Sle 2018
Sle 2018
Lent 2018
These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.
The course will focus on the definition and basic properties of SLE. The key ideas are
conformal invariance and a certain spatial Markov property, which make it possible
to use Itô calculus for the analysis. In particular we will show that, almost surely, for
κ ≤ 4 the curves are simple, for 4 ≤ κ < 8 they have double points but are non-crossing,
and for κ ≥ 8 they are space-filling. We will then explore the properties of the curves
for a number of special values of κ (locality, restriction properties) which will allow us
to relate the curves to other conformally invariant structures.
The fundamentals of conformal mapping will be needed, though most of this will be
developed as required. A basic familiarity with Brownian motion and Itô calculus will
be assumed but recalled.
1
Contents III Schramm–Loewner Evolutions
Contents
0 Introduction 3
1 Conformal transformations 4
1.1 Conformal transformations . . . . . . . . . . . . . . . . . . . . . 4
1.2 Brownian motion and harmonic functions . . . . . . . . . . . . . 5
1.3 Distortion estimates for conformal maps . . . . . . . . . . . . . . 6
1.4 Half-plane capacity . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2 Loewner’s theorem 14
2.1 Key estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2 Schramm–Loewner evolution . . . . . . . . . . . . . . . . . . . . 19
4 Phases of SLE 24
Index 47
2
0 Introduction III Schramm–Loewner Evolutions
0 Introduction
Schramm–Loewner evolution is a random curve in a complex domain D (which
we will often take to be H for convenience), parametrized by a positive real
number κ. This was introduced by Schramm in 1999 to describe various scaling
limits that arise in probability theory and statistical physics.
Recall that if we take a random walk on the integer lattice Zd , and take the
scaling limit as the grid size tends to 0, we converge towards a Brownian motion.
There are other discrete models that admit a scaling limit, and the limit is often
something that is not Brownian motion.
Schramm showed that if the scaling limit satisfies certain conformal invariance
properties, which many models do, then it must be SLEκ for some κ. If we can
identify exactly which κ it belongs to, then this will completely determine what
the scaling limit is, and often conveys a lot of extra information such as the
critical exponents. This has been done for several discrete models:
– The “level sets” of a Gaussian free field are SLE4 ’s (see chapter7).
In this course, we will prove some basic properties of SLEκ , and then establish
the last three results/conjectures.
3
1 Conformal transformations III Schramm–Loewner Evolutions
1 Conformal transformations
1.1 Conformal transformations
Definition (Conformal map). Let U, V be domains in C. We say a holomorphic
function f : U → V is conformal if it is a bijection.
We will write D for the open unit disk, and H for the upper half plane. An
important theorem about conformal maps is the following:
Theorem (Riemann mapping theorem). Let U be a simply connected domain
with U =6 C and z ∈ U be any point. Then there exists a unique conformal
transformation f : D → U such that f (0) = z, and f 0 (0) is real and positive.
i(1 + w)
g(w) = .
1−w
In general, the conformal transformations H → H consist of maps of the form
az + b
f (z) =
cz + d
with a, b, c, d ∈ R and ad − bc 6= 0.
√
Example. For t ≥ 0, we let Ht = H \ [0, 2 ti]. The map Ht → H given by
p
z 7→ z 2 + 4t
4
1 Conformal transformations III Schramm–Loewner Evolutions
∂2u ∂2u
∂ ∂v ∂ ∂v
2
+ 2 = + − = 0.
∂x ∂y ∂x ∂y ∂y ∂x
5
1 Conformal transformations III Schramm–Loewner Evolutions
Then one sees that g is holomorphic and |g(z)| ≤ 1 for all z ∈ ∂D, hence for all
z ∈ D by the maximum modulus principle. If |f (z0 )| = |z0 | for some z0 ∈ D \ {0},
then g must be constant, so f is linear.
6
1 Conformal transformations III Schramm–Loewner Evolutions
By scaling it suffices to prove this for the case r = 1. This follows from the
following result:
Theorem. If f ∈ U, then |a2 | ≤ 2.
The proof of this proposition will involve quite some work. So let us just
conclude the theorem from this.
Proof of Koebe-1/4 theorem. Suppose f : D → D is in U, and z0 6∈ D. We shall
show that |z0 | ≥ 41 . Consider the function
z0 f (z)
f˜(z) = .
z0 − f (z)
f (z) = z + a2 z 2 + · · · ,
then
˜ 1
f (z) = z + a2 + z2 + · · · .
z0
So we obtain the bounds
1
|a2 |, a2 + ≤ 2.
z0
By the triangle inequality, we must have |z0−1 | ≤ 4, hence |z0 | ≥ 14 .
The 1/4 theorem bounds the distortion in terms of the value of f 0 (0). Con-
versely, if we know how much the distortion is, we can use the theorem to derive
bounds on f 0 (0).
Corollary. Let D, D̃ be domains and z ∈ D, z̃ ∈ D̃. If f : D → D̃ is a conformal
transformation with f (z) = z̃, then
d˜ 4d˜
≤ |f 0 (z)| ≤ ,
4d d
where d = dist(z, ∂D) and d˜ = dist(z̃, ∂ D̃).
