Brownian half-plane excursion and critical Liouville quantum gravity
Abstract
In a groundbreaking work, Duplantier, Miller and Sheffield showed that subcritical Liouville quantum gravity (LQG) coupled with Schramm–Loewner evolutions (SLE) can be obtained by gluing together a pair of Brownian motions. In this paper, we study the counterpart of their result in the critical case via a limiting argument. In particular, we prove that as one sends in the subcritical setting, the space-filling SLE in a disk degenerates to the CLE (where CLE is conformal loop ensembles) exploration introduced by Werner and Wu, along with a collection of independent and identically distributed coin tosses indexed by the branch points of the exploration. Furthermore, in the same limit, we observe that although the pair of initial Brownian motions collapses to a single one, one can still extract two different independent Brownian motions from this pair, such that the Brownian motion encodes the LQG distance from the CLE loops to the boundary of the disk and the Brownian motion encodes the boundary lengths of the CLE loops. In contrast to the subcritical setting, the pair does not determine the CLE-decorated LQG surface. Our paper also contains a discussion of relationships to random planar maps, the conformally invariant CLE metric and growth fragmentations.
1 INTRODUCTION
The most classical object of random planar geometry is probably the two-dimensional Brownian motion together with its variants. Over the past 20 years, a plenitude of other interesting random geometric objects have been discovered and studied. Among those we find Liouville quantum gravity (LQG) surfaces [19] and conformal loop ensembles (CLE) [56, 61]. LQG surfaces aim to describe the fields appearing in the study of 2D LQG and can be viewed as canonical models for random surfaces. They can be mathematically defined in terms of volume forms [19, 31, 50] (used in this paper), but recently also in terms of random metrics [17, 26]. CLE is a random collection of loops that correspond conjecturally to interfaces of the -state Potts model and the FK random cluster model in the continuum limit (see, for example, [42]).
In this paper we study a coupling of LQG measures, CLE and Brownian motions, taking a form of the kind first discovered in [18]. On the one hand we consider a ‘uniform’ exploration of drawn on top of an independent LQG surface known as the critical LQG disk. On the other hand, we take a seemingly simpler object: the Brownian half-plane excursion. In this coupling one component of the Brownian excursion encodes the branching structure of the CLE exploration, together with a certain (LQG surface dependent) distance of CLE loops from the boundary. The other component of the Brownian excursion encodes the LQG boundary lengths of the discovered CLE loops.
Our result can be viewed as the critical () analog of Duplantier–Miller–Sheffield's mating of trees theorem for , [18]. The original mating of trees theorem first observes that the quantum boundary length process defined by a space-filling SLE (where SLE is Schramm–Loewner evolutions) curve drawn on a subcritical LQG surface is given by a certain correlated planar Brownian motion. Moreover, it says that one can take the two components of this planar Brownian motion, glue each one to itself (under its graph) to obtain two continuum random trees and then mate these trees along their branches to obtain both the LQG surface and the space-filling SLE curve wiggling between the trees in a measurable way. This theorem has had far-reaching consequences and applications, for example, to the study of random planar maps and their limits [23, 25, 30], SLE and CLE [3, 5, 20, 43], and LQG itself [4, 41]. See the survey [21] for further applications.
Obtaining a critical analog of the mating of trees theorem was one of the main aims of this paper. The problem one faces is that the above-described picture degenerates in many ways as (for example, the correlation of the Brownian motions tends to one and the LQG measure converges to the zero measure). However, it is known that the LQG measure can be renormalized in a way that gives meaningful limits [6], and the starting point of the current project was the observation that the pair of Brownian motions can be renormalized via an affine transformation to give something meaningful as well.
Still, not all the information passes nicely to the limit, and in particular extra randomness appears. Therefore, our limiting coupling is somewhat different in nature to that of [18] (or [2] for the finite volume case of quantum disks). Most notably, one of the key results of [2, 18] is that the CLE decorated LQG determines the Brownian motions, and vice versa. In our case neither statement holds in the same way; see Section 5.2.1 for more details. For example, to define the Brownian excursion from the branching CLE exploration, one needs a binary variable at every branching event to decide on an ordering of the branches.
We believe that in addition to completing the critical version of Duplantier–Miller–Sheffield's mating of trees theorem, the results of this paper are intriguing in their own right. Moreover, as explained below, this paper opens the road for several interesting questions in the realm of SLE theory, about LQG-related random metrics, in the setting of random planar maps decorated with statistical physics models, and about links to growth-fragmentation processes.
1.1 Contributions
Since quite some setup is required to describe our results for precisely, we postpone the detailed statement to Theorem 5.5. Let us state here a caricature version of the final statement. Some of the objects appearing in the statement will also be precisely defined only later, yet should be relatively clear from their names.
Theorem 1.1.Let
- be the field of a critical quantum disk together with associated critical LQG measures (see Section 4.1);
- denote the uniform space-filling in the unit disk parameterized by critical LQG mass, which is defined in terms of a uniform exploration plus a collection of independent coin tosses (see Section 2.1.5);
- and describe a Brownian (right) half-plane excursion (see Section 4.3).
- We show that a in the disk converges to the uniform CLE exploration introduced by Werner and Wu [64], as (Proposition 2.6). Here an extra level of randomness appears in the limit, in the sense that new CLE loops in the exploration are always added at a uniformly chosen point on the boundary, in contrast to the case where the loops are traced by a continuous curve.
- Using a limiting argument, we also show in Section 3 how to make sense of a ‘uniform’ space-filling SLE exploration, albeit no longer defined by a continuous curve. Again extra randomness appears in the limit: contrary to the case, the nested uniform CLE exploration does not uniquely determine this space-filling SLE.
- Perhaps less surprisingly but nonetheless not without obstacles, we show that the nested CLE in the unit disk converges to the nested CLE with respect to Hausdorff distance (Proposition 2.18). We also show that after dividing the associated quantum gravity measures by , a -LQG disk converges to a critical LQG disk.
- First, as stated above in Theorem 1.1, determines , but the opposite does not hold. A natural question is whether there is another natural mating of trees type theorem for where one has measurability in both directions.
- Second, our coupling sheds light on the recent work of Aïdékon and Da Silva [1] who identify a (signed) growth fragmentation embedded naturally in the Brownian half-plane excursion. The cells in this growth fragmentation correspond to very natural observables in our exploration.
- Third, as we have already mentioned, one of the coordinates in our Brownian excursion encodes a certain LQG distance of CLE loops from the boundary. It is reasonable to conjecture that this distance should be related to the CLE distance defined in [64] via a Lamperti transform.†
- Fourth, several interesting questions can be asked in terms of convergence of discrete models. Critical FK-decorated planar maps and stable maps are two immediate candidates.
1.2 Outline
The rest of the paper is structured as follows. In Section 2, after reviewing background material on branching SLE and CLE, we will prove the convergence of the exploration in the disk to the uniform CLE exploration, and also show the convergence of the nested CLE with respect to Hausdorff distance. In Section 3, we use the limiting procedure to give sense to a notion of space-filling SLE. In Section 4, we review the basics of LQG surfaces and of the mating of trees story, and prove convergence of the Brownian motion functionals appearing in [2, 18] after appropriate normalization. We also finalize a certain proof of Section 3, which is interestingly (seemingly) much easier to prove in the mating of trees context. Finally, in Section 5 we conclude the proof of joint convergence of Brownian motions, space-filling SLE and LQG. This allows us to state and conclude the proof of our main theorem. We finish the paper with a small discussion on connections, and an outlook on several interesting open questions.
2 CONVERGENCE OF BRANCHING SLE AND CLE AS
2.1 Background on branching SLE and conformal loop ensembles
2.1.1 Spaces of domains
- for every , and is simply connected planar domain;
- for all
- for every , if is the unique conformal map from to that sends 0 to 0 and has , then .
Recall that a sequence of simply connected domains containing 0 are said to converge to a simply connected domain in the Carathéodory topology (viewed from 0) if we have uniformly in for any , where (respectively, ) are the unique conformal maps from to (respectively, ) sending 0 to 0 and with positive real derivative at 0. Carathéodory convergence viewed from is defined in the analogous way.
2.1.2 Radial Loewner chains
Remark 2.1.Let us further remark that if are a sequence of driving measures as above, such that converges weakly (that is, with respect to the weak topology on measures) to some on for every , then the corresponding Loewner chains are such that in [37, Proposition 6.1]. In particular, one can check that if and for some piecewise continuous functions , and then the corresponding Loewner chains converge in if for any fixed and bounded and continuous, we have
Remark 2.2.In what follows we will sometimes need to consider evolving domains that satisfy the conditions to be an element of up to some finite time . In this case we may extend the definition of for by setting , where is the unique conformal map sending and with .With this extension, defines an element of .
If we have a sequence of such objects, then we say that they converge to a limiting object in if and only if these extensions converge. We will use this terminology without further comment in the rest of the paper.
2.1.3 Radial
Let , and recall the relationship (1.1) between and . Although the use of is somewhat redundant at this point, we do so to avoid redefining certain notations later on.
Usually we will start with , and then we say that the force point is at : everything in the above discussion remains true in this case; see [56]. In this setting we refer to and/or (interchangeably) as simply a radial targeted at 0.



2.1.4 An approximation to radial SLE
We will use the following approximations to in (in order to show convergence to the CLE exploration). Fixing , and taking the processes and as above, the idea is to remove intervals of time where is making tiny excursions away from 0, and then define to be the radial Loewner chain whose driving function is equal to , but with these times cut out.
2.1.5 Uniform exploration targeted at the origin
- sample random variables uniformly and independently on ;
- define for by setting
(2.7)for and ;
- let be the radial Loewner chain with driving function .
With these definitions we have that in as , where the limit process is the uniform CLE exploration introduced in [64], and run until the outermost CLE loop surrounding 0 is discovered.
Since we are considering processes in , we need to reparameterize by seen from the origin. By definition, for each , is a simple loop rooted at a point in that does not surround 0. If we declare the loop to be traversed counterclockwise, we can view it as a curve parameterized so that for all (the choice of direction means that is surrounded by the left-hand side of ). We then define to be the unique process in such that for each with , and all , is the connected component of containing 0. In other words, is a reparameterization of by seen from 0, where instead of loops being discovered instantaneously, they are traced continuously in a counterclockwise direction. The process is defined until time , at which point the origin is surrounded by a loop (the law of this loop is that of the outermost loop surrounding the origin in a nested CLE in ).
With this definition, the same argument as in [64, Section 4] shows that in as . Moreover, this convergence in law holds jointly with the convergence (in particular, has the law of the first time that a reflected Brownian motion started from 0 hits , as was already observed in [52]).
