The Full Scaling Limit of Two-Dimensional Critical Percolation
Federico Camia Research partially supported by a Marie Curie Intra-European Fellowship under contract MEIF-CT-2003-500740.
E-mail: camia@eurandom.tue.nl
EURANDOM, P.O. Box 513, 5600 MB Eindhoven, The Netherlands
Charles M. Newman Research partially supported by the U.S. NSF under grant DMS-01-04278.
E-mail: newman@courant.nyu.edu
Courant Inst. of Mathematical Sciences, New York University, New York, NY 10012, USA
Abstract
We use
paths to construct a process of continuum nonsimple loops in the plane and prove that this process coincides with the full continuum scaling limit of 2D critical site percolation on the triangular lattice – that is, the scaling limit of the set of all interfaces between different clusters. Some properties of the loop process, including conformal invariance, are also proved. In the main body of the paper these results are proved while assuming, as argued by Schramm and Smirnov, that the percolation exploration path converges in distribution to the trace of chordal
. Then, in a lengthy appendix, a detailed proof is provided for this convergence to
, which itself relies on Smirnov's result that crossing probabilities converge to Cardy's formula.
Keywords: continuum scaling limit, percolation, SLE, critical behavior, triangular lattice, conformal invariance.
AMS 2000 Subject Classification: 82B27, 60K35, 82B43, 60D05, 30C35.
1 Introduction and Motivation
In the theory of critical phenomena it is usually assumed that a physical system near a continuous phase transition is characterized by a single length scale (the “correlation length”) in terms of which all other lengths should be measured. When combined with the experimental observation that the correlation length diverges at the phase transition, this simple but strong assumption, known as the scaling hypothesis, leads to the belief that at criticality the system has no characteristic length, and is therefore invariant under scale transformations. This suggests that all thermodynamic functions at criticality are homogeneous functions, and predicts the appearance of power laws. It also means that it should be possible to rescale a critical system appropriately and obtain a continuum model (the “continuum scaling limit”) which may have more symmetries and be easier to study than the original discrete model defined on a lattice.
Indeed, thanks to the work of Polyakov [23] and others [4, 5] , it was understood by physicists since the early seventies that critical statistical mechanical models should possess continuum scaling limits with a global conformal invariance that goes beyond simple scale invariance, as long as the discrete models have “enough” rotation invariance.
This property gives important information, enabling the determination of twoand three-point functions at criticality, when they are nonvanishing. Because the conformal group is in general a finite dimensional Lie group, the resulting constraints are limited in number; however, the situation becomes particularly interesting in two dimensions, since there every analytic function
defines a conformal transformation, at least at points where
. As a consequence, the conformal group in two dimensions is infinite-dimensional.
After this observation was made, a large number of critical problems in two dimensions were analyzed using conformal methods, which were applied, among others, to Ising and Potts models, Brownian motion, Self-Avoiding Walk (SAW), percolation, and Diffusion Limited Aggregation (DLA). The large body of knowledge and techniques that resulted, starting with the work of Belavin, Polyakov and Zamolodchikov [4, 5] in the early eighties, goes under the name of Conformal Field Theory (CFT). In two dimensions, one of the main goals of CFT and its most important application to statistical mechanics is a complete classification of all universality classes via irreducible representations of the infinite-dimensional Virasoro algebra.
Partly because of the success of CFT, work in recent years on critical phenomena seemed to slow down somewhat, probably due to the feeling that most of the leading problems had been resolved. Nonetheless, however powerful and successful it may be, CFT has some limitations and leaves various open problems. First of all, the theory deals primarily with correlation functions of local (or quasi-local) operators, and is therefore not always the best tool to investigate other quantities. Secondly, given some critical lattice model, there is no way, within the theory itself, of deciding to which CFT it corresponds. A third limitation, of a different nature, is due to the fact that the methods of CFT, although very powerful, are generally speaking not completely rigorous from a mathematical point of view.
In a somewhat surprising twist, the most recent developments in the area of two-dimensional critical phenomena have emerged in the mathematics literature and have followed a newdirection, which has provided new tools and a way of coping with at least some of the limitations of CFT. The new approach may even provide a reinterpretation of CFT, and seems to be complementary to the traditional one in the sense that questions that are difficult to pose and/or answer within CFT are easy and natural in this new approach and vice versa.
The main tool of this radically new approach is the Stochastic Loewner Evolution (
), or Schramm Loewner Evolution, as it is also known, introduced by Schramm [28] .
The new approach, which is probabilistic in nature, focuses directly on non-local structures that characterize a given system, such as cluster boundaries in Ising, Potts and percolation models, or loops in the
model. At criticality, these non-local objects become, in the continuum limit, random curves whose distributions can be uniquely identified thanks to their conformal invariance and a certain “Markovian” property. There is a one-parameter family of
s, indexed by a positive real number
, and they appear to be the only possible candidates for the scaling limits of interfaces of two-dimensional critical systems that are believed to be conformally invariant.
In particular, substantial progress has been made in recent years, thanks to
, in understanding the fractal and conformally invariant nature of (the scaling limit of ) large percolation clusters, which has attracted much attention and is of interest both for intrinsic reasons, given the many applications of percolation, and as a paradigm for the behavior of other systems. The work of Schramm [28] and Smirnov [30] has identified the scaling limit of a certain percolation interface with
, providing, along with the work of Lawler-Schramm-Werner [19, 20] and Smirnov-Werner [34] , a confirmation of many results in the physics literature, as well as some new results.
However,
describes a single interface, which can be obtained by imposing special boundary conditions, and is not in itself sufficient to immediately describe the full scaling limit of the system. In fact, not only the nature and properties, but the very existence of the full scaling limit remained an open question. This is true of all models, such as Ising and Potts models, that are represented in terms of clusters. Werner [36] considered this problem in the context of
for values of
between
and
. For percolation (corresponding to
), the same problem was addressed in [7] , where
was used to construct a random process of continuous loops in the plane, which was identified with the full scaling limit of critical two-dimensional percolation, but without detailed proofs.
In this paper, we complete the analysis of [7] , making rigorous the connection between the construction given there and the full scaling limit of percolation, and we prove some properties of the full scaling limit, the Continuum Nonsimple Loop process, including (one version of ) conformal invariance. We do this in two parts. First, we give proofs in which we assume the validity of what we will call statement (S) (see Section 5 ), which is a specific version of the results of Schramm and of Smirnov [28, 30, 31, 32, 33] concerning convergence of percolation exploration paths to
(see the discussion towards the end of Section 4.1 ). Since no detailed proof of statement (S) (or indeed, any version of convergence to
) has been available, in Appendix A we give a proof based only on that part of Smirnov's results about the convergence of crossing probabilities to Cardy's formula [30] (see Theorem 4 in Appendix A ). We note that statement (S) is restricted toJordan domains while no such restriction is indicated in [30, 31] .
The rest of the paper is organized as follows. In Section 2 , we give necessary definitions and introduce
. Section 3 is devoted to the construction of the Continuum Nonsimple Loop process. In Section 4 , we introduce the discrete model and a discrete construction analogous to the continuum one presented in Section 3 . Most of the main results of this paper are stated in Section 5 , while Section 6 contains the proofs of those results, which use (S). The long Appendix A contains the proof of statement (S) (it is a consequence of Corollary A.1 there) and the short Appendix B contains convergence results for sequences of conformal maps which are used throughout the paper.
We remark that although our proof in Appendix A of convergence of exploration paths to
roughly follows Smirnov's outline [30, 31] , based on his proof [30, 31] of convergence of crossing probabilities to Cardy's formula and on the Markovian properties of hulls and tips, there are at least two technically significant modifications. The first is that we use a different sequence of stopping times to obtain a Markov chain approximation to
, which results in a different geometry for the approximation (see Remark A.2 ). The second is that the control of “close encounters” by the exploration path to the domain boundary is not handled by general results for “three-arms” events at the boundary of a half-plane, but rather by an argument based on continuity of crossing probabilities with respect to domain boundaries (see Lemmas A.2 , A.3 , A.4 and A.5 ). Moreover, we cannot use directly Smirnov's result on convergence of crossing probabilities (see Theorem 4 ), but need an extended version which is given in Theorem 6 of Appendix A .
We conclude by noting that the convergence results of Appendix A are sufficient not only for our purposes of obtaining the full scaling limit, but also for obtaining the critical exponents (see [34] ).
2 Preliminary Definitions
We will find it convenient to identify the real plane
and the complex plane
. We will also refer to the Riemann sphere
and the open upper half-plane
(and its closure
), where chordal
will be defined (see Section 2.3 ).
will denote the open unit disc
.
A domain
of the complex plane
is a nonempty, connected, open subset of
; a simply connected domain
is said to be a Jordan domain if its (topological) boundary
is a Jordan curve (i.e., a simple continuous loop).
We will make repeated use of Riemann's mapping theorem, which states that if
is any simply connected domain other than the entire plane
and
, then there is a unique conformal map
of
onto
such that
and
.
2.1 Compactification of
When taking the scaling limit
one can focus on fixed finite regions,
, or consider the whole
at once. The second option avoids dealing with boundary conditions, but requires an appropriate choice of metric.
A convenient way of dealing with the whole
is to replace the Euclidean metric with a distance function
defined on
by
|
(1)
|
where the infimum is over all smooth curves
joining
with
, parametrized by arclength
, and where
denotes the Euclidean norm. This metric is equivalent to the Euclidean metric in bounded regions, but it has the advantage of making
precompact.
Adding a single point at infinity yields the compact space
which is isometric, via stereographic projection, to the two-dimensional sphere.
2.2 The Space of Curves
In dealing with the scaling limit we use the approach of Aizenman-Burchard [2] . Denote by
the complete separable metric space of continuous curves in the closure
of the disc
of radius
with the metric ( 2 ) defined below. Curves are regarded as equivalence classes of continuous functions from the unit interval to
, modulo monotonic reparametrizations.
will represent a particular curve and
a parametrization of
;
will represent a set of curves (more precisely, a closed subset of
).
will denote the uniform metric on curves, defined by
|
(2)
|
where the infimum is over all choices of parametrizations of
and
from the interval
. The distance between two closed sets of curves is defined by the induced Hausdorff metric as follows:
|
(3)
|
The space
of closed subsets of
(i.e., collections of curves in
) with the metric ( 3 ) is also a complete separable metric space. We denote by
its Borel
-algebra.
For each fixed
, the random curves that we consider are polygonal paths on the edges of the hexagonal lattice
, dual to the triangular lattice
. A superscript
is added to indicate that the curves correspond to a model with a “short distance cutoff ” of magnitude
.
We will also consider the complete separable metric space
of continuous curves in
with the distance
|
(4)
|
where the infimum is again over all choices of parametrizations of
and
from the interval
. The distance between two closed sets of curves is again defined by the induced Hausdorff metric as follows:
|
(5)
|
The space
of closed sets of
(i.e., collections of curves in
) with the metric ( 5 ) is also a complete separable metric space. We denote by
its Borel
-algebra.
When we talk about convergence in distribution of random curves, we always mean with respect to the uniform metric ( 2 ), while when we deal with closed collections of curves, we always refer to the metric ( 3 ) or ( 5 ).
Remark 2.1.
In this paper, the space
of closed sets of
is generally used for collections of exploration paths and cluster boundary loops and their scaling limits,
paths and continuum nonsimple loops. There is one place however, in the statements and proofs of Lemmas A.2 , A.4 and A.5 , where we also apply
in essentially the original setting of Aizenman and Burchard [
1,
2]
, i.e., for collections of blue and yellow simple
-paths (see Section 4 for precise definitions) and their scaling limits. The slight modification needed to keep track of both the paths and their colors is easily managed.
2.3 Chordal
in the Upper Half-Plane
The Stochastic Loewner Evolution (
) was introduced by Schramm [28] as a tool for studying the scaling limit of two-dimensional discrete (defined on a lattice) probabilistic models whose scaling limits are expected to be conformally invariant. In this section we define the chordal version of
; for more on the subject, the interested reader can consult the original paper [28] as well as the fine reviews by Lawler [17] , Kager and Nienhuis [14] , and Werner [37] , and Lawler's book [18] .
Let
denote the upper half-plane. For a given continuous real function
with
, define, for each
, the function
as the solution to the ODE
|
(6)
|
with
. This is well defined as long as
, i.e., for all
, where
|
(7)
|
Let
and let
be the unbounded component of
; it can be shown that
is bounded and that
is a conformal map from
onto
. For each
, it is possible to write
as
|
(8)
|
when
. The family
is called the Loewner chain associated to the driving function
.
Definition 2.1.
Chordal
is the Loewner chain
that is obtained when the driving function
is
times a standard real-valued Brownian motion
with
.
For all
, chordal
is almost surely generated by a continuous random curve
in the sense that, for all
,
is the unbounded connected component of
;
is called the trace of chordal
.
2.4 Chordal
in an Arbitrary Simply Connected Domain
Let
(
) be a simply connected domain whose boundary is a continuous curve. By Riemann's mapping theorem, there are (many) conformal maps from the upper half-plane
onto
. In particular, given two distinct points
(or more accurately, two distinct prime ends), there exists a conformal map
from
onto
such that
and
. In fact, the choice of the points
and
on the boundary of
only characterizes
up to a multiplicative factor, since
would also do.
Suppose that
is a chordal
in
as defined above; we define chordal
in
from
to
as the image of the Loewner chain
under
. It is possible to show, using scaling properties of
, that the law of
is unchanged, up to a linear time-change, if we replace
by
. This makes it natural to consider
as a process from
to
in
, ignoring the role of
.
We are interested in the case
, for which
is generated by a continuous, nonsimple, non-self-crossing curve
with Hausdorff dimension
. We will denote by
the image of
under
and call it the trace of chordal
in
from
to
;
is a continuous nonsimple curve inside
from
to
, and it can be given a parametrization
such that
and
, so that we are in the metric framework described in Section 2.2 . It will be convenient to think of
as an oriented path, with orientation from
to
.
3 Construction of the Continuum Nonsimple Loops
3.1 Construction of a Single Loop
As a preview to the full construction, we explain how to construct a single loop using two
paths inside a domain
whose boundary is assumed to have a given orientation (clockwise or counterclockwise). This is done in three steps (see Figure 1 ), of which the first consists in choosing two points
and
on the boundary
of
and “running” a chordal
,
, from
to
inside
. As explained in Section 2.4 , we consider
as an oriented path, with orientation from
to
. The set
is a countable union of its connected components, which are open and simply connected. If
is a deterministic point in
, then with probability one,
is not touched by
[26] and so it belongs to a unique domain in
that we denote
.
The elements of
can be conveniently thought of in terms of how a point
in the interior of the component was first “trapped” at some time
by
, perhaps together with either
or
(the portions of the boundary
from
to
counterclockwise or clockwise respectively): (1) those components whose boundary contains a segment of
between two successive visits at
and
to
(where here and below
), (2) the analogous components with
replaced by the other part of the boundary
, (3) those components formed when
with
winding about
in a counterclockwise direction between
and
, and finally (4) the analogous clockwise components.
We give to the boundary of a domain of type 3 or 4 the orientation induced by how the curve
winds around the points inside that domain. For a domain
of type 1 or 2 which is produced by an “excursion”
from
to
, the part of the boundary that corresponds to the inner perimeter of the excursion
(i.e., the perimeter of
seen from
) is oriented according to the direction of
, i.e., from
to
.
If we assume that
is oriented from
to
clockwise, then the boundaries of domains of type 2 have a well defined orientation, while the boundaries of domains of type 1 do not, since they are composed of two parts which are both oriented from the beginning to the end of the excursion that produced the domain.
Now, let
be a domain of type 1 and let
and
be respectively the starting and ending point of the excursion that generated
. The second step to construct a loop is to run a chordal
,
, inside
from
to
; the third and final step consists in pasting together
and
.