Proof. By translation, scaling and rotation, we may assume that z = z̃ = 0,
d = 1 and f 0 (0) = 1. Then we have
7
1 Conformal transformations III Schramm–Loewner Evolutions
Proof. In the ideal world, we will have something that helps us directly compute
area(f (D)). However, for the derivation to work, we need to talk about what
f does to the boundary of D, but that is not necessarily well-defined. So we
compute area(f (rD)) for r < 1 and then take the limit r → 1.
So fix r ∈ (0, 1), and define the curve γ(θ) = f (reiθ ) for θ ∈ [0, 2π]. Then we
can compute
Z Z
1 1
z̄ dz = (x − iy)(dx + i dy)
2i γ 2i γ
Z
1
= (x − iy) dx + (ix + y) dy
2i γ
ZZ
1
= 2i dx dy
2i f (rD)
= area(f (rD)),
using Green’s theorem. We can also compute the left-hand integral directly as
1 2π
Z Z
1
z̄ dz = f (reiθ )f 0 (reiθ )ireiθ dθ
2i γ 2i 0
∞
! ∞ !
1 2π X
Z X
= ān rn e−iθn an nrn−1 eiθ(n−1) ireiθ dθ
2i 0 n=1 n=1
∞
X
=π r2n |an |2 n.
n=1
8
1 Conformal transformations III Schramm–Loewner Evolutions
Proof. The proof is essentially the same as last time. Let r > 1, and let
Kr = F (rD̄) (or, if you wish, C \ F (C \ rD̄)), and γ(θ) = F (reiθ ). As in the
previous proposition, we have
Z
1
area(Kr ) = z̄ dz
2i γ
1 2π
Z
= F (reiθ )F 0 (reiθ )ireiθ dθ
2i 0
∞
!
X
2 2 −2n
=π r − n|bn | r .
n=1
9
1 Conformal transformations III Schramm–Loewner Evolutions
Since g maps the real line to the real line, all bi must be real. Moreover, by
injectivity, considering large z shows that N = 1. In other words, we can write
∞
X bn
g(z) = b−1 z + b0 + ,
n=1
zn
10
1 Conformal transformations III Schramm–Loewner Evolutions
Since b0 and b−1 are both real, this is still a conformal transformation, and
|gA (z) − z| → 0 as z → ∞.
0 0 −1
To show uniqueness, suppose gA , gA are two such functions. Then gA ◦ gA :
H → H is such a function for A = ∅. Thus, it suffices to show that if g : H → H
is a conformal mapping such that g(z) − z → 0 as z → ∞, then in fact g = z.
But we can expand g(z) − z as
∞
X cn
g(z) − z = ,
n=1
zn
11
1 Conformal transformations III Schramm–Loewner Evolutions
Observe that these results together imply that if A ∈ Q and A ⊆ r(D̄ ∩ H),
then
hcap(A) ≤ hcap(r(D̄ ∩ H)) ≤ r2 hcap(D̄ ∩ H) = r2 .
So we know that hcap(A) ≤ diam(A)2 .
Compared to the above proofs, it seems much less straightforward to prove
non-negativity of the half-plane capacity. Our strategy for doing so is to relate
the half-plane capacity to Brownian motion! This is something we will see a lot
in this course.
Proposition. Let A ∈ Q and Bt be complex Brownian motion. Define the
stopping time
τ = inf{t ≥ 0 : Bt 6∈ H \ A}.
Then
(i) For all z ∈ H \ A, we have
(ii)
hcap(A) = lim y Ey [im(Bτ )].
y→∞
In particular, hcap(A) ≥ 0.
(iii) If A ⊆ D̄ ∩ H, then
Z π
2
hcap(A) = Eeiθ [im(Bτ )] sin θ dθ.
π 0
Proof.
(i) Let φ(z) = im(z − gA (z)). Since z − gA (z) is holomorphic, we know φ is
harmonic. Moreover, ϕ is continuous and bounded, as it → 0 at infinity.
These are exactly the conditions needed to solve the Dirichlet problem
using Brownian motion.
Since im(gA (z)) = 0 when z ∈ ∂(H \ A), we know im(Bτ ) = im(Bt −
gA (Bτ )). So the result follows.
(ii) We have
where we use the fact that hcap(A) is real, so we can take the limit of the
real part instead.
12
1 Conformal transformations III Schramm–Loewner Evolutions
τ = inf{t ≥ 0 : Bt 6∈ D}
τ̃ = inf{t ≥ 0 : B̃t 6∈ D̃}
Set
Z τ
τ0 = |f 0 (Bs )|2 ds
0
Z s
0 2
σ(t) = inf s ≥ 0 : |f (Br )| dr = t
0
Bt0 = f (Bσ(t) ).
1 − |z|2
iθ 1
f (e ) = .
2π |eiθ − z|
Similarly, on H, starting from z = x + iy, the exit distribution is
1 y
f (u) = .
π (x − u)2 + y 2
Note that if x = 0, y = 1, then this is just
1 1
f (u) = .
π u2 + 1
This is the Cauchy distribution!
13
2 Loewner’s theorem III Schramm–Loewner Evolutions
2 Loewner’s theorem
2.1 Key estimates
Before we prove Loewner’s theorem, we establish some key identities and esti-
mates. As before, a useful thing to do is to translate everything in terms of
Brownian motion.
Proposition. Let A ∈ Q and B be a complex Brownian motion. Set
τ = inf{t ≥ 0 : Bt 6∈ H \ A}.
