Part III Percolation and Random Walks on
Graphs
Based on lectures by P. Sousi
Notes taken by Dexter Chua
Michaelmas 2017
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.
A phase transition means that a system undergoes a radical change when a continuous
parameter passes through a critical value. We encounter such a transition every day
when we boil water. The simplest mathematical model for phase transition is percolation.
Percolation has a reputation as a source of beautiful mathematical problems that are
simple to state but seem to require new techniques for a solution, and a number of
such problems remain very much alive. Amongst connections of topical importance
are the relationships to so-called Schramm–Loewner evolutions (SLE), and to other
mo dels from statistical physics. The basic theory of percolation will be described in
this course with some emphasis on areas for future development.
Our other major topic includes random walks on graphs and their intimate connection
to electrical networks; the resulting discrete potential theory has strong connections
with classical potential theory. We will develop tools to determine transience and
recurrence of random walks on infinite graphs. Other topics include the study of
spanning trees of connected graphs. We will present two remarkable algorithms to
generate a uniform spanning tree (UST) in a finite graph
G
via random walks, one
due to Aldous-Broder and another due to Wilson. These algorithms can be used to
prove an important property of uniform spanning trees discovered by Kirchhoff in the
19th century: the probability that an edge is contained in the UST of
G
, equals the
effective resistance between the endpoints of that edge.
Pre-requisites
There are no essential pre-requisites beyond probability and analysis at undergraduate
levels, but a familiarity with the measure-theoretic basis of probability will be helpful.
Contents
0 Introduction
1 Percolation
1.1 The critical probability
1.2 Correlation inequalities
1.3 Two dimensions
1.4 Conformal invariance and SLE in d = 2
2 Random walks
2.1 Random walks in finite graphs
2.2 Infinite graphs
3 Uniform spanning trees
3.1 Finite uniform spanning trees
3.2 Infinite uniform spanning trees and forests
0 Introduction
This course is naturally divided into two parts percolation, and random
walks on graphs. Percolation is one of the simplest models that experience phase
transition — an abrupt change in quantitative feature due to a continuous change
of a parameter. More sophisticated examples of percolation the boiling of water
and the loss of long-range correlation in magnets when temperature increases.
But we are not physicists. So let’s talk about percolation. For reasons that
become clear later, consider an n × (n + 1) lattice connected by edges:
We now fix some
p
[0
,
1], and for each edge in the graph, we either keep it
or remove it with probability
p
. There are many questions we may ask. For
example, we may ask for the probability that there is a left-to-right crossing of
open edges.
For example, we have f
n
(0) = 0 and f
n
(1) = 1.
An interesting choice of
p
to consider is
p
=
1
2
. We can argue that
f
n
(
1
2
)
must be
1
2
, by symmetry. More precisely, consider the dual lattice:
Note that this lattice is isomorphic to the original lattice we were thinking about,
by applying a rotation. Now each edge in the dual lattice crosses exactly one edge
in the original lattice, and we can set the edge to be open iff the corresponding
edge in the original lattice is open. This gives rise to a percolation on the dual
lattice with p =
1
2
as well.
Now notice that that there is a left-right crossing of open edges in the original
lattice iff there is no top-bottom crossing in the dual graph. But since the dual
and original lattices are isomorphic, it follows that the probability that there is
a left-right crossing in the original lattice is equal to the probability that there
is no left-right crossing in the original lattice. So both of these must be
1
2
.
The ability to talk about the dual graph is a very important property that
is only true in 2 dimensions. In general, there are many things known for 2
dimensions via the dual, which do not generalize to higher dimensions.
The other topic we are going to discuss is random walks in graphs. In IB
Markov chains, and maybe IA Probability, we considered random walks on the
integer lattice
Z
d
. Here we shall consider random walks on any graph. We
shall mostly think about finite graphs, but we will also talk about how certain
results can be extended to infinite graphs. It turns out a rather useful way of
thinking about this is to think of the graph as representing an electrical network.
Then many concepts familiar from high school physics translate to interesting
properties about the graph, and importing well-known (and elementary) results
from electrical networks helps us understand graphs better.
1 Percolation
1.1 The critical probability
There are two models of percolation bond percolation and site percolation. In
this course, we will focus on bond percolation, but we will look at site percolation
in the example sheets.
The very basic set up of percolation theory involves picking a graph
G
=
(
V, E
), where
V
is the set of vertices and
E
is the set of edges. We also pick a
percolation probability
p
[0
,
1]. For each edge
e E
, we keep it with probability
p
and throw it with probability 1
p
. In the first case, we say the edge is open,
and in the latter, we say it is closed.
More precisely, we define the probability space to be Ω =
{
0
,
1
}
E
, where 0
denotes a closed edge and 1 denotes an open one (in the case of site percolation,
we have Ω =
{
0
,
1
}
V
). We endow with the
σ
-algebra generated by cylinder
sets
{ω Ω : ω(e) = x
e
for all e A},
where
A
is a finite set and
x
e
{
0
,
1
}
for all
e
. In other words, this is the
product
σ
-algebra. As probability measure, we take the product measure
P
p
, i.e.
every edge is 1 with probability
p
and 0 with probability 1
p
. We will write
η
p
{0, 1}
E
for the state of the system.
Now what can we say about the graph resulting from this process? One
question we may ask is whether we can connect two points in the graphs via the
edges that remain. To further the discussion, we introduce some notation.
Notation. We write x y if there is an open path of edges from x to y.
Notation. We write C(x) = {y V : y x}, the cluster of x.
Notation. We write x if |C(x)| = .
From now on, we shall take
G
=
L
d
= (
Z
d
, E
(
Z
d
)), the
d
-dimensional integer
lattice. Then by translation invariance,
|C
(
x
)
|
has the same distribution as
|C
(0)
|
for all x. We now introduce a key piece of notation:
Definition (θ(p)). We define θ(p) = P
p
(|C(0)| = ).
Most of the questions we ask surround this
θ
(
p
). We first make the most
elementary observations:
Example. θ(0) = 0 and θ(1) = 1.
A natural question to ask is then if we can find
p
(0
,
1) such that
θ
(
p
)
>
0.
But even before answering that question, we can ask a more elementary one
is θ an increasing function of p?
Intuitively, it must be. And we can prove it. The proof strategy is known as
coupling. We have already seen coupling in IB Markov Chains, where we used it
to prove the convergence to the invariant distribution under suitable conditions.
Here we are going to couple all percolation processes for different values of P .
Lemma. θ is an increasing function of p.
Proof.
We let (
U
(
e
))
eE(Z
d
)
be iid
U
[0
,
1] random variables. For each
p
[0
,
1],
we define
η
p
(e) =
(
1 U(e) p
0 otherwise
Then
P
(
η
p
(
e
) = 1) =
P
(
U
(
e
)
< p
) =
p
. Since the
U
(
e
) are independent, so are
η
p
. Thus η
p
has the law of bond percolation with probability p.
Moreover, if p q, then η
p
(e) η
q
(e). So the result follows.
Note that this is not only useful as a theoretical tool. If we want to simulate
percolation with different probabilities
p
, we can simply generate a set of
U
[0
,
1]
variables, and use it to produce a percolation for all p.
If we wish, we can provide an abstract definition of what coupling is, but the
detailed definition is not of much practical use:
Definition
(Coupling)
.