Running
inside
from
to
partitions
into new domains. Notice that if we assign an orientation to the boundaries of these domains according to the same rules used above, all of those boundaries have a well defined orientation, so that the construction of loops just presented can be iterated inside each one of these domains (as well as inside each of the domains of type 2, 3 and 4 generated by
in the first step). This will be done in the next section.
3.2 The Full Construction Inside The Unit Disc
In this section we define the Continuum Nonsimple Loop process inside the unit disc
via an inductive procedure. Later, in order to define the continuum nonsimple loops in the whole plane, the unit disc will be replaced by a growing sequence of large discs,
, with
(see Theorem 2 ). The basic ingredient in the algorithmic construction, given in the previous section, consists of a chordal
path
between two points
and
of the boundary
of a given simply connected domain
.
We will organize the inductive procedure in steps, each one corresponding to one
inside a certain domain generated by the previous steps. To do that, we need to order the domains present at the end of each step, so as to choose the one to use in the next step.
For this purpose, we introduce a deterministic countable set of points
that are dense in
and are endowed with a deterministic order (here and below by deterministic we mean that they are assigned before the beginning of the construction and are independent of the
's).
The first step consists of an
path,
, inside
from
to
, which produces many domains that are the connected components of the set
. These domains can be priority-ordered according to the maximal
or
coordinate distances between points on their boundaries and using the rank of the points in
(contained in the domains) to break ties, as follows. For a domain
, let
be the maximal
or
-distance between points on its boundary, whichever is greater. Domains with larger
have higher priority, and if two domains have the same
, the one containing the highest ranking point of
from those two domains has higher priority. The priority order of domains of course changes as the construction proceeds and new domains are formed.
The second step of the construction consists of an
path,
, that is produced in the domain with highest priority (after the first step). Since all the domains that are produced in the construction are Jordan domains, as explained in the discussion following Corollary 5.1 , for all steps we can use the definition of chordal
given in Section 2.4 .
As a result of the construction, the
paths are naturally ordered:
. It will be shown (see especially the proof of Theorem 1 below) that every domain that is formed during the construction is eventually used (this is in fact one important requirement in deciding how to order the domains and therefore how to organize the construction).
So far we have not explained how to choose the starting and ending points of the
paths on the boundaries of the domains. In order to do this, we give an orientation to the boundaries of the domains produced by the construction according to the rules explained in Section 3.1 . We call monochromatic a boundary which gets, as a consequence of those rules, a well defined (clockwise or counterclockwise) orientation; the choice of this term will be clarified when we discuss the lattice version of the loop construction below.
We will generally take our initial domain
(or
) to have a monochromatic boundary (either clockwise or counterclockwise orientation).
It is easy to see by induction that the boundaries that are not monochromatic are composed of two “pieces” joined at two special points (call them A and B, as in theexample of Section 3.1 ), such that one piece is a portion of the boundary of a previous domain, and the other is the inner perimeter of an excursion (see again Section 3.1 ). Both pieces are oriented in the same direction, say from A to B (see Figure 1 ).
For a domain whose boundary is not monochromatic, we make the “natural” choice of starting and ending points, corresponding to the end and beginning of the excursion that produced the domain (the points B and A respectively, in the example above). As explained in Section 3.1 , when such a domain is used with this choice of points on the boundary, a loop is produced, together with other domains, whose boundaries are all monochromatic.
For a domain whose boundary is monochromatic, and therefore has a well defined orientation, there are various procedures which would yield the “correct” distribution for the resulting Continuum Nonsimple Loop process; one possibility is as follows.
Given a domain
,
and
are chosen so that, of all pairs
of points in
, they maximize
if
, or else they maximize
. If the choice is not unique, to restrict the number of pairs one looks at those pairs, among the ones already obtained, that maximize the other of
. Notice that this leaves at most two pairs of points; if that's the case, the pair that contains the point with minimal real (and, if necessary, imaginary) part is chosen. The iterative procedure produces a loop every time a domain whose boundary is not monochromatic is used. Our basic loop process consists of the collection of all loops generated by this inductive procedure (i.e., the limiting object obtained from the construction by letting the number of steps
), to which we add a “trivial” loop for each
in
, so that the collection of loops is closed in the appropriate sense [2] . The Continuum Nonsimple Loop process in the whole plane is introduced in Theorem 2 , Section 5 . There, a “trivial” loop for each
has to be added to make the space of loops closed.
4 Lattices and Paths
We will denote by
the two-dimensional triangular lattice, whose sites we think of as the elementary cells of a regular hexagonal lattice
embedded in the plane as in Figure 2 . A sequence
of sites of
such that
and
are neighbors in
for all
and
whenever
will be called a
-path and denoted by
.
If the first and last sites of the path are neighbors in
, the path will be called a
-loop.
We say that a finite subset
of
is simply connected if both
and
are connected (by the edges of
). For a simply connected set
of hexagons, we denote by
its external site boundary, or s-boundary (i.e., the set of hexagons that do not belong to
but are adjacent to hexagons in
), and by
the topological boundary of
when
is considered as a domain of
. We will call a bounded, simply connected subset
of
a Jordan set if its s-boundary
is a
-loop.
For a Jordan set
, a vertex
that belongs to
can be either of two types, according to whether the edge incident on
that is not in
belongs to a hexagon in
or not. We call a vertex of the second type an e-vertex (e for “external” or “exposed”).
Given a Jordan set
and two e-vertices
in
, we denote by
the portion of
traversed counterclockwise from
to
, and call it the right boundary; the remaining part of the boundary is denote by
and is called the left boundary. Analogously, the portion of
of
whose hexagons are adjacent to
is called the right s-boundary and the remaining part the left s-boundary.
A percolation configuration
on
is an assignment of
(equivalently, yellow) or
(blue) to each site of
. For a domain
of the plane, the restriction to the subset
of
of the percolation configuration
is denoted by
.
On the space of configurations
, we consider the usual product topology and denote by
the uniform measure, corresponding to Bernoulli percolation with equal density of yellow (minus) and blue (plus) hexagons, which is critical percolation in the case of the triangular lattice.
A (percolation) cluster is a maximal, connected, monochromatic subset of
; we will distinguish between blue (plus) and yellow (minus) clusters. The boundary of a cluster
is the set of edges of
that surround the cluster (i.e., its Peierls contour); it coincides with the topological boundary of
considered as a domain of
. The set of all boundaries is a collection of “nested” simple loops along the edges of
.
Given a percolation configuration
, we associate an arrow to each edge of
belonging to the boundary of a cluster in such a way that the hexagon to the right of the edge with respect to the direction of the arrow is blue (plus). The set of all boundaries then becomes a collection of nested, oriented, simple loops. A boundary path (or b-path)
is a sequence
of distinct edges of
belonging to the boundary of a cluster and such that
and
meet at a vertex of
for all
. To each b-path, we can associate a direction according to the direction of the edges in the path. Given a b-path
, we denote by
(respectively,
) the set of blue (resp., yellow) hexagons (i.e., sites of
) adjacent to
; we also let
.
4.1 The Percolation Exploration Process and Path
For a Jordan set
and two e-vertices
in
, imagine coloring blue all the hexagons in
and yellow all those in
. Then, for any percolation configuration
inside
, there is a unique b-path
from
to
which separates the blue cluster adjacent to
from the yellow cluster adjacent to
. We call
a percolation exploration path (see Figure 3 ).
An exploration path
can be decomposed into left excursions
, i.e., maximal b-subpaths of
that do not use edges of the left boundary
. Successive left excursions are separated by portions of
that contain only edges of the left boundary
.
Analogously,
can be decomposed into right excursions, i.e., maximal b-subpaths of
that do not use edges of the right boundary
. Successive right excursions are separated by portions of
that contain only edges of the right boundary
.
Notice that the exploration path
only depends on the percolation configuration
inside
and the positions of the e-vertices
and
; in particular, it does not depend on the color of the hexagons in
, since it is defined by imposing fictitious
boundary conditions on
. To see this more clearly, we next show how to construct the percolation exploration path dynamically, via the percolation exploration process defined below.
Given a Jordan set
and two e-vertices
in
, assign to
a counterclockwise orientation (i.e., from
to
) and to
a clockwise orientation. Call
the edge incident on
that does not belong to
and orient it in the direction of
; this is the “starting edge” of an exploration procedure that will produce an oriented path inside
along the edges of
, together with two nonsimple monochromatic paths on
. From
, the process moves along the edges of hexagons in
according to the rules below. At each step there are two possible edges (left or right edge with respect to the current direction of exploration) to choose from, both belonging to the same hexagon
contained in
or
.
-
∙
If
belongs to
and has not been previously “explored,” its color is determined by flipping a fair coin and then the edge to the left (with respect to the direction in which the exploration is moving) is chosen if
is blue (plus), or the edge to the right is chosen if
is yellow (minus).
-
∙
If
belongs to
and has been previously explored, the color already assigned to it is used to choose an edge according to the rule above.
-
∙
If
belongs to the right external boundary
, the left edge is chosen.
-
∙
If
belongs to the left external boundary
, the right edge is chosen.
-
∙
The exploration process stops when it reaches
.
We can assign an arrow to each edge in the path in such a way that the hexagon to the right of the edge with respect to the arrow is blue; for edges in
, we assign the arrows according to the direction assigned to the boundary. In this way, we get an oriented path,whose shape and orientation depend solely on the color of the hexagons explored during the construction of the path.
When we present the discrete construction, we will encounter Jordan sets
with two e-vertices
assigned in some way to be discussed later. Such domains will have either monochromatic (plus or minus) boundaries or
boundary conditions, corresponding to having both
and
monochromatic, but of different colors.
As explained, the exploration path
does not depend on the color of
, but the interpretation of
does. For domains with
boundary conditions, the exploration path represents the interface between the yellow cluster containing the yellow portion of the s-boundary of
and the blue cluster containing its blue portion.
For domains with monochromatic blue (resp., yellow) boundary conditions, the exploration path represents portions of the boundaries of yellow (resp., blue) clusters touching
and adjacent to blue (resp., yellow) hexagons that are the starting point of a blue (resp., yellow) path (possibly an empty path) that reaches
, pasted together using portions of
.
In order to study the continuum scaling limit of an exploration path, we introduce the following definitions.
Definition 4.1.
Given a bounded, simply connected domain
of the plane, we denote by
the largest Jordan set of hexagons of the scaled hexagonal lattice
that is contained in
, and call it the
-approximation of
.
It is clear that if
is a Jordan domain, then as
,
converges to
in the metric ( 2 ).
Definition 4.2.
Let
be a bounded domain of the plane and
its
-approximation. For
, choose the pair
of e-vertices in
closest to, respectively,
and
(if there are two such vertices closest to
, we choose, say, the first one encountered going clockwise along
, and analogously for
). Given a percolation configuration
, we define the exploration path
.
For a fixed
, the measure
on percolation configurations
induces a measure
on exploration paths
. In the continuum scaling limit,
, one is interested in the weak convergence of
to a measure
supported on continuous curves, with respect to the uniform metric ( 2 ) on continuous curves.
One of the main tools in this paper is the result on convergence to
announced by Smirnov [30] (see also [31] ), whose detailed proof is to appear [32] : The distribution of
converges, as
, to that of the trace of chordal
inside
from
to
, with respect to the uniform metric ( 2 ) on continuous curves.
Actually, we will rather use a slightly stronger conclusion, given as statement (S) at the beginning of Section 5 below, a version of which, according to [34] (see p. 734 there), and [33] , will be contained in [32] . This stronger statement is that the convergence of the percolation process to
takes place locally uniformly with respect to the shape of the domain
and the positions of the starting and ending points
and
on its boundary
.
We will use this version of convergence to
to identify the Continuum Nonsimple Loop process with the scaling limit of all critical percolation clusters. Statement (S) is a direct consequence of Corollary A.1 , which is proved in Appendix A . Although the convergence statements in Corollary A.1 and in (S) are stronger than those in [30, 31] , we note that they are restricted to Jordan domains, a restriction not present in [30, 31] .
Before concluding this section, we give one more definition. Consider the exploration path
and the set
. The set
is the union of its connected components (in the lattice sense), which are simply connected. If the domain
is large and the e-vertices
are not too close to each other, then with high probability the exploration process inside
will make large excursions into
, so that
will have more than one component. Given a point
contained in
, we will denote by
the domain corresponding to the unique element of
that contains
(notice that for a deterministic
,
is well defined with high probability for
small, i.e., when
and
).
4.2 Discrete Loop Construction
Next, we show how to construct, by twice using the exploration process described in Section 4.1 , a loop
along the edges of
corresponding to the external boundary of a monochromatic cluster contained in a large, simply connected, Jordan set
with monochromatic blue (say) boundary conditions (see Figures 4 and 5 ).
Consider the exploration path
and the sets
and
(see Figure 4 ).
The set
is the union of its connected components (in the lattice sense), which are simply connected. If the domain
is large and the e-vertices
are chosen not too close to each other, with large probability the exploration process inside
will make large excursions into
, so that
will have many components.
There are four types of components which may be usefully thought of in terms of their external site boundaries: (1) those components whose site boundary contains both sites in
and
, (2) the analogous components with
replaced by
and
by
, (3) those components whose site boundary only contains sites in
, and finally (4) the analogous components with
replaced by
.
Notice that the components of type 1 are the only ones with
boundary conditions, while all other components have monochromatic s-boundaries. For a given component
of type 1, we can identify the two edges that separate the yellow and blue portions of its s-boundary. The vertices
and
of
where those two edges intersect
are e-vertices and are chosen to be the starting and ending points of the exploration path
inside
.
If
are respectively the ending and starting points of the left excursion
of
that “created”
, by pasting together
and
with the help of the edges of
contained between
and
and between
and
, we get a loop
which corresponds to the boundary of a yellow cluster adjacent to
(see Figure 5 ). Notice that the path
in general splits
into various other domains, all of which have monochromatic boundary conditions.
4.3 Full Discrete Construction
We now give the algorithmic construction for discrete percolation which is the analogue of the continuum one. Each step of the construction is a single percolation exploration process; the order of successive steps is organized as in the continuum construction detailed in Section 3.2 . We start with the smallest Jordan set
of hexagons that covers the unit disc
. We will also make use of the countable set
of points dense in
that was introduced earlier.
The first step consists of an exploration process inside
. For this, we need to select two points
and
in
(which identify the starting and ending edges). We choose for
the e-vertex closest to
, and for
the e-vertex closest to
(if there are two such vertices closest to
, we can choose, say, the one with smallest real part, and analogously for
). The first exploration produces a path
and, for
small, many new domains of all four types. These domains are ordered according to the maximal
or
distance
between points on their boundaries and, if necessary, with the help of points in
, as in the continuum case, and that order is used, at each step of the construction, to determine the next exploration process. With this choice, the exploration processes and paths are naturally ordered:
.
Each exploration process of course requires choosing a starting and ending vertex and edge. For domains of type 1, with a
or
boundary condition, the choice is the natural one, explained before.
For a domain
(used at the
th step) of type other than 1, and therefore with a monochromatic boundary, the starting and ending edges are chosen with a procedure that mimics what is done in the continuum case. Once again, the exact procedure used to choose the pair of points is not important, as long as they are not chosen too close to each other. This is clear in the discrete case because the procedure that we are presenting is only “discovering” the cluster boundaries. In more precise terms, it is clear that one could couple the processes obtained with different rules by means of the same percolation configuration, thus obtaining exactly the same cluster boundaries.