Then
– If x > Rad(A), then
1
gA (x) = lim πy − Piy [Bτ ∈ [x, ∞)] .
y→∞ 2
where the first equality follows from the fact that Brownian motion exits through
the positive reals with probability 12 ; the second equality follows from the
previously computed exit distribution; and the last follows from dominated
convergence.
Now suppose A 6= ∅. We will use conformal invariance to reduce this to the
case above. We write gA = uA + ivA . We let
Then we know
Piy [Bτ ∈ [x, ∞)] = PgA (iy) [Bσ ∈ [gA (x), ∞)]
= PivA (iy) [Bσ ∈ [gA (x) − uA (iy), ∞)].
14
2 Loewner’s theorem III Schramm–Loewner Evolutions
hcap(A)
h(z) = z + − gA (z).
z
We aim to control the imaginary part of this, and then use the Cauchy–Riemann
equations to control the real part as well. We let
im(z)
v(z) = im(h(z)) = im(z − gA (z)) = hcap(A).
|z|2
σ = inf{t ≥ 0 : Bt 6∈ H \ D̄}
τ = inf{t ≥ 0 : Bt 6∈ H}.
Let p(z, eiθ ) be the density with respect to the Lebesgue measure at eiθ for Bσ .
Then by the strong Markov property at the time σ, we have
Z π
im(z − gA (z)) = Eeiθ [im(Bτ )]p(z, eiθ ) dθ.
0
We also have Z π
2
hcap(A) = Eeiθ [im(Bτ )] sin θ dθ.
π 0
15
2 Loewner’s theorem III Schramm–Loewner Evolutions
So
im(z)
|v(z)| = im(z − gA (z)) − hcap(A)
|z|2
Z π
im(z) 2 π
Z
= Eeiθ [im(Bτ )]p(z, eiθ ) dθ − · E iθ [im(Bτ )] sin θ dθ .
0 |z|2 π 0 e
By applying (∗), we get
c hcap(A) im(z)
|v(z)| ≤ c .
|z|3
where c is a constant.
Recall that v is harmonic as it is the imaginary part of a holomorphic function.
By example sheet 1 Q9, we have
c hcap(A) c hcap(A)
|∂x v(z)| ≤ , |∂y v(z)| ≤ .
|z|3 |z|3
By the Cauchy–Riemann equations, re(h(z)) satsifies the same bounds. So we
know that
c hcap(A)
|h0 (z)| ≤ .
|z|3
Then Z ∞
h(iy) = h0 (is) ds,
y
To get the desired bound for a general z, integrate h0 along the boundary of the
circle of radius |z| to get
c hcap(A)
|h(z)| = |h(reiθ )| ≤ + h(iz).
|z|2
The following is a very useful fact about Brownian motion:
Theorem (Beurling estimate). There exists a constant c > 0 so that the
following holds. Let B be a complex Brownian motion, and A ⊆ D̄ be connected,
0 ∈ A, and A ∩ ∂ D̄ 6= ∅. Then for z ∈ D, we have
Pz [B[0, τ ] ∩ A = ∅] ≤ c|z|1/2 ,
16
2 Loewner’s theorem III Schramm–Loewner Evolutions
where
d = diam(Ã \ A), r = sup{im(z) : z ∈ Ã}.
Proof. By scaling, we can assume that r = 1.
– If d ≥ 1, then the result follows since
and so
τ = inf{t ≥ 0 : Bt 6∈ H \ A}.
17
2 Loewner’s theorem III Schramm–Loewner Evolutions
(ii) locally growing if for all T > 0 and ε > 0, there exists δ > 0 such that
whenever 0 ≤ s ≤ t ≤ s + δ ≤ T , we have
diam(gAs (At \ As )) ≤ ε.
This is a continuity condition.
(iii) parametrized by half-plane capacity if hcap(At ) = 2t for all t ≥ 0.
We write A be the set of families of compact H-hulls which satisfy (i) to (iii).
We write AT for the set of such families defined on [0, T ].
Example. If γ is is a simple curve in H starting from 0 and At = γ[0, t]. This
clearly satisfies (i), and the previous proposition tells us this satsifies (ii). On the
first example sheet, we show that we can reparametrize γ so that hcap(At ) = 2t
for all t ≥ 0. Upon doing so, we have At ∈ A.
Theorem. Suppose that (At ) ∈ A. Let gt = gAt . Then there exists a continuous
function U : [0, ∞) → R so that
2
∂t gt (z) = , g0 (z) = z.
gt (z) − Ut
This is known as the chordal Loewner ODE , and U is called the driving
function.
Proof.
T First note that since the hulls are locally growing, the intersection
g
s≥t (As ) consists of exactly one point. Call this point Ut . Again by the
t
locally growing property, Ut is in fact a continuous function in t.
Recall that if A ∈ Q, then
hcap(A) hcap(A)Rad(A)
gA (z) = z + +O .
z |z|2
If x ∈ R, then as gA+x (z) − x = gA (z − x), we have
hcap(z) hcap(A)Rad(A + x)
gA (z) = gA (z + x) − x = z + +O . (∗)
z+x |z + x|2
Fix ε > 0. For 0 ≤ s ≤ t, let
gs,t = gt ◦ gs−1 .
Note that
hcap(gT (At+ε \ At )) = 2ε.