Let
µ
and
ν
be two probability measures on (potentially)
different probability spaces. A coupling is a pair of random variables (
X, Y
)
defined on the same probability space such that the marginal distribution of
X
is µ and the marginal distribution of Y is ν.
With the lemma, we can make the definition
Definition (Critical probability). We define p
c
(d) = sup{p [0, 1] : θ(p) = 0}.
Recall we initially asked whether
θ
(
p
) can be non-zero for
p
(0
,
1). We
can now rephrase and strengthen this question by asking for the value of p
c
(d).
There are a lot more questions we can ask about p
c
and θ(p).
For example, we know that
θ
(
p
) is a
C
function on (
p
c
,
1]. However, we
do not know if
θ
is continuous at
p
c
in
d
= 3. We will see soon that
p
c
=
1
2
in
d = 2, but the exact value of p
c
is not known in higher dimensions.
Let’s start actually proving things about p
c
. We previously noted that
Proposition. p
c
(1) = 1.
The first actually interesting theorem is the following:
Theorem. For all d 2, we have p
c
(d) (0, 1).
We shall break this up into two natural parts:
Lemma. For d 2, p
c
(d) > 0.
Proof.
Write Σ
n
for the number of open self-avoiding paths of length
n
starting
at 0. We then note that
P
p
(|C(0)| = ) = P
p
(n 1 : Σ
n
1) = lim
n→∞
P
p
n
1) lim
n→∞
E
p
n
].
We can now compute
E
p
n
]. The point is that expectation is linear, which
makes this much easier to compute. We let
σ
n
be the number of self-avoiding
paths of length n from 0. Then we simply have
E
p
n
] = σ
n
p
n
.
We can bound
σ
n
by 2
d ·
(2
d
1)
n1
, since we have 2
d
choices of the first step,
and at most 2d 1 choices in each subsequent step. So we have
E
p
n
] 2d(2d 1)
n1
p
n
=
2d
2d 1
(p(2d 1))
n
.
So if p(2d 1) < 1, then θ(p) = 0. So we know that
p
c
(d)
1
2d 1
.
Before we move on to the other half of the theorem, we talk a bit more about
self-avoiding paths.
Definition
(
σ
n
)
.
We write
σ
n
for the number of self-avoiding paths of length
n
starting from 0.
In the proof, we used the rather crude bound
σ
n
2d · (2d 1)
n1
.
More generally, we can make the following bound:
Lemma. We have σ
n+m
σ
n
σ
m
.
Proof.
A self-avoiding path of length
n
+
m
can be written as a concatenation of
self-avoiding paths of length
n
starting from 0 and another one of length
m
.
Taking the logarithm, we know that
log σ
n
is a subadditive sequence. It turns
out this property alone is already quite useful. It is an exercise in analysis to
prove the following lemma:
Lemma
(Fekete’s lemma)
.
If (
a
n
) is a subadditive sequence of real numbers,
then
lim
n→∞
a
n
n
= inf
n
a
k
k
: k 1
o
[−∞, ).
In particular, the limit exists.
This allows us to define
Definition (λ and κ). We define
λ = lim
n→∞
log σ
n
n
, κ = e
λ
.
κ is known as the connective constant.
Then by definition, we have
σ
n
= e
(1+o(1))
= κ
n+o(n)
.
as n .
Thus, asymptotically, the growth rate of
σ
n
is determined by
κ
. It is natural
to then seek for the value of
κ
, but unfortunately we don’t know the value of
κ
for the Euclidean lattice. A bit more on the positive side, the value for the
hexagonal lattice has been found recently:
Theorem (Duminil-Copin, Smirnov, 2010). The hexagonal lattice has
κ
hex
=
q
2 +
2.
We might want to be a bit more precise about how
σ
n
grows. For
d
5, we
have the following theorem:
Theorem
(Hara and Slade, 1991)
.
For
d
5, there exists a constant
A
such
that
σ
n
=
n
(1 + O(n
ε
))
for any ε <
1
2
.
We don’t really know what happens when
d <
5, but we have the following
conjecture:
Conjecture.
σ
n
n
11/32
κ
n
d = 2
n
γ
κ
n
d = 3
(log n)
1/4
κ
n
d = 4
One can also instead try to bound
σ
n
from above. We have the following
classic theorem:
Theorem (Hammersley and Welsh, 1962). For all d 2, we have
σ
n
Cκ
n
exp(c
0
n)
for some constants C and c
0
.
In fact, a better bound was recently found:
Theorem (Hutchcroft, 2017). For d 2, we have
σ
n
Cκ
n
exp(o(
n)).
This will be proved in the example sheet.
What would be a good way to understand self-avoiding walks? Fixing an
n
,
there are only finitely many self-avoiding walks of length
n
. So we can sample
such a self-avoiding walk uniformly at random. In general, we would expect
the total displacement of the walk to be
n
. Thus, what we can try to do
is to take
n
while simultaneously shrinking space by a factor of
1
n
. We
would then hope that the result converges toward some Brownian motion-like
trajectory. If we can characterize the precise behaviour of this scaling limit,
then we might be able to say something concrete about
κ
and the asymptotic
behaviour of σ
n
.
But we don’t really know what the scaling limit is. In the case
d
= 2, it is
conjectured to be
SLE
(
8
3
). Curiously, it was proven by Gwynne and Miller in
2016 that if we instead looked at self-avoiding walks on a random surface, then
the scaling limit is SLE(
8
3
).
That’s enough of a digression. Let’s finish our proof and show that
p
c
(
d
)
<
1.
A first observation is that it suffices to show this for
d
= 2. Indeed, since
Z
d
embeds into
Z
d+1
for all
d
, if we can find an infinite cluster in
Z
d
, then the same
is true for
Z
d+1
. Thus, it is natural to restrict to the case of
d
= 2, where duality
will be prove to be an extremely useful tool.
Definition
(Planar graph)
.
A graph
G
is called planar if it can be embedded
on the plane in such a way that no two edges cross.
Definition
(Dual graph)
.
Let
G
be a planar graph (which we call the primal
graph). We define the dual graph by placing a vertex in each face of
G
, and
connecting 2 vertices if their faces share a boundary edge.
Example. The dual of Z
2
is isomorphic to Z
2
The dual lattice will help us prove a lot of properties for percolation in Z
2
.
Lemma. p
c
(d) < 1 for all d 2.
Proof.
It suffices to show this for
d
= 2. Suppose we perform percolation on
Z
2
.
Then this induces a percolation on the dual lattice by declaring an edge of the
dual is open if it crosses an open edge of Z
2
, and closed otherwise.
Suppose
|C
(0)
| <
in the primal lattice. Then there is a closed circuit in
the dual lattice, given by the “boundary” of
C
(0). Let
D
n
be the number of
closed dual circuits of length
n
that surround 0. Then the union bound plus
Markov’s inequality tells us
P
p
(|C(0)| < ) = P
p
(n D
n
1)
X
n=4
E
p
[D
n
],
using the union bound and Markov’s inequality.
It is a simple exercise to show that
Exercise.
Show that the number of dual circuits of length
n
that contain 0 is
at most n · 4
n
.
From this, it follows that
P
p
(|C(0)| < )
X
n=4
n · 4
n
(1 p)
n
.