As in the continuum case, we can choose the following procedure. (In Theorem 1 we will slightly reorganize the procedure by using a coupling to the continuum construction to guarantee that the order of exploration of domains of the discrete and continuum procedures match despite the rules for breaking ties.) Given a domain
,
and
are chosen so that, of all pairs
of points in
, they maximize
if
, or else they maximize
. If the choice is not unique, to restrict the number of pairs one looks at those pairs, among the ones already obtained, that maximize the other of
. Notice that this leaves at most two pairs of points; if that's the case, the pair that contains the point with minimal real (and, if necessary, imaginary) part is chosen.
The procedure continues iteratively, with regions that have monochromatic boundaries playing the role played in the first step by the unit disc. Every time a region with
boundary conditions is used, a new loop, corresponding to the outer boundary contour ofa cluster, is formed by pasting together, as explained in Section 3.1 , the new exploration path and the excursion containing the region where the last exploration was carried out.
All the new regions created at a step when a loop is formed have monochromatic boundary conditions.
5 Main Results
In this section we collect our main results about the Continuum Nonsimple Loop process. Before doing that, we state a precise version, called statement (S), on convergence of exploration paths to
that we will use in the proofs of these results, presented in Section 6 . (A proof of statement (S) is given in Appendix A ; it is an immediate consequence of Corollary A.1 there. The proof relies, among other things, on the result of Smirnov [30] concerning convergence of crossing probabilities to Cardy's formula [10, 11] – see Theorem 4 .) We note that (S) or (Corollary A.1 ) is both more general and more special than the convergence statements in [30, 31] — more general in that the domain can vary with
as
, but more special in the restriction to Jordan domains.
Given a Jordan domain
with two distinct points
on its boundary, let
denote the law of
, the trace of chordal
, and let
denote the law of the percolation exploration path
. Let
be the space of continuous curves inside
from
to
. We define
(where
denotes the open ball of radius
centered at
in the metric ( 2 )) and denote by
the Prohorov distance; weak convergence is equivalent to convergence in the Prohorov metric. Statement (S) is the following; it is used in the proofs of all the results of this section except for Lemmas 5.1 - 5.2 .
-
(S)
For Jordan domains, there is convergence in distribution of the percolation exploration path to the trace of chordal
that is locally uniform in the shape of the boundary with respect to the uniform metric on continuous curves ( 2 ), and in the location of the starting and ending points with respect to the Euclidean metric; i.e., for
a Jordan domain with
,
,
and
such that for all
with
Jordan and with
and
,
.
5.1 Preliminary Results
We first give some important results which are needed in the proofs of the main theorems. We start with two lemmas which are consequences of [2] , of standard bounds on the probability of events corresponding to having a certain number of monochromatic crossings of an annulus (see Lemma 5 of [16] and Appendix A of [20] ), but which do not depend on statement (S).
Lemma 5.1.
Let
be the percolation exploration path on the edges of
inside (the
-approximation of )
between (the e-vertices closest to)
and
. For any fixed point
, chosen independently of
, as
,
and the boundary
of the domain
that contains
jointly have limits in distribution along subsequences of
with respect to the uniform metric ( 2 ) on continuous curves. Moreover, any subsequence limit of
is almost surely a simple loop [
3]
.
Lemma 5.2.
Using the notation of Lemma 5.1 , let
be the limit in distribution of
as
along some convergent subsequence
and
the boundary of the domain
of
that contains
.
Then, as
,
converges in distribution to
.
The two lemmas above are important ingredients in the proof of Theorem 1 below. The second one says that, for every subsequence limit, the discrete boundaries converge to the boundaries of the domains generated by the limiting continuous curve. If we use statement (S), then the limit
of
is the trace of chordal
for every subsequence
, and we can use Lemmas 5.2 and 5.1 to deduce that all the domains produced in the continuum construction are Jordan domains. The key step in that direction is represented by the following result, our proof of which relies on (S).
Corollary 5.1.
For any deterministic
, the boundary
of a domain
of the continuum construction is almost surely a Jordan curve.
The corollary says that the domains that appear after the first step of the continuum construction are Jordan domains. The steps in the second stage of the continuum construction consist of
paths inside Jordan domains, and therefore Corollary 5.1 , combined with Riemann's mapping theorem and the conformal invariance of
, implies that the domains produced during the second stage are also Jordan. By induction, we deduce that all the domains produced in the continuum construction are Jordan domains.
We end this section with one more lemma which is another key ingredient in the proof of Theorem 1 ; we remark that its proof requires (S) in a fundamental way.
Lemma 5.3.
Let
denote a random Jordan domain, with
two points on
. Let
, be a sequence of random Jordan domains with points on their boundaries such that, as
,
converges in distribution to
with respect to the uniform metric ( 2 ) on continuous curves, and the Euclidean metric on
. For any sequence
with
as
,
converges in distribution to
with respect to the uniform metric ( 2 ) on continuous curves.
5.2 The Main Theorems
In this section we state the main theorems of this paper and a corollary, our most important result, that the Continuum Nonsimple Loop process is the scaling limit of the set of all cluster boundaries for critical site percolation on the triangular lattice. The corollary is obtained by combining the first two theorems. The proofs of all these results rely on statement (S). As noted before, statement (S) is proved in Appendix A .
Theorem 1.
For any
, the first
steps of (a suitably reorganized version of ) the full discrete construction inside the unit disc (of Section 4.3 ) converge, jointly in distribution, to the first
steps of the full continuum construction inside the unit disc (of Section 3.2 ). Furthermore, the scaling limit of the full (original or reorganized) discrete construction is the full continuum construction.
Moreover, if for any fixed
we let
denote the number of steps needed to find all the cluster boundaries of Euclidean diameter larger than
in the discrete construction, then
is bounded in probability as
; i.e.,
. This is so in both the original and reorganized versions of the discrete construction.
The second part of Theorem 1 means that both versions of the discrete construction used in the theorem find all large contours in a number of steps which does not diverge as
. This, together with the first part of the same theorem, implies that the continuum construction does indeed describe all macroscopic contours contained inside the unit disc (with blue boundary conditions) as
.
The construction presented in Section 3.2 can of course be repeated for the disc
of radius
, for any
, so we should take a “thermodynamic limit” by letting
. In this way, we would eliminate the boundary (and the boundary conditions) and obtain a process on the whole plane. Such an extension from the unit disc to the plane is contained in the next theorem.
Let
be the (limiting) distribution of the set of curves (all continuum nonsimple loops) generated by the continuum construction inside
(i.e., the limiting measure, defined by the inductive construction, on the complete separable metric space
of collections of continuous curves in
).
For a domain
, we denote by
the mapping (on
or
) in which all portions of curves that exit
are removed. When applied to a configuration of loops in the plane,
gives a set of curves which either start and end at points on
or form closed loops completely contained in
. Let
be the same mapping lifted to the space of probability measures on
or
.
Theorem 2.
Theorem 1 implies that there exists a unique probability measure
on the space
of collections of continuous curves in
such that
as
in the sense that for every bounded domain
, as
,
.
Remark 5.1.
We remark that we will generally take monochromatic blue boundary conditions on the disc
of radius
. But one could also takemonochromatic boundary conditions with color depending on
or even non-monochromatic boundary conditions without any essential change in the results or the proofs.
Corollary 5.2.
The Continuum Nonsimple Loop process
in the plane defined in Theorem 2 is the scaling limit of the collection of all boundary contours for critical site percolation on the triangular lattice.
The next theorem states some properties of the Continuum Nonsimple Loop process in the plane.
Theorem 3.
The Continuum Nonsimple Loop process in the plane has the following properties, the first three of which are valid with probability one:
-
1.
The Continuum Nonsimple Loop process is a random collection of noncrossing continuous loops in the plane. The loops can and do touch themselves and each other many times, but there are no triple points; i.e. no three or more loops can come together at the same point, and a single loop cannot touch the same point more than twice, nor can a loop touch a point where another loop touches itself.
-
2.
Any deterministic point (i.e., chosen independently of the loop process) of the plane is surrounded by an infinite family of nested loops with diameters going to both zero and infinity; any annulus about that point with inner radius
and outer radius
contains only a finite number of those loops. Consequently, any two distinct deterministic points of the plane are separated by loops winding around each of them.
-
3.
Any two loops are connected by a finite “path” of touching loops.
-
4.
The Continuum Nonsimple Loop process is conformally invariant in the sense that, given a Jordan domain
and a conformal homeomorphism
onto
, the scaling limits,
and
, of the loops inside
and
taken with, say, blue boundary conditions are related by
. (Here
denotes the probability distribution of the loop process
when
is distributed by
.)
To conclude this section, we show how to recover chordal
from the Continuum Nonsimple Loop process, i.e., given a (deterministic) Jordan domain
with two boundary points
and
, we give a construction that uses the continuum nonsimple loops of
to generate a process distributed like chordal
inside
from
to
.
Remember, first of all, that each continuum nonsimple loop has either a clockwise or counterclockwise direction, with the set of all loops surrounding any deterministic point alternating in direction. For convenience, let us suppose that
is at the “bottom” and
is at the “top” of
so that the boundary is divided into a left and right part by thesetwo points. Fix
and call
the set of all the directed segments of loops that connect from the left to the right part of the boundary touching
at a distance larger than
from both
and
, and
the analogous set of directed segments from the right to the left portion of
. For a fixed
, there is only a finite number of such segments, and, if they are ordered moving along the left boundary of
from
to
, they alternate in direction (i.e., a segment in
is followed by one in
and so on).
Between a segment in
and the next segment in
, there are countably many portions of loops intersecting
which start and end on
and are maximal in the sense that they are not contained inside any other portion of loop of the same type; they all have counterclockwise direction and can be used to make a “bridge” between the right-to-left segment and the next one (in
). This is done by pasting the portions of loops together with the help of points in
and a limit procedure to produce a connected (nonsimple) path.
If we do this for each pair of successive segments on both sides of the boundary of
, we get a path that connects two points on
. By letting
and taking the limit of this procedure, since almost surely
and
are surrounded by an infinite family of nested loops with diameters going to zero, we obtain a path that connects
with
; this path is distributed as chordal
inside
from
to
. The last claim follows from considering the analogous procedure for percolation on the discrete lattice
, using segments of boundaries. It is easy to see that in the discrete case this procedure produces exactly the same path as the percolation exploration process. By Corollary 5.2 , the scaling limit of this discrete procedure is the continuum one described above, therefore the claim follows from (S).
6 Proofs
In this section we present the proofs of the results stated in Section 5 .
Proof of Lemma 5.1 . The first part of the lemma is a direct consequence of [2] ; it is enough to notice that the (random) polygonal curves
and
satisfy the conditions in [2] and thus have a scaling limit in terms of continuous curves, at least along subsequences of
.
To prove the second part, we use standard percolation bounds (see Lemma 5 of [16] and Appendix A of [20] ) to show that, in the limit
, the loop
does not collapse on itself but remains a simple loop.
Let us assume that this is not the case and that the limit
of
along some subsequence
touches itself, i.e.,
for
with positive probability.
If that happens, we can take
small enough so that the annulus
is crossed at least four times by
(here
is the ball of radius
centered at
).
Because of the choice of topology, the convergence in distribution of
to
implies that we can find coupled versions of
and
on some probability space
such that
, for all
as
(see, for example, Corollary 1 of [6] ).
Using this coupling, we can choose
large enough (depending on
) so that
stays in an
-neighborhood
of
. This event however would correspond to (at least) four paths of one color (corresponding to the four crossings by
) and two of the other color of the annulus
.
As
, we can let
, in which case the probability of seeing the event just described somewhere inside
goes to zero [16, 20] , leading to a contradiction.
Proof of Lemma 5.2 . Let
be a convergent subsequence for
and
the limit in distribution of
as
. For simplicity of notation, in the rest of the proof we will drop the
and write
instead of
. Because of the choice of topology, the convergence in distribution of
to
implies that we can find coupled versions of
and
on some probability space
such that
, for all
as
(see, for example, Corollary 1 of [6] ). Using this coupling, our first task will be to prove the following claim:
-
(C)
For two (deterministic) points
, the probability that
but
or vice versa goes to zero as
.
Let us consider first the case of
such that
but
. Since
is an open subset of
, there exists a continuous curve
joining
and
and a constant
such that the
-neighborhood
of the curve is contained in
, which implies that
does not intersect
. Now, if
does not intersect
, for
small enough, then there is a
-path
of unexplored hexagons connecting the hexagon that contains
with the hexagon that contains
, and we conclude that
. This shows that the event that
but
implies the existence of a curve
whose
-neighborhood
is not intersected by
but whose
-neighborhood
is intersected by
. This implies that
,
such that
. But the right hand side goes to zero for every
as
, which concludes the proof of one direction of the claim.
To prove the other direction, we consider two points
such that
but
. Assume that
is trapped before
by
and suppose for the moment that
is a domain of type 3 or 4; the case of a domain of type 1 or 2 is analogous and will be treated later. Let
be the first time
is trapped by
with
the double point of
where the domain
containing
is “sealed off.” At time
, a new domain containing
is created and
is disconnected from
.
Choose
small enough so that neither
nor
is contained in the ball
of radius
centered at
, nor in the
-neighborhood
of the portion of
which surrounds
. Then it follows from the coupling that, for
small enough, there are appropriate parameterizations of
and
such that the portion
of
is inside
, and
and
are contained in
.
For
and
to be contained in the same domain in the discrete construction, there must be a
-path
of unexplored hexagons connecting the hexagon that contains
to the hexagon that contains
. From what we said in the previous paragraph, any such
-path connecting
and
would have to go though a “bottleneck” in
.
Assume now, for concreteness but without loss of generality, that
is a domain of type 3, which means that
winds around
counterclockwise, and consider the hexagons to the “left” of
. Those hexagons form a “quasi-loop” around
since they wind around it (counterclockwise) and the first and last hexagons are both contained in
. The hexagons to the left of
belong to the set
, which can be seen as a (nonsimple) path by connecting the centers of the hexagons in
by straight segments. Such a path shadows
, with the difference that it can have double (or even triple) points, since the same hexagon can be visited more than once. Consider
as a path
with a given parametrization
, chosen so that
is inside
when
is, and it winds around
together with
.
Now suppose that there were two times,
and
, such that
and
winds around
. This would imply that the “quasi-loop” of explored yellow hexagons around
is actually completed, and that
.
Thus, for
and
to belong to the same discrete domain, this cannot happen.
For any
, if we take
small enough,
will be contained inside
, due to the coupling. Following the considerations above, the fact that
and
belong to the same domain in the discrete construction but to different domains in the continuum construction implies, for
small enough, that there are four disjoint yellow
-paths crossing the annulus
(the paths have to be disjoint because, as we said,
cannot, when coming back to
after winding around
, touch itself inside
). Since
is also crossed by at least two blue
-paths from
, there is a total of at least six
-paths, not all of the same color,crossing the annulus
.
Let us call
the event described above, where
; a standard bound [16] on the probability of six disjoint crossings (not all of the same color) of an annulus gives that the probability of
scales as
with
. As
, we can let
go to zero (keeping
fixed); when we do this, the probability of
goes to zero sufficiently rapidly with
to conclude, like in the proof of Lemma 5.1 , that the probability to see such an event anywhere in
goes to zero.
In the case in which
belongs to a domain of type 1 or 2, let
be the excursion that traps
and
be the point on the boundary of
where
starts and
the point where it ends. Choose
small enough so that neither
nor
is contained in the balls
and
of radius
centered at
and
, nor in the
-neighborhood
of the excursion
. Because of the coupling, for
small enough (depending on
),
shadows
along
, staying within
. If this is the case, any
-path of unexplored hexagons connecting the hexagon that contains
with the hexagon that contains
would have to go through one of two “bottlenecks,” one contained in
and the other in
.
Assume for concreteness (but without loss of generality) that
is in a domain of type 1, which means that
winds around
counterclockwise. If we parameterize
and
so that
and
,
forms a “quasi-excursion” around
since it winds around it (counterclockwise) and it starts inside
and ends inside
. Notice that if
touched
, inside both
and
, this would imply that the “quasi-excursion” is a real excursion and that
.