The exploration can be continued after this first loop exploration time by iteration. More precisely, given the process up to time , one next samples an independent exploration in the interior of the discovered loop containing 0, but now with loops traced clockwise instead of counterclockwise. When the next-level loop containing 0 is discovered, the procedure is repeated, but going back to counterclockwise tracing. Continuing in this way, we define the whole uniform CLE exploration targeted at 0: . Note that by definition is then just the process , stopped at time .
Remark 2.3.The ‘clockwise/counterclockwise’ switching defined above is consistent with what happens in the picture when . Indeed, it follows from the Markov property of (in the case) that after time , the evolution of until it next hits 0 is independent of the past and equal in law to . This implies that the future of the curve after time has the law of an in the connected component of the remaining domain containing 0, but now with force point starting infinitesimally counterclockwise from the tip, until 0 is surrounded by a clockwise loop. This procedure alternates, just as in the case.
2.1.6 Exploration of the (nested) CLE
In the previous subsections, we have seen how to construct processes, denoted by () from 1 to 0 in , and that these are generated by curves . We have also seen how to construct a uniform exploration, , targeted at 0 in . The 0 in the subscripts here is to indicate that 0 is a special target point. But we can also define the law of an , or a exploration process, targeted at any point in the unit disk. To do this we simply take the law of or , where is the unique conformal map sending 0 to and 1 to 1. We will denote these processes by , where the are also clearly generated by curves for . By definition, the time parameterization for is such that for all (similarly for ).
In fact, both and the uniform exploration satisfy a special target invariance property; see, for example, [53] for and [64, Lemma 8] for CLE. This means that they can be targeted at a countable dense set of point in simultaneously, in such a way that for any distinct , the processes targeted at and agree (modulo time reparameterization) until the first time that and lie in different connected components of the yet-to-be-explored domain. We will choose our dense set of points to be , and for refer to the coupled process (or ) as the branching in . Similarly, we refer to the coupled process as the branching exploration in .
Note that in this setting we can associate a process to each : we consider the image of under the unique conformal map from sending and , and define to be the unique process such that this new radial Loewner chain is related to via Equations (2.6) and (2.3). Note that has the same law as for each fixed (by definition), but the above procedure produces a coupling of .
We will use the following property connecting chordal and radial SLE (that is closely related to target invariance).
Lemma 2.4. ([[53], Theorem 3])Consider the radial with force point at for , stopped at the first time that and 0 are separated. Then its law coincides (up to a time change) with that of a chordal SLE from 1 to in , stopped at the equivalent time.
We remark that from , we can almost surely define a curve for any fixed , by taking the almost sure limit (with respect to the supremum norm on compacts of time) of the curves , where is a sequence tending to as . This curve has the law of an from 1 to in [45, Section 2.1]. Let us caution at this point that such a limiting construction does not work simultaneously for all . Indeed, there are almost surely certain exceptional points , the set of which almost surely has Lebesgue measure zero, for which the limit of does not exist for some sequence ; see Figure 4.

Let us now explain how, for each , we can use the branching to define a (nested) . The conformal loop ensemble in is a collection of non-crossing (nested) loops in the disk, [61], whose law is invariant under Möbius transforms . The ensemble can therefore be defined in any simply connected domain by conformal invariance, and the resulting family of laws is conjectured (in some special cases proved, for example, [8, 16, 22, 33, 63]) to be a universal scaling limit for collections of interfaces in critical statistical physics models.
- Let be the first time that hits , and let be the last time before this that is equal to 0.
- Let . In fact, point is one of the exceptional points for which the limit of does not exist for all sequences , so it is not immediately clear how to define ; see Figure 4. However, the limit is well defined if we insist that the sequence is such that 0 and are separated by at time for each .
- Define to be the limit of the curves as . In particular the condition on the sequence means that is almost surely a double point of . With this definition of , it follows that
- Set .
We write for the connected component of containing : note that this is equal to . We will call this the (outermost) interior bubble containing .
We define the sequence of nested loops for by iteration (so ), and denote the corresponding sequence of nested domains (interior bubbles) containing by . More precisely, the th loop is defined inside in the same way that the first loop is defined inside , after mapping conformally to and considering the curve rather than .
The uniform exploration defines a nested in a similar but less complicated manner; see [64]. For any , to define (the outermost loop containing ) we consider the Loewner chain and define the times and (according to ) as in the case. Then between times and the Loewner chain is tracing a simple loop — starting and ending at a point . This loop is what we define to be . We define to be the interior of : note that this is also equal to . Finally, we define the nested collection of loops containing and their interiors by iteration, denoting these by (so and ).
2.1.7 Space-filling SLE
Now, for we can also use the branching SLE, , to define a space-filling curve known as space-filling SLE. This was first introduced in [18, 39]; see also [10, Appendix A.3] for the precise definition of the space-filling loop that we will use. The presentation here closely follows [21].
In our definition, the branches of are all processes started from point 1, and with force points initially located infinitesimally clockwise from 1. This means that the associated space-filling SLE will be a so-called counterclockwise space-filling loop from 1 to 1 in .†
Given an instance of a branching SLE, to define the associated space-filling SLE, we start by defining an ordering on the points of . For this we use a coloring procedure. First, we color the boundary of blue. Then, for each , we can consider the branch of the branching SLE targeted toward . We color the left-hand side of red, and the right-hand side of blue. Whenever disconnects one region of from another, we can then label the resulting connected components as monocolored or bicolored, depending on whether the boundaries of these components are made up of one or two colors, respectively.
For and distinct elements of , we know (by definition of the branching SLE) that and will agree until the first time that and are separated. When this occurs, it is not hard to see that precisely one of or will be in a newly created monocolored component. If this is we declare that , and otherwise that . In this way, we define a consistent ordering on ; see Figure 5.

- We can think of as a version of ordinary that iteratively fills in bubbles, or disconnected components, as it creates them. The ordering means that it will fill in monocolored components first, and come back to bicolored components only later.
- The word counterclockwise in the definition refers to the fact that the boundary of is covered up by in a counterclockwise order.
2.2 Convergence of the SLE branches
- the branch toward to the exploration branch toward ; and
- the nested loops surrounding to the nested loops surrounding .
Let us assume without loss of generality that our target point is the origin. We first consider the radial branch targeting 0, , up until the first time that 0 is surrounded by a counterclockwise loop. The basic result is as follows.
Proposition 2.5. in as .
By Remark 2.3 and the iterative definition of the exploration toward 0, the convergence for all time follows immediately from the above.
Proposition 2.6. in as .
Our proof of Proposition 2.5 will go through the approximations and . Namely, we will show that for any fixed level of approximation, as , equivalently . Broadly speaking, this holds since the macroscopic excursions of the underlying processes converge, and in between these macroscopic excursions we can show that the location of the tip of the curve distributes itself uniformly on the boundary of the unexplored domain. We combine this with the fact that the approximations converge to as , uniformly in , to obtain the result.
The heuristic explanation for the mixing of the curve tip on the boundary is that the force point in the definition of an causes the curve to ‘whizz’ around the boundary more and more quickly as . This means that in any fixed amount of time (for example, between macroscopic excursions), it will forget its initial position and become uniformly distributed in the limit. Making this heuristic rigorous is the main technical step of this subsection, and is achieved in Section 2.2.3.
2.2.1 Excursion measures converge as
The first step toward proving Proposition 2.5 is to describe the sense in which the underlying process for the branch converges to the process for the CLE exploration. It is convenient to formulate this in the language of excursion theory; see Lemma 2.8.
To begin we observe, and record in the following remark, that when is very small, it behaves much like a Bessel process of a certain dimension.
Remark 2.7.Suppose that . By Girsanov's theorem, if the law of is weighted by the martingale
Now, observe that by the Markov property of , we can define its associated (infinite) excursion measure on excursions from 0. We define to be the image of this measure under the operation of stopping excursions if and when they reach height .
For , we write for restricted to excursions with maximum height exceeding , and normalized to be a probability measure. It then follows from the strong Markov property that the excursions of during the intervals are independent samples from , and is the index of the first of these samples that actually reaches height . We also write for the measure restricted to excursions that reach , again normalized to be a probability measure.
Finally, we consider the excursion measure on excursions from 0 for Brownian motion. We denote the image of this measure, after stopping excursions when they hit , by . Analogously to above, we write for conditioned on the excursion exceeding height . We write for conditioned on the excursion reaching height .
Lemma 2.8.For any , in law as , with respect to . The same holds with in place of .
Proof.For , set , and equip it with the metric . Set , recalling the definition . We first state and prove the analogous result for Bessel processes.
Lemma 2.9.Let be a sample from the Bessel- excursion measure away from 0, conditioned on exceeding height , and stopped on the subsequent first hitting of 0 or . Let be a sample from the Brownian excursion measure with the same conditioning and stopping.† Then for any , as , in the space .
Proof of Lemma 2.9.For any , can be sampled (see [18, Section 3]) by
- first sampling from the probability measure on with density proportional to ;
- then running a Bessel- process from 0 to ;
- stopping this process at if ; or
- placing it back to back with the time reversal of an independent Bessel- from 0 to if .
Now we continue the proof of Lemma 2.8. Recalling the Radon–Nikodym derivative of Remark 2.7 (note that as ), we conclude that if and are sampled from and , respectively, and stopped upon hitting for the first time after hitting , then in law as , in the space .
2.2.2 Strategy for the proof of Proposition 2.5
With Lemma 2.8 in hand the strategy to prove Proposition 2.5 is to establish the following two lemmas.
Lemma 2.10.Let be a continuous bounded function on . Then as , uniformly in , equivalently .
Proof.Fix as above, and let us assume that the processes as varies and are coupled together in the natural way: using the same underlying and . By Remark 2.1, in particular (2.4), it suffices to prove that
To do this, let us consider the total (that is, cumulative) duration of such excursions of , before the first time that reaches . The reason for restricting to this time interval is to use the final observation in Remark 2.7: that the integrand in the definition of is deterministically bounded up to time . This will allow us to transfer the question to one about Bessel processes. And, indeed, since the number of times that will reach before time is a geometric random variable with success probability uniformly bounded away from 0 (due to Lemma 2.8), it is enough to show that tends to 0 in probability as , uniformly in .
For this, we first note that by Remark 2.7, for any we can write
To this end, we begin by using Cauchy–Schwarz to obtain the upper bound
Recall that under , has the law of times a Bessel process of dimension . Now, by [47, Theorem 1] we can construct a dimension Bessel process by concatenating excursions from a Poisson point process with intensity times Lebesgue measure on , where is a probability measure on Bessel excursions with maximum height for each . Moreover, by Brownian scaling, , for . (For proofs of these results, see, for example, [47].)
Now, if we let , then conditionally on , we can write as the sum of the excursion lifetimes over points in a (conditionally independent) Poisson point process with intensity
Lemma 2.11.For any fixed , converges to in law as , with respect to the Carathéodory Euclidean topology.