Apply (∗) with A = gt (At+ε \ At ), x = −Ut , and use that Rad(A + x) =
Rad(A − Ut ) ≤ diam(A) to see that
2ε 1
gA (z) = gt,t+ε (z) = z + + 2ε diam(gt (At+ε \ At ))O .
z − Ut |z − Ut |2
So
gt+ε (z) − gt (z) = (gt,t+ε − gt,t ) ◦ gt (z)
2ε 1
= + 2ε diam(gt (At+ε \ At ))O .
gt (z) − Ut |gt (z) − Ut |2
Dividing both sides by ε and taking the limit as ε → 0, the desired result follows
since diam(gt (At+ε \ At )) → 0.
18
2 Loewner’s theorem III Schramm–Loewner Evolutions
At = H \ domain(gt ).
On the example sheet, we show that this is indeed a family in A, and gt are
indeed conformal transformations.
19
2 Loewner’s theorem III Schramm–Loewner Evolutions
20
3 Review of stochastic calculus III Schramm–Loewner Evolutions
Stochastic integral
If Xt = Mt + At is a continuous semi-martingale, and Ht is a previsible process,
we set Z t Z t Z t
Hs dXs = Hs dMs + Hs dAs ,
0 0 0
where the first integral is the Itô integral, and the second is the Lebesgue–Stieltjes
integral. The first term is a continuous local martingale, and the second is a
continuous bounded variation process.
The Itô integral is defined in the same spirit as the Riemann integral. The
key thing that makes it work is that there is “extra cancellation” in the definition,
from the fact that we are integrating against a martingale, and so makes the
integral converge even though the process we are integrating against is not of
bounded variation.
Quadratic variation
If M is a continuous local martingale, then the quadratic variation is
d2n te−1
X
[M ]t = lim (M(k+1)2−n − Mk2−n )2 ,
n→∞
k=0
21
3 Review of stochastic calculus III Schramm–Loewner Evolutions
Itô’s formula
Itô’s formula is the stochastic calculus’ analogue of the fundamental theorem of
calculus. It takes a bit of work to prove Itô’s formula, but it is easy to understand
what the intuition is. If f ∈ C 2 , then we have
n
X
f (t) = f (0) + (f (tk ) − f (tk−1 ))
k=1
Taking the limit, and using that E[(Btk − Btk−1 )2 ] = tk − tk−1 , we get
Z t Z t
1
f (Bt ) = f (0) + f 0 (Bs ) dBs + f 00 (Bs ) ds.
0 2 0
22
3 Review of stochastic calculus III Schramm–Loewner Evolutions
for all t ≥ 0. At the end of the stochastic calculus course, we will see that there
is a unique solution to this equation provided b, σ are Lipschitz functions.
This in particular implies we can solve
2
∂t gt (z) = , g0 (z) = z
gt (z) − Ut
√
for Ut = κBt , where Bt is a standard Brownian motion.
23
4 Phases of SLE III Schramm–Loewner Evolutions
4 Phases of SLE
We now study some basic properties of SLEκ . Our first goal will be to prove the
following theorem:
24
4 Phases of SLE III Schramm–Loewner Evolutions
[Z]t = 4Zt dt
U02−d = E[Uτ2−d
a ∧τb
] = a2−d P[τa < τb ] + b2−d P[τb < τa ].
25
4 Phases of SLE III Schramm–Loewner Evolutions
This is then the time γ cuts x off from ∞. We thus want to understand P[τx < ∞].
We can calculate
2 √ 2 √
dVtx = d(gt (x) − Ut ) = dt − κ dBt = x dt − κ dBt .
gt (x) − Ut Vt
This looks almost like the Bessel SDE, but we need to do some rescaling to write
this as x
V 2/κ
d √t = × √ dt + dB̃t , B̃t = −Bt .
κ Vt / κ
√
So we get that Vt× / κ ∼ BESd , with d = 1 + 4/κ. Thus, our previous result
implies (
1 κ>4
P[τx < ∞] = .
0 κ≤4
26
4 Phases of SLE III Schramm–Loewner Evolutions
≥ P[gm (Am ) − Um ⊇ [−n, n] and gm (γ(t + m)) − Um ∈ [−n, n] for some t].
Since this is true for all m and n, we can take the limit m → ∞, then the limit
n → ∞ to obtain the desired result.
It turns out there are two further regimes when κ > 4.
Theorem. If κ ≥ 8, then SLEκ is space-filling, but not if κ ∈ (4, 8).
This will be shown in the second example sheet. Here we prove a weaker
result:
Proposition. SLEκ fills ∂H iff κ ≥ 8.
To phrase this in terms of what we already have, we fix κ > 4. Then for any
0 < x < y, we want to understand the probability
g(x, y) = P[τx = τy ].
where we assume 0 < x < y. If τx = τy , this means the curve γ cuts x and y from
∞ ft the same time, and hence doesn’t hit [x, y]. Conversely, if P[τx = τy ] = 0,
then with probability one, γ hits (x, y), and this is true for all x and y. Thus,
we want to show that g(x, y) > 0 for κ ∈ (4, 8), and g(x, y) = 0 for κ ≥ 8.
To simplify notation, observe that g(x, y) = g(1, y/x) as SLEκ is scale-
invariant. So we may assume x = 1. Moreover, we know
g(1, r) → 0 as r → ∞,
P[A \ B] = P[B \ A] = 0,
Vtr − Vt1
sup < ∞. (∗)
t<τ1 Vt1
27
4 Phases of SLE III Schramm–Loewner Evolutions
Proof. If (∗) happens, then we cannot have that τ1 < τr , or else the supremum
is infinite. So (∗) ⊆ {τ1 = τr }. The prove the proposition, we have to show that
Vtr − Vt1
P τ1 = τr , sup = ∞ = 0.
t<τ1 Vt1
Vr −V1
σM = inf t ≥ 0 : t 1 t ≥ M .