For any
, if we take
small enough,
will be contained inside
, due to the coupling. Therefore, the fact that
implies, with probability going to one as
, that for
fixed and any
,
enters the ball
and does not touch
inside the larger ball
, for
or
. This is equivalent to having at least two yellow and one blue
-paths (contained in
) crossing the annulus
. Let us call
the event described above, where
; a standard bound [20] (this bound can also be derived from the one obtained in [16] ) on the probability of disjoint crossings (not all of the same color) of a semi-annulus in the upper half-plane gives that the probability of
scales as
with
. (We can apply the bound to our case because the unit disc is a convex subset of the half-plane
and therefore the intersection of an annulus centered at say
with the unit disc is a subset of the intersection of the same annulus with the half-plane
.) As
, we can let
go to zero (keeping
fixed), concluding that the probability that such an event occurs anywhere on the boundary of the disc goes to zero.
We have shown that, for two fixed points
, having
but
or vice versa implies the occurrence of an event whose probability goes to zero as
, and the proof of the claim is concluded.
We now introduce the Hausdorff distance
between two closed nonemptysubsets of
:
|
(9)
|
With this metric, the collection of closed subsets of
is a compact space. We will next prove that
converges in distribution to
as
, in the topology induced by ( 9 ). (Notice that the coupling between
and
provides a coupling between
and
, seen as boundaries of domains produced by the two paths.) We will now use Lemma 5.1 and take a further subsequence
of the
's that for simplicity of notation we denote by
such that, as
,
converge jointly in distribution to
, where
is a simple loop. For any
, since
is a compact set, we can find a covering of
by a finite number of balls of radius
centered at points on
. Each ball contains both points in the interior
of
and in the exterior
of
, and we can choose (independently of
) one point from
and one from
inside each ball.
Once again, the convergence in distribution of
to
implies the existence of a coupling such that, for
large enough, the selected points that are in
are contained in
, and those that are in
are contained in the complement of
. But by claim (C), each one of the selected points that is contained in
is also contained in
with probability going to
as
; analogously, each one of the selected points contained in the complement of
is also contained in the complement of
with probability going to
as
. This implies that
crosses each one of the balls in the covering of
, and therefore
. From this and the coupling between
and
, it follows immediately that, for
large enough,
with probability close to one.
A similar argument (analogous to the previous one but simpler, since it does not require the use of
), with the roles of
and
inverted, shows that
with probability going to
as
. Therefore, for all
,
as
, which implies convergence in distribution of
to
, as
, in the topology induced by ( 9 ). But Lemma 5.1 implies that
converges in distribution (using ( 2 )) to a simple loop, therefore
must also be a simple loop; and we have convergence in the topology induced by ( 2 ).
It is also clear that the argument above is independent of the subsequence
, so the limit of
is unique and coincides with
. Hence, we have convergence in distribution of
to
, as
, in the topology induced by ( 2 ), and indeed joint convergence of
to
.
Proof of Corollary 5.1 . The corollary follows immediately from Lemma 5.1 and Lemma 5.2 , as already seen in the proof of Lemma 5.2 .
Proof of Lemma 5.3 . First of all recall that the convergence of
to
in distribution implies the existence of coupled versions of
and
onsome probability space
such that
,
,
for all
as
(see, for example, Corollary 1 of [6] ). This immediately implies that the conditions to apply Radó's theorem (see Theorem 9 of Appendix B ) are satisfied. Let
be the conformal map that takes the unit disc
onto
with
and
, and let
be the conformal map from
onto
with
and
. Then, by Theorem 9 ,
converges to
uniformly in
, as
.
Let
(resp.,
) be the chordal
inside
(resp.,
) from
to
(resp., from
to
),
,
,
, and
,
,
.
We note that, because of the conformal invariance of chordal
,
(resp.,
) is distributed as chordal
in
from
to
(resp., from
to
). Since
and
for all
, and
uniformly in
, we conclude that
and
for all
.
Later we will prove a “continuity” property of
(Lemma 6.1 ) that allows us to conclude that, under these conditions,
converges in distribution to
in the uniform metric ( 2 ) on continuous curves. Once again, this implies the existence of coupled versions of
and
on some probability space
such that
, for all
as
. Therefore, thanks to the convergence of
to
uniformly in
,
, for all
as
. But since
is distributed as
and
is distributed as
, we conclude that, as
,
converges in distribution to
in the uniform metric ( 2 ) on continuous curves.
We now note that (S) implies that, as
,
converges in distribution to
uniformly in
, for
large enough. Therefore, as
,
converges in distribution to
, and the proof is concluded.
Lemma 6.1.
Let
be the unit disc,
and
two distinct points on its boundary, and
the trace of chordal
inside
from
to
. Let
and
be two sequences of points in
such that
and
. Then, as
, the trace
of chordal
inside
from
to
converges in distribution to
in the uniform topology ( 2 ) on continuous curves.
Proof. Let
be the (unique) linear fractional transformation that takes the unit disc
onto itself, mapping
to
,
to
, and a third point
distinct from
and
to itself.
and
depend continuously on
and
. As
, since
and
,
converges uniformly to the identity in
.
Using the conformal invariance of chordal
, we couple
and
by writing
. The uniform convergence of
to the identity implies that
as
, which is enough to conclude that
converges to
in distribution.
Proof of Theorem 1 . Let us prove the second part of the theorem first. We will do this for the original version of the discrete construction, but essentially the same proof works for the reorganized version we will describe below, as we will explain later. Suppose that at step
of this discrete construction an exploration process
is run inside a domain
, and write
, where
are the maximal connected domains of unexplored hexagons into which
is split by removing the set
of hexagonsexplored by
.
Let
and
be respectively the maximal
and
-distances between pairs of points in
. Suppose, without loss of generality, that
, and consider the rectangle
(see Figure 6 ) whose vertical sides are aligned to the
-axis, have length
, and are each placed at
-distance
from points of
with minimal or maximal
-coordinate in such a way that the horizontal sides of
have length
; the bottom and top sides of
are placed in such a way that they are at equal
-distance from the points of
with minimal or maximal
-coordinate, respectively.
It follows from the Russo-Seymour-Welsh lemma [27, 29] (see also [15, 13] ) that the probability to have two vertical
-crossings of
of different colors is bounded away from zero by a positive constant
that does not depend on
(for
small enough). If that happens, then
. The same argument of course applies to the maximal
-distance when
. We can summarize the above observation in the following lemma.
Lemma 6.2.
Suppose that at step
of the full discrete construction an exploration process
is run inside a domain
. If
, then for
small enough (i.e.,
for some constant
),
with probability at least
independent of
. The same holds for the maximal
-distances when
.
Here is another lemma that will be useful later on.
Lemma 6.3.
Two “daughter” subdomains,
and
, either have disjoint s-boundaries, or else their common s-boundary consists of exactly two adjacent hexagons (of the same color) where the exploration path
came within
hexagons of touching itself just when completing the s-boundary of one of the two subdomains.
Proof. Suppose that the two daughter subdomains have s-boundaries
and
that are not disjoint and let
be the set of (sites of
that are the centers of the) hexagons that belong to both s-boundaries.
can be partitioned into subsets consisting of single hexagons that are not adjacent to any another hexagon in
and groups of hexagons that form simple
-paths (because the s-boundaries of the two subdomains are simple
-loops). Let
be such a subset of hexagons of
that form a simple
-path
. Then there is a
-path
of hexagons in
that goes from
to
without using any other hexagon of
and a different
-path
in
that goes from
to
without using any other hexagon of
. But then, all the hexagons in
other than
and
are “surrounded” by
and therefore cannot have been explored by the exploration process that produced
and
, andcannot belong to
or
, leading to a contradiction, unless
. Similar arguments lead to a contradiction if
is partitioned into more than one subset.
If
is not adjacent to any other hexagon in
, then it is adjacent to two other hexagons of
and two hexagons of
. Since
has only six neighbors and neither the two hexagons of
adjacent to
nor those of
can be adjacent to each other, each hexagon of
is adjacent to one of
. But then, as before,
is “surrounded” by
and therefore cannot have been explored by the exploration process that produced
and
, and cannot belong to
or
, leading once again to a contradiction. The proof is now complete, since the only case remaining is the one where
consists of a single pair of adjacent hexagons as stated in the lemma.
With these lemmas, we can now proceed with the proof of the second part of the theorem. Lemma 6.2 tells us that large domains are “chopped” with bounded away from zero probability (
), but we need to keep track of domains of diameter larger than
in such a way as to avoid “double counting” as the lattice construction proceeds.
More accurately, we will keep track of domains
having
, since only these can have diameter larger than
. To do so, we will associate with each domain
having
that we encounter as we do the lattice construction a non-negative integer label. The first domain is
(see the beginning of Section 4.3 ) and this gets label
. After each exploration process in a domain
with
, if the number
of “daughter” subdomains
with
is
, then the label of
is no longer used, if instead
, then one of these
subdomains (chosen by any procedure – e.g., the one with the highest priority for further exploration) is assigned the same label as
and the rest are assigned the next
integers that have never before been used as labels. Note that once all domains have
, there are no more labelled domains.
Lemma 6.4.
Let
denote the total number of labels used in the above procedure; then for any fixed
,
is bounded in probability as
; i.e.,
.
Proof. Except for
, every domain comes with (at least) a “physically correct” monochromatic “half-boundary” (notice that we are considering s-boundaries and that a half-boundary coming from the “artificially colored” boundary of
is not considered a physically correct monochromatic half-boundary). Let us assume, without loss of generality, that
.
If we associate with each label the “last” (in terms of steps of the discrete construction) domain which used that label (its daughter subdomains all had
), then we claim that it follows from Lemma 6.3 that (with high probability) any two such last domains that are labelled have disjoint s-boundaries. This is a consequence of the fact that the two domains are subdomains of two “ancestors” that are distinct daughter subdomains of the same domain (possibly
) and whose s-boundaries are therefore (by Lemma 6.3 ) either disjoint or else overlap at a pair of hexagons where an exploration path had a close encounter of distance two hexagons with itself. But since we are dealing only with macroscopic domains (of diameter at least order
), such a close encounter would imply,like in Lemmas 5.1 and 5.2 , the existence of six crossings, not all of the same color, of an annulus whose outer radius can be kept fixed while the inner radius is sent to zero together with
. The probability of such an event goes to zero as
and hence the unit disc
contains, with high probability, at least
disjoint monochromatic
-paths of diameter at least
, corresponding to the physically correct half-boundaries of the
labelled domains.
Now take the collection of squares
of side length
centered at the sites
of a scaled square lattice
of mesh size
, and let
be the number of squares of side
needed to cover the unit disc. Let
and consider the event
, which implies that, with high probability, the unit disc contains at least
disjoint monochromatic
-paths of diameter at least
and that, for at least one
, the square
intersects at least six disjoint monochromatic
-paths of diameter larger that
, so that the “annulus”
is crossed by at least six disjoint monochromatic
-paths contained inside the unit disc.
If all these
-paths crossing
have the same color, say blue, then since they are portions of boundaries of domains discovered by exploration processes, they are “shadowed” by exploration paths and therefore between at least one pair of blue
-paths, there is at least one yellow
-path crossing
. Therefore, whether the original monochromatic
-paths are all of the same color or not,
is crossed by at least six disjoint monochromatic
-paths not all of the same color contained in the unit disc. Let
denote the
as
of the probability that such an event happens anywhere inside the unit disc. We have shown that the event
implies a “six-arms” event unless not all labelled domains have disjoint s-boundaries. But the latter also implies a “six-arms” event, as discussed before; therefore
|
(10)
|
Since
, bounds in [16] imply that, for
fixed,
as
, which shows that
|
(11)
|
and concludes the proof of the lemma.
Now, let
denote the number of distinct domains that had label
(this is equal to the number of steps that label
survived). Let us also define
to be the smallest integer
such that
and
to be the random variable corresponding to how many Bernoulli trials (with probability
of success) it takes to have
successes.
Then, we may apply (sequentially) Lemma 6.2 to conclude that for any
|
(12)
|
where
is an independent copy of
. Now let
be i.i.d. random variables equidistributed with
.
Let
be the number of steps needed so that all domains left to explore have
.
Then, for any positive integer
,
|
(13)
|
Notice that, for fixed
,
as
. Moreover, for any
, by Lemma 6.4 , we can choose
large enough so that
. So, for any
, it follows that
|
(14)
|
which implies that
|
(15)
|
To conclude this part of the proof, notice that the discrete construction cannot “skip” a contour and move on to explore its interior, so that all the contours with diameter larger than
must have been found by step
if all the domains present at that step have diameter smaller than
. Therefore,
, which shows that
is bounded in probability as
.
For the first part of the theorem, we need to prove, for any fixed
, joint convergence in distribution of the first
steps of a suitably reorganized discrete construction to the first
steps of the continuum one. Later we will explain why this reorganized construction has the same scaling limit as the one defined in Section 4.3 . For each
, the first
steps of the reorganized discrete construction will be coupled to the first
steps of the continuum one with suitable couplings in order to obtain the convergence in distribution of those steps of the discrete construction to the analogous steps of the continuum one; the proof will proceed by induction in
. We will explain how to reorganize the discrete construction as we go along; in order to explain the idea of the proof, we will consider first the cases
,
and
, and then extend to all
.
. The first step of the continuum construction consists of an
from
to
inside
. Correspondingly, the first step of the discrete construction consists of an exploration path
inside
from the e-vertex closest to
to the e-vertex closest to
.
The convergence in distribution of
to
is covered by statement (S).
. The convergence in distribution of the percolation exploration path to chordal
implies that we can couple
and
generating them as random variables on some probability space
such that
for all
as
(see, for example, Corollary 1 of [6] ).
Now, let
be the domain generated by
that is chosen for the second step of the continuum construction, and let
be the highest ranking point of
contained in
.For
small enough,
is also contained in
; let
be the unique connected component of the set
containing
(this is well-defined with probability close to
for small
);
is the domain where the second exploration process is to be carried out.
From the proof of Lemma 5.2 , we know that the boundaries
and
of the domains
and
produced respectively by the path
and
are close with probability close to one for
small enough.
For the next step of the discrete construction, we choose the two e-vertices
and
in
that are closest to the points
and
of
selected for the coupled continuum construction (if the choice is not unique, we can select the e-vertices with any rule to break the tie) and call
the percolation exploration path inside
from
to
. It follows from [2] that
converge jointly in distribution along some subsequence to some limit
. We already know that
is distributed like
and we can deduce from the joint convergence in distribution of
to
(Lemma 5.2 ), that
is distributed like
. Therefore, if we call
the
path inside
from
to
, Lemma 5.3 implies that
is distributed like
and indeed that, as
,
converge jointly in distribution to
.
. So far, we have proved the convergence in distribution of the (paths and boundaries produced in the) first two steps of the discrete construction to the (paths and boundaries produced in the) first two steps of the discrete construction. The third step of the continuum construction consists of an
path
from
to
, inside the domain
with highest priority after the second step has been completed. Let
be the highest ranking point of
contained in
,
the domain of the discrete construction containing
after the second step of the discrete construction has been completed (this is well defined with probability close to
for small
), and choose the two e-vertices
and
in
that are closest to the points
and
of
selected for the coupled continuum construction (if the choice is not unique, we can select the e-vertices with any rule to break the tie). The third step of the discrete construction consists of an exploration path
from
to
inside
.