2.2.3 Convergence at a fixed level of approximation as
The remainder of this section will now be devoted to proving Lemma 2.11. This is slightly trickier, and so we will break down its proof further into Lemmas 2.12 and 2.13.
Thus, live in the space of sequences of finite length, taking values in . We equip this space with topology such that as if and only if the vector length of is equal to the length of for all large enough, and such that every component of (for ) converges to the corresponding component of with respect to the Euclidean distance. Similarly, live in the space of sequences of finite length, taking values in the space of excursions away from .
We equip this sequence space with topology such that as if and only if the vector length of is equal to the vector length of for all large enough, together with component-wise convergence with respect to .
Lemma 2.12.For any , as .
Proof.This is a direct consequence of Lemma 2.8 and the definition of .
Lemma 2.13.For any , in law as .
This second lemma will take a bit more work to prove. However, we can immediately see how the two together imply Lemma 2.11.
Proof of Lemma 2.11.Lemmas 2.12 and 2.13 imply that the driving functions of converge in law to the driving function of with respect to the Skorokhod topology. This implies the result by Remark 2.1.
Our new goal is therefore to prove Lemma 2.13. The main ingredient is the following (recall that is the start time of the first excursion of away from 0 that reaches height ).
Lemma 2.14.For any and fixed,
For the proof of Lemma 2.14, we are going to use Remark 2.7. That is, the fact that behaves very much like times a Bessel process of dimension . The Bessel process is much more convenient to work with (in terms of exact calculations) because of its scaling properties. Indeed, for Bessel processes we have the following lemma:
Lemma 2.15.Let be times a Bessel process of dimension (started from 0) and be the start time of the first excursion in which it exceeds . Then for ,
(The assumption that is sufficiently large here is made simply for convenience of proof.)
Proof.By changing the value of appropriately, we can instead take to be a Bessel process of dimension (that is, we forget about the multiplicative factor of ). Note that for and as . By standard Itô excursion theory, can be formed by gluing together the excursions of a Poisson point process with intensity , where is the Bessel- excursion measure. As mentioned previously, it is a classical result that we can decompose (there is a multiplicative constant that we can set to one without loss of generality) where is a probability measure on excursions with maximum height exactly for each and that moreover by Brownian scaling, , for .
Let
The real part of is bounded above by . Then using the Brownian scaling property of explained before, we can bound by . Using the fact that , which can be obtained from a direct calculation, it follows that for all , where does not depend on (say). This allows us to take expectations over in (2.12) (recall the distribution of from (2.11)) to obtain that
We now fix and for the rest of the proof. Our aim is to show that the final expression in (2.13) converges to 0 as (equivalently ). To do this, we use the Brownian scaling property of again to write for each . We also observe that
With this in hand, the proof of Lemma 2.14 follows in a straightforward manner.
Proof of Lemma 2.14.In order to do a Bessel process comparison and use Lemma 2.15, we need to replace the fixed in (2.10) by some which is very large (so we are only dealing with time intervals where is tiny). However, this is not a problem, since for we can write
So we can write, for any
Proof of Lemma 2.13.Equation (2.10) implies that the law of converges to the uniform distribution on the unit circle as . The full result then follows by the Markov property of .
2.2.4 Summary
So, we have now tied up all the loose ends from the proof of Proposition 2.5. Recall that this proposition asserted the convergence in law of a single branch in , targeted at 0, to the corresponding uniform CLE exploration branch. Let us conclude this subsection by noting that the same result holds when we change the target point.
For not necessarily equal to 0, we define to be the space of evolving domains whose image after applying the conformal map from , , lies in .
From the convergence in Proposition 2.6, plus the target invariance of radial and the uniform CLE exploration, it is immediate that
Corollary 2.16.For any , in as .
Recall that is the last time that hits 0 before first hitting and is the time interval during which traces the outermost CLE loop surrounding . Note that is equal to the length of the excursion and similarly is the length of the excursion (for every ), so that by Lemma 2.12 the following extension holds.
Corollary 2.17.For any fixed
2.3 Convergence of the CLE loops
In this subsection we will prove the convergence of with respect to the Hausdorff distance. That this might be non-obvious is illustrated by the following difference: in the limit , whereas this is not at all the case for . Nevertheless, we have
Proposition 2.18.For any
Given (2.14), and that we already know the convergence of as , the proof of Proposition 2.18 boils down to the following lemma.
Lemma 2.19.Suppose that is a subsequential limit in law of as (with the topology of Proposition 2.18). Then we have almost surely.
Proof of Proposition 2.18 given Lemma 2.19.By conformal invariance we may assume that . Observe that by Corollary 2.16, we already know that as , with respect to the product ( Carathéodory ) topology. Indeed, if one takes a sequence converging to 0, and a coupling of and so that almost surely as , it is clear due to (2.14) that each also converges to almost surely. Also note that is tight in with respect to the Hausdorff topology, since all the sets in question are almost surely contained in . Thus is tight in the desired topology, and the limit is uniquely characterized by the above observation and Lemma 2.19. This yields the proposition.
2.3.1 Strategy for the proof of Lemma 2.19
At this point, we know the convergence in law of as , and we know that is the connected component of containing 0 for every . Given a subsequential limit in law of , the difficulty in concluding that lies in the fact that Carathéodory convergence (which is what we have for ) does not ‘see’ bottlenecks; see Figure 6.

To proceed with the proof, we first show that any part of the supposed limit that does not coincide with must lie outside of .
Lemma 2.20.With the setup of Lemma 2.19, we have almost surely.
Once we have this ‘one-sided’ result, it suffices to prove that the laws of and coincide.
Lemma 2.21.Suppose that is as in Lemma 2.19. Then the law of is equal to the law of .
The first lemma follows almost immediately from the Carathéodory convergence of (see the next subsection). To prove the second lemma, we use the fact that for is inversion invariant: more correctly, a whole-plane version of is invariant under the mapping . Roughly speaking, this means that for whole-plane CLE, we can use inversion invariance to obtain the complementary result to Lemma 2.20, and deduce Hausdorff convergence in law of the analogous loops. We then have to do a little work, using the relation between whole-plane CLE and CLE in the disk (a Markov property), to translate this back to the disk setting and obtain Lemma 2.21.
2.3.2 Preliminaries on Carathéodory convergence
We first record the following standard lemma concerning Carathéodory convergence, which will be useful in what follows.
Lemma 2.22. (Carathéodory kernel theorem)Suppose that is a sequence of simply connected domains containing 0, and for each , write for the connected component of the interior of containing 0. Define the kernel of to be if this is non-empty, otherwise declare it to be .
Suppose that and are simply connected domains containing 0. Then with respect to the Carathéodory topology (viewed from 0) if and only if every subsequence of the has kernel .
One immediate consequence of this is the following.
Corollary 2.23.Suppose that as for the product (Hausdorff Carathéodory topology), where for each fixed , the coupling of and is such that is a simply connected domain with , and is a compact subset of with almost surely. Then almost surely.
Proof.By Skorokhod embedding, we may assume without loss of generality that almost surely as .
For write for the connected component of containing 0. By assumption, for every almost surely, which means that for all almost surely. Since converges to in the Hausdorff topology, we have for each , which implies that almost surely. Finally, because in the Carathéodory topology, the Carathéodory kernel theorem gives that almost surely. Hence almost surely, as desired.
In particular:
Proof of Lemma 2.20.This is a direct consequence of Corollary 2.23.
Now, if are such that is a simply connected domain containing 0 for each , we say that with respect to the Carathéodory topology seen from , if and only if with respect to the Carathéodory topology seen from 0. It is clear from this definition and the above arguments (or similar) that the following properties hold.
Lemma 2.24.Suppose that are simply connected domains such that is simply connected containing 0 for each . Then
- if jointly with respect the product (Carathéodory seen from Hausdorff) topology, for some compact sets with for each , then almost surely;
- if jointly with respect the product (Carathéodory seen from Carathéodory seen from 0) topology, for some simply connected domains with for each , then almost surely.
Proof.The first bullet point follows from Corollary 2.23 by considering and . For the second, let us assume by Skorohod embedding that almost surely in the claimed topology. Then the compact sets are tight for the Hausdorff topology, and hence have some subsequential limit . (The argument of) Corollary 2.23 implies that and almost surely. Since is an open simply connected domain containing and is an open simply connected domain containing 0, this implies that almost surely.
2.3.3 Whole-plane CLE and conclusion of the proofs
As mentioned previously, we would now like to use some kind of symmetry argument to prove Lemma 2.21. However, the symmetry we wish to exploit is not present for CLE in the unit disk, and so we have to go through an argument using whole-plane CLE instead. Whole-plane CLE was first introduced in [34] and is, roughly speaking, the local limit of CLE in (any) sequence of domains with size tending to . The key symmetry property of whole-plane CLE that we will use is its invariance under applying the inversion map [27, 34]. More precisely:
Lemma 2.25.Let be a whole-plane with .
- (Inversion invariance) The image of under has the same law as .
- (Markov property) Consider the collection of loops in that lie entirely inside and surround 0. Write (with as usual) for the connected component containing 0 of the complement of the outermost loop in this collection. Write for the second outermost loop in this collection. Then the image of under the conformal map sending to 0 with positive derivative at 0 has the same law as the outermost loop surrounding 0 for a in .
Proof.The inversion invariance is shown in [34, Theorem 1.1] for and [27, Theorem 1.1] for . The Markov property follows from [27, Lemma 2.9] when and [34, Theorem 1] when .
Let us now state the convergence result that we will prove for whole-plane CLE as , and show how it implies Lemma 2.21.
For , we extend the above definitions and write for the largest and second largest whole-plane loops containing 0, which are entirely contained in the unit disk. We let be the connected component of containing 0 for and let be the connected component containing . When we write for the corresponding loops of a whole-plane , and for the corresponding domains containing 0 and . Note that in this case we have and for .
Lemma 2.26. as , with respect to the product
Carathéodory (seen from in the four coordinates) topology.
Proof of Lemma 2.21 given Lemma 2.26.Suppose that converges in law to along some subsequence, with respect to the product (Carathéodory seen from 0 Hausdorff) topology. By the above lemma, we can extend this convergence to the joint convergence of . But then Corollary 2.23 and Lemma 2.24 imply that and almost surely. This implies that almost surely. Moreover, it is not hard to see (using the definition of Hausdorff convergence) that , else would not disconnect 0 from for small . So almost surely.
Now consider, for each , the unique conformal map that sends and has . Then the above considerations imply that if converges in law along some subsequence, with respect to the Hausdorff topology, then the limit must have the law of , where is defined in the same way as but with replaced by . Since the law of is the same as that of for every and the law of has the law of , this proves Lemma 2.21.
Proof of Lemma 2.19 and Proposition 2.18.Combining Lemmas 2.20 and 2.21 yields Lemma 2.19. As explained previously, this implies Proposition 2.18.
