k}=\#\left\{\text{crossing endpoints of the future curve which lie in the interval }[X_{k+1},X_{k}]\right\}. In other words, we are looking at gτJ(η|[τJ,))UτJg_{\tau_{J}}(\eta|_{[\tau_{J},\infty)})-U_{\tau_{J}} as it does these left-right crossings, conditioned on the past, and for each interval we are keeping track of how many endpoints it contains. We wish to show that for every sequence of deterministic integers {nk}k\{n_{k}\}_{k\in{\mathbb{N}}}, we have that

[Nk=nk;k]=0.{\mathbb{P}}[N_{k}=n_{k};\,\forall\,k]=0. (4)

It suffices to show that there are arbitrarily large kk such that [Nk=nk]{\mathbb{P}}[N_{k}=n_{k}] is bounded away from 11. Indeed, the event {Nk=nk for infinitely many k}\{N_{k}=n_{k}\text{ for infinitely many }k\} is a tail event for the Brownian motion driving the SLESLE, and the Blumenthal 010-1 law implies that this has probability 0 or 11. Thus, being bounded away from 11 guarantees that we have (4). We do this in cases as follows:

Case 1: Assume \exists arbitrarily large kk such that nk0n_{k}\neq 0. We claim that q>0\exists\,q>0 such that

[Nk=n]1qn1.{\mathbb{P}}[N_{k}=n]\leq 1-q\quad\forall\,n\geq 1.

To see this, we consider the segment of the curve, just after the (n1)th(n-1)^{th} crossing is completed. Let 𝒯n={nthtime we have a crossing in the interval[Xk,Xk+1]}\mathcal{T}_{n}=\{n^{th}\,\text{time we have a crossing in the interval}\,[X_{k},X_{k+1}]\,\}. Thus 𝒯n\mathcal{T}_{n} is a stopping time, and conditioned on what we have seen up until this time, the future of the curve is still SLESLE.

Refer to caption
Figure 4: We stop the SLESLE after it has made its n1n-1th crossing in the interval shown. Under the map g~\tilde{g}, we send the tip of the curve to the origin, and analyze the likelihood of either observing two more crossings in the red interval of length aa, or no more crossings, in which case the interval is swallowed.

The goal is to have an upper bound on the probability that there are exactly nn crossings, and we do so by comparing the harmonic measure (from \infty) of the interval [Xk+1,η(𝒯n1)][X_{k+1},\eta(\mathcal{T}_{n-1})], to that of the outer boundary of the curve η[0,𝒯n1]\eta[0,\mathcal{T}_{n-1}](and more precisely, this is the harmonic measure from \infty in η[0,𝒯n1]{\mathbb{H}}\setminus\eta[0,\mathcal{T}_{n-1}]) . These quantities are denoted aa and bb respectively, as shown in Figure 3.

The proof relies on the following intuitive argument which we formalize later: If aa is larger than bb, then with positive probability we observe 22 further crossings, hence n+1n+1 total crossings. If aa is smaller than bb, then, with positive probability, we expect the interval [Xk+1,η(𝒯n1)][X_{k+1},\eta(\mathcal{T}_{n-1})] to be covered before we observe the next crossing. In other words, there is always a positive chance that we observe either n1n-1 crossings or n+1n+1 crossings, and so

[Nkn]>0.{\mathbb{P}}[N_{k}\neq n]>0.
Proposition 4.1.

Let η\eta be an SLEκSLE_{\kappa} from 0 to \infty in {\mathbb{H}} with κ>4\kappa>4. For marked points a<0<ca<0<c along the real line, let Ea,cE_{a,c} be the event that the chordal SLEκSLE_{\kappa} trace visits [c,)[c,\infty) before (,a](-\infty,a]. Then

[Ea,c]=F(aca)where F(x)=1Zκ0xduu4κ(1u)4κ{\mathbb{P}}\left[E_{a,c}\right]=F\left(\dfrac{-a}{c-a}\right)\qquad\text{where }F(x)=\dfrac{1}{Z_{\kappa}}\displaystyle\int_{0}^{x}\dfrac{du}{u^{\frac{4}{\kappa}}(1-u)^{\frac{4}{\kappa}}}

and ZκZ_{\kappa} is chosen so that F(1)=1F(1)=1.

