The topology of is random for
Abstract
We study the topology of curves for . More precisely, we show that, a.s., there is no homeomorphism , taking the range of one independent curve to another for . Furthermore, we extend the result to by showing that there is no homeomorphism taking one curve to another, when viewed as curves modulo parametrization.
1 Introduction
1.1 Initial Overview
The Schramm-Loewner Evolutions (), introduced by Oded Schramm [14], describes a family of probability distributions, parameterized by 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 exhibits a universality in its definition.
is the random growth of a set , as described through a conformal map on the complement of . This map is the solution of the Loewner differential equation driven by a Brownian motion, whose ‘speed’ is determined by a single parameter . Rhodes and Schramm in [13] showed that for , a.s. there is a (unique) continuous path such that for each the set is the union of and the bounded connected components of . This has also been shown for , but was dealt with separately. We call the path the SLE trace or curve. It was shown as well in [13] that a.s. We will need the following facts about the curve which are proven in [13]:
-
•
If , then is simple with .
-
•
If , then has double points and intersects .
-
•
If , the curve is space-filling.
There are three (well studied) variants of : chordal , which connects two boundary points (prime ends) in a given domain, radial , which connects a boundary point to an interior point, and whole-plane , which connects two points on the Riemann sphere. In this paper, we will focus on chordal , but we expect that our results can be generalized to other cases. One can also see [7, 2, 15] for some expository work on which go into details beyond the scope of this paper.
Most works on 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 : namely, is the topology of the curve deterministic? Said differently, if we have two independent chordal SLEκ curves and (viewed as curves modulo time parametrization), does there a.s. exist a homeomorphism taking to ?
Since is a simple curve for , the answer to the above question is clearly affirmative in this case. For , 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 or 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 and are compact, non-empty, totally disconnected subsets of without isolated points (e.g., Cantor-type sets), then there is a homeomorphism from to which takes to [12]. The main results of this paper show that the topology of SLEκ is random for . For , 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 curve for is not determined by its range. Both this paper and [10] answer a seemingly simple question about whose answer is much less obvious than one might initially expect.
1.2 Summary of results
Theorem 1.
The topology of chordal is not deterministic in the following sense: Fix and consider two instances of , and in . Then a.s. there is no homeomorphism on taking the range of to the range of .
We consider the left and right boundaries of an curve (which are boundary-touching curves). These curves form ‘bubbles’ in (which we characterize explicitly in a later section) which we use as the primary observable to prove Theorem 1.
The result also holds for . The proof is similar, though a bit more work is needed in the setup.
Theorem 2.
The topology of chordal is not deterministic for in the following sense: Consider two curves, and in . Then a.s. there is no homeomorphism such that 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 in this instance is plane filling.
As a natural extension, one can think about the behavior of these curves for varying .
Conjecture 1.1.
Let be distinct. Let (resp. ) be a chordal SLE (resp. SLE) in . Almost surely, there is no homeomorphism such that viewed as curves modulo time parametrization. If one of or is in , a.s., there is no such homeomorphism which takes the range of to the range of .
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 -dependent.
2 Preliminaries
Here we discuss a few basics as well as how one defines the more general processes. Recall that we write If is a bounded closed subset of such that is simply connected, then we call a hull in w.r.t. . For such , there is a unique that maps conformally onto such that as . The quantity is known as the half plane capacity of , and is denoted by . It can be shown that . The map is said to satisfy the hydrodynamic normalization at infinity. For a real interval , let denote the real-valued continuous functions on I. Suppose for some . The chordal Loewner equation driven by is as follows:
(1) |
For , let and be the chordal Loewner hulls and maps, respectively, driven by . Suppose that for every ,
exists, and , is a continuous curve. Then for every is the complement of the unbounded component of . We call the chordal Loewner trace driven by . In general, however, such a curve may not exist depending on the choice of driving function.
An in from to is defined by the random family of conformal maps obtained by solving the Loewner ODE driven by Brownian motion. In particular, we let , where is a standard Brownian motion. An connecting boundary points and of an arbitrary simply connected Jordan domain can be constructed as the image of an on under a conformal transformation sending to and to . 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].
, which is often written as , is the stochastic process one obtains by solving (1) with a modification on the driving process , which we now discuss. It is a natural generalization of in which one keeps track of additional marked points which are called force points. Fix and . We associate with each for a weight . An process with force points is the measure on continuously growing compact hulls generated by the Loewner chain with replaced by the solution to the system of SDEs given by
(2) |
(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 that where for (we call this time the continuation threshold). In particular,if for all for , then (1) has a unique solution for all times . This even holds when one or both of the are zero. Note that in [8] it is proved that 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 , there is also significant interest in the hulls that are generated by the curves. Duplantier conjectured in [4, 5] the duality between and , which says that the boundary of an hull behaves like an curve, for . Many versions of this duality have also been shown in [16, 17, 3, 8, 9].
Lemma 4.9 in [8] asserts that, for , the outer boundary of an curve is an process. The curves are realized as flow lines of the Gaussian free field (i.e curves coupled with the Gaussian free field in ), 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 , some of which we rely on to prove the main results. We state one such fact as follows:
Lemma 2.1.
Fix . Suppose that is an process in from to with force points located at with and (possibly by taking for ). Assume that . Fix such that and . There exists depending only on , and such that if for , and then the following is true. Suppose that is a simple curve starting from , terminating in , and otherwise does not hit . Let be the - neighborhood of and let
. Then .
Intuitively, Lemma 2.1 tells us that an 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 curve , which are boundary-touching curves, with force points starting at . In fact, the left boundary of turns out to be and by symmetry, the right boundary is . 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 . 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.