So, we are left only to prove Lemma 2.26, concerning whole-plane . We will build up to this with a sequence of lemmas: first proving convergence of nested loops in very large domains, then transferring this to whole-plane CLE and finally appealing to inversion invariance to obtain the result.
Lemma 2.27.Fix . For and a in , denote by the sequence of nested loops containing 0, starting with the second smallest loop to fully enclose the unit disk (set equal to the boundary of if only one or no loops in actually surround ) and such that surrounds for all . Write for the connected components containing 0 of the complements of the . Then converges in law to its CLE counterpart as , with respect to the product Carathéodory topology viewed from 0.
Proof.By Corollary 2.16 and scale invariance of CLE, together with the iterative nature of the construction of nested loops, we already know that the sequence of nested loops in containing 0, starting from the outermost one, converges as , with respect to the product Carathéodory topology viewed from 0. Taking a coupling where this convergence holds almost surely, it suffices to prove that the index of the smallest loop containing the unit disk also converges almost surely. This is a straightforward consequence of the kernel theorem — Lemma 2.22 — plus the fact that the smallest loop in that contains actually contains for some strictly positive almost surely.
Lemma 2.28.The statement of the above lemma holds true if we replace the CLEs in with whole-plane versions.
Proof.For fixed , let , denote whole-plane CLE and on , respectively. The key to this lemma is [46, Theorem 9.1], which states (in particular) that rapidly converges to in the following sense. For some , and can be coupled so that for any and , with probability at least , there is a conformal map from some to , which maps the nested loops of — starting with the smallest containing — to the corresponding nested loops of , and has low distortion in the sense that on .
In fact, it is straightforward to see that and (which in principle depend on ) may be chosen uniformly for (say). Indeed, it follows from the proof in [46] that they depend only on the law of the log conformal radius of the outermost loop containing 0 for a in , and this varies continuously in , [52]. Hence, the result follows by letting in Lemma 2.27 and noting that the second smallest loop containing is contained in with arbitrarily high probability as , uniformly in .
Proof of Lemma 2.26.Lemmas 2.28 and 2.25 (inversion invariance) imply that and as . This ensures that is tight in , so we need only prove that if is a subsequential limit of , then and almost surely. Note that has the same law as , and since for all , Lemma 2.24 implies that . In other words almost surely. Then because and have the same law, we may deduce that they are equal almost surely. Similarly, we see that almost surely.
2.3.4 Conclusion
Recall that for , (respectively, ) denotes the sequence of nested (respectively, ) bubbles and loops containing . By the Markov property and iterative nature of the construction, it is immediate from Proposition 2.18 that
Corollary 2.29.For fixed
3 THE UNIFORM SPACE-FILLING SLE
In this section we show that the ordering on points (with rational coordinates) in the disk, induced by space-filling with , converges to a limiting ordering as . We call this the uniform space-filling SLE.† Nonetheless, we can describe explicitly the law of this ordering, which for any two fixed points comes down to the toss of a fair coin. As for , there would be other ways to define a space-filling SLE process, by considering different explorations of CLE.
Let us now recall some notation in order to properly state the result. For and , we define to be the indicator function of the event that the space-filling hits before (see Section 2.1.7). By convention we set this equal to 1 when .
- for all almost surely;
- is a Bernoulli() random variable for all with ;
- for all with almost surely;
- for all with , if separates from at the same time as it separates from then , otherwise and are independent.
Lemma 3.1.There is a unique joint law on satisfying the above requirements, and such that the marginal law of is that of a branching uniform exploration. With this law, almost surely defines an order on any finite subset of by declaring that if and only if .
We will prove the lemma in just a moment. The main result of this section is the following.
Proposition 3.2. converges to , in law as , with respect to the product topology , where is as defined in Lemma 3.1.
Proof of Lemma 3.1.The main observation is that if a joint law as in the lemma exists, then for all we almost surely have
We now show why this implies that for any with distinct, there exists a unique a conditional law on given , satisfying the requirements of the lemma. We argue by induction on the number of points. Indeed, suppose it is true with for some and take in distinct. We construct the conditional law of given as follows.
- To define :
- – partition the indices into equivalence classes such that if and only if separates from and at the same time;
- – for each equivalence class sample an independent Bernoulli random variable; and
- – set to be the random variable associated with class for every .
- Given and , define with by setting it equal to if and are separated from at the same time, or if and are separated from at the same time.
- For each consider the connected component in the branching exploration that contains points with when they are separated from . The explorations inside these components are mutually independent, independent of the exploration before this separation time, and each has the same law as after mapping to the unit disk. Thus, since each equivalence class contains strictly less than points, using the induction hypothesis, we can define for such that
- – the collections for different are mutually independent; and
- – for each is independent of the CLE exploration outside of , and after conformally mapping everything to the unit disk, is coupled the exploration inside as in the statement of Lemma 3.1.
Using the induction hypothesis, it is straightforward to see that this defines a conditional law on given that satisfies the conditions of the Lemma. Moreover, note that the first two bullet points above, together with (3.1), define the law of and (satisfying the requirements) uniquely. Combining with the uniqueness in the induction hypothesis, it follows easily that the conditional law of given (satisfying the requirements) is unique.
Consequently, given , there exists a unique conditional law on the product space equipped with the product -algebra, such that if has this law then it satisfies the conditions above Lemma 3.1.
This concludes the existence and uniqueness statement of the lemma. The property (3.1) implies that does almost surely define an order on any finite subset of .
In the coming subsections we will prove Proposition 3.2. Since tightness of all the random variables in question is immediate (either by definition or from our previous work) it suffices to characterize any limiting law. We begin in Section 3.1 by showing this for the order of two points; see just below for an outline of the strategy. Then, we will prove that the time at which they are separated by the converges (for the parameterization with respect to either of the points). This is important for characterizing joint limits, when there are three or more points being considered. It also turns out to be non-trivial, due to pathological behavior that cannot be ruled out when one only knows convergence of the SLE branches in the spaces . We conclude the proof in a third subsection, and finally combine this with the results of Section 2 to summarize the ‘Euclidean’ part of this paper in Proposition 3.12.
3.1 Convergence of order for two points
In this section we show that for two distinct points , the law of the order in which they are visited by the space-filling SLE , converges to the result of a fair coin toss as . That is, converges to a Bernoulli random variable as . The rough outline of the proof is as follows
Recall that is determined by an branching tree, in which denotes the SLE branch toward (parameterized according to minus log conformal radius as seen from ). If we consider the time at which separates and , then for every , is actually measurable with respect to . So what we are trying to show is that this measurability turns to independence in the limit. This means that we will not get very far if we consider the conditional law of given , so instead we have to look at times just before . Namely, we will consider the times that is sent first sent to within distance of the boundary by the Loewner maps associated with . We will show that for any fixed , the conditional probability that , given , converges to as . Knowing this for every allows us to reach the desired conclusion.
To show that these conditional probabilities do tend to for fixed , we apply the Markov property at time . This tells us that after mapping to the unit disc, the remainder of evolves as a radial with a force point somewhere on the unit circle. And we know the law of this curve: initially it evolves as a chordal targeted at the force point, and after the force point is swallowed, it evolves as a radial in the to-be-discovered domain with force point starting adjacent to the tip. So we need to show that for such a process, the behavior is ‘symmetric’ in an appropriate sense. In fact, we have to deal with two scenarios, according to whether the images of and are separated or not when the force point is swallowed. If they are separated, our argument becomes a symmetry argument for chordal . If they are not, our argument becomes a symmetry argument for space-filling . For a more detailed outline of the strategy, and the bulk of the proof, see Lemma 3.8.
At this point, let us just record the required symmetry property of space-filling in the following lemma.
Lemma 3.3.Let be a space-filling in , as above. Then for any :
Proof.For this we use a conformal invariance argument. Namely, we note that by conformal invariance of , applying the map
However, this is easily justified, because we can couple an from 1 to 0 and another from to 0, so that they successfully couple (that is, coincide for all later times) before 0 is separated from with arbitrarily high probability (uniformly in ) as . This follows from Lemma 2.14, target invariance of the SLE and (2.9); that is, because in an arbitrarily small amount of time as , the will have swallowed every point on .
Now we proceed with the setup for the main result of this section (Proposition 3.4). Recall that is the sequence of domains formed by the branch of the uniform CLE exploration toward in . For , we write for the first time that separates from and let be a Bernoulli random variable (taking values each with probability ) that is independent of .
Proposition 3.4.Fix . Then if is a subsequential limit in law of (with respect to the product discrete topology), must satisfy the following property. If is equal to stopped at the first time that is separated from , then
Note that this does not yet imply that the times at which and are separated converge.
To set up for the proof of this proposition, we define for , to be the first time that, under the conformal map , the image of is at distance from ; see Figure 7 for an illustration. Define in the same way for . Write and for the same things as and , but with the time now cut off at and , respectively.
Lemma 3.5.
- (a) as for every fixed .
- (b) as .

Proof.For (a) we use that in . Taking a coupling such that this convergence is almost sure, it is clear from the definition of convergence in that, under this coupling, almost surely for every . Statement (b) holds because almost surely as . Indeed, is almost surely increasing in and bounded above by so must have a limit as . On the other hand, cannot be mapped anywhere at positive distance from the boundary under , so it must be that .
Thus, we can reduce the proof of Proposition 3.4 to the following lemma.
Lemma 3.6.For any continuous bounded function with respect to , and any fixed , we have that
Proof of Proposition 3.4 given Lemma 3.6.Consider a subsequential limit as in Proposition 3.4. Write for stopped at the first time that is sent within distance of under the Loewner flow. Then it is clear (by taking a coupling where the convergence holds almost surely) that is equal to the limit in law of as along the subsequence.
On the other hand, Lemma 3.6 implies that the law of such a limit is that of together with an independent Bernoulli random variable. Indeed, any continuous bounded function with respect to the product topology on is of the form for bounded and continuous with respect to . Moreover, and we already know that as .
So has the law of plus an independent Bernoulli random variable for each . Combining with (b) of Lemma 3.5 yields the proposition.
The proof of Lemma 3.6 will take up the remainder of this subsection. An important ingredient is the following result of [32], about the convergence of to as .
Theorem 3.7. ([[32], Theorem 1.10])Chordal between two boundary points in the disk converges in law to chordal as . This is with respect to supremum norm on curves viewed up to time reparameterization.
Proof of Lemma 3.6.Since is bounded, subsequential limits of always exist. Therefore, we need only to show that such a limit must be equal to . For this, we apply the map : recall that this is the unique conformal map from to that sends to 0 and has positive real derivative at ; see Figure 7. We then use the Markov property of . This tells us that conditionally on , the image of under this map is that of an started at some with a force point at (where are measurable with respect to ). Let us call this curve . Let be the image of under , which is also measurable with respect to and has almost surely. Then the conditional expectation of given can be written as a probability for . Namely, it is just the probability that when first separates and 0, the component containing 0 either has boundary made up of entirely of the left-hand side of and the clockwise arc from to , or the right-hand side of and the complementary counterclockwise arc. We denote this event for by .