It follows from [2] that
converge jointly in distribution along some subsequence to some limit
. We already know that
is distributed like
,
like
and
like
, and we would like to apply Lemma 5.3 to conclude that
is distributed like
and indeed that, as
,
converges in distribution to
. In order to do so, we have to first show that
is distributed like
. If
is a subset of
, this follows from Lemma 5.2 , as in the previous case, but if the s-boundary of
contains hexagons of
, then we cannot use Lemma 5.2 directly, although the proof of the lemma can be easily adapted to the present case, as we now explain.
Indeed, the only difference is in the proof of claim (C) and is due to the fact that, when dealing with a domain of type 1 or 2, we cannot use the bound on the probability of three disjoint crossings of a semi-annulus because the domains we are dealing with may not be convex (like the unit disc). On the other hand, the discrete domains like
and
where we have to run exploration processes at various steps of the discreteconstruction are themselves generated by previous exploration processes, so that any hexagon of the s-boundary of such a domain has three adjacent hexagons which are the starting points of three disjoint
-paths (two of one color and one of the other). Two of these
-paths belong to the s-boundary of the domain, while the third belongs to the adjacent percolation cluster (see Figure 7 ). This allows us to use the bound on the probability of six disjoint crossings of an annulus.
To see this, let
be the
-paths contained in the s-boundary of the discrete domain (i.e.,
in the present context) and
the
-path belonging to the adjacent cluster, all starting from hexagons adjacent to some hexagon
(centered at
) in the s-boundary of
. For
and
small enough, let
be the event that the exploration path
enters the ball
without touching
inside the larger ball
.
implies having (at least) three disjoint
-paths (two of one color and one of the other),
and
, contained in
and crossing the annulus
, with
and
disjoint from
and
. Hence,
implies the event that there are (at least) six disjoint crossings (not all of the same color) of the annulus
.
Once claim (C) is proved, the rest of the proof of Lemma 5.2 applies to the present case. Therefore, we have convergence in distribution of
to
, which allows us to use Lemma 5.3 and conclude that
converges in distribution to
.
. We proceed by induction in
, iterating the steps explained above; there are no new difficulties; all steps for
are analogous to the case
.
To conclude the proof of the theorem, we need to show that the scaling limit of the original full discrete construction defined in Section 4.3 is the same as that of the reorganized one just used in the proof of the first part of the theorem. In order to do so, we can couple the two constructions by using the same percolation configuration for both, so that the two constructions have at their disposal the same set of loops to discover. We proved above that the original discrete construction finds all the “macroscopic” loops, so we have to show that this is true also for the reorganized version of the discrete construction. This is what we will do next, using essentially the same arguments as those employed for the original discrete construction; we present these arguments for the sake of completeness since there are some changes.
Consider the reorganized discrete construction described above, where the starting and ending points of the exploration processes at each step are chosen to be close to those of the corresponding (coupled) continuum construction. Suppose that at step
of this discrete construction an exploration process
is run inside a domain
, and write
, where
are the connected domains into which
is split by the set
of hexagons explored by
.
Let
(resp.,
) and
(resp.,
) be respectively the maximal
and
-distance between pairs of points in
(resp.,
). If
and the e-vertices on
are chosen to be closest to two points of
with maximal
-distance, then the same construction and argument spelled out earlier in the first part of the proof (corresponding to the second part of the theorem) show that
with bounded away from zero probability.
If the e-vertices on
are chosen to be closest to two points of
with maximal
-distance but
, then consider the rectangle
whose vertical sides are aligned to the
-axis, have length
, and are each placed at the same
-distance from the points of
with minimal or maximal
-coordinate in such a way that the horizontal sides of
have length
; the bottom and top sides of
are placed in such a way that they touch the points of
with minimal or maximal
-coordinate, respectively. Notice that, because of the coupling between the continuum and discrete constructions, for any
, for
large enough,
and
. Since in the case under consideration we have
and
, for
large enough, we must also have
.
Once again, it follows from the Russo-Seymour-Welsh lemma that the probability to have two vertical
-crossings of
of different colors is bounded away from zero by a positive constant that does not depend on
(for
small enough). If that happens, then
.
All other cases are handled in the same way, implying that the maximal
and
-distances of domains that appear in the discrete construction have a positive probability (bounded away from zero) to decrease by (approximately) a factor
at each step of the discrete construction in which an exploration process is run in that domain.
With this result at our disposal, the rest of the proof, that for any
the number of steps needed to find all the loops of diameter larger than
is bounded in probability as
(which implies that all the “macroscopic” loops are discovered), proceeds exactly like for the original discrete construction.
Proof of Theorem 2 . First of all, we want to show that
does not depend on
, provided
is strictly contained in
and
. In order to do this, we assume that the above conditions are satisfied for the pair
and show that
for all
.
Take two copies of the scaled hexagonal lattice,
and
, their dual lattices
and
, and two percolation configurations,
and
, both with blue boundary conditions and coupled in such a way that
. The laws of the boundaries of
and
are also coupled, in such a way that the boundaries or portions of boundaries contained inside
are identical for all small enough
. Therefore, letting
and using the convergence of the percolation boundaries inside
and
to the continuum nonsimple loop processes
and
respectively, we conclude that
.
From what we have just proved, it follows that the probability measures
on
, for
, satisfy the consistency conditions
for all
.
Since
,
are complete separable metric spaces, the measurable spaces
,
are standard Borel spaces and so we can apply Kolmogorov's extension theorem (see, for example, [12] ) and conclude that there exists a unique probability measure on
with
for all
. It follows that, for
and all
strictly contained in
and such that
,
, which concludes the proof.
Proof of Corollary 5.2 . The corollary is an immediate consequence of Theorems 1 and 2 , where the full scaling limit is intended in the topology induced by ( 5 ).
Proof of Theorem 3 . 1. The fact that the Continuum Nonsimple Loop process is a random collection of noncrossing continuous loops is a direct consequence of its definition.
The fact that the loops touch themselves is a consequence of their being constructed out of
, while the fact that they touch each other follows from the observation that a chordal
path
touches
with probability one. Therefore, each new loop in the continuum construction touches one or more previous ones (many times).
The nonexistence of triple points follows directly from Lemma 5 of [16] on the number of crossings of an annulus, combined with Corollary 5.2 , which allows to transport discrete results to the continuum case. In fact, a triple point would imply, for discrete percolation, at least six crossings (not all of the same color) of an annulus whose ratio of inner to outer radius goes to zero in the scaling limit, leading to a contradiction.
2. This follows from straightforward Russo-Seymour-Welsh type arguments for percolation (for more details, see, for example, Lemma 3 of [16] ), combined with Corollary 5.2 .
3. Combining Russo-Seymour-Welsh type arguments for percolation (see, for example, Lemma 3 of [16] ) with Corollary 5.2 , we know that
-a.s. there exists a (random)
, with
, such that
is surrounded by a continuum nonsimple loop contained in
. From (the proof of ) Theorem 2 , we also know that
for all
. This implies that by taking
large enough and performing the continuum construction inside
, we have a positive probability of generating a loop
contained in the annulus
, with
. If that is the case, all the loops contained inside
are connected, by construction, to the loop
surrounding
by a finite sequence (a “path”) of loops (remember that in the continuum construction each loop is generated by pasting together portions of
paths inside domains whose boundaries are determined by previously formed loops or excursions). Therefore, any two loops contained inside
are connected to each other by a “path” of loops.
Using again the fact that
for all
, and letting first
and then
go to
, we see from the discussion above (with
as well) that any two loops are connected by a finite “path” of intermediate loops,
-a.s.
4. In order to prove the claim, we will define a discrete construction inside
coupled to the continuum construction inside
, by means of the conformal map
from
to
.
Roughly speaking, this new discrete construction for
is one in which the
pairs at each step are chosen to be closest to the
points in
mapped from
via
,where the pairs
are those that appear at the corresponding steps of the continuum construction inside
.
More precisely, let
be the first
path in
from
to
. Because of the conformal invariance of
, the image
of
under
is a path distributed as the trace of chordal
in
from
to
. Therefore, the exploration path
inside
from
to
, chosen to be closest to
and
respectively, converges in distribution to
, as
, which means that there exist a coupling between
and
such that the paths stay close for
small.
We see already that one can use the same strategy as in the proof of the first part of Theorem 1 , and obtain a discrete construction whose exploration paths are coupled to the
paths
that are the images of the paths
in
. Then, for this discrete construction, the scaling limit of the exploration paths will be distributed as the images of the
paths in
.
In order to conclude the proof, we just have to show that the discrete construction inside
defined above finds all the boundaries in a number of steps that is bounded in probability as
(this is equivalent to the second part of Theorem 1 ). To do that, we use the second part of Theorem 1 , which implies that, for any fixed
and
, the number of steps of the discrete construction in
that are necessary to ensure that only domains with diameter less than
are present is bounded in probability as
.
Since
(can be extended to a function that) is continuous in the compact set
,
is uniformly continuous and so we can now choose
such that any subdomain of
of diameter at most
is mapped by
to a subdomain of
of diameter at most
. This, combined with the coupling between
paths and exploration paths inside
, assures that the number of steps necessary for the new discrete construction inside
to find all the loops of diameter at least
is bounded in probability as
.
Therefore, the scaling limit, as
, of this new discrete construction for
gives the measure
. It follows by construction that
, which concludes the proof. Appendix A: Convergence of the Percolation Exploration Path In this first appendix, we provide a detailed proof of statement (S). The existence of subsequential limits for the percolation exploration path, which follows from the work of Aizenman and Burchard [2] , means that the proof can be divided into two parts: first we will give a characterization of chordal
in terms of two properties that determine it uniquely; then we will show that any subsequential scaling limit of the percolation exploration path satisfies these two properties.
The characterization part will follow from known properties of hulls and of
(see [22] and [35] ). The second part will follow from an extension of Smirnov's result about the convergence of crossing probabilities to Cardy's formula [10, 11] (see Theorem 6 below) for sequences of Jordan domains
, with the domain
changing together with the mesh
of the lattice, combined with the proof of a certain spatial Markov property forsubsequential limits of percolation exploration hulls (Theorem 7 ). We note that although Theorem 6 represents only a slight extension to Smirnov's result on convergence of crossing probabilities, this extension and its proof play a major role in the technically important Lemmas A.2 , A.4 and A.5 , which control the “close encounters” of exploration paths to domain boundaries. The proof of Theorem 6 is modelled after a simpler geometric argument involving only rectangles used in [8] .
Let
be a bounded simply connected domain containing the origin whose boundary
is a continuous curve. Let
be the (unique) conformal map from the unit disc to
with
and
; note that by Theorem 10 of Appendix B ,
has a continuous extension to
. Let
be four points of
(or more accurately, four prime ends) in counterclockwise order – i.e., such that
, with
in counterclockwise order. Also, let
. Cardy's formula [10, 11] for the probability
of a “crossing” inside
from the counterclockwise arc
to the counterclockwise arc
is
|
(16)
|
where
is a hypergeometric function.
For a given mesh
, the probability of a crossing inside
from the counterclockwise arc
to the counterclockwise arc
is the probability of the existence of a blue
-path contained in
, the
-approximation of
(see Definition 4.1 above), that starts at a hexagon adjacent to one intersecting
and ends at a hexagon adjacent to one intersecting
. Smirnov proved the following major theorem, concerning the conjectured behavior [10, 11] of crossing probabilities in the scaling limit.
Theorem 4.
(Smirnov [
30]
) In the limit
, the crossing probability becomes conformally invariant and converges to Cardy's formula ( 16 ).
Remark A.1.
We actually only need Theorem 4 for Jordan domains, as can be seen by a careful reading of the proof of Theorem 6 . We note that Smirnov does not restrict attention to that case.
Let us now specify the objects that we are interested in. Suppose
is a simply connected domain whose boundary
is a continuous curve, and
are two distinct points in
(or more accurately, two distinct prime ends), and let
be a probability measure on continuous (non-self-crossing) curves
with
and
(we remark that the use of
instead of
for the time parametrization is purely for convenience). Let
denote the (unique) connected component of
whose closure contains
, where
, the filling of
, is a closed connected subset of
.
is called a hull if it satisfies the condition
|
(17)
|
We will generally be interested in curves
such that
is a hull for each
, although we normally only consider
at certain stopping times
. (An example of such a curve that we are particularly interested in is the trace of chordal
.) Let
be a closed subset of
such that
,
, and
is a bounded simply connected domain whose boundary contains the counterclockwise arc
that does not belong to
(except for its endpoints
and
– see Figure 8 ).
Let
be the first time that
hits
and assume that the filling
of
is a hull; we denote by
the distribution of
. To explain what we mean by the distribution of a hull, consider the set
of closed subsets
of
that do not contain
and such that
is a simple (continuous) curve contained in
starting at a point on
and ending at a point on
(see Figure 9 ).
Let
be the set of closed subsets of
of the form
, where
and
.
For a given
and corresponding
, let
be the set whose elements are possible hulls at time
; we claim that the events
, for
, form a
-system
(i.e., they are closed under finite intersections; we also include the empty set in
), and we consider the
-algebra
generated by these events. To see that
is closed under pairwise intersections, notice that, if
, then
and
(or else
is empty). We are interested in probability spaces of the form
.
We say that the exit distribution of
is determined by Cardy's formula if, for any
and any counterclockwise arc
of
, the probability that
hits
at time
on
is given by
|
(18)
|
It is easy to see that if the hitting distribution of
is determined by Cardy's formula, then the probabilities of events in
are also determined by Cardy's formula in the following way. Let
be the union of
, with
given by a curve from
to
and
given by a curve from
to
; then, assuming that
,
,
,
,
are ordered counterclockwise around
,
|
(19)
|
Since
is a
-system, the probabilities of the events in
determine uniquely the distribution of the hull in the sense described above. Therefore, if we let
denote the trace of chordal
inside
from
to
,
its hull up to time
, and
the first time that
hits
, we have the following simple but useful lemma.
Lemma A.1.
With the notation introduced above, if the hitting distribution of
is determined by Cardy's formula and
is a hull, then
is distributed like the hull
of chordal
.
Proof. It is enough to note that the hitting distribution for chordal
is determined by Cardy's formula [19] .
Now let
be a conformal map from
to the upper half-plane
such that
and
(these two conditions determine
only up to a multiplicative constant). For
fixed, let
denote the semi-ball of radius
centered at
on the real line and let
denote the first time
hits
, where
. Define recursively
as the first time
hits
, where
,
, and
is a conformal map from
to
that maps
to
and
to
. We also define
, so that
. We note that, like
, the conformal maps
are only defined up to a multiplicative factor.
Notice that
is a bounded simply connected domain chosen so that the conformal transformation which maps
to
maps
to the semi-ball
centered at the origin on the real line. With these definitions, consider the (discrete-time) stochastic process
for
; we say that
satisfies the spatial Markov property if each
is a hull and
for
is a Markov chain (for any choice of the multitplicative factors for
). Notice that the trace of chordal
satisfies the spatial Markov property, due to the conformal invariance and Markovian properties [28] of
.
Next, we give the main characterization theorem.
Theorem 5.
If the filling process
of
satisfies the spatial Markov property and its hitting distribution is determined by Cardy's formula, then
is distributed like the trace
of chordal
inside
started at
and aimed at
.
Proof. Since the trace
of chordal
in a bounded Jordan domain
is defined (up to a linear time change) as
, where
is the trace of chordal
in the upper half-plane started at
and
is any conformal map from the upper half-plane
to
such that
and
, it is enough to show that
is distributed like the trace of chordal
in the upper half-plane. Let
denote the filling of
at time
and let
be the unique conformal transformation that maps
onto
with the following expansion at infinity:
|
(20)
|
We choose to parametrize
so that
(this is often called parametrization by capacity,
being the capacity of the filling up to time
). We want to compare
with the trace
of chordal
in the upper half-plane parametrized in the same way (i.e., with
), so that, if
denotes the filling of
at time
,
is mapped onto
by a conformal transformation with the following expansion at infinity:
|
(21)
|
Our strategy will be to construct suitable polygonal approximations
and
of
and
which converge, as
, to the original curves (in the uniform metric on continuous curves ( 2 )), and show that
and
have the same distribution. This implies that the distributions of
and
must coincide.