Note that 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
uniquely, modulo index shift. For concreteness, we choose the indexing for the sequence so that 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 and let (resp. ) be the last time before at which hits the left (resp. right) boundary. Then determines the set of bubbles (i.e. connected components of the region between the left and right boundaries) which are formed before time as well as their types.
Proof.
This follows trivially from the fact that cannot cross itself and disconnects all of the bubbles formed before time from ∎
Between pairs of consecutive type 3 bubbles, and , we may either observe a type 1 or 2 bubble, or we may not. Let be the event that there is a type 1 or type 2 bubble between and , and define
the bi-infinite sequence of 0’s and 1’s consisting of the indicators of the ’s.
Lemma 3.2.
For any fixed deterministic bi-infinite sequence of ’s and ’s , we have .
Proof.
Consider a left-infinite sequence . For , let be the event that . We wish to show that . We will argue this by contradiction, but we first require a bit of setup. For , , let be the th smallest such that the Euclidean diameter of is at least . Now, we claim that for all . We argue to the contrary, and so we assume that there exists some such that . Note that by scale invariance, is independent of , and so depends only on . Consider the event , which is a tail event for the Brownian motion that drives the , for every choice of . To see this, note that Lemma 3.1 implies that for each , determines for each which is small enough so that the bubble is formed before time . Thus, by continuity from above, we note that
and so the Blumenthal law implies that, a.s., there exists a sequence such that the events occur for all . This implies that there exist infinitely many such that occurs. Thus, it follows that a.s., infinitely many such that
forcing the sequence to be periodic. We claim that this implies that the sequence is periodic.
Indeed, Let be the period of . Since there are arbitrarily large for which and is periodic, it follows that with probability tending to as , the sequence
is equal to for some .
By scale invariance, the probability that this is the case for all values of is equal to 1. Thus, as , we see that the entire sequence is equal to , shifted by some . This means that if we observe for some , we can determine the rest of the sequence , forcing this sequence to be itself periodic.
For , we have that by Lemma 3.1 determines the sequence for some , which by periodicity is enough to determine the sequence . Thus, by Lemma 3.1, determines modulo an index shift for each , and hence the sequence 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 be a finite sequence of ’s and ’s which does not appear in , with . Then it must hold that
Note that the existence of such a follows from the periodicity of . With this result, we can conclude that the sequence can contain any finite sequence of ’s and ’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 processes, and more specifically the right boundary, , conditioned on the left boundary, , has distribution of (see Lemma 7.1 in [8]):

Lemma 3.4.
Let be a stopping time for given , at which forms a type 3 bubble denoted . Let be the event that there is a type 1 or type 2 bubble between and , as defined previously. Then,
Proof.
With some setup, this is a straightforward application of Lemma 2.1. Indeed, let and define to be the connected component of containing . Set
By Lemma 2.1, we have that
where the second inequality follows from symmetry considerations. Indeed, we can simply apply Lemma 2.1 to the curve , under the conditional law given . In this case, an interval on the left boundary corresponds to a segment of . Note that these probabilities are strictly less than as they are both positive and complementary. With this, and appealing to the setting of Fig. 2, we have that , conditioned on , will either first intersect the left boundary and form a type 1 bubble before forming another type 3 bubble, or it will intersect before hitting the left boundary again, forming another type 3 bubble. In particular, the event that a type 1 bubble is formed after occurs with probability strictly between and as desired. ∎
Proof of Proposition 3.3.
We define a sequence of stopping times as follows: For a given bubble , let be the corresponding time at which is formed. By our choice of indexing of the type 3 bubbles, we have that
Note that is measurable with respect to and , and for each , we have that by Lemma 3.4,
Thus, it follows that
To finish the proof, we note that since is determined by and for , so
The probability within the expectation on the right hand side is always positive, and so inducting on (and setting as a final step) yields the desired result. ∎
By Proposition 3.3, we see that can contain any finite sequence of ’s and ’s not contained in , implying that cannot be deterministic modulo index shift. This is a contradiction. Thus, for every .
Thus, by scale invariance we see that for every and . Note that every is equal to for some rational and some . Indeed, every th bubble has some positive diameter, and there are at most finitely many bubbles before it of larger diameter. Thus, we can set to be the number of bubbles before the th bubble with diameter exceeding that of the th bubble, and simply let be any rational number slightly smaller than this diameter. From this, it follows that
In particular, we have that .
∎
Proof of Theorem.
Now let and be two independent ’s. In order for and to be homeomorphic via a homeomorphism that takes to , it must be the case that the corresponding bi-infinite sequences and 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 is equal to any of the countably many possible index shifted versions of is zero. Hence the probability that and are homeomorphic, via a homeomorphism that takes to , 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 and recall , in this instance, is plane filling. Let be an instance of in . We are interested in the successive crossing times (about the origin) of the curve , i.e., the times at which 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 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 along the boundary, which accumulate only at the tip of the curve. Via the corresponding Loewner map , we may conformally map this configuration as shown in Fig. 3, so that the tip goes to , 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 by the Markov property.
A bit more care is needed in defining these quantities. Let be the last time before such that . Define the sets
and set . Notice that is a discrete set since is continuous, and so it cannot cross back and forth between and infinitely many times during any compact time interval contained in . Thus, we may index the elements of as a countable sequence of well defined crossing times .
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 , which is the th left-right crossing around that we observe. For , let be the th smallest for which the Euclidean diameter of is at least . It is not difficult to see that the set of times is indeed a set of stopping times. To see this, let . If one sees , then one can determine the set (but not necessarily its indexing). This follows from the definition of the times as the intersection points of and , as shown previously. Hence determines the set of excursions . We have if and only if this set of excursions includes at least elements which have Euclidean diameter at least . Hence is determined by , which holds for any choice of .