Therefore, by dominated convergence, Lemma 3.6 follows from Lemma 3.8 stated and proved below.
Lemma 3.8.Let be an started at some with a force point at . Fix . Let be the event that when first separates and 0, the component containing 0 either has boundary made up of entirely of the left-hand side of and the clockwise arc from to , or the right-hand side of and the complementary counterclockwise arc.
Another way to describe the event is the following. If the clockwise boundary arc from to together with the left-hand side of is colored red, and the counterclockwise boundary arc together with the right-hand side of is colored blue (as in Figures 7 and 8) then is the event that when 0 and are separated, the component containing 0 is ‘monocolored’.

Outline for the proof of Lemma 3.8. Note that until the first time that 0 is separated from , has the law (up to time reparameterization) of a chordal from to in ; see Lemma 2.4. Importantly, we know by Theorem 3.7 that this converges to chordal as .
- In the latter case (left-hand side of Figure 8), note that is very unlikely to return anywhere near to 0 or before swallowing the force point at . Hence, whether or not occurs depends only on whether the curve goes on to hit the boundary ‘just to the left’ of , or ‘just to the right’. Indeed, hitting on one side will correspond to 0 being in a monocolored red bubble when it is separated from , meaning that will occur, while hitting on the other side will correspond to being in a monocolored blue bubble, and it will not. By the Markov property and symmetry, we will argue that each of these happen with (conditional) probability close to .
- In the former case (right-hand side of Figure 8), will go on to swallow the force point before separating 0 and , with high probability as . Once this has occurred, will continue to evolve in the cut-off component containing 0 and , as an with force point initially adjacent to the tip. But then by mapping to the unit disk again, the conditional probability of becomes the probability that a space-filling visits one particular point before another. This converges to as by Lemma 3.3.
Proof of Lemma 3.8.Let us now proceed with the details. For small, let be run until the first entry time of . By Theorem 3.7, the probability that separates 0 or from before time tends to 0 as for any fixed . We write for this event.
We also fix a , chosen such that and are contained in the closure of . Again from the convergence to SLE we can deduce that
- for the intersection of and the event that separates 0 and in , with 0 on the left of ;
- for the same thing but with left replaced by right; and
- for the intersection of and the event that does not separate 0 and in .
Let us now describe what is going on with the terms and . The term corresponds to the left-hand side scenario of Figure 8, and the term corresponds to the same scenario, but when 0 and lie on opposite sides of the curve to those illustrated in the figure. We will show that
Proof of (3.5). First, by (3.3), we can pick small enough such that the differences
We use this to observe, by conformally mapping to that
Proof of (3.6). Write for the event that swallows the force point before separating 0 and . Then we can rewrite 4 as
To do this, we condition on run up to the time that the force point is swallowed. Conditioned on this initial segment we can use the Markov property of to describe the future evolution of . Indeed, it is simply that of a radial started from and targeted toward 0, with force point located infinitesimally close to the starting point. Viewing the evolution of after time as one branch of a space-filling we then have
This is precisely the statement of Lemma 3.3. Thus we conclude the proof of (3.6), and therefore Lemma 3.8.
3.2 Convergence of separation times
We now want to prove that for the actual separation times converge to the separation time in law (jointly with the exploration) as . The difficulty is as follows. Suppose we are on a probability space where converges almost surely to . Then we can deduce (by Lemma 3.5) that any limit of must be greater than or equal to . But it still could be the case that and are ‘almost separated’ at some sequence of times that converge to as , but that the then go on to do something else for a macroscopic amount of time before coming back to finally separate and . Note that in this situation the would be creating ‘bottlenecks’ at the almost separation times, so it would not contradict Proposition 3.4).
The main result of this subsection is the following.
Proposition 3.9.For any
Remark 3.10.It is easy to see that is tight in for any fixed . For example, this follows from Corollary 2.29, which implies that minus the log conformal radius, seen from , of the first loop containing and not , is tight. Since is bounded above by this minus log conformal radius, tightness of follows.
There is one situation where convergence of the separation times is already easy to see from our work so far. Namely, when and are separated (in the limit) at a time when a CLE loop has just been drawn. More precisely:
Lemma 3.11.Suppose that is such that
- ;
- is equal to (modulo time reparameterization), up to the time that is separated from ;
- ; and
- conditionally on the above event occurring, is independent of and has the law of a Bernoulli random variable.
Proof.Without loss of generality, by switching the roles of and if necessary and by the Markov property of the explorations, it suffices to consider the case that is the outermost loop (generated by ) containing .
By Skorokhod embedding together with Corollary 2.17 and Proposition 2.18, we may assume that we are working on a probability space where the convergence assumed in the lemma holds almost surely, jointly with the convergence (in the Hausdorff sense), (in the Carthéodory sense) and . (Recall the definitions of these times from Section 2.1.6). We may also assume that the convergence holds almost surely as for all rational .
Now we restrict to the event that separates from at time , so that in particular is at positive distance from . The Hausdorff convergence thus implies that for all large enough (that is, is outside of the first loop containing ), and therefore that for all large enough (that is, separation occurs no later than this loop closure time). Since is defined to be the almost sure limit of as , and we have assumed that almost surely, this implies that almost surely on the event . On the other hand, we know that and as for all rational positive , so that for all and therefore almost surely. Together this implies that on the event .
Next, observe that by the same argument as in the penultimate sentence above, we have with probability 1. Moreover, we saw that on the event , for all large enough. But we also have that , so that and therefore for all large enough. Hence,
For the final bullet point, if we write for stopped at time , we already know from the previous subsection that the law of given is fair Bernoulli. Moreover, since and are independent for every , it follows that is independent of . So in general (that is, without restricting to the event ) we can say that, given and , has the conditional law of a Bernoulli random variable. Since the event (that ) is measurable with respect to , and we have already seen that on this event, we deduce the final statement of the lemma.
Proof of Proposition 3.9.By tightness (Remark 3.10), and since we already know the convergence in law of to , it suffices to prove that any joint subsequential limit in law of has almost surely. So let us assume that we have such a subsequential limit (along some sequence ) and that we are working on a probability space where the convergence holds almost surely. As remarked previously, since for each and , we already know that almost surely. So we just need to prove that , or alternatively, that for any fixed. Since and are the almost sure limits of and as , it is sufficient to prove that for each
More precisely, let us assume from now on that is fixed, and write for the collection of faces (squares) of that intersect . We write for the event that there exists that is separated by from during the interval and such that is visited by the space-filling before . We write for the same event but with the interval [ instead. So if the event occurs, then either occurs or does not. Hence, for any :
Let us start with (3.9). First, Lemma 3.11 tells us that since many will be separated from by the CLE exploration during the time interval as , the same will be true for the space-filling SLE on the time interval when are small. More precisely, for any fixed , , the lemma implies that
This is almost exactly what we need. However, recall that although only requires one to be disconnected from by , it also requires that this is visited by the space-filling SLE before . This is why we ask for squares to be separated because then by Lemma 3.11, whether they are visited before or after converges to a sequence of independent coin tosses. Namely, for any ,
Hence, to conclude the proof of the proposition, it suffices to justify (3.10). Although this is a statement purely about SLE, it turns out to be somewhat easier to prove using the connection with LQG in [18]. Thus we postpone the proof of (3.10) to Section 4.4, at which point we will have introduced the necessary objects and stated the relevant theorem of [18]. Let us emphasize that this proof will rely only on [18] and basic properties of LQG (and could be read immediately by someone already familiar with the theory) so it is safe from now on to treat Proposition 3.9 as being proved.
3.3 Convergence of the partial order: Proof of Proposition 3.2
Proof of Proposition 3.2.The following three claims are the main ingredients.
Claim 1..
Proof.This follows from Corollary 2.16, Proposition 3.9 and the fact that for every and , and agree (up to time change) until and are separated, and then evolve independently.
Claim 2.For any , .
Proof.As usual, due to tightness, it is enough to show that any subsequential limit of , along a sequence , has the correct joint distribution. In fact, we may assume that
Now, Proposition 3.4 implies that, given and stopped at times , respectively, the conditional law of is fair Bernoulli. On the other hand, since
Claim 3.For any , .
Proof.The same argument as for Claim 2 extends directly to this slightly more general setting (we omit the details).
With Claim 1 in hand (and the argument proving Lemma 3.1) all we need to show is that for any subsequential limit in law of as , the conditional law of given satisfies the bullet points above Lemma 3.1. That is, (a) for all ; (b) for all distinct; (c) is (conditionally) Bernoulli for any such ; and (d) for all with , if separates from at the same time as it separates from then ; otherwise and are (conditionally) independent.
Observe that (a) and (b) follow by definition of the , and (c) follows from Claim 3. The first case of (d) also follows by definition, and the second follows from the definition of together with the branching property of and the convergence of the separation times.
3.4 Joint convergence of SLE, CLE and the order variables
The results of Sections 2 and 3 give the final combined result:
Proposition 3.12.
4 LIOUVILLE QUANTUM GRAVITY AND MATING OF TREES
4.1 Liouville quantum gravity
Let denote the infinite strip. By, for example, [18, Lemma 4.3], has an orthogonal decomposition , where is the subspace of consisting of functions (modulo constants) which are constant on vertical lines of the form and is the subspace of consisting of functions which have mean zero on all such vertical lines. This leads to a decomposition of the free boundary GFF on , where (respectively, ) is the projection of onto (respectively, ). We call the lateral component of .
The LQG disk is an LQG surface of special interest, since it arises in scaling limit results concerning random planar maps, for example, [13, 24]. The following is our definition of the unit boundary length -LQG disk in the subcritical case. Our field is equal to plus the field defined in, for example, [18]: this is because we want it to have boundary length 1 for our definition of (which is times the usual one).
Definition 4.1. (Unit boundary length -LQG disk for )Let be a field on the strip with the law of the lateral component of a free boundary GFF on . Let be a function on such that , where
- (i) has the law of conditioned to be negative for all time, for a standard Brownian motion started from 0; and
- (ii) is independent of and satisfies .
See [30, Definition 2.4 and Remark 2.5] for a proof that the above does correspond to + the unit boundary length disk of [18]. Note that (see, for example, [18, Lemma 4.20]) is finite for each fixed , so that the above definition makes sense. In fact, we can say something stronger, namely Lemma 4.2. We remark that the power in the lemma has not been optimized.
Lemma 4.2.There exists not depending on such that
Finally, if is defined in the same way as above but instead letting have the law of times a three-dimensional Bessel process, then we also have that
Proof.Let us first deal with the subcritical measures. In this case, we write
Finally, exactly the same proof works in the case of the critical measure, using [49, Section 1.1.1] to see that has a finite th moment, which does not depend on by translation invariance.