Proof.

This is Theorem 10 in [1]. ∎

Remark 2.

It is possible to get an estimate which is weaker than Theorem 3 above, but which is still sufficient for our purposes, via the following elementary argument. For xx\in{\mathbb{R}}, let tx:=inf[t0:η(t)=x].t_{x}:=\inf[t\geq 0:\eta(t)=x]. If we let P(n)=[tn<t1]P(n)={\mathbb{P}}[t_{n}<t_{-1}], a bit of thought shows that

P(n)P(n1)[1P(n)]P(n)\geq P(n-1)[1-P(n)]

which thus implies that

P(n)P(n1)1+P(n1).P(n)\geq\dfrac{P(n-1)}{1+P(n-1)}.

The equality case can be realized as P(n)=1n+1P(n)=\dfrac{1}{n+1}, the details of which we omit. By considering f(x)=xx+1f(x)=\dfrac{x}{x+1}, which is increasing on 0{\mathbb{R}}_{\geq 0}, we find that

P(n)f(P(n1))f(2)(P(n2)f(n)(12)=1n+1P(n)\geq f(P(n-1))\geq f^{(2)}(P(n-2)\dots\geq f^{(n)}\left(\frac{1}{2}\right)=\dfrac{1}{n+1}

which gives a rough (yet easy to compute) estimate. Note, for our purposes, we only require a positive probability.

We return to the notation introduced in Fig. 4, and we consider the the behavior of the SLESLE curve given the relative quantities aa and bb. In particular, we require the following two key lemmas to prove the original claim:

Lemma 4.2.

If aba\leq b, it holds with conditional probability at least 12\frac{1}{2}, given η|[0,𝒯n1]\eta|_{[0,\mathcal{T}_{n-1}]}, that η|[𝒯n1,)\eta|_{[\mathcal{T}_{n-1},\infty)} hits Xk+1X_{k+1} before [b,).[b,\infty).

Proof.

Notice that by symmetry, there is a positive chance that we disconnect [Xk+1,η(𝒯n1)][X_{k+1},\eta(\mathcal{T}_{n-1})] before hitting bb. Indeed, this follows from the fact that [t1<t1]=12.{\mathbb{P}}[t_{-1}<t_{1}]=\frac{1}{2}.

Lemma 4.3.

There exists a deterministic κ\kappa-dependent constant c>0c>0 such that if a>ba>b, it holds with conditional probability at least cc given η|[0,𝒯n1]\eta|_{[0,\mathcal{T}_{n-1}]} that η|[𝒯n1,)\eta|_{[\mathcal{T}_{n-1},\infty)} crosses between (,0)(-\infty,0) and (0,)(0,\infty) at least twice before hitting Xk+1.X_{k+1}.

Proof.

If a>ba>b, then we can apply the estimate given in Proposition 4.1 via a two step process. We retain the notation from Remark 2, and define tat_{-a} and tbt_{b} as discussed, after having mapped η([0,𝒯n1])\eta([0,\mathcal{T}_{n-1}]) to the real line via the map g~\tilde{g}. Note that Proposition 4.1 implies that p>0\exists\,p>0 such that [tb<ta/2]p{\mathbb{P}}[t_{b}<t_{-a/2}]\geq p. In fact, we have assumed a>ba>b, so pp in this instance can be thought of as a universal bound. We condition on this event occurring, and we look at the harmonic measure of the outer boundary curve η~\tilde{\eta} of this most recent crossing. Note that hmη~(,η~)\text{hm}_{{\mathbb{H}}\setminus\tilde{\eta}}(\infty,\tilde{\eta}) is bounded above by the harmonic measure of the outer boundary at the time we hit a2-\dfrac{a}{2}. This follows from the fact that the harmonic measure can only increase, as we observe more of the curve. Moreover, the law of this harmonic measure, divided by aa, is independent of aa by scale invariance, and is almost surely finite. This implies that C=C(p)>0\exists\,C=C(p)>0 such that

[hmη~(,η~)Ca]1p2{\mathbb{P}}\left[\text{hm}_{{\mathbb{H}}\setminus\tilde{\eta}}(\infty,\tilde{\eta})\leq Ca\right]\geq 1-\dfrac{p}{2}

from which it follows that

[hmη~(,η~)Ca,tb<ta/2]p2.{\mathbb{P}}\left[\text{hm}_{{\mathbb{H}}\setminus\tilde{\eta}}(\infty,\tilde{\eta})\leq Ca,\,t_{b}<t_{-a/2}\right]\geq\dfrac{p}{2}.