We fix some and some , and set . Within , i.e. between the outer boundaries of this crossing, we can keep track of the times at which sequentially hits these boundaries. More precisely, we let be the outer boundary of . We define our sequence of crossing times inductively as follows:
and so on. The sequences and define two discrete sets of times that our curve successively hits the outer boundaries and respectively. We assume without loss of generality that the th excursion goes from left to right. By considering only the outer boundary (as a priori is a well-defined stopping time), we can construct a sequence of marked points along the negative real axis, via the (shifted) Loewner map which sends to . That is to say, . As we are considering a fixed , we may write for ease.
We let In other words, we are looking at 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 , we have that
(4) |
It suffices to show that there are arbitrarily large such that is bounded away from . Indeed, the event is a tail event for the Brownian motion driving the , and the Blumenthal law implies that this has probability or . Thus, being bounded away from guarantees that we have (4). We do this in cases as follows:
Case 1: Assume arbitrarily large such that . We claim that such that
To see this, we consider the segment of the curve, just after the crossing is completed. Let . Thus is a stopping time, and conditioned on what we have seen up until this time, the future of the curve is still .

The goal is to have an upper bound on the probability that there are exactly crossings, and we do so by comparing the harmonic measure (from ) of the interval , to that of the outer boundary of the curve (and more precisely, this is the harmonic measure from in ) . These quantities are denoted and respectively, as shown in Figure 3.
The proof relies on the following intuitive argument which we formalize later: If is larger than , then with positive probability we observe further crossings, hence total crossings. If is smaller than , then, with positive probability, we expect the interval to be covered before we observe the next crossing. In other words, there is always a positive chance that we observe either crossings or crossings, and so
Proposition 4.1.
Let be an from to in with . For marked points along the real line, let be the event that the chordal trace visits before . Then
and is chosen so that .
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 , let If we let , a bit of thought shows that
which thus implies that
The equality case can be realized as , the details of which we omit. By considering , which is increasing on , we find that
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 curve given the relative quantities and . In particular, we require the following two key lemmas to prove the original claim:
Lemma 4.2.
If , it holds with conditional probability at least , given , that hits before
Proof.
Notice that by symmetry, there is a positive chance that we disconnect before hitting . Indeed, this follows from the fact that ∎
Lemma 4.3.
There exists a deterministic -dependent constant such that if , it holds with conditional probability at least given that crosses between and at least twice before hitting
Proof.
If , 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 and as discussed, after having mapped to the real line via the map . Note that Proposition 4.1 implies that such that . In fact, we have assumed , so 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 of this most recent crossing. Note that is bounded above by the harmonic measure of the outer boundary at the time we hit . 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 , is independent of by scale invariance, and is almost surely finite. This implies that such that
from which it follows that
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 and . In particular, This yields a positive -dependent constant lower bound for the probability that has at least two crossings before hitting . ∎
Case 2: for all but finitely many .
This condition implies that the travels a positive distance of time without any left-right crossings, which happens with probability . This shows that for any fixed deterministic sequence with only finitely many non-zero elements, we have that

Proof of theorem.
Consider two instances of in , and , with corresponding sequences of points and respectively, for fixed indicies , corresponding to the th crossing of and th crossing of respectively. Here, we indicate objects associated with for by a superscript . Note that by construction, and for some and (rational) . Each sequence of points generates a sequence for and so by the independence of and , as well as (4), we have that for any choice of and number
This implies that
(5) |
as there are countably many possible choices of , meaning we can apply this very argument for each fixed choice of , and apply the union bound.
Observe that a homeomorphism from to itself taking to , modulo time parametrization, must preserve the number of left right crossings of the ‘future’ curves, which correspond to the sequences , and it must take to for some . In particular, as in the setting of Figure 4, for any fixed and there is no homeomorphism which takes to and to by (5). As the set of crossing times is discrete, this holds for any choice of indices and , where there are only countably many choices. Thus, it must hold that,
∎
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 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.