Let us describe first the construction for
; we use exactly the same construction for
. We remark that the important features in the construction of the polygonal approximations are the spatial Markov property of the fillings and Cardy's formula, which are valid for both
and
.
For
fixed, as above let
denote the semi-ball of radius
centered at
on the real line. Let
denote the first time
hits
, where
, and define recursively
as the first time
hits
, where
and
. Notice that
is a bounded simply connected domain chosen so that the conformal transformation which maps
to
maps
to the semi-ball
centered at the point of the real line where the “tip”
of the hull
is mapped. The spatial Markov property and the conformal invariance of the hull of
imply that if we write
, with
, the
's are i.i.d. random variables, and also that the distribution of
is the same as the distribution of
, where
is equidistributed with
, but also is independent of
. The polygonal approximation
is obtained by joining, for all
,
to
with a straight segment, where the speed
is chosen to be constant.
Now let
denote the first time
hits
, where
, and define recursively
as the first time
hits
, where
and
. We also define
, so that
.
Once again,
is a bounded simply connected domain chosen so that the conformal transformation which maps
to
maps
to the semi-ball
centered at the point on the real line where the “tip”
of the hull
is mapped. The polygonal approximation
is obtained by joining, for all
,
to
with a straight segment, where the speed
is chosen to be constant.
Consider the sequence of times
defined in the natural way so that
and the (discrete-time) stochastic processes
and
related by
. If for
we let
denote the translation that maps
to
and define the family of conformal maps
from
to
, then
sends
to
and
to
, and
is the first time
hits
,where
and
. Therefore,
is a sequence of stopping times like those used in the definition of the spatial Markov property and, thanks to the relation
, the fact that
satisfies the spatial Markov property implies that
is a Markov chain. We also note that the fact that the hitting distribution of
is determined by Cardy's formula implies the same for the hitting distribution of
, thanks to the conformal invariance of Cardy's formula. We next use these properties to show that
is distributed like
.
To do so, we first note that the conformal transformations
and
are random and that their distributions are functionals of the distributions of the hulls
and
, since there is a one-to-one correspondence between hulls and conformal maps (with the normalization we have chosen in ( 20 )–( 21 )). Therefore, since
is distributed like
(see Lemma A.1 ),
and
have the same distribution, which also implies that
is distributed like
because, due to the parametrization by capacity of
and
,
is exactly the coefficient of the term
in the expansion at infinity of
, and
is exactly the coefficient of the term
in the expansion at infinity of
. Moreover, it is also clear that
is distributed like
, because their distributions are both determined by Cardy's formula, and so
is distributed like
. Notice that the law of the hull
is conformally invariant because, by Lemma A.1 , it coincides with the law of the
hull
.
Using now the Markovian character of
, which implies that, conditioned on
,
and
are determined by Cardy's formula in
, from the fact that
is equidistributed with
and therefore
is equidistributed with
, we obtain that the hull
is distributed like
and its “tip”
is distributed like the “tip”
of the hull
, and we can conclude that the joint distribution of
is the same as the joint distribution of
. It also follows immediately that
is equidistributed with
and
is equidistributed with
or indeed with
.
By repeating this argument recursively, using at each step the Markovian character of the hulls and tips, we obtain that, for all
, the joint distribution of
is the same as the joint distribution of
. This immediately implies that
has the same distribution as
.
In order to conclude the proof, we just have to show that, as
,
converges to
and
to
in the uniform metric on continuous curves ( 2 ). This, however, follows easily from properties of continuous curves, if we can show that the time intervals
and
go to
as
. To see this, we recall that
and
are distributed like
and use Lemma 2.1 of [21] , which implies the (deterministic) bound
, which follows from the well-know bound
for the capacity
of ( 21 ).
We will next prove a version of Smirnov's result (Theorem 4 above) extended to cover the convergence of crossing probabilities to Cardy's formula for the case of sequences of domains. The statement of Theorem 6 below is certainly not optimal, but it is sufficient for our purposes. We remark that a weaker statement restricted, for instance, only to Jordan domains would not be sufficient – see Figure 19 and the discussion referring to it in the proof of Theorem 7 below. First, we introduce some definitions that will simplify the notation in the rest of the paper.
We will consider bounded simply connected domains
whose boundaries
are continuous curves. Let
be the (unique) conformal map from the unit disc to
with
and
; note that by Theorem 10 of Appendix B ,
has a continuous extension to
. Let
be three points of
(or more accurately, three prime ends) in counterclockwise order – i.e., such that
,
and
, with
,
and
in counterclockwise order. We will call
admissible with respect to
if the counterclockwise arcs
,
and
are simple curves,
does not touch the interior of either
or
, and from each point in
there is a path to infinity that does not cross
.
Notice that, according to our definition, the interiors of the arcs
and
can touch.
If that happens, the double-points of the boundary (belonging to both arcs) are counted twice and considered as two distinct points (and are two different prime ends). The significance of the notion of admissible is that certain domains arising naturally in the proof of Theorem 7 (and thus in the proof of convergence to
) are not Jordan but are admissible – see Figure 19 .
Consider a sequence of admissible domains
with
distinct selected points
on
on each of their boundaries. If
is an admissible domain with
selected distinct points
on its boundary such that, as
,
and
, we say that
converges to
and write
.
Theorem 6.
Consider a sequence
of domains containing the origin, admissible with respect to
, and with
belonging to the interior of the counterclockwise arc
. Assume that
converges, as
, to
, where
is a domain containing the origin, admissible with respect to
, and
belongs to the interior of the counterclockwise arc
. Then, for any sequence
, the probability
of a blue crossing inside
from the counterclockwise segment
of
to the counterclockwise segment
of
converges, as
, to Cardy's formula
(see ( 16 )) for a crossing inside
from the counterclockwise segment
of
to the counterclockwise segment
of
.
Proof. We will construct for each small
, two domains with boundary points,
and
, approximating
in such a way that not only does
(by the continuity of Cardy's formula – see Lemma B.2 ) and the same for
, but also so that
|
(22)
|
This yields the desired result. The construction of the approximating domains uses fairly straightforward conformal mapping arguments. We provide details for
; the construction of
is analogous. Before providing the construction details for
, we give several paragraphs of overview.
To construct
, we will first need to take an inner approximation
of
, where
is a simple loop surrounded by
and with
. We will then construct a Jordan domain
whose boundary
is composed of pieces of
and of
, plus four segments joining
with
(see Figure 10 ), as we explain below.
Next, we will take an outer approximation
of
, where
is a simple loop surrounding
and with
. We will also need four simple curves
in the interior of the (topological) annulus between
and
and connecting their endpoints
on
and
with each of the four curves touching
at exactly one point which is either in the interior of the counterclockwise segment
(for
and
) or else the counterclockwise segment
(for
and
). Furthermore each of these connecting curves is close to its corresponding point
, or
; i.e.,
, etc. (see Figure 10 where each connecting segment represents “half ” of one of these connecting curves).
We will take
,
,
, and
with
the concatenation of:
from
to
, the portion of
from
to
counterclockwise,
from
to
, the portion of
from
to
counterclockwise,
from
to
, the portion of
from
to
counterclockwise,
from
to
, and the portion of
from
to
counterclockwise. It is important that (for fixed
and
) there is a strictly positive minimal distance between
and
, and between
and the union of the two counterclockwise segments
and
of
. These features will guarantee (as we explain with more detail below) that for fixed
, once
is large enough, a continuous curve within
that corresponds to the crossing event whose probability is
must have a subpath corresponding to the crossing event in
whose probability is
. This is the key feature of
, which will yield the first inequality of ( 22 ).
We now give a more detailed explanation of the construction of
. We first construct the parts of
that are inside
. Let
be the conformal map from
onto
with
and
, and consider the image
of the circle
under
and the inverse images,
, under
of
. Let
be the straight segment between
on the unit circle
, and
on the circle
, and define
, and
similarly, but using clockwise rotations by
and
for
and
(see Figure 11 ).
is a candidate for
and
is a candidate for half of
(where
or
or
or
), so we must choose
and the
's small enough so that
,
, etc. We then define
and similarly for
,
and
(see Figure 12 ).
Consider now the Jordan domain
whose boundary
is given by the concatenation of:
from
to
, the portion of
from
to
counterclockwise,
from
to
, the portion of
from
to
counterclockwise,
from
to
, the portion of
from
to
counterclockwise,
from
to
, and the portion of
from
to
counterclockwise (see Figure 13 ).
The exterior of this new Jordan domain
is a connected domain for which we can do a construction analogous to the one for the original domain
using a conformal mapfrom
to obtain candidates for
and for the exterior halves of the
's. To do this, we use a conformal map from
to the exterior of
and use
as replacements for
(see Figure 14 ).
Finally, we use the freedom to choose the exterior replacements for
and the
's differently from the interior values to make sure that the interior and exterior halves of the
's match up. We also choose the exterior values for
small enough so that
stays between
and
, between
and
, between
and
, between
and
(see Figure 15 ).
Once all the pieces of
are available, they are put together as explained above, and as show in Figures 16 and 17 below.
It should be clear that for a given approximation
of
constructed as described above there is a strictly positive
such that the distance between
and the portions of
that belong to
and
is not smaller than
, and the distance between the union
, two counterclockwise segments of
, and
is also not smaller than
. On the other hand, for any
, there exists
such that for all
,
is contained inside the
-neighborhood of
with the counterclockwise segment
(resp.,
) in the
-neighborhood of the counterclockwise segment
(resp.,
). This implies that for
large enough, any blue path crossing inside
from the counterclockwise segment
of
to the counterclockwise segment
of
must have a subpath that stays inside
and crosses between the counterclockwise segment
of
and thecounterclockwise segment
of
. Therefore the crossing probability
is a lower bound for
for all
, so that
|
(23)
|
as desired (the equality uses Smirnov's result, Theorem 4 , for fixed
).
We now note that as
,
. This allows us to use the continuity of Cardy's formula (Lemma B.2 in Appendix B ) to obtain
|
(24)
|
From this and ( 23 ) it follows that
|
(25)
|
The remaining part of the proof involves defining a domain
analogous to
but with the property that the probability
of an appropriate crossing, such that
|
(26)
|
is an upper bound for
for all
large enough. (The details of the construction of
are analogous to those of
; we leave them to the reader.) This shows that
|
(27)
|
which, combined with ( 25 ), implies
|
(28)
|
and concludes the proof.
We note that Theorem 6 , combined with the continuity of Cardy's formula in the shape of the domain (for admissible domains) and positions of the four points on the boundary (Lemma B.2 ), implies the convergence of crossing probabilities to Cardy's formula locally uniformly in the shape of the domain with respect to the uniform metric on curves, and in the location of the four points on the boundary with respect to the Euclidean metric; i.e., for
an admissible domain with
(with the notation used in Theorem 6 ),
,
and
such that for all admissible domains
with
and
,
, where
is Cardy's formula and
is the corresponding crossing probability.
Our next task is to show that the filling of any subsequential scaling limit of the percolation exploration process satisfies the spatial Markov property. Let us start withsome notation. First of all, suppose that
is a Jordan domain, and
and
are two points on
. Define the discrete filling (or simply filling) at time
of a percolation exploration path
(with a given parametrization) inside (the
-approximation
of )
from (the e-vertex closest to)
to (the e-vertex closest to)
to be the union of the hexagons explored up to time
and those unexplored hexagons from which it is not possible to reach
without crossing an explored hexagon or
(in other words, this is the set of hexagons that at time
have been explored or are disconnected from
by the exploration path).
Consider a sequence
of Jordan domains such that
, where
are two distinct points on
. Denote by
the percolation exploration path inside (the
-approximation
of )
from (the e-vertex closest to)
to (the e-vertex closest to)
.
Notice that we can couple the paths
simultaneously for all values of
and
by using the same percolation configuration to generate all of them. We can then apply the results of [2] to conclude that there exists a subsequence
such that the law of
converges to some limiting law for a process
supported on (Hölder) continuous curves inside
from
to
. The filling
of
, appearing in the next theorem, is defined just above Equation ( 17 ).
Theorem 7.
For any subsequential limit process
of the percolation exploration path
defined above, the filling
of
, as a process, satisfies the spatial Markov property.
Proof. Let
be a subsequence such that the law of
converges to some limiting law supported on continuous curves
inside
from
to
. We will prove the spatial Markov property by showing that
as defined in the proof of Theorem 5 are jointly distributed like the corresponding
hull variables, which do have the spatial Markov property. For each
, let
denote the filling at time
of
(with some parametrization – we do not need to worry about the choice of parametrization here). It follows from the Markovian character of the percolation exploration process that, for all
, the filling
of the percolation exploration path
satisfies a suitably adapted (to the discrete setting) spatial Markov property. (In fact, the percolation exploration path satisfies a stronger property – roughly speaking, that the future of the path given the filling of the past is distributed as a percolation exploration path in the original domain from which the filling of the past has been removed.) To be more precise, let
be a conformal transformation that maps
to
such that
and
and let
denote the first exit time of
from
defined as the first time an explored hexagon intersects the image under
of the semi-circle
. Define recursively
as the first exit time of
from
, where
is a conformal map from
to
that maps
to
and
to
. We also define
, so that
, and the (discrete-time) stochastic process
for
. The Markovian character of the percolation exploration process impliesthat, for every
,
is a Markov process (in
). In order to study the limit as
of
, we first need to analyze in more detail the mappings
.
The conformal transformations
are defined as follows, with an arbitrary multiplicative factor
. We choose
to be the composition
of two maps, where
,
is the conformal transformation that maps
onto
with
and
(we are assuming for simplicity that the domain
contains the origin; if that is not the case, one can think of a translated domain that does contain the origin), while
is the inverse of the transformation
|
(29)
|
that maps
onto
, where
is chosen so that
and
can be chosen so that
,
and
, which means that
maps
to
and
to
, so that
indeed maps
to
and
to
.
Since
converges in distribution to
, we can find two coupled versions of
and
on some probability space
such that
converges to
for all
; in the rest of the proof we work with these new versions which, with a slight abuse of notation, we denote with the same names as the original ones. Let
be a conformal transformation that maps
to
such that
and
and let
denote the first time
hits
, with
. Define recursively
as the first time
hits
, with
, where
is a conformal map from
to
that maps
to
and
to
. We also define
, so that
, and the (discrete-time) stochastic process
. As above, we choose the conformal transformation
to be the composition
of two maps, where
,
is the conformal transformation that maps
onto
with
and
(once again, we are assuming for simplicity that the domain
contains the origin; if that is not the case, one can think of a translated domain that does contain the origin), while
is the inverse of the transformation
|
(30)
|
that maps
onto
, where
is chosen so that
and
can be chosen so that
,
and
, which means that
maps
to
and
to
, so that
indeed maps
to
and
to
.
Analogous quantities can be defined for the trace of chordal
. For clarity, they will be indicated here by the superscript
; e.g.,
,
,
and
.We want to show recursively that, for any
, as
,
converge jointly in distribution to
. By recursively applying Theorem 6 and Lemma A.1 , we can then conclude that
are jointly equidistributed with the corresponding
hull variables (at the corresponding stopping times)
. Since the latter do satisfy the spatial Markov property, so will the former, as desired.
The zeroth step consists in noticing that the convergence of
to
as
allows us to apply Radó's theorem (i.e., Theorem 9 of Appendix B ) to show that
converges to
uniformly in
. This, together with the convergence of
to
and
to
, implies that
converges to
and
to
. Therefore, we also have the convergence of
to
and we can conclude that
converges to
uniformly on compact subsets of
, which implies that the boundary
of
converges to the boundary
of
in the uniform metric on continuous curves.