Thus, if
p
is sufficiently near 1, then
P
p
(
|C
(0)
| <
) is bounded away from 1.
By definition, if
p < p
c
(
d
), then 0 is almost surely not contained in an infinite
cluster. If
p > p
c
(
d
), then there is a positive probability that 0 is contained
in an infinite cluster. However, it is of course not necessarily the case that
0 is connected to
with probability 1. In fact, there is at least probability
(1
p
)
2d
that 0 is not connected to
, since 0 cannot be connected to
if all
its neighbouring edges are closed. However, it is still possible that there is some
infinite cluster somewhere. It’s just that it does not contain 0.
Proposition. Let A
be the event that there is an infinite cluster.
(i) If θ(p) = 0, then P
p
(A
) = 0.
(ii) If θ(p) > 0, then P
p
(A
) = 1.
Proof.
(i) We have
P
p
(A
) = P
p
(x : |C(x)| = )
X
xZ
d
P
p
(|C(x)| = ) =
X
θ(p) = 0.
(ii)
We need to apply the Kolmogorov 0-1 law . Recall that if
X
1
, X
2
, . . .
are
independent random variables, and
F
n
=
σ
(
X
k
:
k n
),
F
=
T
n0
F
n
,
Then F
is trivial, i.e. for all A F
, P(A) {0, 1}.
So we order the edges of Z
d
as e
1
, e
2
, . . . and denote their states
w(e
1
), w(e
2
), . . . .
These are iid random variables. We certainly have
P
p
(
A
)
θ
(
p
)
>
0.
So if we can show that
A
F
, then we are done. But this is clear,
since changing the states of a finite number of edges does not affect the
occurrence of A
.
The next follow up questions is how many infinite clusters do we expect to
get?
Theorem
(Burton and Keane)
.
If
p > p
c
, then there exists a unique infinite
cluster with probability 1.
This proof is considerably harder than the ones we have previously done. We
might think we can use the Kolmogorov 0-1 law, but we can’t, since changing
a finite number of edges can break up or join together infinite clusters, so the
event that there are
k
infinite clusters for
k >
0 is not in
F
. However, we can
exploit the fact that N is translation invariant.
Exercise.
Let
A
be an event that is translation invariant. Then
P
p
(
A
) = 0 or 1
almost surely.
Proof.
Let
N
be the number of infinite clusters. Then by the lemma, we know
N
is constant almost surely. So there is some
k N {∞}
such that
P
p
(
N
=
k
) = 1.
First of all, we know that
k 6
= 0, since
θ
(
p
)
>
0. We shall first exclude 2
k <
,
and then exclude k = .
Assume that
k <
. We will show that
P
p
(
n
= 1)
>
0, and hence it must
be the case that P
p
(n = 1) = 1.
To bound this probability, we let
B
(
n
) = [
n, n
]
d
Z
d
(which we will
sometimes write as B
n
), and let B(n) be its boundary. We know that
P
p
(all infinite clusters intersect B(n)) 1
as
n
. This is since with probability 1, there are only finitely many clusters
by assumption, and for each of these configurations, all infinite clusters intersect
B(n) for sufficiently large n.
In particular, we can take n large enough such that
P
p
(all infinite clusters intersect B(n))
1
2
.
We can then bound
P
p
(N = 1) P
p
(all infinite clusters intersect B(n)
and all edges in B(n) are open).
Finally, note that the two events in there are independent, since they involve
different edges. But the probability that all edges in
B
(
n
) are open is just
p
E(B(n))
. So
P
p
(N = 1)
1
2
p
E(B(n))
> 0.
So we are done.
We now have to show that
k 6
=
. This involves the notion of a trifurcation.
The idea is that we will show that if k = , then the probability that a vertex
is a trifurcation is positive. This implies the expected number of trifurcations
is
n
d
. We will then show deterministically that the number of trifurcations
inside
B
(
n
) must be
|B
(
n
)
|
, and so there are
O
(
n
d1
) trifurcations, which is
a contradiction.
We say a vertex x is a trifurcation if the following three conditions hold:
(i) x is in an infinite open cluster C
;
(ii) There exist exactly three open edges adjacent to x;
(iii) C
\ {x} contains exactly three infinite clusters and no finite ones.
This is clearly a translation invariant notion. So
P
p
(0 is a trifurcation) = P
p
(x is a trifurcation)
for all x Z
d
.
Claim. P
p
(0 is a trifurcation) > 0.
We need to use something slightly different from
B
(
n
). We define
S
(
n
) =
{x Z
d
: kxk
1
n}.
The crucial property of this is that for any
x
1
, x
2
, x
3
S
(
n
), there exist three
disjoint self-avoiding paths joining
x
i
to 0 (exercise!). For each triple
x
1
, x
2
, x
3
,
we arbitrarily pick a set of three such paths, and define the event
J(x
1
, x
2
, x
3
) = {all edges on these 3 paths are open
and everything else inside S(n) is closed}.
Next, for every possible infinite cluster in
Z
d
\ S
(
n
) that intersects
S
(
n
) at at
least one point, we pick a designated point of intersection arbitrarily.
Then we can bound
P
p
(0 is a trifurcation) P
p
(∃C
1
, C
2
, C
3
Z
d
\ S(n)
infinite clusters which intersect S(n) at x
1
, x
2
, x
3
, and J(x
1
, x
2
, x
3
)).
Rewrite the right-hand probability as
P
p
(J(x
1
, x
2
, x
3
) | ∃C
1
, C
2
, C
3
Z intersecting S(n))
× P
p
(∃C
1
, C
2
, C
3
Z
d
\ S(n))
We can bound the first term by
min(p, 1 p)
E(S(n))
.
To bound the second probability, we have already assumed that
P
p
(
N
=
) = 1.
So
P
p
(
∃C
1
, C
2
, C
3
Z
d
\ S
(
n
)
intersecting S
(
n
))
1 as
n
. We can
then take
n
large enough such that the probability is
1
2
. So we have shown
that c P
p
(0 is a trifurcation) > 0.
Using the linearity of expectation, it follows that
E
p
[number of trifurcations inside B(n))] c|B(n)| n
d
.
On the other hand, we can bound the number of trifurcations in
B
(
n
) by
|B
(
n
)
|
.
To see this, suppose
x
1
is a trifurcation in
B
(
n
). By definition, there exists
3 open paths to
B
(
n
). Fix three such paths. Let
x
2
be another trifurcation.
It also has 3 open paths to the
B
(
n
), and its paths to the boundary could
intersect those of
x
1
. However, they cannot create a cycle, by definition of a
trifurcation. For simplicity, we add the rule that when we produce the paths for
x
2
, once we intersect the path of x
1
, we continue following the path of x
1
.
Exploring all trifurcations this way, we obtain a forest inside B(n), and the
boundary points will be the leaves of the forest. Now the trifurcations have
degree 3 in this forest. The rest is just combinatorics.
Exercise.
For any tree, the number of degree 3 vertices is always less than the
number of leaves.
1.2 Correlation inequalities
In this section, we are going to prove some useful inequalities and equalities,
and use them to prove some interesting results about
θ
(
p
) and the decay of
P
p
(0 B
n
).
To motivate our first inequality, suppose we have 4 points
x, y, u, v
, and we
want to ask for the conditional probability
P
p
(x y | u v).