Vt
But at time σM , we have Vtr = (M + 1)Vt1 , and τ1 = τr if these are cut off at
the same time. Thus, the conformal Markov property implies
PM = g(1, M + 1).
So we are done.
So we need to show that
(
Vtr − Vt1
> 0 κ ∈ (4, 8)
P sup 1 <∞ .
t<τ1 Vt =0 k≥8
We will prove this using stochastic calculus techniques. Since ratios are hard, we
instead consider r
Vt − Vt1
Zt = log .
Vt1
We can then compute
d − 1 Vtr − Vt1
3 1 1
dZt = −d + dt − 1 dBt , Z0 = log(r − 1).
2 (Vt1 )2 2 (Vt1 )2 Vtr Vt
dσ(t)
σ(0) = 0, dt = 1 )2 .
(Vσ(t)
So the map
Z σ(t)
1
t 7→ B̃t = − dBs
0 Vs1
is a continuous local martingale, with
" Z #
σ(t) Z σ(t)
1 1
[B̃]t = − 1
dBs = ds = t.
0 Vs 0 (Vs1 )2
t
28
4 Phases of SLE III Schramm–Loewner Evolutions
So we find that
sup Z̃t ≥ Z̃0 + sup B̃t = +∞.
t t
So we find that
Vtr − Vt1
sup = ∞.
t<τ1 Vt1
So g(x, y) = 0 for all 0 < x < y if κ ≥ 8.
Now if κ ∈ (4, 8), we pick ε > 0 and set
ε
r =1+ .
2
Then we have Z̃0 = log(r − 1) = log(ε/2). We will show that for small ε, we have
g(1, r) > 0. In fact, it is always positive, which is to be shown on the example
sheet.
We let
τ = inf{t > 0 : Z̃t = log ε}.
Then
r 1
d − 1 t∧τ Vσ(s) − Vσ(s)
Z
3
Z̃t∧τ = Z̃0 + B̃t∧τ + −d t∧τ + r ds
2 2 0 Vσ(s)
d − 1 t∧τ Z̃s
Z
3
≤ Z̃0 + B̃t∧τ + −d t∧τ + e ds
2 2 0
3 d−1
≤ Z̃0 + B̃t∧τ + −d t∧τ − (t ∧ τ )ε
2 2
3 d−1
= Z̃0 t + B̃t∧τ + −d + ε (t ∧ τ ).
2 2
We let
3 d−1
Zt∗ = Z̃0 t + B̃t + −d+ ε t.
2 2
29
4 Phases of SLE III Schramm–Loewner Evolutions
So we know that
P sup Z̃t < log ε > 0.
t≥0
So we have ε
g 1, 1 + > 0.
2
This concludes the proof of the theorem.
30
5 Scaling limit of critical percolation III Schramm–Loewner Evolutions
Then there exists a unique interface γε that connects x to y with the property
that the black hexagons on its left and white on its right. It was conjectured
(and now proved by Smirnov) that the limit of the law of γε exists in distribution
and is conformally invariant.
This means that if D̃ is another simply connected domain, and x̃, ỹ ∈ ∂ D̃ are
distinct, then ϕ : D → D̃ is a conformal transformation with ϕ(x) = x̃, ϕ(y) = ỹ,
then ϕ(γ) is equal in distribution of the scaling limit of percolation in D̃ from x̃
to ỹ.
Also, percolation also satisfies a natural Markov property: if we condition on
γε up to a given time t, then the rest of γε is a percolation exploration in the
remaining domain. The reason for this is very simple — the path only determines
the colours of the hexagons right next to it, and the others are still randomly
distributed independently.
If we assume these two properties, then the limiting path γ satisfies the
conformal Markov property, and so must be an SLEκ . So the question is, what
is κ?
To figure out this κ, we observe that the scaling limit of percolation has a
locality property, and we will later see that SLE6 is the only SLE that satisfies
this locality property.
To explain the locality property, fix a simply-connected domain D in H (for
simplicity), and assume that 0 ∈ ∂D. Fixing a point y ∈ ∂D, we perform the
percolation exploration as before. Then the resulting path would look exactly
the same as if we performed percolation on H (with black boundary on R<0 and
white boundary on R>0 ), up to the point we hit ∂D \ ∂H. In other words, γ
31
5 Scaling limit of critical percolation III Schramm–Loewner Evolutions
doesn’t feel the boundary conditions until it hits the boundary. This is known
as locality.
It should then be true that the corresponding SLEκ should have an analogous
property. To be precise, we want to find a value of κ so that the following is
true: If γ is an SLEκ in H from 0 to ∞, run up until it first hits ∂D \ ∂H, then
ψ(γ) is a (stopped) SLEκ in H from 0 to ∞ where ψ : D → H is a conformal
transformation with ψ(0) = 0, ψ(y) = ∞. This is the locality property. We will
show that locality holds precisely for κ = 6.
Suppose that (At ) ∈ A has Loewner driving function Ut . We define
Ãt = ψ(At ).
Then Ãt is a family of compact H-hulls that are non-decreasing, locally growing,
and A0 = ∅. However, in general, this is not going to be parametrized by capacity.