Starting from there, the first step of our recursion argument is organized as follows:
-
(1)
by “number of arms” percolation bounds [16] and Lemma A.2 below, but also
by Lemma A.1 (Theorem 6 is used here).
-
(2)
, but also
, by (1).
-
(3)
, but also
, by Corollary B.1 .
-
(4)
, but also
, by (3).
At this point, we are in the same situation as at the zeroth step, but with
,
and
replaced by
,
and
respectively, and we can proceed by recursion. As explained above, the theorem then follows from the fact that the
hull variables do posses the spatial Markov property.
In what follows we will show that the “number of arms” bounds [16] and Lemma A.2 below imply the convergence of
to
, and that the conditions to apply Theorem 6 , Lemma A.2 and Corollary B.1 are always satisfied. (This last point boils down to showing that the domains
and
are admissible for all
.) We begin by showing first that, as
,
converges in distribution to
(which also implies that
converges in distribution to
), from which it follows that
converges in distribution to
.
Consider
and
; they are both composed of two domains (which “meet” at
and
respectively), which we denote by
and
and by
and
, respectively (see Figure 18 ).
It follows from [2] that for some further subsequence
of the
's (which we denote by simply replacing
by
),
converge jointly in distribution to some limit;we already know that
must converge to
and want to use this fact and a suitably adapted triangular array version of Lemma 5.2 (whose validity does not rely on statement (S)) to conclude that the limit is unique and coincides with
. (Notice that
and
are two of the (many) domains in
produced by the exploration process started at
and stopped when it first hits the image under
of the semi-circle
, so we are in a context close to that of Lemma 5.2 ).
First of all, we need to show that the scaling limit of
touches the image of the semi-circle
under
at a single point. This follows immediately from the definition of the stopping time
for every fixed
(with the map
), but it could fail to be true in the limit
. The fact that it holds true in the limit is a direct consequence of Lemma A.2 below (in fact, a simpler version concerning Jordan domains would suffice here, but not when we iterate the argument – see below), which also implies that the single point at which the scaling limit of
touches the image of the semi-circle
under
coincides with the limit of
and with
.
Therefore, if we remove the single point
, the scaling limit of the boundary of
splits into a left and a right part (corresponding to the scaling limit of the leftmost yellow and the rightmost blue
-paths of hexagons explored by
, respectively) that do not touch the image of the semi-circle
under
.
Moreover, Lemma A.3 below implies that if
has a “close encounter” with
, then it touches
. Analogously, the standard bound on the probability of six crossings of an annulus [16] , used repeatedly before, implies that wherever
has a “close encounter” with itself, there is touching (see the proof of Lemma 5.1 ). These two observations assure that the scaling limit of
is almost surely a filling (of
), i.e., a closed connected set whose complement in
is simply connected. From the same bound on the probability of six crossings of an annulus, we can also conclude that the scaling limits of the left and right boundaries of
are almost surely simple (continuous) curves, as in the proof of Lemma 5.1 .
It is also possible to conclude that the intersection of the scaling limit of the left and right boundaries of
with the boundary of
almost surely does not contain arcs of positive length. In fact, if that were the case, it would be possible to find a subdomain
with three counterclockwise points
on its boundary such that the probability that an exploration path started at
and stopped when it first hits the arc
of
has a positive probability, in the scaling limit, of hitting at
or
, contradicting Cardy's formula (which, by Theorem 6 , holds for all subsequential scaling limits). This means that the scaling limit of
almost surely satisfies the condition in ( 17 ) and is therefore a hull. It says as well that, almost surely, the scaling limit
of
does not “stick” to the boundary of
, which implies that also
satisfies the condition in ( 17 ) and is therefore a hull.
It also implies that
and
are admissible domains since the part of the boundary of either
or
that belongs to the boundary of
can be split up, by removing the single point
, into two pieces which are, by an application of theproof of Lemma 5.1 , simple continuous curves, while the remaining part of the boundary of either
or
is a Jordan arc whose interior does not touch the hull
.
(Notice however, that they need not be Jordan domains because
has cut-points with positive probability – see Figure 19 ). This will be important later, when we need to apply Lemma A.1 (and therefore Theorem 6 ), Corollary B.1 and Lemmas A.2 - A.3 again.
Then, since hulls are characterized by their “envelope” (see Lemma A.1 and the discussion preceding it), the joint convergence in distribution of
to
would be enough to conclude that
converges to
as
, and in fact that
converges in distribution to
(and this will be valid also for the original subsequence
and not just for the further subsequence
). In order to get that, as explained before, we can use the convergence in distribution of
to
and apply almost the same arguments as used in the proof of Lemma 5.2 . The only difference is that, in proving claim (C), we cannot use the bound on the probability of three crossings of an annulus centered at a boundary point because we are not necessarily dealing with a convex domain. To replace that bound we use once again Lemmas A.2 - A.3 below (a simpler version concerning Jordan domains would again suffice here, but not when we iterate the argument – see below).
We can then conclude that
converges in distribution to
, which in turn implies the joint convergence in distribution of
to
and concludes the first step of the argument.
We next need to prove that
converges in distribution to
. Since we have already proved the convergence of
to
, we claim that all we really need to prove is the convergence of
to
. To see this, notice that
is distributed like the hull of a percolation exploration path inside
. Besides, the convergence in distribution of
to
implies that we can find versions of
and
on some probability space
such that
converges to
and
converges to
for all
. These two observations imply that, if we work with the coupled versions of
and
, we are in the same situation as before, but with
(resp.,
) replaced by
(resp.,
) and
(resp.,
) by
(resp.,
). As already remarked,
and
are admissible domains, which allows us to use Theorem 6 (and therefore Lemma A.1 ), Corollary B.1 and Lemmas A.2 - A.3 .
Then the conclusion that
converges in distribution to
follows from the same arguments as before, again using Corollary B.1 and Lemmas A.2 - A.3 . In order to get claim (C), in places where the exploration path comes close to the boundary of the past hull we can use the bound on the probability of six crossings of an annulus in the plane (as already seen in the case
of the proof of Theorem 1 ), while in places where it comes close to the remaining portionof the boundary (i.e.,
or the Jordan arc
in Figure 19 ) we can use Lemmas A.2 - A.3 .
We can now iterate those same arguments
times, for any
. It is in fact easy to see by induction that the domains
and
that appear in the successive steps are admissible for all
. Therefore we can keep using Theorem 6 (and therefore Lemma A.1 ), Corollary B.1 and Lemmas A.2 - A.3 . If we keep track at each step of the previous ones, in the spirit of Theorem 1 , this provides the joint convergence of all the curves and fillings involved at each step and concludes the proof of Theorem 7 .
Lemma A.2.
Let
be a sequence of domains admissible with respect to
and let
be the percolation exploration path in
started at
and stopped when it first hits the counterclockwise arc
of
. Assume that, as
,
converges to
, where
is a domain admissible with respect to
and
. Let
, where
is the event that
contains a segment that stays within
and has a double crossing of the annulus
without that segment touching
, and
is the event that
enters
, but is stopped outside
and does not touch
. Then, for any
,
|
(31)
|
Essentially, this means that as
(and
), it becomes increasingly unlikely that the exploration path ever comes close to
without quickly touching
nearby.
Lemma A.2 easily implies the following result, used in Theorem 7 to show that if
touches
, then
touches
nearby, for
large enough.
Lemma A.3.
With the notation and assumptions of Lemma A.2 ,
|
(32)
|
Proof. First of all notice that for
in
but not in
and not in
, the events
and
are exactly the same because the exploration path is, by definition, stopped on
. Therefore, we only have to prove that the event corresponding to the union over
of
has probability going to zero as
. We already know from Lemma A.2 that
happens with small probability for those points. This is, however, not sufficient because the exploration path could enter
, then exit
, and then re-enter it and touch
inside
,which is not an event in
. But such an event would imply that
first touches
inside one of the two balls of radius
centered at
and
, and by an application of Cardy's formula the probability that the latter happens goes to zero as
.
The proof of Lemma A.2 is partly based on relating the failure of ( 31 ) to the occurrence with strictly positive probability of certain continuum limit “mushroom” events (see Lemma A.5 ) that we will show must have zero probability because otherwise there would be a contradiction to Lemma A.4 , which itself is a consequence of the continuity of Cardy's formula with respect to the domain boundary. In both of the next two lemmas, we denote by
any subsequence limit of the probability measures for the collection of all colored (blue and yellow)
-paths on all of
, in the Aizenman-Burchard sense (see Remark 2.1 ). We recall that in our notation,
represents an open domain and
,
represent closed segments of its boundary. In Lemma A.4 below, we restrict attention to a Jordan domain
since that case suffices for the use of Lemma A.4 in the proof of Lemma A.2 .
Lemma A.4.
For
, with
a Jordan domain, consider the following crossing events,
, where
denotes either blue or yellow and
:
Then
.
Proof. The proof is similar to that of Theorem 6 , but easier because
is here a Jordan domain. Indeed, it is enough to construct a new Jordan domain
(with appropriately selected points
on the boundary and corresponding events
) such that the occurrence of
in
implies the occurrence of
in
and with
as
. The continuity of Cardy's formula (Lemma B.2 in Appendix B ) does the rest.
Lemma A.5.
For
as in Lemma A.2 ,
, and
, we define
, the yellow “mushroom” event (at
), to be the event that there is a yellow path in
from
to
and a blue path in
, between some pair of distinct points
in
, that passes through
and such that this blue path is between
and the yellow path (see Figure 20 ). We similarly define
with the colors interchanged and
where
denotes blue or yellow. Then for any deterministic domain
and any
,
.
Proof. If
for some
, then there is some segment
of
of diameter not larger than
such that
|
(33)
|
Choose any point
and consider the new domain
whose boundary consists of the correctly chosen (as we explain below) segment of the circle
between the two points
where
first hits
on either side of
, together with the segment from
to
of
(see Figure 21 ).
The correct circle segment between
and
is the (counter) clockwise one if
is between
and
along
when
is oriented (counter) clockwise. It is also not hard to see that since
,
is a Jordan domain, so that Lemma A.4 can be applied. In the new domain
,
is the same curve segment as it was in the old domain
, but
is now a segment of the circle
. It should be clear that
|
(34)
|
which yields a contradiction of Lemma A.4 if
.
Proof of Lemma A.2 . We first note that since the probability in ( 31 ) is nonincreasing in
, we may assume that
, as requested by Lemma A.5 .
Let us first consider the simpler case of
in which
. We follow the exploration process until time
, when it first touches
for some
, and consider the annulus
. Let
be the leftmost yellow
-path and
the rightmost blue
-path in
at time
that cross
.
and
split the annulus
into three sectors that, for simplicity, we will call the central sector, containing the crossing segment of the exploration path, the yellow (left) sector, with
as part of its boundary, and the blue (right) sector, the remaining one, with
as part of its boundary.
We then look for a yellow “lateral” crossing within the yellow sector from
to
and a blue lateral crossing within the blue sector from
to
. Notice that the yellow sector may contain “excursions” of the exploration path coming off
, producing nested yellow and blue excursions off
, and the same for the blue sector. But for topological reasons, those excursions are such that for every group of nested excursions, the outermost one is always yellow in the yellow sector and blue in the blue sector. Therefore, by standard percolation theory arguments, the conditional probability (conditioned on
at time
) to find a yellow lateral crossing of the yellow sector from
to
is bounded below by the probability to find a yellow circuit in an annulus with inner radius
and outer radius
. An analogous statement holds for the conditional probability (conditioned on
at time
and also on the entire percolation configuration in theyellow sector) to find a blue lateral crossing of the blue sector from
to
. Thus for any fixed
, by an application of the Russo-Seymour-Welsh lemma [27, 29] , the conditional probability to find both a yellow lateral crossing within the yellow sector from
to
and a blue lateral crossing within the blue sector from
to
goes to one as
.
But if such yellow and blue crossings are present, the exploration path is forced to touch
before exiting
, and if that happens, the exploration process is stopped, so that it will never exit
and the union over
of
cannot occur.
This concludes the proof of this case.
Let us now consider the remaining case in which
. The basic idea of the proof is then that by straightforward weak convergence and related coupling arguments, the failure of ( 31 ) would imply that some subsequence limit
would satisfy
, which would contradict Lemma A.5 . This is essentially because the close approach of an exploration path on the
-lattice to
without quickly touching nearby yields one two-sided colored
-path (the “perimeter” of the portion of the hull of the exploration path seen from a boundary point of close approach) and a one-sided
-path of the other color belonging to the percolation cluster not seen from the boundary point (i.e., shielded by the two-sided path). Both the two-sided path and the one-sided one are subsets of
.
Assume by contradiction that ( 31 ) is false, so that close encounters without touching happen with bounded away from zero probability. Consider for concreteness an exploration path
that has a close approach to a point
in the counterclockwise arc
. The exploration path may have multiple close approaches to
with differing colors of the perimeter as seen from
, but for topological reasons, the last time the exploration path comes close to
, it must do so in such a way as to produce a yellow
-path
(seen from
) that crosses
twice, and a blue path
that crosses it once (see Figure 22 ). This is so because the exploration process that produced
ended somewhere on
(and outside
), which is to the right of (i.e., clockwise to)
.
The presence of
implies that there are a yellow leftmost
-path
and a yellow rightmost
-path
(looking at
from inside
) crossing the annulus
.
The paths
and
split the annulus
into three sectors, that we will call the central sector, containing
, the left sector, with
as part of its boundary, and the right sector, with
as part of its boundary. Again for topological reasons, all other monochromatic crossings of the annulus are contained in the central sector, including at least one blue path
. As in the previous case, the left and right sectors can contain nested monochromatic excursions off
, but this time for every group of excursions, the outermost one is yellow in both sectors. Now consider the annulus
. We look for a yellow lateral crossing within the left sector from
to
and a yellow lateral crossing within the right sector from
to
. Since the outermost excursions in both sectors are yellow, the conditional probability to find a yellow lateral crossing within the left sector from
to
is bounded below by the probability to find a yellow circuit in an annulus with inner radius
and outer radius
, and an analogous statement holds for the conditional probability to find a yellow lateral crossing within the right sector from
to
. Thus for any fixed
, by an application of the Russo-Seymour-Welsh lemma [27, 29] , the conditional probability to find both yellow lateral crossings goes to one as
. But the presence of such yellow crossings would produce a (blue) mushroom event, leading to a contradiction with Lemma A.5 .
We are finally ready to prove the main result of this section which implies statement (S) at the beginning of Section 5 .
Corollary A.1.
Consider a sequence
of Jordan domains with two distinct selected points
on their boundaries
. Assume that
, where
is a Jordan domain with two distinct selected points on its boundary
. Denote by
the percolation exploration path inside (the
-approximation
of )
from (the e-vertex closest to)
to (the e-vertex closest to)
. Then for any sequence
, as
,
converges in distribution to the trace
of chordal
inside
from
to
.
Proof. It follows from [2] that
converges in distribution along subsequence limits
.
Since we have proved that the filling of any such subsequence limit
satisfies the spatial Markov property (Theorem 7 ) and the exit distribution of
is determined by Cardy's formula (Theorem 6 ), we can deduce from Theorem 5 that the limit is unique and that the law of
converges, as
, to the law of the trace
of chordal
inside
from
to
.
Appendix B: Sequences of Conformal Maps In this appendix, we give some results about sequences of conformal maps that are used in various places throughout the paper. For more details, the interested reader should consult [24] .
Definition B.1.
(see Section 1.4 of [
24]
) Let
be given and let
be domains with
. We say that
as
with respect to
in the sense of kernel convergence if
-
1.
either
, or else
is a domain
with
such that some neighborhood of every
lies in
for large
; and
-
2.
for
there exist
such that
as
.