Intuitively, we expect this to be greater than
P
p
(
x y
), since
u v
tells us
there are some open edges around, which is potentially helpful. The key property
underlying this intuition is that having more edges is beneficial to both of the
events. To quantify this, we need the notion of an increasing random variable.
Again, let
G
= (
V, E
) be a graph and Ω =
{
0
,
1
}
E
. We shall assume that
E
is countable.
Notation (). Given ω, ω
0
Ω, we write ω ω
0
if ω(e) ω
0
(e) for all e E.
This defines a partial order on Ω.
Definition
(Increasing random variable)
.
A random variable
X
is increasing if
X(ω) X(ω
0
) whenever ω ω
0
, and is decreasing if X is increasing.
Definition
(Increasing event)
.
An event
A
is increasing (resp. decreasing) if
the indicator 1(A) is increasing (resp. decreasing)
Example. {|C(0)| = ∞} is an increasing event.
An immediate consequence of the definition is that
Theorem. If N is an increasing random variable and p
1
p
2
, then
E
p
1
[N] E
p
2
[N],
and if an event A is increasing, then
P
p
1
(A) P
p
2
(A).
Proof. Immediate from coupling.
What we want to prove is the following result, which will be extremely useful.
Theorem
(Fortuin–Kasteleyn–Ginibre (FKG) inequality)
.
Let
X
and
Y
be
increasing random variables with E
p
[X
2
], E
p
[Y
2
] < . Then
E
p
[XY ] E
p
[X]E
p
[Y ].
In particular, if A and B are increasing events, then
P
p
(A B) P
p
(A)P
p
(B).
Equivalently,
P
p
(A | B) P
p
(A).
Proof.
The plan is to first prove this in the case where
X
and
Y
depend on a
finite number of edges, and we do this by induction. Afterwards, the desired
result can be obtained by an application of the martingale convergence theorem.
In fact, the “real work” happens when we prove it for
X
and
Y
depending on
a single edge. Everything else follows from messing around with conditional
probabilities.
If
X
and
Y
depend only on a single edge
e
1
, then for any
ω
1
, ω
2
{
0
,
1
}
, we
claim that
(X(ω
1
) X(ω
2
))(Y (ω
1
) Y (ω
2
)) 0.
Indeed,
X
and
Y
are both increasing. So if
ω
1
> ω
2
, then both terms are
positive; if
ω < ω
2
, then both terms are negative, and there is nothing to do if
they are equal.
In particular, we have
X
ω
1
2
∈{0,1}
(X(ω
1
) X(ω
2
))(Y (ω
1
) Y (ω
2
))P
p
(ω(e
1
) = ω
1
)P
p
(ω(e
1
) = ω
2
) 0.
Expanding this, we find that the LHS is 2(
E
p
[
XY
]
E
p
[
X
]
E
p
[
Y
]), and so we
are done.
Now suppose the claim holds for
X
and
Y
that depend on
n < k
edges. We
shall prove the result when they depend on k edges e
1
, . . . , e
k
. We have
E
p
[XY ] = E
p
[E
p
[XY | ω(e
1
), . . . , ω(e
k1
)]].
Now after conditioning on
ω
(
e
1
)
, . . . , ω
(
e
k1
), the random variables
X
and
Y
become increasing random variables of ω(e
k
). Applying the first step, we get
E
p
[XY | ω(e
1
), . . . , ω(e
k1
)]
E
p
[X | ω(e
1
), . . . , ω(e
k1
)]E
p
[Y | ω(e
1
), . . . , ω(e
k1
)]. ()
But
E
p
[
X | ω
(
e
1
)
, . . . , ω
(
e
k1
)] is a random variable depending on the edges
e
1
, . . . , e
k1
, and moreover it is increasing. So the induction hypothesis tells us
E
p
E
p
[X | ω(e
1
), . . . , ω(e
k1
)]E
p
[Y | ω(e
1
), . . . , ω(e
k1
)]
E
p
E
p
[X | ω(e
1
), . . . , ω(e
k1
)]
E
p
E
p
[Y | ω(e
1
), . . . , ω(e
k1
)]
Combining this with (the expectation of) () then gives the desired result.
Finally, suppose
X
and
Y
depend on the states of a countable set of edges
e
1
, e
2
, . . .. Let’s define
X
n
= E
p
[X | ω(e
1
), . . . , ω(e
n
)]
Y
n
= E
p
[Y | ω(e
1
), . . . , ω(e
n
)]
Then
X
n
and
Y
n
are martingales, and depend only on the states of only finitely
many edges. So we know that
E
p
[X
n
Y
n
] E
p
[X
n
]E
p
[Y
n
] = E
p
[X]E
p
[Y ].
By the
L
2
-martingale convergence theorem,
X
n
X
,
Y
n
Y
in
L
2
and almost
surely. So taking the limit n , we get
E
p
[XY ] E
p
[X]E
p
[Y ].
What we want to consider next is the notion of disjoint occurrence. For
example, we want to able to ask the probability that there exists two disjoint
paths connecting a pair of points.
To formulate this disjointness, suppose we have an event
A
, and let
ω A
.
To ask whether this occurrence of
A
depends only on some set
S E
of edges,
we can look at the set
[ω]
S
= {ω
0
Ω : ω
0
(e) = ω(e) for all e S}.
If [
ω
]
S
A
, then we can rightfully say this occurrence of
A
depends only on
the edges in
S
. Note that this depends explicitly and importantly on
ω
, i.e. the
“reason”
A
happened. For example, if
A
=
{x y}
, and
ω A
, then we can
take
S
to be the set of all edges in a chosen path from
x
to
y
in the configuration
of ω. This choice will be different for different values of ω.
Using this, we can define what it means for two events to occur disjointly.
Definition
(Disjoint occurrence)
.
Let
F
be a set and =
{
0
,
1
}
F
. If
A
and
B
are events, then the event that A and B occurs disjointly is
A B = {ω Ω : S F s.t. [ω]
S
A and [ω]
F \S
B}.
Theorem
(BK inequality)
.
Let
F
be a finite set and Ω =
{
0
,
1
}
F
. Let
A
and
B be increasing events. Then
P
p
(A B) P
p
(A)P
p
(B).
This says if
A
and
B
are both events that “needs” edges to occur, then requir-
ing that they occur disjointly is more difficult than them occurring individually.
The proof is completely magical. There exist saner proofs of the inequality,
but they are rather longer.
Proof (Bollob´as and Leader).
We prove by induction on the size
n
of the set
F
.
For n = 0, it is trivial.
Suppose it holds for
n
1. We want to show it holds for
n
. For
D {
0
,
1
}
F
and i = 0, 1, set
D
i
= {(ω
1
, . . . , ω
n1
) : (ω
1
, . . . , ω
n1
, i) D}.
Let A, B {0, 1}
F
, and C = A B. We check that
C
0
= A
0
B
0
, C
1
= (A
0
B
1
) (A
1
B
0
).
Since
A
and
B
are increasing,
A
0
A
1
and
B
0
B
1
, and
A
i
and
B
i
are also
increasing events. So
C
0
(A
0
B
1
) (A
1
B
0
)
C
1
A
1
B
1
.