On the second example sheet, we show that this has half plane capacity
Z t
ã(t) = hcap(Ãt ) = γ(ψt0 (Us ))2 ds,
0
We then have
∂t ãt
∂t g̃t (z) = , g̃0 (z),
g̃t (z) − Ũt
To see this, simply recall that if At is γ([0, t]) for a curve γ, then Ut = gt (γ(t)).
To understand Ũt , it is convenient to know something about ψt :
In particular, at z = Ut , we have
Then we have
2
∂t g̃σ(t) (z) = , g̃0 (z) = z.
g̃σ(t) − Ũσ(t)
32
5 Scaling limit of critical percolation III Schramm–Loewner Evolutions
It then remains to try to understand √ dŨσ(t) . Note that so far what we said
works for general Ut . If we put Ut = κBt , where B is a standard Brownian
motion, then Itô’s formula tells us before the time change, we have
where Z σ(t)
B̃t = ψs0 (Us ) dBs
0
is a standard Brownian motion by the Lévy characterization and the definition
of σ(t).
The point is that when κ = 6, the drift term goes away, and so
√
Ũσ(t) = 6B̃t .
33
6 Scaling limit of self-avoiding walks III Schramm–Loewner Evolutions
Brownian excursion
When studying the restriction property, we will encounter another type of process,
known as Brownian excursion. Fix a simply connected domain D ⊆ C and
points x, y ∈ ∂D distinct. Then roughly speaking, a Brownian excursion is a
Brownian motion starting at x conditioned to leave D at ∂D. Of course, we
cannot interpret these words literally, since we want to condition on a zero
probability event.
To make this rigorous, we first use conformal invariance to say we only have
to define this for H with x = 0, y = ∞.
To construct it in H, we start with a complex Brownian motion B = B 1 +iB 2 ,
with B01 = 0 and B02 = ε > 0. We then condition B on the event that B 2 hits
R 0 before hitting 0. This is a positive probability event. We then take limits
R → ∞ and ε → 0. On the second example sheet, we will show that this makes
sense, and the result is called Brownian excursion.
It turns out the limiting object is pretty simple. It is given by
B̂ = (B̂ 1 , B̂ 2 )
where B̂ 1 is a standard Brownian motion and B̂ 2 ∼ BES3 , and the two are
independent.
The key property about Brownian excursion we will need is the following:
Proposition. Suppose A be a compact H-hull and gA is as usual. If x ∈ R \ A,
then
0
Px [B̂[0, ∞) ∩ A = ∅] = gA (x).
34
6 Scaling limit of self-avoiding walks III Schramm–Loewner Evolutions
Pz [B[0, σR ] ∩ (A ∪ R) = ∅]
lim lim .
ε→0 R→∞ P[B[0, σR ] ∩ R = ∅]
So we get
im(gA (z)) im(gA (z))
≤ numerator ≤ .
R + 3Rad(A) R − 3Rad(Z)
Combining, we find that the desired probability is
im(gA (x + iε)) 0
lim = gA (x).
ε→0 ε
Q+ = {A ∈ Q : Ā ∩ (−∞, 0] = ∅}
Q− = {A ∈ Q : Ā ∩ [0, ∞) = ∅}
For A ∈ Q± = Q+ ∪ Q− , we define ψA : H → A → H by
This is the version of gA that fixes 0, which is what we need. This is the unique
conformal transformation with
ψA (z)
ψA (0) = 0, lim = 1.
z→∞ z
We will need the following fact about SLE:
35
6 Scaling limit of self-avoiding walks III Schramm–Loewner Evolutions
This is not very difficult to prove, but the proof is uninteresting and we have
limited time. Since SLEκ is simple for κ ≤ 4 and is also transient, it follows that
for all A ∈ Q± ,
0 < P[γ[0, ∞) ∩ A = ∅] < 1.
This is useful because we want to condition on this event. Write
VA = {γ[0, ∞) ∩ A = ∅}.
M̃t = P[VA | Ft ]
M̃t → 1VA
36
6 Scaling limit of self-avoiding walks III Schramm–Loewner Evolutions
then we get
Mt = E[1VA | Ft ] = M̃t .
37
6 Scaling limit of self-avoiding walks III Schramm–Loewner Evolutions
sup{im(ω) : ω ∈ A} = 1.
Note that σr < ∞ almost surely for all r > 0 since SLE8/3 is transient.
Lemma. Mt∧τ → 1 on VA as t → ∞.
Proof. Let B̂ be a Brownian excursion in H \ γ[0, σr ] from γ(σr ) to ∞. Let B
be a complex Brownian motion, and
We will show that this expression is ≤ Cr−1/2 for some constant C > 0. Then we
know that Mσr ∧τ → 1 as r → ∞ on VA . This is convergence along a subsequence,
but since we already know that Mt∧τ converges this is enough.
We first tackle the denominator, which we want to bound from below. The
idea is to bound the probability that the Brownian motion reaches the lime
im(z) = r + 1 without hitting R ∪ γ[0, σr ]. Afterwards, the gambler’s ruin
estimate tells us the probability of reaching im(z) = R without going below the
1
im(z) = r line is R−r .