It is clear from the definition that every subsequence limit also converges to
and it is also easy to see that the limit is uniquely determined. With this definition we can now state Carathéodory's kernel theorem [9] .
Theorem 8.
(see Theorem 1.8 of [
24]
) Let
map
conformally onto
with
and
. If
, let
; otherwise let
map
conformally onto
with
and
. Then, as
,
locally uniformly in
if and only if
with respect to
.
The next result, Radó's theorem [25] , deals with sequences of Jordan domains and is used in the main body of the paper. In this case the conformal maps have a continuous extension to
.
Theorem 9.
(see Theorem 2.11 of [
24]
) For
, let
and
be Jordan curves parametrized respectively by
and
,
, and let
and
be conformal maps from
onto the inner domains of
and
such that
and
,
for all
. If
as
uniformly in
then
as
uniformly in
.
The type of convergence of sequences of Jordan domains
to a Jordan domain
encountered in the main body of the paper (i.e., in the sense that
converges, as
, to
in the uniform metric ( 2 ) on continuous curves) is clearly sufficient to apply Theorem 9 . In Appendix A , however, we have to deal with domains that are not Jordan, and therefore we cannot use Radó's theorem. The tools needed to deal with those situations are described below.
Definition B.2.
(see Section 2.2 of [
24]
) The closed set
is called locally connected if for every
there is
such that, for any two points
with
, we can find a continuum
with diameter smaller than
and with
.
In the definition above, a continuum denotes a compact connected set with more than one point. We remark that every continuous curve (with more than one point) is a locally connected continuum (the converse is also true: every locally connected continuum is a curve). The concept of local connectedness gives a topological answer to the problem of global extension of a conformal map to the domain boundary, as follows.
Theorem 10.
(see Continuity Theorem in Section 2.1 of [
24]
) Let
map the unit disk
conformally onto
. Then the function
has a continuous extension to
if and only if
is locally connected.
When
has a continuous extension to
, we do not distinguish between
and its extension. This is always the case for the conformal maps considered in this paper. The problem wether this extension is injective on
has also a topological answer, as follows.
Theorem 11.
(see Carathhéodory Theorem in Section 2.1 of [
24]
) In the notation of Theorem 10 , the function
has a continuous and injective extension if and only if
is a Jordan curve.
When considering sequences of domains whose boundaries are locally connected the following definition is useful.
Definition B.3.
(see Section 2.2 of [
24]
) The closed sets
are uniformly locally connected if, for every
, there exists
independent of
such that any two points
with
can be joined by continua
of diameter smaller than
.
The convergence of domains used in this paper (i.e.,
if
in the uniform metric ( 2 ) on continuous curves) clearly implies kernel convergence, which immediately allows us to use Theorem 8 . However, we need uniform convergence in
.
This is guaranteed by Radó's theorem in the case of Jordan domains; in the non-Jordan case, sufficient conditions to have uniform convergence are stated in the next theorem.
Theorem 12.
(see Corollary 2.4 of [
24]
) Let
be a sequence of bounded domains such that, for some
,
for all
and such that
is uniformly locally connected. Let
map
conformally onto
with
. If
as
for each
, then the convergence is uniform in
.
In order to use Theorem 12 in Appendix A we need the following lemma. The definitions of admissible domain and the related notion of convergence are given just before Theorem 6 in Appendix A .
Lemma B.1.
Let
be a sequence of domains admissible with respect to
and assume that, as
,
, where
is a domain admissible with respect to
. Then the sequence of closed sets
is uniformly locally connected.
Proof. In order to prove the lemma, we claim that it suffices to focus on pairs of points on the boundaries, i.e., to show that:
for every
, there exists
independent of
such that any two points
with
can be joined by a continuum of diameter
contained in the complement
of
.
To verify our claim, let us assume
for the moment, and consider two points
(but not necessarily in
) with
, where
.
If (at least) one of the two points, say
, is at distance greater than
from
, then we can connect
and
using the closed ball of radius
centered at
, since
. If both points are at distance smaller than
from
, we can connect each point to a closest point on
by a straight segment of length smaller than
. Those two points on
can then be connected to each other by a continuum
of diameter
contained in
, and the union of
with the two straight segments gives a continuum of diameter
connecting
with
and contained in
.
We now prove
. Since
in the uniform metric ( 2 ) on continuous curves, for every
there exists
such that for all
,
. The admissibility of
implies that we can split its boundary into three Jordan arcs,
,
,
, such that
does not touch the interior of either
or
. We can do the same with
, letting
,
and
. Let
and
,
,
be parametrizations of
and
respectively, with
for
and
.
Let us assume, by contradiction, that
is false. Then there are indices
(actually
, but we abuse notation a bit) and points
with
(as
) that cannot be joined by a continuum of diameter
contained in
. By compactness considerations, we may assume that
and
as
, with
. Suppose that
and
belong to the interior of the same Jordan arc
for all
large enough.
Let
,
,
and
. It follows that
and
, and since
is a Jordan arc,
. For
large enough, the function
maps the closed segment of
between
and
onto a continuum in
containing
and
whose diameter tends to zero as
, leading to a contradiction with our assumption.
Similar reasoning gives a contradiction if
and
both belong to
or both belong to
for all
large enough, since the concatenation of
with
or of
with
is still a Jordan arc. The above reasoning applies except when
is on both
and
. When
, one can paste together small Jordan arcs on
and
to get a suitable continuum leading to a contradiction. The sole remaining case is when for all
large enough,
belongs to the interior of
and
belongs to the interior of
.
(Notice that we are ignoring the “degenerate” case in which
coincides with the “last” [from
] double-point on
, and
is a simple loop. In that case
and
could converge to
and
or
could still belong to
for arbitrarily large
's. However, in that case one can find two distinct points on
,
and
, such that
is admissible with respect to
, and points
and
on
converging to
and
respectively, and define accordingly new Jordan arcs,
and
, so that
and
for
large enough. We assume that this has been done if necessary, and for simplicity of notation drop the primes.) In this case let
denote the closed straight line segment in the plane between
and
. Imagine that
is oriented from
to
and let
be the first point of
intersected by
and
be the previous intersection of
with
. Clearly,
. For
large enough,
cannot belong to
either, or otherwise in the limit
,
would touch the interior of
and
. We deduce that for all
large enough,
. Since
and
are continuous curves and therefore locally connected,
and
belong to a continuum
contained in
whose diameter goes to zero as
, and the same for
and
(with
and
replaced by
and
).
Since the interior of
does not intersect any portion of
, it is either contained in
or in its complement
. If
, we have a contradiction since the union of
with
and
is contained in
and is a continuum containing
and
whose diameter goes to zero as
.
If the interior of
is contained in
, let us consider a conformal map
from
onto
. Since
is locally connected, the conformal map
extends continuously to the boundary of the unit disc. Let
,
,
,
and
. The points
are in counterclockwise order on
, so that any curve in
from
to the counterclockwise arc
must cross the curve from
to
whose image under
is
. This implies that any curve in
going from
to the counterclockwise arc
of
must cross the (interior of the) line segment
.
Then, in the limit
, any curve in
from
to the counterclockwise arc
must contain the limit point
. On the other hand, except for its starting and ending point, any such curve is completely contained in
, which implies that either
or else that (in the limit
) the counterclockwise arc
is the single point at
. We have already dealt with the former case. In the latter case, one can paste together small Jordan arcs from
to
, from
to
, and from
to
, and take the union with
and
(defined above) to get a suitable continuum in
containing
and
, leading to a contradiction. This concludes the proof.
Theorem 12 , together with Theorem 8 and Lemma B.1 , implies the following result, which is used in Appendix A .
Corollary B.1.
With the notation and assumptions of Lemma B.1 (and also assuming that
and
contain the origin), let
map
conformally onto
with
and
, and
map
conformally onto
with
and
. Then, as
,
uniformly in
.
Proof. As already remarked, the convergence of
to
in the uniform metric ( 2 ) on continuous curves (which is part of the definition of
) easily implies that the conditions in Carathéodory's kernel theorem (Theorem 8 ) are satisfied and therefore that
converges to
locally uniformly in
, as
. By an application of Lemma B.1 , the sequence
is uniformly locally connected, so that we can apply Theorem 12 to conclude that, as
,
converges to
uniformly in
.
We conclude this appendix with a simple lemma, used in the proof of Theorem 6 , about the continuity of Cardy's formula with respect to the shape of the domain and the positions of the four points on the boundary.
Lemma B.2.
For
and
as in Theorem 6 , let
denote Cardy's formula (see ( 16 )) for a crossing inside
from the counterclockwise segment
of
to the counterclockwise segment
of
and
the corresponding Cardy's formula for the limiting domain
.
Then, as
,
.
Proof. Let
be the conformal map that takes
onto
with
and
, and let
denote the conformal map from
onto
with
and
; let
,
,
,
,
,
,
, and
. We can apply Corollary B.1 to conclude that, as
,
converges to
uniformly in
. This, in turn, implies that, as
,
,
,
, and
.
Cardy's formula for a crossing inside
from the counterclockwise segment
of
to the counterclockwise segment
of
is given by
|
(35)
|
where
|
(36)
|
Because of the continuity of
in
,
,
,
, and the continuity of Cardy's formula ( 35 ) in
, the convergence of
,
,
and
immediately implies the convergence of
to
.
Acknowledgements. We are grateful to Greg Lawler, Oded Schramm and Wendelin Werner for various interesting and useful conversations and to Stas Smirnov for communications about a paper in preparation. We note in particular that a discussion with Oded Schramm at the November 2004 Northeast Probability Seminar at the CUNY Graduate Center, about dependence of exploration paths with respect to small changes of domain boundaries, pointed us in a direction that eventually led to Lemmas A.2 - A.5 . F. C. thanks Wendelin Werner for an invitation to Université Paris-Sud 11, and Vincent Beffara and Luiz Renato Fontes for many helpful discussions. We are especially grateful to Vincent Beffara for pointing out a gap in a preliminary version of Appendix A . We thank Lai-Sang Young for comments about the presentation of our results. F. C. acknowledges the kind hospitality of the Courant Institute where part of this work was completed.
References
-
M. Aizenman, Scaling limit for the incipient spanning clusters, in Mathematics of Multiscale Materials; the IMA Volumes in Mathematics and its Applications (K. Golden, G. Grimmett, R. James, G. Milton and P. Sen, eds.), Springer (1998).
-
M. Aizenman and A. Burchard, Hölder regularity and dimension bounds for random curves, Duke Math. J. 99, 419-453 (1999).
-
M. Aizenman, B. Duplantier and A. Aharony, Connectivity exponents and the external perimeter in
independent percolation, Phys. Rev. Lett. 83, 1359-1362 (1999).
-
A. A. Belavin, A. M. Polyakov and A. B. Zamolodchikov, Infinite conformal symmetry of critical fluctuations in two dimensions, J. Stat. Phys. 34, 763-774 (1984).
-
A. A. Belavin, A. M. Polyakov and A. B. Zamolodchikov, Infinite conformal symmetry in two-dimensional quantum field theory, Nucl. Phys. B 241, 333-380 (1984).
-
P. Billingsley, Weak Convergence of Measures: Applications in Probability, Society for Industrial and Applied Mathematics, Philadelphia (1971).
-
F. Camia and C. M. Newman, Continuum Nonsimple Loops and 2D Critical Percolation, J. Stat. Phys. 116, 157-173 (2004).
-
F. Camia, C. M. Newman and V. Sidoravicius, Cardy's formula for some dependent percolation models, Bull. Brazilian Math. Soc. 33, 147-156 (2002).
-
C. Carathéodory, Untersuchungen über die konformen Abbildungen von festen und veränderlichen Gebieten, Math. Ann. 72, 107-144 (1912).
-
J. L. Cardy, Critical percolation in finite geometries, J. Phys. A 25, L201-L206 (1992).
-
J. Cardy, Lectures on Conformal Invariance and Percolation, available at arXiv:math-ph/0103018 (2001).
-
R. Durrett, Probability: Theory and Examples, Third Edition, Duxbury Advanced Series (2004).
-
G. R. Grimmett, Percolation, second edition, Springer, Berlin (1999).
-
W. Kager and B. Nienhuis, A Guide to Stochastic Löwner Evolution and Its Applications, J. Phys. A 115, 1149-1229 (2004).
-
H. Kesten, Percolation Theory for Mathematicians, Birkhäuser, Boston (1982).
-
H. Kesten, V. Sidoravicius and Y. Zhang, Almost all words are seen in critical site percolation on the triangular lattice, Electr. J. Probab. 3, paper no. 10 (1998).
-
G. Lawler, Conformally Invariant Processes, in Lecture notes for the 2002 ICTP School and Conference on Probability, ICTP Lecture Notes Series, Vol. XVII (2004).
-
G. F. Lawler, Conformally Invariant Processes in the Plane, to be published by the American Mathematical Society (2005).
-
G. Lawler, O. Schramm and W. Werner, Values of Brownian intersection exponents I: Half-plane exponents, Acta Math. 187, 237-273 (2001).
-
G. Lawler, O. Schramm and W. Werner, One arm exponent for critical 2D percolation, Electronic J. Probab. 7, paper no. 2 (2002).
-
G. Lawler, O. Schramm and W. Werner, Conformal invariance of planar loop-erased random walk and uniform spanning trees, Ann. Prob. 32, 939-995 (2004).
-
G. Lawler, O. Schramm and W. Werner, Conformal restriction: the chordal case, J. Amer. Math. Soc. 16, 917-955 (2003).
-
A. M. Polyakov, Conformal symmetry of critical fluctuations, JETP Letters 12, 381-383 (1970).
-
Ch. Pommerenke, Boundary Behaviour of Conformal Maps, Springer-Verlag, Berlin (1992).
-
T. Radó, Sur la représentation conforme de domaines variables, Acta Sci. Math. (Szeged) 1, 180-186 (1923).
-
S. Rohde and O. Schramm, Basic properties of SLE, Ann. Math., to appear, preprint arXiv:math.PR0106036 (2003).
-
L. Russo, A note on percolation, Z. Wahrsch. Ver. Geb. 43, 39-48 (1978).
-
O. Schramm, Scaling limits of loop-erased random walks and uniform spanning trees, Israel J. Math. 118, 221-288 (2000).
-
P. D. Seymour, D. J. A. Welsh, Percolation probabilities on the square lattice, in Advances in Graph Theory (B. Bollobás ed.), Annals of Discrete Mathematics 3, North-Holland, Amsterdam, pp. 227-245 (1978).
-
S. Smirnov, Critical percolation in the plane: Conformal invariance, Cardy's formula, scaling limits, C. R. Acad. Sci. Paris 333, 239-244 (2001).
-
S. Smirnov, Critical percolation in the plane. I. Conformal invariance and Cardy's formula. II. Continuum scaling limit. (long version of [30] , dated Nov. 15, 2001), available at http://www.math.kth.se/
stas/papers/index.html.
-
S. Smirnov, in preparation.
-
S. Smirnov, private communication.
-
S. Smirnov and W. Werner, Critical exponents for two-dimensional percolation, Math. Rev. Lett. 8, 729-744 (2001).
-
W. Werner, Critical exponents, conformal invariance and planar Brownian motion, in Proceedings of the 3rd Europ. Congress of Math., Prog. Math., Vol. 202, 87-103 (2001).
-
W. Werner, SLEs as boundaries of clusters of Brownian loops, C. R. Math. Acad. Sci. Paris 337, 481–486 (2003).
-
W. Werner, Random planar curves and Schramm-Loewner Evolutions, in Lectures on probability theory and statistics, Lecture Notes in Math., Vol. 1840, Springer, Berlin, pp. 107-195 (2004).