By the induction hypothesis, we have
P
p
(C
0
) = P
p
(A
0
B
0
) P
p
(A
0
)P
p
(B
0
)
P
p
(C
1
) P
p
(A
1
B
1
) P
p
(A
1
)P
p
(B
1
)
P
p
(C
0
) + P
p
(C
1
) P
p
((A
0
B
1
) (A
1
B
0
)) + P
p
((A B
1
) (A
1
B
0
))
= P
p
(A
0
B
1
) + P
p
(A
1
B
0
)
P
p
(A
0
)P
p
(B
1
) + P
p
(A
1
)P
p
(B
0
).
Now note that for any D, we have
P
p
(D) = pP
p
(D
1
) + (1 p)P
p
(D
0
).
By some black magic, we multipy the first inequality by (1
p
)
2
, the second by
p
2
and the third by p(1 p). This gives
pP
p
(C
1
) + (1 p)P
p
(C
0
) (pP
p
(A
1
) + (1 p)P
p
(A
0
))(pP
p
(B
1
) + (1 p)P
p
(B
0
)).
Expand and we are done.
It turns out the increasing hypothesis is not necessary:
Theorem
(Reimer)
.
For all events
A, B
depending on a finite set, we have
P
p
(A B) P
p
(A)P
p
(B).
But the proof is much harder, and in all the case where we want to apply
this, the variables are increasing.
As an application of the BK inequality, we first prove a preliminary result
about the decay of
P
p
(0
B
(
n
)). To prove our result, we will need a stronger
condition than p < p
c
. Recall that we defined
θ(p) = P
p
(|C(0)| = ).
We also define
χ(p) = E
p
[|C(0)|].
If χ(p) is finite, then of course θ(p) = 0. However, the converse need not hold.
Theorem.
If
χ
(
p
)
<
, then there exists a positive constant
c
such that for all
n 1,
P
p
(0 B(n)) e
cn
.
Later, we will show that in fact this holds under the assumption that
p < p
c
.
However, that requires a bit more technology, which we will develop after this
proof.
The idea of the proof is that if we want a path from, say, 0 to
B
(2
n
), then
the path must hit a point on
B
(
n
). So there is a path from 0 to a point on
B
(
n
), and a path from that point to
B
(2
n
). Moreover, these two paths are
disjoint, which allows us to apply the BK inequality.
Proof. Let
X
n
=
X
xB(n)
1(0 x).
Now consider
X
n=0
E[X
n
] =
X
n
X
xB(n)
P
p
(0 x) =
X
xZ
d
P
p
(0 x) = χ(p).
Since
χ
(
p
) is finite, we in particular have
E
p
[
X
n
]
0 as
n
. Take
m
large
enough such that E
p
[X
m
] < δ < 1.
Now we have
P
p
(0 B(m + k)) = P
p
(x B(m) : 0 x and x B(m + k) disjointly)
X
xB(m)
P
p
(0 x)P
p
(x B(m + k)) (BK)
X
xB(m)
P
p
(0 x)P
p
(0 B(k)) (trans. inv.)
P
p
(0 B(k))E
p
[X
m
].
So for any
n > m
, write
n
=
qm
+
r
, where
r
[0
, m
1]. Then iterating the
above result, we have
P
p
(0 B(n)) P
p
(0 B(mq)) δ
q
δ
1+
n
m
e
cn
.
To replace the condition with the weaker condition
p < p
c
, we need to
understand how
θ
(
p
) changes with
p
. We know
θ
(
p
) is an increasing function in
p
. It would be great if it were differentiable, and even better if we could have an
explicit formula for
dθ
dp
.
To do so, suppose we again do coupling, and increase
p
by a really tiny bit.
Then perhaps we would expect that exactly one of the edges switches from being
closed to open. Thus, we want to know if the state of this edge is pivotal to the
event |C(0)| = , and this should determine the rate of change of θ.
Definition
(Pivotal edge)
.
Let
A
be an event and
ω
a percolation configuration.
The edge e is pivotal for (A, ω) if
1(ω A) 6= 1(ω
0
A),
where ω
0
is defined by
ω
0
(f) =
(
ω(f) f 6= e
1 ω(f) f = e
.
The event that
e
is pivotal for
A
is defined to be the set of all
ω
such that
e
is
pivotal for (A, ω).
Note that whether or not e is pivotal for (A, ω) is independent of ω(e).
Theorem
(Russo’s formula)
.
Let
A
be an increasing event that depends on the
states of a finite number of edges. Then
d
dp
P
p
(A) = E
p
[N(A)],
where N (A) is the number of pivotal edges for A.
Proof.
Assume that
A
depends the states of
m
edges
e
1
, . . . , e
m
. The idea is to
let each
e
i
be open with probability
p
i
, whree the
{p
i
}
may be distinct. We then
vary the p
i
one by one and see what happens.
Writing ¯p = (p
1
, . . . , p
m
), we define
f(p
1
, . . . , p
m
) = P
¯p
(A),
Now
f
is the sum of the probability of all configurations of
{e
1
, . . . , e m}
for
which
A
happens, and is hence a finite sum of polynomials. So in particular, it
is differentaible.
We now couple all percolation process. Let (
X
(
e
) :
e L
d
) be iid
U
[0
,
1]
random variables. For a vector ¯p = (p(e) : e L
d
), we write
η
¯p
(e) = 1(X(e) p(e)).
Then we have P
¯p
(A) = P(η
¯p
A).
Fix an edge
f
and let
¯p
0
= (
p
0
(
e
)) be such that
p
0
(
e
) =
p
(
e
) for all
e 6
=
f
, and
p
0
(f) = p(f ) + δ for some δ > 0. Then
P
¯p
0
(A) P
¯p
(A) = P(η
¯p
0
A) P(η
¯p
A)
= P(η
¯p
0
A, η
¯p
A) + P(η
¯p
0
A, η
¯p
A) P(η
¯p
A).
But we know
A
is increasing, so
P
(
η
¯p
0
A, η
¯p
A
) =
P
(
η
¯p
A
). So the first
and last terms cancel, and we have
P
¯p
0
(A) P
¯p
(A) = P(η
¯p
0
A, η
¯p
6∈ A).
But we observe that we simply have
P(η
¯p
0
A, η
¯p
6∈ A) = δ · P
¯p
(f is pivotal for A).
Indeed, we by definition of pivotal edges, we have
P(η
¯p
0
A, η
¯p
6∈ A) = P
¯p
(f is pivotal for A, p(f) < X(f) p(f) + δ).
Since the event
{f is pivotal for A}
is independent of the state of the edge
f
,
we obtain
P
¯p
(f is pivotal, p(f) < X(f) p(f) + δ) = P
¯p
(f is pivotal) · δ.
Therefore we have
p(f)
P
¯p
(A) = lim
δ0
P
¯p
0
(A) P
¯p
(A)
δ
= P
¯p
(f is pivotal for A).
The desired result then follows from the chain rule:
d
dp
P
p
(A) =
m
X
i=1
p(e
i
)
P
¯p
(A)
¯p=(p,...,p)
=
m
X
i=1
P
p
(e
i
is pivotal for A)
= E
p
[N(A)].
If
A
depends on an infinite number of edges, then the best we can say is that
lim inf
δ0
P
p+δ
(A) P
p
(A)
δ
E
p
[N(A))].