We may now define the critical unit boundary length LQG disk as follows.
Definition 4.3. (Unit boundary length 2-LQG disk)Letting be as in Lemma 4.2 we define the unit boundary length 2-LQG disk to be the surface , where
Note that is finite by Lemma 4.2.
Remark 4.4.Readers may have previously encountered the above as the definition of a quantum disk with two marked boundary points. A quantum surface with marked points is an equivalence class of with , using the equivalence relation described by (4.1), but with the additional requirement that maps marked points to marked points. In this paper we will use Definitions 4.1 and 4.3 to define specific equivalence class representatives of quantum disks, but we will always consider them as quantum surfaces without any marked points. That is, we will consider their equivalence classes under the simple relation (4.1).
The following lemma says that the subcritical disk converges to the critical disk as (equivalently, ). We say that a sequence of measures on a metric space (equipped with the Borel -algebra) converges weakly to a measure if for all such that we have .
Lemma 4.5.For let be the field of Definition 4.1 and be the field of Definition 4.3. Then , where the first coordinate is equipped with the topology and the second and third coordinates are equipped with the weak topology of measures on and , respectively.
Proof.To conclude it is sufficient to prove the following, for an arbitrary sequence :
- (i) we have convergence in law along the sequence if we replace by , and by for every ; and
- (ii) there exists a coupling of the such that in as .
We may therefore couple and so that their lateral components are identical, and the components that are constant on vertical lines converge almost surely on compacts as . For this coupling, the result of [6] implies that
For (ii), first observe that
Remark 4.6.We reiterate that and almost surely. Moreover, we have the convergence as .
Remark 4.7.For we define the -boundary length disk to be a surface with the law of , where for as in Definition 4.1 or 4.3. Lemma 4.5 also holds if we assume all the disks are -boundary length disks.
The fields that appear in the statement of our main theorem are defined as follows.
Definition 4.8.We define fields (respectively, ) to be parameterizations of unit boundary length -LQG disks (respectively, the 2-LQG disk) by instead of . More specifically we take to be the conformal map from to that sends to , respectively. Then we set
Remark 4.9.Lemma 4.5 clearly also implies the convergence
In fact, it implies the convergence of various embeddings of quantum disks. Of particular use to us will be the following:
Lemma 4.10.Suppose that for each , is as in Remark 4.7 for some and that is defined by choosing a point from in , defining conformal such that and , and setting
Suppose similarly that is defined by taking the field in Remark 4.7 with the same , picking a point from ; taking conformal with and ; and setting
Then as , we have that
Proof.We assume that ; the result for other and the uniform convergence in (4.5) follows immediately from the definition in Remark 4.7.
The proof then follows from Lemma 4.5. We take a coupling where the convergence is almost sure: in particular, the fields converge almost surely to in and the measures converge weakly almost surely to in . This means that we can sample a sequence of from the and from , such that almost surely. Since is at positive distance from , this implies that the conformal maps converge to almost surely on compacts of and therefore that in and converges weakly to . Finally, (4.5) follows from the convergence proved above, and the fact that it holds for the limit measure .
Later, we will also need to consider fields obtained from the field of Lemma 4.10 via a random rotation. For this purpose, we record the following remark.
Remark 4.11.Suppose that are a sequence of fields coupled with some rotations such that has the law of from Lemma 4.10 with , for some , . Suppose further that in as . Then and almost surely. Indeed, is tight in , and any subsequential limit has coupled as above. Since for every and it follows from Lemma 4.10 that and almost surely. On the other hand, it is not hard to see that must be equal to almost surely, which implies the result.
4.2 Mating of trees
Mating of trees theory, [18], provides a powerful encoding of LQG and SLE in terms of Brownian motion. We will state the version in the unit disk below.
Let and let be times a standard planar Brownian motion with correlation , started from (1,0) or (0,1). Condition on the event that first leaves the first quadrant at the origin (0,0); this is a zero probability event but can be made sense of via a limiting procedure; see, for example, [2, Proposition 4.2]. We call the resulting conditioned process (restricted until the time at which the process first leaves the first quadrant) a Brownian cone excursion with correlation . Note that we use the same terminology for the resulting process for any and either choice of (1,0) or (0,1) for the starting point.
To state the mating of trees theorem (disk version) we first introduce some notation. Let denote a unit boundary length -LQG disk for , embedded as described in Definition 4.8. Let denote a space-filling SLE in , starting and ending at 1, which is independent of . Recall that this is defined from a branching SLE as described in Section 2.1.7, where the branch targeted toward is denoted by (one can obtain from by deleting time intervals on which is exploring regions of that have been disconnected from ). Parameterize by the area measure induced by . Let denote the process started at (0,1) and ending at (0,0) which encodes the evolution of the left-hand side and right-hand side boundary lengths of ; see Figure 9.

The following theorem follows essentially from [18]. For precise statements, see [40, Theorem 2.1] for the law of and see [40, Theorem 7.3] for the law of the monocolored components.
Theorem 4.12. ([[18, 40]])In the setting above, has the law of a Brownian cone excursion with correlation . The pair is measurable with respect to the -algebra generated by . Furthermore, if is sampled from renormalized to be a probability measure, then the monocolored complementary components of define independent -LQG disks conditioned on their -LQG boundary lengths and areas, that is, if we condition on the ordered sequence of boundary lengths and areas of the monocolored domains disconnected from by then the corresponding LQG surfaces are independent -LQG disks with the given boundary lengths and areas.
Remark 4.13.In fact, we now know from [4] that the variance of the Brownian motion from which the law of can be constructed is equal to , where . In particular, the variance is of order .
For each fixed there is a natural parameterization of called its quantum natural parameterization which is defined in terms of as follows. First define to be the time at which first hits . Then let denote the set of for which we cannot find a cone excursion (that is, such that on , and either or ) such that . We call the times in ancestor-free times relative to time . It is possible to show (see [18, Section 1.4.2]) that the local time of is well defined.† Let denote the increasing function describing the local time of such that and for . Then let for denote the right-continuous inverse of .
Definition 4.14. (Quantum natural parameterization)With the above definitions
4.3 Convergence of the mating of trees Brownian functionals

We will prove in this subsection that all the quantities defined above have a joint limit in law as . Namely, let us consider an uncorrelated Brownian excursion in the right half-plane from (0,1) to (0,0); the process can, for example, be constructed via a limiting procedure where we condition a standard planar Brownian motion from (0,1) to (0,0) on first leaving at a point where . For less than the total duration of , let denote the set of times at which has a backward running infimum relative to time , that is, if for all . Let denote the increasing function describing the local time of such that and for . Then let denote the right-continuous inverse of , and define by .
Then we have the following convergence.
Lemma 4.15. as , where we use the Hausdorff topology on the second coordinate and the Skorokhod topology on the remaining coordinates.
Proof.First we consider the infinite volume case where is a two-sided planar Brownian motion started from 0, with the same variance and covariance as before, namely variance and covariance 0. In this infinite volume setting we define similar to before, such that for , is the set of ancestor-free times relative to time , is an increasing process given by the local time of satisfying for , is the right-inverse of and . We make a similar adaptation of the definition to the infinite volume setting for ; in particular, is ( times) a standard uncorrelated two-sided Brownian motion planar motion. By translation invariance in law of and , and since and determine the rest of the objects in question, it is sufficient to show convergence for .
First we claim that for all we can sample by considering a PPP in the second quadrant with intensity for , such that points of this PPP are in bijection with the complementary components of with representing the length of the component and representing the relative ordering of the components. (In the case , refers to .) For the claim follows since restricted to the complementary components of has law given by the Brownian excursion measure. For the claim follows from [18]: It is explained in [18, Section 1.4.2] that has the law of the zero set of some Bessel process, which verifies the claim modulo the formula for . The dimension of is [20, Table 1 and Example 2.3], and we get the formula for by adding 1 to this number.
Next we argue that the marginal law of converges to the marginal law of . Consider the definition of these sets via PPP as described in the previous paragraph. Since , the PPP for converge in law to the PPP for on all sets bounded away from . This implies that for any compact interval we have convergence in law of to for the Hausdorff distance.
Now we will argue that if denotes the backward running infima of relative to time 0, then
Next we will argue that , assuming we choose the multiplicative constant consistently when defining and . The convergence result follows again from the construction of and via a PPP, since the coordinate of the PPP defines the local time (modulo multiplication by a deterministic constant).
Using that , that and determine the other two elements in this tuple and that , we get
To conclude the proof we will transfer from the infinite volume setting to the finite volume setting. Let us start by recalling that there is a natural infinite measure on Brownian excursions in the cone which is uniquely characterized (modulo multiplication by a constant) by the following property. Let be as in the previous paragraph, let and let be the interval with largest left end point of length at least during which makes an excursion in the cone . Here a cone excursion in is a path starting at for some and , ending at , and otherwise staying inside . Define
The measure allows a disintegration , where a path sampled from almost surely starts at . Furthermore, for , a path sampled from and rescaled by so it ends at (and with Brownian scaling of time), has law . Finally, an excursion sampled from is equal in law to the excursion in the statement of the lemma; see [2].
Let us now use these facts to complete the proof. We define a function such that for a two-sided planar Brownian motion as above we have almost surely. For a Brownian cone excursion in starting at we define such that is equal in law to the tuple in the theorem statement. We also extend the definition of to the case of Brownian excursions in starting at for general in the natural way.
Now let be coupled with as in (4.6) for some fixed , and let be the event that starts at for . Define similarly for . We claim that
With as in the previous paragraph let denote a random variable which is obtained by conditioning on and then applying a Brownian rescaling of so that starts at . We get from (4.7) that . Note that if we condition the excursions in the statement of the lemma to have duration at least , then these have exactly the same laws as conditioned to have duration at least . Thus the lemma follows upon taking , since the probability that the considered excursions have duration at least tends to 1, uniformly in .
4.4 Proof of (3.10)
The mating of trees theorem (Theorem 4.12) together with the convergence proved in the previous subsection now make the proof of this statement reasonably straightforward. Indeed, in plain language, it says that the probability of an branch almost separating two points and (where ‘almost’ is encoded by a small parameter ) but then going on to separate a bicolored component of macroscopic size from at some time strictly before separating from , goes to 0 as , uniformly in . The idea is to couple this SLE with an independent -LQG disk and note that if the event mentioned above were to occur, then the component containing and at time would have a small ‘bottleneck’ and hence define a very strange distribution of -LQG mass when viewed as a -LQG surface. On the other hand, if we sample several points from the -LQG area measure on the disk, then one of these is likely to be in the bicolored component separated from and at time . So the mating of trees theorem says that should really look like a quantum disk, and in particular, have a rather well behaved distribution of -LQG mass without bottlenecks. This contradiction will lead us to the proof of (3.10).