This bound guarantees a positive probability that, after we have observed the first crossing, the harmonic measure of the outer boundary is not too large. Now we condition on this event, and we apply Proposition 4.1 to the quantities CaCa and a2\dfrac{a}{2}. In particular, This yields a positive κ\kappa-dependent constant lower bound for the probability that η|[𝒯n1,)\eta|_{[\mathcal{T}_{n-1},\infty)} has at least two crossings before hitting Xk+1X_{k+1}. ∎

Case 2: nk=0n_{k}=0 for all but finitely many kk.

This condition implies that the SLESLE travels a positive distance of time without any left-right crossings, which happens with probability 0. This shows that for any fixed deterministic sequence {nk}k\{n_{k}\}_{k\in{\mathbb{N}}} with only finitely many non-zero elements, we have that

[{Nk}={nk}]=0.{\mathbb{P}}[\{N_{k}\}=\{n_{k}\}]=0.
Refer to caption
Figure 5: We observe two instances of SLESLE, η1\eta^{1} and η2\eta^{2}, stopped after the m1m_{1}th and m2m_{2}th crossings respectively. Any homeomorphism between the two should send one tip to the other, and retain the structure of the future crossings (i.e., preserve the corresponding sequences {Nkj}\{N_{k}^{j}\}).
Proof of theorem.

Consider two instances of SLEκSLE_{\kappa} in {\mathbb{H}}, η1\eta^{1} and η2\eta^{2}, with corresponding sequences of points {Xm1,k1}k\{X_{m_{1},k}^{1}\}_{k\in{\mathbb{N}}} and {Xm2,k2}k\{X_{m_{2},k}^{2}\}_{k\in{\mathbb{N}}} respectively, for fixed indicies m1,m2𝒮m_{1},m_{2}\in\mathcal{S}, corresponding to the m1m_{1}th crossing of η1\eta^{1} and m2m_{2}th crossing of η2\eta^{2} respectively. Here, we indicate objects associated with ηj\eta^{j} for j{1,2}j\in\{1,2\} by a superscript jj. Note that by construction, m1=Jr1(n1),1m_{1}=J_{r_{1}}^{(n_{1}),1} and m2=Jr2(n2),2m_{2}=J_{r_{2}}^{(n_{2}),2} for some n1,n2n_{1},n_{2} and (rational) r1,r2r_{1},r_{2}. Each sequence of points {Xkj}k\{X_{k}^{j}\}_{k\in{\mathbb{N}}} generates a sequence {Nkj}k\{N_{k}^{j}\}_{k\in{\mathbb{N}}} for j{1,2}j\in\{1,2\} and so by the independence of η1\eta^{1} and η2\eta^{2}, as well as (4), we have that for any choice of m1,m2m_{1},m_{2} and number ll

[Nk1=Nk+l2;k]=[Nk1=Nk+l2;k|η2]=0.{\mathbb{P}}\left[N_{k}^{1}=N_{k+l}^{2};\,\,\forall\,k\right]={\mathbb{P}}\left[N_{k}^{1}=N_{k+l}^{2};\,\,\forall\,k\,|\,\eta^{2}\right]=0.

This implies that

[l s.t Nk1=Nk+l2;k]=0{\mathbb{P}}\left[\exists\,l\text{ s.t }N_{k}^{1}=N_{k+l}^{2};\,\forall\,k\right]=0 (5)

as there are countably many possible choices of ll, meaning we can apply this very argument for each fixed choice of ll, and apply the union bound.