To see this, again set B(n) = [n, n]
d
Z
d
. Define ¯p
n
by
¯p
n
(e) =
(
p e 6∈ B(n)
p + δ e B(n)
.
Then since a is increasing, we know
P
p+δ
(A) P
p
(A)
δ
P
¯p
n
(A) P
p
(A)
δ
.
We can then apply the previous claim, take successive differences, and take
n .
Corollary.
Let
A
be an increasing event that depends on
m
edges. Let
p q
[0, 1]. Then P
q
(A) P
p
(A)
q
p
m
.
Proof.
We know that
{f is pivotal for A}
is independent of the state of
f
, and
so
P
p
(ω(f) = 1, f is pivotal for A) = pP
p
(f is pivotal for A).
But since
A
is increasing, if
ω
(
f
) = 1 and
f
is pivotal for
A
, then
A
occurs.
Conversely, if f is pivotal and A occurs, then ω(f) = 1.
Thus, by Russo’s formula, we have
d
dp
P
p
(A) = E
p
[N(A)]
=
X
e
P
p
(e is pivotal for A)
=
X
e
1
p
P
p
(ω(e) = 1, e is pivotal for A)
=
X
e
1
p
P
p
(e is pivotal | A)P
p
(A)
= P
p
(A)
1
p
E
p
[N(A) | A].
So we have
d
dp
P
p
(A)
P
p
(A)
=
1
p
E
p
[N(A) | A].
Integrating, we find that
log
P
q
(A)
P
p
(A)
=
Z
q
p
1
u
E
u
[N(A) | A] du.
Bounding E
u
[N(A) | A] m, we obtain the desired bound.
With Russo’s formula, we can now prove the desired theorem.
Theorem. Let d 2 and B
n
= [n, n]
d
Z
d
.
(i)
If
p < p
c
, then there exists a positive constant
c
for all
n
1,
P
p
(0
B
n
) e
cn
.
(ii) If p > p
c
, then
θ(p) = P
p
(0 )
p p
c
p(1 p
c
)
.
This was first proved by Aizenman and Barsky who looked at the more
general framework of long-range percolation. Menshikov gave an alternative
proof by analyzing the geometry of pivotal edges. The proof we will see is by
Duminil-Copin and Tassion. Recall that we defined
χ(p) = E
p
[|C(0)|].
We saw that if
χ
(
p
)
<
, then
P
p
(0
B
n
)
e
cn
. We now see that
χ
(
p
)
<
iff p < p
c
.
The strategy of the proof is to define a new percolation probability
˜p
c
, whose
definition makes it easier to prove the theorem. Once we have established the
two claims, we see that (i) forces ˜p
c
p
c
, and (ii) forces ˜p
c
p
c
. So they must
be equal.
Proof (Duminil-Copin and Tassion). If S V is finite, we write
S = {(x, y) E : x S, y 6∈ S}.
We write
x
S
y
if there exists an open path of edges from
x
to
y
all of whose
end points lie in S.
Now suppose that 0 S. We define
ϕ
p
(S) = p
X
(x,y)∂S
P
p
(0
S
x).
Define
˜p
c
= sup{p [0, 1] : exists a finite set S with 0 S and ϕ
p
(S) < 1}.
Claim. It suffices to prove (i) and (ii) with p
c
replaced by ˜p
c
.
Indeed, from (i), if
p < ˜p
c
, then
P
p
(0
B
n
)
e
cn
. So taking the limit
n
, we see
θ
(
p
) = 0. So
˜p
c
p
c
. From (ii), if
p > ˜p
c
, then
θ
(
p
)
>
0. So
p
c
˜p
c
. So p
c
= ˜p
c
.
We now prove (i) and (ii):
(i)
Let
p < ˜p
c
. Then there exists a finite set
S
containing 0 with
ϕ
p
(
S
)
<
1.
Since
S
is finite, we can pick
L
large enough so that
S B
L1
. We will
prove that P
p
(0 B
kL
) (ϕ
p
(S))
k1
for k 1.
Define C = {x S : 0
S
x}. Since S B
L1
, we know S B
kL
= .
Now if we have an open path from 0 to
B
kL
, we let
x
be the last element
on the path that lies in
C
. We can then replace the path up to
x
by a path
that lies entirely in
S
, by assumption. This is then a path that lies in
C
up
to
x
, then takes an edge on
S
, and then lies entirely outside of
C
c
. Thus,
P
p
(0 B
kL
)
X
AS
0A
X
(x,y)∂A
P
p
(0
A
x, (x, y) open, C = A, y
A
C
B
kL
).
Now observe that the events
{C
=
A,
0
S
x}
,
{
(
x, y
)
is open}
and
{y
A
c
B
kL
} are independent. So we obtain
P
p
(0 B
kL
)
X
AS,0A
X
(x,y)∂S
p P
p
(0
S
x, C = A) P
p
(y
A
c
B
kL
).
Since we know that y B
L
, we can bound
P
p
(y
A
c
B
kL
) P
p
(0 B
(k1)L
).
So we have
P
p
(0 B
kL
) p P
p
(0 B
(k1)L
)
X
AS,0A
X
(x,y)∂S
P
p
(0
S
x, C = A)
= P
p
(0 B
(k1)L
) p
X
(x,y)∂S
P
p
(0
S
x)
= P
p
(0 B
(k1)L
)ϕ
p
(S).
Iterating, we obtain the deseired result.
(ii) We want to use Russo’s formula. We claim that it suffices to prove that
d
dp
P
p
(0 B
n
)
1
p(1 p)
inf
SB
n
,0S
ϕ
p
(S)(1 P
p
(0 B
n
)).
Indeed, if
p > ˜p
c
, we integrate from
˜p
c
to
p
, use in this range
ϕ
p
(
S
)
1,
and then take the limit as n .
The event
{
0
B
n
}
is increasing and only dependson a finite number of
edges. So we can apply Russo’s formula
d
dp
P
p
(0 B
n
) =
X
eB
n
P
p
(e is pivotal for {0 B
n
})
Since being pivotal and being open/closed are independent, we can write
this as
=
X
eB
n
1
1 p
P
p
(e is pivotal for {0 B
n
}, e is closed)
=
X
eB
n
1
1 p
P
p
(e is pivotal for {0 B
n
}, 0 6↔ B
n
)
Define S = {x B
n
: x 6↔ B
n
}. Then {0 6↔ B
n
} implies 0 S. So
d
dp
P
p
(0 B
n
) =
1
1 p
X
eB
n
X
AB
n
,0A
P
p
(e is pivotal, S = A)
Given that
S
=
A
, an edge
e
= (
x, y
) is pivotal iff
e A
and 0
A
x
. So
we know
d
dp
P
p
(0 B
n
) =
1
1 p
X
AB
n
,0A
X
(x,y)∂A
P
p
(0
A
x, S = A).
Observe that
{
0
A
x}
and
{S
=
A}
are independent, since to determine if
S
=
A
, we only look at the edges on the boundary of
A
. So the above is
equal to
1
1 p
X
AB
n
,0A
X
(x,y)∂A
P
p
(0
A
x)P
p
(S = A)
=
1
p(1 p)
X
AB
n
,0A
ϕ
p
(A)P
p
(S = A)
1
p(1 p)
inf
SB
n
,0S
ϕ
P
(S)P
p
(0 6↔ B
n
),
as desired.