Let us now get on with the details. For we consider a CLE exploration alongside an independent unit boundary length quantum disk as in Definition 4.8. We write for its associated LQG area measure and let be a point in sampled from normalized to be a probability measure. We let be fixed.
Corollary 4.16.Consider the event that
- (that is, the component containing when and are separated is monocolored);
- when (this monocolored component) is mapped to , with a point in the interior chosen proportionally to sent to 0, the resulting quantum mass of is greater than .
Proof.Theorem 4.12 says that the monocolored components separated from by are quantum disks conditionally on their boundary lengths and areas. Moreover, we know that the total mass of the original disk converges in law to something almost surely finite as , by Lemma 4.5 and Remark 4.6. Recalling the definition of from Section 4.3, we also know that the largest quantum boundary length among all monocolored components separated from has law given by the largest jump of , for chosen uniformly in . Indeed, if corresponds to as in the paragraph above Definition 4.14, then is a uniform time in and the jumps of are precisely the quantum boundary lengths of the monocolored components disconnected from . By Lemma 4.15 we may deduce that the law of this largest jump converges to something almost surely finite as . Thus, by choosing large enough, we may work on an event with arbitrarily high probability (uniformly in ) where there are fewer than monocolored components separated for with mass at least , and where they all have boundary length less than . Lemma 4.10 then provides the result.
We also need one more elementary property of radial Loewner chains to assist with the proof of (3.10).
Lemma 4.17.Consider the image of a point under the radial Loewner flow corresponding to . Then with probability one, is a non-decreasing function of time (until point is swallowed).
Proof.From the radial Loewner equation one can compute directly that, until point is swallowed,
Proof of (3.10).Fix and suppose that for some . Recall that is the event that there exists that is separated by from during the interval and such that the disconnected component containing is monocolored. Let be as above Corollary 4.16, and let . Then is strictly positive, due to the convergence result Lemma 4.8, plus the fact that when is picked from the critical LQG area measure for a critical unit boundary length disk. By independence, we then have , where is the event that and .
We can also choose small enough and large enough that on an event with probability , uniformly in :
- (respectively, ) where (respectively, ) is the first nested bubble containing (respectively, ) that is entirely contained in (respectively, ;
- and have -mass greater than or equal to ;
- if we map (respectively, ) to with (respectively, ) sent to 0, then the images of and are contained in ; and
- .
- (i) and are contained in for all ;
- (ii) for any and conformal map sending to with sent to 0, the image of is contained in a neighborhood of .
To finish the proof, we consider the event which has probability by construction. Conditionally on this event, if we sample a point from according to the measure , then this point will lie in with conditional probability . If this happens, then upon mapping to the unit disk with this point sent to the origin, a set of mass (namely ) will necessarily be sent to (see point (ii) above). Note that is a function of only (and in particular does not depend on ).
So in summary, if , then for some depending only on , where is as in Corollary 4.16. This means that if (3.10) does not hold, then for some . This contradicts Corollary 4.16, and hence (3.10) is proved.
5 MATING OF TREES FOR AND JOINT CONVERGENCE OF CLE, LQG AND BROWNIAN MOTIONS AS
The processes are each parameterized by conformal radius seen from , and equipped with the topology of for every . The loop ensembles are equipped with the topology of Hausdorff convergence for the countable collection of loops surrounding each .
Then by Remark 4.9, Proposition 3.12 and the independence of and (respectively, and ), we have that
Proposition 5.1. as .
Proposition 5.2. as .
Here, are equipped with the Hausdorff topology, and the stochastic processes in the definition of are equipped with the Skorokhod topology.
We now wish to describe the joint limit of as . For this, we first need to introduce a little notation.
Remark 5.3.For , by Theorem 4.12 and the definitions above, we have that
- the duration of is equal to , hence for all almost surely;
- for , the time at which visits is almost surely given by ;
- the ordered boundary lengths of the components of are almost surely equal to the ordered jumps of , and the sign of each jump is equal to the sign of the corresponding element of ; and
- the ordered masses of the components of are almost surely equal to the ordered jumps of .
We can also define analogous objects associated with the CLE exploration: if and are separated at time by the exploration branch toward , and we set ; if we set . The set is then defined in exactly the same way. Note that in this case the elements of are ordered by declaring that comes before if and only if and for such that . We now say that is ordered clockwise (respectively, counterclockwise) if there is an even (respectively, odd) number of loops which enclose , and write (respectively, ).
The main ingredient that will allow us to describe the joint limit of is the following:
Proposition 5.4.Given , denote by a point sampled from in (normalized to be a probability measure) and given , denote by a point sampled in the same way from . For given , write for the ordered components of with area , and define similarly for the ordered components of with area . Suppose that for (respectively, for ) are sampled from (respectively, ) normalized to be probability measures, and (respectively, ) are the conformal maps that send to 0 (respectively, to 0) with positive real derivative at (respectively, ). Set (respectively, ) and (respectively, ) to be the 0 function for (respectively, ). Then
In other words, the ordered and signed sequence of monocolored quantum surfaces separated from converges almost surely, as a sequence of quantum surfaces (see above (4.1)) to the ordered sequence of monocolored quantum surfaces separated from as .
From this, we can deduce our main theorem.
Theorem 5.5. converges jointly in law to a tuple as . In the limiting tuple, have marginal laws as above, and are independent, and determines .
Furthermore, we have the following explicit description of the correspondence between and in the limit. Suppose that is sampled from the critical Liouville measure normalized to be a probability measure. Then
- for all almost surely and the conditional law of
(5.1)given is uniform on ,
- satisfies the following for a deterministic constant :
(5.2)almost surely, where for , is the number of domains such that ,
- the ordered collection is almost surely equal to the ordered collection of jumps of (where are defined from as in Section 4.3).
Duration of | |
‘ ordered before ' | |
‘quantum area of points ordered before ' | |
Quantum natural distance of from | |
Jumps of | LQG boundary lengths of ‘components ordered before ' |
Sign of jump | Parity of |
Jumps of | LQG areas of ‘components ordered before ' |
CRT encoded by | exploration branches parameterized by quantum natural distance |
Proof of Theorem 5.5 given Proposition 5.4.Since we know the marginal convergence of each component of , we know that the triple is tight in . Thus our task is to characterize any subsequential limit of . Note that Proposition 5.1 already tells us that are independent, and Proposition 5.2 tells us that the marginal law of is that of a Brownian half-plane excursion plus associated observables.
To characterize the law of we will prove that if is sampled according to in , conditionally independently of the rest of then
- (i) the duration of is equal to almost surely;
- (ii) defined by (5.1) is conditionally uniform on given ;
- (iii) the ordered collection is almost surely equal to the ordered collection of jumps of (defined from as in Section 4.3); and
- (iv) satisfy (5.2) almost surely.
The same argument implies that the law of any subsequential limit is unique. More concretely, suppose that , are two sequences tending to 0 as , such that and as . Then we can also take a joint subsequential limit of ; call it where necessarily and , since we already know the convergence . Repeating the argument of the previous paragraph gives that almost surely. In particular, the marginal law of is the same as that of .
So we are left to justify the above claim. To this end, let
We next argue that the convergence (5.5) necessarily implies the joint convergence
So to summarize, if we have any subsequential limit of we can couple it with (whose conditional law given is that of a sample from ) and with for every positive , such that the joint convergence (5.6) holds along some subsequence . By Skorokhod embedding we may assume that this convergence is almost sure, and so just need to prove that (i)–(iv) hold for the limit. This essentially follows from Remark 5.3 and the convergence of the final coordinates in (5.6); we give the details for each point below.
- (i) This holds since for all almost surely for every , and almost surely.
- (ii) The convergence of the areas in (5.6) implies that
converges almost surely to defined in (5.1) along the subsequence . On the other hand, is conditionally uniform on given for every .
- (iii) The ordered collection of jumps of converge almost surely to the ordered collection of jumps of on the one hand, by definition of the convergence (and by considering a sequence converging to ). On the other hand, they are equal to the ordered collection for every . Since this latter collection converges almost surely to the ordered collection , we obtain (iii).
- (iv) This follows from (iii) and the fact that the marginal law of is that of a Brownian excursion in the right half-plane. Specifically, the first coordinate of at a given time can almost surely be recovered from the jumps of its inverse local time at backward running infima with respect to time , see (5.3), and the second coordinate can also be recovered from the collection of its signed jumps when reparameterized by this inverse local time. When , the values are recovered exactly using the formula (5.2) after using (iii) to translate between and .
5.1 Proof of Proposition 5.4
- (1) The tuples on the left-hand side in Proposition 5.4 are tight in , so we may take a subsequential limit (that we will work with for the remainder of the proof).
- (2) (that is, is not on any nested CLE loop) for all almost surely.
- (3) If are conformal with and , then for each almost surely.†
- (4) Given , the are conditionally independent and distributed according to in each .
- (5) almost surely, where the set on the left is ordered as usual.
- (6) for each almost surely.
These clearly suffice for the proposition.
Proof of (1).Tightness of the first five components follows from the fact that as , plus the tightness of the quantum boundary lengths in (recall that these converge when converges). To see the tightness of we note that there are at most non-zero terms, where is tight in . Moreover, each non-zero has the law of , where is as in Lemma 4.10, are random rotations (which automatically form a tight sequence in ) and are some tight sequence of real numbers. This implies the result by Lemma 4.10.
Proof of (2).Suppose that are sampled conditionally independently according to in , normalized to be a probability measure. Then where the are sampled conditionally independently from and almost surely all lie in . On the other hand, since and are independent, one can sample by taking and then setting for each , with .
Proof of (3).By Skorokhod's theorem, we may work on a probability space where we have the almost sure convergence
To carry out the careful argument, let us fix . Since almost surely by (2), there exists and such that . By taking smaller if necessary, we can also find with . Note that by definition. Due to the almost sure convergence , , and it then follows that , and for all large enough. Moreover, we know that the maps with , converge on compacts of to sending 0 to and with .
On the other hand, where sends and has , and for each , where has and . Since almost surely, and the are uniformly close to and bounded away from the boundary of , this implies that converges to uniformly on compacts of . In turn, this implies that restricted to any compact of is equal to , which verifies that almost surely.
Proof of (4).For this it suffices to prove that for each ,
Proof of (5).As in (3) we assume that we are working on a probability space where we have almost sure convergence along a sequence , so we need to show that the limiting domains are precisely the elements of that have area greater than or equal to . The same argument as for (4) gives that each is a component of with area greater than or equal to . So it remains to show that they are the only such elements of .
For this, suppose that has . Then for some with probability 1. Choosing , such that it is easy to see that is the almost sure Carathéodory limit seen from of as . Using the convergence of to and Corollary 2.23, we therefore see that and so for some and all large enough. From here we may argue as in the proof of (3) to deduce that the Carathéodory limit of is equal to . Thus, since is the Carathéodory limit of which is equal to for all large enough, we conclude that .