In fact, we shall consider the box S = [−1, 1]2 +γ(σr ) of side length 2 centered
at γ(σr ). Let η be the first time B leaves S, and we want this to leave via the
top edge `. By symmetry, if we started right at γ(σr ), then the probability of
leaving at ` is exactly 14 . Thus, if we are at z = γ(σr ) + iε, then the probability
of leaving via ` is > 41 .
What we would want to show is that
1
Pz [B(η) ∈ ` | B[0, η] ∩ γ[0, σr ] = ∅] > . (†)
4
We then have the estimate
1 1
denominator ≥ · Pz [B[0, η] ∩ γ[0, σr ] = ∅] · .
4 R−r
Intuitively, (†) must be true, because γ[0, σr ] lies below im(z) = r, and so if
B[0, η] doesn’t hit γ[0, σr ], then it is more likely to go upwards. To make this
38
6 Scaling limit of self-avoiding walks III Schramm–Loewner Evolutions
rigorous, we write
1
< Pz [B(η) ∈ `]
4
= Pz [B(η) ∈ ` | B[0, η] ∩ γ[0, σr ] = ∅] P[B[0, η] ∩ γ[0, σr ] = ∅]
+ Pz [B(η) ∈ ` | B[0, η] ∩ γ[0, σr ] 6= ∅] P[B[0, η] ∩ γ[0, σr ] 6= ∅]
To prove (†), it suffices to observe that the first factor of the second term is is
≤ 14 , which follows from the strong Markov property, since Pw [B(η) ∈ `] ≤ 41
whenever im(w) ≤ r, which in particular is the case when w ∈ γ[0, σr ].
To bound the numerator, we use the strong Markov property and the Beurling
estimate to get
Pz [B hits A without hitting R ∪ γ[0, σr ]] ≤ Pz [B[0, η] ∩ γ[0, σr ]] · Cr−1/2 .
Combining, we know the numerator in (∗) is
1
≤ C · r−1/2 · P[B[0, η] ∩ γ[0, σr ] = ∅].
R
These together give the result.
Lemma. Mt∧τ → 0 as t → ∞ on VAc .
This has a “shorter” proof, because we outsource a lot of the work to the
second example sheet.
Proof. By the example sheet, we may assume that A is bounded by a smooth,
simple curve β : (0, 1) → H.
Note that γ(τ ) = β(s) for some s ∈ (0, 1). We need to show that
lim ψt0 (Ut ) = 0.
t→τ
For m ∈ N, let
1
tm = inf t ≥ 0 : |γ(t) − β(s)| =
m
Since β is smooth, there exists δ > 0 so that
` = [β(s), β(s) + δn] ⊆ A,
where n is the unit inward pointing normal at β(s). Let
Lt = gt (`) − Ut .
Note that a Brownian motion starting from a point on ` has a uniformly positive
chance of exiting H \ γ[0, tm ] on the left side of γ[0, tm ] and on the right side as
well.
On the second example sheet, we see that this implies that
Ltm ⊆ {w : im(w) ≥ a| Re(w)|}
for some a > 0, using the conformal invariance of Brownian motion. Intuitively,
this is because after applying gt − Ut , we have uniformly positive probability of
exiting via the positive or real axis, and so we cannot be too far away in one
directionn.
Again by the second example sheet, the Brownian excursion in H from 0 to
∞ hits Ltm with probability → 1 as m → ∞.
39
6 Scaling limit of self-avoiding walks III Schramm–Loewner Evolutions
We thus conclude
Note that this is the same as the probability that the hull of γ1 , . . . , γ8 des not
intersect A, where Sthe hull is the union of the γj ’s together with the bounded
components of H \ j γj .
In the same manner, if B̂1 , . . . , B̂5 are independent Brownian excursions,
then
0
P[B̂j [0, ∞) ∩ A = ∅ for all j] = (ψA (0))5 .
Thus, the hull of γ1 , . . . , γ8 has the same distribution as the hull of B̂1 , . . . , B̂5 .
Moreover, if we take a boundary point of the hull of, say, γ1 , . . . , γ8 , then
we would expect it to belong to just one of the SLE8/3 ’s. So the boundary of
the hull of γ1 , . . . , γ8 looks locally like an SLE8/3 , and the same can be said for
the hull fo B̂1 , . . . , B̂5 . Thus, we conclude that the boundary of a Brownian
excursion looks “locally” like SLE8/3 .
40
7 The Gaussian free field III Schramm–Loewner Evolutions
Then
H01 (D) = Hsupp (U ) ⊕ Hharm (U )
is an orthogonal decomposition of H01 (D). This is going to translate to a
Markov property of the Gaussian free field.
41
7 The Gaussian free field III Schramm–Loewner Evolutions
Moreover, if x, y ∈ Rn , then (h, x) and (h, y) are jointly Gaussian with covariance
(x, y).
Thus, we can associate with h a family of Gaussian random variables (h, x)
indexed by x ∈ Rn with mean zero and covariance given by the inner product
on Rn . This is an example of a Gaussian Hilbert space.
We now just do the same for the infinite-dimensional vector space H01 (D).
One can show that this is separable, and so we can pick an orthonormal basis
(fn ). Then the Gaussian free field h on D is defined by
∞
X
h= αj fj ,
j=1
where the αj ’s are iid N (0, 1) random variables. Thus, if f ∈ H01 (D), then
More generally, if f, g ∈ H01 (D), then (h, f )∇ and (h, g)∇ are jointly Gaussian
with covariance (f, g)∇ . Thus, the Gaussian free field is a family of Gaussian
variables (H, f )∇ indexed by f ∈ H01 (D) with mean zero and covariance ( · , · )∇ .