We might ask if
P
p
(0
B
n
)
e
cn
is the best convergence rate when
p < p
c
, but we cannot do better, since P
p
(0 B
n
) p
n
.
Also, if p < p
c
, then we can easily bound
P
p
(|C(0)| n) P
p
(0 B
n
1/d
) exp(cn
1/d
).
However, this is not a good bound. In fact,
n
1/d
can be replaced by
n
, but we
will not prove it here. This tells us the largest cluster in
B
n
will have size of
order log n with high probability.
1.3 Two dimensions
We now focus on 2 dimensions. As discussed previously, we can exploit duality to
prove a lot of things specific to two dimensions. In particular, we will show that,
at
p
=
p
c
=
1
2
, certain probabilities such as
P
1
2
(0
B
(
n
)) exhibit a power law
decay. This is in contrast to the exponential decay for subcritical percolation
(and being bounded away from zero for supercritical percolation).
First we establish that p
c
is actually
1
2
.
Theorem. In Z
2
, we have θ
1
2
= 0 and p
c
=
1
2
.
It is conjectured that for all
d
2, we have
θ
(
p
c
(
d
)) = 0. It is known to be
true only in d = 2 and d 11.
This was proved first by Harris, Kesten, Russo, Seymour, Welsh in several
iterations.
Proof. First we prove that θ
1
2
= 0. This will imply that p
c
1
2
.
Suppose not, and θ
1
2
> 0. Recall that B(n) = [n, n]
2
, and we define
C(n) = [(n 1), (n 1)]
2
+
1
2
,
1
2
in the dual lattice. The appearance of the
1 is just a minor technical inconve-
nience. For the same n, our B(n) and C(n) look like
B(n)C(n)B(n)
We claim that for large
n
, there is a positive probability that there are open
paths from the left and right edges of
B
(
n
) to
, and also there are closed
paths from the top and bottom edges of
C
(
n
) to
. But we know that with
probability 1, there is a unique infinite cluster in both the primal lattice and
the dual lattice. To connect up the two infinite open paths starting from the
left and right edges of
B
(
n
), there must be an open left-right crossing of
B
(
n
).
To connect up the two infinite closed paths starting from the top and bottom
of
C
(
n
), there must be a closed top-bottom crossing. But these cannot both
happen, since this would require an open primal edge crossing a closed dual edge,
which is impossible.
To make this an actual proof, we need to show that these events do happen
with positive probability. We shall always take
p
=
1
2
, and will not keep repeating
it.
First note that since there is, in particular, an infinite cluster with probability
1, we have
P(B(n) ) 1.
So we can pick n large enough such that
P(B(n) ), P(C(n) ) 1
1
8
4
.
Let
A
`
/
A
r
/
A
t
/
A
b
be the events that the left/right/top/bottom side of
B
(
n
) is
connected to
via an open path of edges. Similarly, let
D
`
be the event that
the left of C(n) is connected to via a closed path, and same for D
r
, D
r
, D
b
.
Of course, by symmetry, for
i, j {`, r, t, b}
, we have
P
(
A
i
) =
P
(
A
j
). Using
FKG, we can bound
P(S
n
6↔ ) = P(A
c
`
A
c
r
A
c
t
A
c
b
) (P(A
c
`
))
4
= (1 P(A
`
))
4
.
Thus, by assumption on n, we have
(1 P(A
`
))
4
1
8
4
,
hence
P(A
`
)
7
8
.
Of course, the same is true for other A
i
and D
j
.
Then if G = A
`
A
r
D
t
D
b
, which is the desired event, then we have
P(G
c
) P(A
c
`
) + P(A
c
r
) + P(D
c
t
) + P(D
c
b
)
1
2
.
So it follows that
P(G)
1
2
,
which, as argued, leads to a contradiction.
So we have
p
c
1
2
. It remains to prove that
p
c
1
2
. Suppose for contradiction
that
p
c
>
1
2
. Then
p
=
1
2
is in the subcritical regime, and we expect exponential
decay. Thus, (again with
p
=
1
2
fixed) there exists a
c >
0 such that for all
n 1,
P(0 B(n)) e
cn
.
Consider
C
n
= [0
, n
+ 1]
×
[0
, n
], and define
A
n
to be the event that there exists
a left-right crossing of C
n
by open edges.
Again consider the dual box D
n
= [0, n] × [1, n] +
1
2
,
1
2
.
C
n
D
n
Define
B
n
to be the event that there is a top-bottom crossing of
D
n
by closed
edges of the dual.
As before, it cannot be the case that
A
n
and
B
n
both occur. In fact,
A
n
and
B
n
partition the whole space, since if
A
n
does not hold, then every path from
left to right of C
n
is blocked by a closed path of the dual. Thus, we know
P(A
n
) + P(B
n
) = 1.
But also by symmetry, we have P(A
n
) = P(B
n
). So
P(A
n
) =
1
2
.
On the other hand, for any point on the left edge, the probability of it reaching
the right edge decays exponentially with n. Thus,
P(A
n
) n(n + 1)P(0 B
n
) (n + 1)e
cn
which is a contradiction. So we are done.
So we now know that p
c
=
1
2
. We now want to prove that
P
1
2
(0 B(n)) An
α
for some
A, α
. To prove this, we again consider the dual lattice. Observe that if
we have a closed dual circuit around the origin, then this prevents the existence
of an open path from the origin to a point far far away. What we are going to
do is to construct “zones” in the dual lattice like this:
The idea is to choose these zones in a way such that the probability that each
zone contains a closed circuit around the origin is
ζ
for some fixed
ζ
. Thus, if
B
(
n
) contains
m
many of these zones, then the probability that 0 is connected
to
B
(
n
) is bounded above by (1
ζ
)
m
. We would then be done if we can show
that m log n.
The main strategy to bounding these probabilities is to use FKG. For example,
if we want to bound the probability that there is a closed circuit in a region
then we cut it up into the pieces
If we can bound the probability of there being a left-to-right crossing of a
horizontal piece, and hence by symmetry the probability of there being a top-to-
bottom crossing of a vertical piece, then FKG gives us a bound on the probability
of there being a closed circuit.
For convenience of notation, we prove these bounds for open paths in the
primal lattice. We make the following definitions:
B(k`, `) = [`, (2k 1)`] × [`, `]
B(`) = B(`, `) = [`, `]
2
A(`) = B(3`) \ B(`)
LR(k`, `) = {there exists left-right crossing of B(k`, `) of open edges}
LR(`) = LR(`, `)
O(`) = {there exists open circuit in A(`) that contains 0 in its interior}.
We first note that we have already proven the following:
Proposition. P
1
2
(LR(`))
1
2
for all `.
Proof.
We have already seen that the probability of there being a left-right
crossing of [0
, n
+ 1]
×
[0
, n
] is at least
1
2
. But if there is a left-right crossing of
[0, n + 1] × [0, n], then there is also a left-right crossing of [0, n] × [0, n]!
For a general
p
, Russo–Symour–Welsh (RSW) lets us bound
P
p
(
O
(
`
)) by
P
p
(LR(`)):
Theorem (Russo–Symour–Welsh (RSW) theorem). If P
p
(LR(`)) = α, then
P
p
(O(`))
α
1
1 α
4
12
.