The fact that the orders of the collections in (3) coincide follows from the convergence of the order variables as part of (and the argument we have now used several times that allows one to transfer from to points in : we omit the details).
Proof of (6).Let us work under almost sure convergence as in the proof of (3), fix and define as in the proof of (3). By Proposition 3.2, we know that almost surely as , and that is determined by the number of loops nested around which discovers before or at time (see the definition of CLE loops from the space-filling/branching in Section 2.1.6). If occurs between two such times for , it is clear from the almost sure convergence of and that the number of loop closure times for occurring before or at converges to the number of loop closure times for occurring before or at time . If is a loop closure time for , the result follows from Lemma 3.11.
5.2 Discussion and outlook
- 1. Can one obtain a version of critical mating of trees where there is bi-measurability between the decorated LQG surface and the pair of Brownian motions (with possibly additional information included)?
- 2. There is an interesting relation to growth-fragmentation processes studied in [1]. Can one combine these two point of views in a fruitful way?
- 3. The Brownian motion encodes a distance of each point to the boundary, and in particular between any CLE loop and the boundary. What is its relation to the CLE metric introduced in [59]?
- 4. Can one prove convergence of observables in critical FK-decorated random planar maps toward the observables in the critical mating of trees picture?
Let us finally mention that there are also other interesting questions in the realm of critical LQG, for example, the behavior of height functions on top of critical planar maps, which are certainly worth exploring too.
5.2.1 Measurability
In the subcritical mating of trees, that is, when , and we consider the coupling described in the introduction or in Section 5 (for simplicity without subscripts), [18] proves that in the infinite-volume setting the pair determines and vice versa. In particular, can be obtained from via a measurable map. This result is extended to the finite volume case of LQG disks in [2].
By contrast, some of this measurability gets lost when we consider our critical setting. The easier direction to consider is whether determine . In the subcritical case this comes basically from the construction, and it does not matter what we really mean by : the nested CLE, the space-filling SLE and the radial exploration tree of CLE are all measurable with respect to one another. This, however, gets more complicated in the critical case. First, the question of whether the nested CLE determines the uniform exploration tree of CLE is already not straightforward; this is a theorem of an unpublished work [59]. Moreover, the nested CLE no longer determines the space-filling exploration from Section 3: indeed, we saw that to go from the uniform exploration tree to the ordering on points, some additional order variables are needed. These order variables are, however, the only missing information when going from to : the conclusion of Theorem 5.5 is that when we include the order variables in (in other words consider the space-filling exploration) then indeed is measurable with respect to .
In the converse direction, things are trickier. In the coupling considered in this paper, does not determine the pair ; however, we conjecture that is determined modulo a countable number of ‘rotations’. Informally, one can think of these rotations as follows: a rotation is an operation where we stop the CLE exploration at a time when the domain of exploration is split into two domains and , we consider the LQG surfaces and , and we conformally weld these two surfaces together differently. The field and loop ensemble of the new surface will be different than the pair of the original surface, but their law is unchanged if we choose the new welding appropriately (for example, if we rotate by a fixed amount of LQG length), and is pathwise unchanged. Therefore performing such a rotation gives us two different pairs and with the same law, and which are associated with the same . We believe that these rotations are the only missing part needed to obtain measurability in this coupling. In fact, by considering a different CLE exploration, where loops are pinned in a predetermined way (for example, where all loops are pinned to some trunk, such as in, for example, [36]), one could imagine obtaining a different coupling of , where does determine .
5.2.2 Growth fragmentation
We saw below the statement of Theorem 5.5 how certain observables in the Brownian excursion map to observables (for example, quantum boundary lengths and areas of discovered loops) in , when we restrict to a single uniform CLE exploration branch. Given the definition of the branching exploration (recall that the explorations toward any two points coincide exactly until they are separated by the discovered loops and then evolve independently) this is one way to define an entire branching process from the Brownian excursion.
In fact, this embedded branching process was already described completely, and independently, in an earlier work of Aïdekon and Da Silva [1]. Namely, given with law as in Theorem 5.5, one can consider for any the countable collection of excursions of to the right side of the vertical line with horizontal component . Associated with each such excursion is a total displacement (the difference between the vertical coordinate of the start and end points) and a sign (depending on which of these coordinates is larger). In [1], the authors prove that if one considers the evolution of these signed displacements as increases, then one obtains a signed growth fragmentation process with completely explicit law. The fact that this process is a growth fragmentation means, roughly speaking, that it can be described by the evolving ‘mass’ of a family of cells: the mass of the initial cell evolves according to a positive self-similar Markov process, and every time this mass has a jump, a new cell with exactly this mass is introduced into the system. Each such new cell initiates an independent cell system with the same law. In the setting of signed growth fragmentations, masses may be both positive and negative.
In the coupling , such a growth fragmentation is therefore naturally embedded in . It corresponds to a parameterization of the branching uniform exploration by quantum natural distance from the boundary (that is, by the value of the component), and branching occurs whenever components of the disk become disconnected in the exploration. At any given time, the absolute mass of a fragment is equal to the quantum boundary length of the corresponding component, and the sign of the fragment is determined by the number of loops that surround this component.
Let us also mention that growth fragmentations in the setting of CLE on LQG were also studied in [43, 44], and coincide with the growth fragmentations obtained as scaling limits from random planar map explorations in [12]. Taking in these settings (either in [43] or in [44]) is also very natural and would give other insights about than those obtained in this paper. Lehmkuehler takes this approach in [36].
5.2.3 Link with the conformally invariant metric on
Recall the uniform CLE exploration from Section 2.1.5, which was introduced by Werner and Wu [64]. Werner and Wu interpret the time at which a loop of the CLE is added, with the time parameterization (2.8), as the distance of to the boundary ; we refer to it here as the CLE exploration distance of to . In an unpublished work, Sheffield, Watson and Wu [59] prove that this distance is the distance as measured by a conformally invariant metric on . This metric is conjectured to be the limit of the adjacency metric on CLE loops as . It is also argued in [59] that the uniform exploration of is determined by .
5.2.4 Discrete models
The mating of trees approach to LQG coupled with CLE is inspired by certain random walk encodings of random planar maps decorated by statistical physics models. The first such encoding is the hamburger/cheeseburger bijection of Sheffield [58] for random planar maps decorated by the critical Fortuin–Kasteleyn random cluster model (FK-decorated planar map).
In the FK-decorated planar map each configuration is a planar map with an edge subset, whose weight is assigned according to the critical FK model with parameter . Sheffield encodes this model by five-letter words whose symbol set consists of hamburger, cheeseburger, hamburger order, cheeseburger order and fresh order. The fraction of fresh orders within all orders is given by . As we read the word, a hamburger (respectively, cheeseburger) will be consumed by either a hamburger (respectively, cheeseburger) order or a fresh order, in a last-come-first-serve manner. In this setting, the discrete analog of our Brownian motion is the net change in the burger count and the burger discrepancy since time zero, which we denote by .
It was proved in [58] that converges in law to , where are independent standard one-dimensional Brownian motions and . When , the correlation of is the same as for the left and right boundary length processes of space-filling decorated -LQG (cf. Theorem 4.12) where and . This is consistent with the conjecture that under these parameter relations, LQG coupled with (equivalently, space-filling ) is the scaling limit of the FK-decorated planar map for . Indeed, based on the Brownian motion convergence in [58], it was shown in [22, 28, 29] that geometric quantities such as loop lengths and areas converge as desired.
When and , we have , just as in the limit of LQG coupled with , where the correlation of the left and right boundary length processes tend to 1. We believe that the process converges in law to ; moreover, based on this convergence and results in our paper, it should be possible to extract the convergence of the loop lengths and areas for FK decorated planar map to the corresponding observables in critical LQG coupled with . We leave this as an open question. It would also be very interesting to identify the order of the normalization , which is related to the asymptotic of the partition function of the FK-decorated planar map with .
Another model of decorated random planar maps that is believed to converge (after uniformization) to CLE decorated LQG is the O() loop model, where the critical case corresponds to . It is therefore also interesting to ask whether our Brownian half-plane excursion can be obtained as a scaling limit of a suitable boundary length exploration process in this discrete setting. In fact, a very closely related question was considered in [15], where the authors identify the scaling limit of the perimeter process in peeling explorations of infinite volume critical Boltzmann random planar maps (see [14] for the relationship between these maps and the O(2) model). Modulo finite/infinite volume differences, this scaling limit, which is a Cauchy process, corresponds to a single ‘branch’ in our Brownian motion (see Section 5.2.2).
ACKNOWLEDGEMENTS
J. Aru was supported by Eccellenza grant 194648 of the Swiss National Science Foundation. N. Holden was supported by grant 175505 of the Swiss National Science Foundation, along with Dr. Max Rössler, the Walter Haefner Foundation and the ETH Zürich Foundation. E. Powell was supported by grant 175505 of the Swiss National Science Foundation. X. Sun was supported by the NSF grant DMS-2027986 and the NSF Career grant DMS-2046514. J. Aru and N. Holden were both part of SwissMAP. We all thank Wendelin Werner and ETH for their hospitality. We also thank Elie Aïdékon, Nicolas Curien, William Da Silva, Ewain Gwynne, Laurent Ménard, Avelio Sepúlveda and Samuel Watson for useful discussions. Finally, we thank the anonymous referee for their careful reading of this paper, and helping to improve the exposition in numerous places.
JOURNAL INFORMATION
The Journal of the London Mathematical Society is wholly owned and managed by the London Mathematical Society, a not-for-profit Charity registered with the UK Charity Commission. All surplus income from its publishing programme is used to support mathematicians and mathematics research in the form of research grants, conference grants, prizes, initiatives for early career researchers and the promotion of mathematics.
REFERENCES
- † We thank N. Curien for explaining this relation to us.
- † That is, if is the Brownian excursion measure then the integral is finite for -almost all excursions; see [64, Section 2].
- † Variants of this process, for example, chordal/whole-plane versions, a clockwise version, and version with another starting point, can be defined by modifying the definition of the branching SLE; see, for example, [2, 21].
- † Of course this depends on , but we drop this from the notation for simplicity.
- † This name is partially inspired from the fact that the process is constructed via a uniform CLE exploration, and partly since, every time the domain of exploration is split into two components, the components are ordered uniformly at random.
- † This local time (and the corresponding local time for defined below) is defined only up to a deterministic multiplicative constant. We fix this constant in the proof of Lemma 4.15.
- † With respect to the Euclidean topology in the third coordinate, and the topology in the final coordinates defined such that as if and only if the number of non-zero components on the left-hand side is equal to the number of non-zero components on the right-hand side for all large enough, and the first components converge in the product discrete Euclidean topology.
- † Once we have point (5), it follows that these are equal to the .