42
7 The Gaussian free field III Schramm–Loewner Evolutions
P∞
We can’t actually quite make this definition, because the sum h = j=1 αj fj
does not converge. So h is not really a function, but a distribution. However,
this difference usually does not matter.
We can translate the properties of H01 into analogous properties of the
Gaussian free field.
Proposition.
(i) If ϕ : D → D̃ is a conformal transformation and h is a Gaussian free field
on D, then h ◦ ϕ−1 is a Gaussian free field on D̃.
(ii) Markov property: If U ⊆ D is open, then we can write h = h1 + h2 with
h1 and h2 independent where h1 is a Gaussian free field on U1 and h2 is
harmonic on U .
Proof.
(i) Clear.
(ii) Take h1 to be the projection onto Hsupp (U ). This works since we can take
the orthonormal basis (fn ) to be the union of an orthonormal basis of
Hsupp (U ) plus an orthonormal basis of Hharm (U ).
Often, we would rather think about the L2 inner product of h with something
else. Observe that integration by parts tells us
−1
(h, f )∇ =
(h, ∆f )L2 .
2π
Thus, we would be happy if we can invert ∆, which we can by IB methods.
Recall that
∆(− log |x − y|) = −2πδ(y − x),
where ∆ acts on x, and so − log |x − y| is a Green’s function for ∆. Given a
domain D, we wish to obtain a version of the Green’s function that vanishes on
the boundary. To do so, we solve ∆G̃x = 0 on D with the boundary conditions
G̃x (y) = − log |x − y| if y ∈ ∂D.
We can then set
G(x, y) = − log |x − y| − G̃x (y).
With this definition, we can define
Z
−1 1
∆ ϕ(x) = − G(x, y)ϕ(y) dy.
2π
Then ∆∆−1 ϕ(x) = ϕ(x), and so
(h, ϕ) ≡ (h, ϕ)L2 = −2π(h, ∆−1 ϕ)∇ .
Then (h, ϕ) is a mean-zero Gaussian with variance
(2π)2 k∆−1 ϕk2∇ = (2π)2 (∆−1 ϕ, ∆−1 ϕ)∇
= −2π(∆−1 ϕ, ∆∆−1 ϕ)
= (−2π∆−1 ϕ, ϕ)
ZZ
= ϕ(x)G(x, y)ϕ(y) dx dy.
43
7 The Gaussian free field III Schramm–Loewner Evolutions
It is not hard to show that the Green’s function is in fact conformally invariant:
Proposition. Let D, D̃ be domains in C and ϕ is a conformal transformation
D → D̃. Then GD (x, y) = GD̃ (ϕ(x), ϕ(y)).
One way to prove this is to use the conformal invariance of Brownian motion,
but a direct computation suffices as well.
Schramm and Sheffield showed that the level sets of h, i.e. {x : h(x) = 0}
are SLE4 ’s. It takes some work to make this precise, since h is not a genuine
function, but the formal statement is as follows:
Theorem (Schramm–Sheffield). Let λ = π2 . Let γ ∼ SLE √4 in H from 0 to ∞.
Let gt its Loewner evolution with driving function Ut = κBt = 2Bt , and set
ft = gt − Ut . Fix W ⊆ H open and let
τ = inf{t ≥ 0 : γ(t) ∈ W }.
(h + 𝒽)(x) = sgn(x)λ.
44
7 The Gaussian free field III Schramm–Loewner Evolutions
θ2
h i
E eiθ((h+𝒽)◦ft∧τ ,ϕ) = exp iθm0 (ϕ) − σ02 (ϕ) .
2
If we knew that
θ2 2
exp iθmt (ϕ) − σt (ϕ)
2
is a martingale, then taking the expectation of the above equation yields the
desired results.
Note that this looks exactly like the form of an exponential martingale, which
in particular is a martingale. So it suffices to show that mt (ϕ) is a martingale
with
[m· (ϕ)]t = σ02 (ϕ) − σt2 (ϕ).
To check that mt (ϕ) is a martingale, we expand it as
2λ 2λ
𝒽 ◦ ft (z) = λ − arg(ft (z)) = λ − im(log(gt (z) − Ut )).
π π
So it suffices to check that log(gt (z) − Ut ) is a martingale. We apply Itô’s formula
to get
1 2 1 κ/2
d log(gt (z)−Ut ) = · dt− dUt − dt,
gt (z) − Ut gt (z) − Ut gt (z) − Ut (gt (z) − Ut )2
45
7 The Gaussian free field III Schramm–Loewner Evolutions
To finish the proof, we need to show that dσt2 (ϕ) takes the same form. Recall
that the Green’s function can be written as
Since we have
log(ft (x) − ft (y)) = log(gt (x) − gt (y)),
we can compute
1 2 2
d log(gt (x) − gt (y)) = − dt
gt (x) − gt (y) gt (x) − Ut gt (y) − Ut
−2
= dt.
(gt (x) − Ut )(gt (y) − Ut )
Similarly, we have
−2
d log(gt (x) − gt (y)) = dt.
(gt (x) − Ut )(gt (y) − Ut )
So we have
2 2
dGt (x, y) = − im im dt.
gt (x) − Ut gt (y) − Ut
46
Index III Schramm–Loewner Evolutions
Index
47