Observe that a homeomorphism from ¯\overline{{\mathbb{H}}} to itself taking η1\eta^{1} to η2\eta^{2}, modulo time parametrization, must preserve the number of left right crossings of the ‘future’ curves, which correspond to the sequences {Nkj}\{N_{k}^{j}\}, and it must take η1(τm11)\eta^{1}(\tau_{m_{1}}^{1}) to η2(τm22)\eta^{2}(\tau_{m_{2}}^{2}) for some m2m_{2}. In particular, as in the setting of Figure 4, for any fixed m1m_{1} and m2m_{2} there is no homeomorphism which takes η1\eta^{1} to η2\eta^{2} and η1(τm11)\eta^{1}(\tau_{m_{1}}^{1}) to η2(τm22)\eta^{2}(\tau_{m_{2}}^{2}) by (5). As the set 𝒮\mathcal{S} of crossing times is discrete, this holds for any choice of indices m1m_{1}and m2m_{2} , where there are only countably many choices. Thus, it must hold that,

[ a homeomorphism Φ:¯¯ taking η1 to η2]=0.{\mathbb{P}}\left[\exists\,\text{ a homeomorphism }\Phi:\overline{{\mathbb{H}}}\to\overline{{\mathbb{H}}}\text{ taking }\eta^{1}\text{ to }\eta^{2}\right]=0.

Acknowledgements

I would like to thank Prof. Ewain Gwynne for suggesting this problem, and for answering my many questions about the material presented here, as well as SLESLE in general. I would also like to thank Prof. Gregory Lawler for suggesting readings and proof techniques to supplement this paper.

References

  • [1] Vincent Beffara. Schramm-Loewner Evolution and other conformally invariant objects. Probability and Statistical Physics in Two and More Dimensions, 15:1–48, 2012.
  • [2] N. Berestycki and J.R. Norris. Lectures on Schramm-Loewner Evolution, 2016. Available at http://www.statslab.cam.ac.uk/~james/Lectures/.
  • [3] Julien Dubédat. Duality of Schramm-Loewner evolutions. Ann. Sci. Éc. Norm. Supér. (4), 42(5):697–724, 2009.
  • [4] Bertrand Duplantier. Conformally invariant fractals and potential theory. Physical Review Letters, 84(7):1363–1367, Feb 2000.
  • [5] Bertrand Duplantier. Higher conformal multifractality. Journal of Statistical Physics, 110(3–6):691–738, 2003.
  • [6] Gregory Lawler, Oded Schramm, and Wendelin Werner. Conformal restriction: the chordal case. J. Amer. Math. Soc., 16(4):917–955 (electronic), 2003.
  • [7] Gregory F. Lawler. Conformally invariant processes in the plane, volume 114 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2005.
  • [8] Jason Miller and Scott Sheffield. Imaginary geometry I: interacting SLEs. Probab. Theory Related Fields, 164(3-4):553–705, 2016.
  • [9] Jason Miller and Scott Sheffield. Imaginary geometry IV: interior rays, whole-plane reversibility, and space-filling trees. Probab. Theory Related Fields, 169(3-4):729–869, 2017.
  • [10] Jason Miller, Scott Sheffield, and Wendelin Werner. Non-simple SLE curves are not determined by their range. J. Eur. Math. Soc. (JEMS), 22(3):669–716, 2020.
  • [11] Jason Miller and Hao Wu. Intersections of SLE Paths: the double and cut point dimension of SLE. Probab. Theory Related Fields, 167(1-2):45–105, 2017.
  • [12] Edwin Moise. Geometric Topology in Dimensions 2 and 3. Springer, New York, NY, 1977.
  • [13] Steffen Rohde and Oded Schramm. Basic properties of SLE. Ann. of Math. (2), 161(2):883–924, 2005.
  • [14] Oded Schramm. Scaling limits of loop-erased random walks and uniform spanning trees. Israel J. Math., 118:221–288, 2000.
  • [15] Wendelin Werner. Random planar curves and Schramm-Loewner evolutions. In Lectures on probability theory and statistics, volume 1840 of Lecture Notes in Math., pages 107–195. Springer, Berlin, 2004.
  • [16] Dapeng Zhan. Duality of chordal SLE. Invent. Math., 174(2):309–353, 2008.
  • [17] Dapeng Zhan. Duality of chordal SLE, II. Ann. Inst. Henri Poincaré Probab. Stat., 46(3):740–759, 2010.