A large part of the proof is done by the cut-and-paste argument we sketched
above. However, to successfully do cut-and-paste, it turns out we need bounds
on the probability of a left-right crossing on something that is not a square. The
key, non-trivial estimate is the following:
Lemma. If P
p
(LR(`)) = α, then
P
p
LR
3
2
`, `

(1
1 α)
3
.
To prove this, we need a result from the first example sheet:
Lemma
(
n
th root trick)
.
If
A
1
, . . . , A
n
are increasing events all having the same
probability, then
P
p
(A
1
) 1
1 P
p
n
[
i=1
A
i
!!
1/n
.
Proof sketch.
Observe that the proof of FKG works for decreasing events as well,
and then apply FKG to A
c
i
.
We can now prove our initial lemma.
Proof sketch.
Let
A
be the set of left-right crossings of
B
(
`
) = [
`, `
]
2
. Define a
partial order on
A
by
π
1
π
2
if
π
1
is contained in the closed bounded region of
B(`) below π
2
.
Note that given any configuration, if the set of open left-right crossings is
non-empty, then there exists a lowest one. Indeed, since
A
must be finite, it
suffices to show that meets exist in this partial order, which is clear.
For a left-right crossing
π
, let (0
, y
π
) be the last vertex on the vertical axis
where
π
intersects, and let
π
r
be the path of the path that connects (0
, y
π
) to
the right.
O
(0, y
π
)
π
r
Let
A
= {π A : y
π
0}
A
+
= {π A : y
π
0}
Letting
B
(
`
)
0
= [0
,
2
`
]
×
[
`, `
], our goal is to find a left-right crossing of the
form
O
More precisely, we want the following paths:
(i) Some π A
(ii) Some top-bottom crossing of B(`
0
) that crosses π
r
.
(iii)
Some left-right crossing of
B
(
`
0
) that starts at the positive (i.e. non-
negative) y axis.
To understand the probabilities of these events happening, we consider the
“mirror” events and then apply the square root trick.
Let π
0
r
be the reflection of π
r
on {(`, k) : k Z}. For any π A, we define
V
π
= {all edges of π are open}
M
π
= {exists open crossing from top of B(`)
0
to π
r
π
0
r
}
M
π
= {exists open crossing from top of B(`)
0
to π
r
}
M
+
π
= {exists open crossing from top of B(`)
0
to π
0
r
}
L
+
= {exists open path in A
+
}
L
= {exists open path in A
}
L
π
= {π is the lowest open LR crossing of B(`)}
N = {exists open LR crossing of B(`)
0
}
N
+
= {exists open LR crossing in B(`)
0
starting from positive vertical axis}
N
= {exists open LR crossing in B(`)
0
starting from negative vertical axis}
In this notation, our previous observation was that
[
π∈A
(V
π
M
π
)
| {z }
G
N
+
LR
3
2
`, `
So we know that
P
p
LR
3
2
`, `

P
p
(G N
0
) P
p
(G)P
p
(N
+
),
using FKG.
Now by the “square root trick”, we know
P
p
(N
+
) 1
q
1 P
p
(N
+
N
).
Of course, we have P
p
(N
+
N
) = P
p
(LR(`)) = α. So we know that
P
p
(N
+
) 1
1 α.
We now have to understand
G
. To bound its probability, we try to bound it by
the union of some disjoint events. We have
P
p
(G) = P
p
[
π∈A
(V
π
M
π
)
P
p
[
πA
(M
π
L
π
)
=
X
π∈A
P
p
(M
π
| L
π
)P
p
(L
π
).
Claim.
P
p
(M
π
| L
π
) 1
1 α.
Note that if
π
intersects the vertical axis in one point, then
P
p
(
M
π
| L
π
) =
P
p
(
M
π
| V
π
), since
L
π
tells us what happens below
π
, and this does not affect
the occurrence of M
π
.
Since M
π
and V
π
are increasing events, by FKG, we have
P
p
(M
π
| V
π
) P
p
(M
π
) 1
q
1 P
p
(M
π
M
+
π
) = 1
q
1 P
p
(M
π
).
Since P
p
(M
π
) P
p
(LR(`)) = α, the claim follows.
In the case where
π
is more complicated, we will need an extra argument,
which we will not provide.
Finally, we have
P
p
(G)
X
π∈A
P
p
(L
π
)(1
1 α) = (1
1 α)P
p
(L
).
But again by the square root trick,
P
p
(L
) 1
q
1 P
p
(L
+
L
) = 1
1 α,
and we are done.
We now do the easy bit to finish off the theorem:
Lemma.
P
p
(LR(2`, `)) P
p
(LR(`))
P
p
LR
3
2
`, `

2
P
p
(LR(3`, `)) P
p
(LR(`)) (P
p
(LR (2`, `)))
2
P
p
(O(`)) P
p
(LR(3`, `))
4
Proof. To prove the first inequality, consider the box [0, 4`] × [`, `].
3``
We let
LR
1
= {exists left-right crossing of [0, 3`] × [`, `]}
LR
2
= {exists left-right crossing of [`, 4`] × [`, `]}
T B
1
= {exists top-bottom crossing of [`, 3`] × [`, `]}.
Then by FKG, we find that
P
p
(LR(2`, `)) P
p
(LR
1
LR
2
T B
1
) P
p
(LR
1
)P
p
(LR
2
)P
p
(T B
1
).
The others are similar.
Theorem. There exists positive constants α
1
, α
2
, α
3
, α
4
, A
1
, A
2
, A
4
such that
P
1
2
(0 B(n)) A
1
n
α
1
P
1
2
(|C(0)| n) A
2
n
α
2
E(|C(0)|
α
3
)
Moreover, for p > p
c
=
1
2
, we have
θ(p) A
4
p
1
2
α
4
.
It is an exercise on the example sheet to prove that
P
1
2
(0
B
(
n
))
1
2
n
using the BK inequality. So the true decay of
P
1
2
(0
B
(
n
)) is indeed a power
law.
Proof.
(i) We first prove the first inequality. Define dual boxes
B(k)
d
= B(k) +
1
2
,
1
2
.
The dual annuli A(`)
d
are defined by
A(`)
d
= B(3`)
d
\ B(`)
d
.
We let
O
(
`
)
d
be the event that there is a closed dual circuit in
A
(
`
)
d
containing
1
2
,
1
2
. Then RSW tells us there is some ζ (0, 1) such that
P
1
2
(O(`)
d
) ζ,
independent of `. Then observe that
P
1
2
(0 B(3
k
+ 1)) P
p
(O(3
r
)
d
does not occur for all r < k).
Since the annuli (
A
(3
r
)
d
) are disjoint, the events above are independent.
So
P
1
2
(0 B(3
k
+ 1)) (1 ζ)
k
,
and this proves the first inequality.
(ii)
The second inequality follows from the first inequality plus the fact that
|C(0)| n implies 0 B(g(n)), for some function g(n)
n.
(iii)
To show that
E
1
2
[
|C
(0)
|
α
3
]
<
for some
α
3
, we observe that this expecta-
tion is just
X
n
P
1
2
(|C(0)|
α
3
n).
(iv) To prove the last part, note that
θ(p) = P
p
(|C(0)| = ) P
p
(0 B
n
)
for all
n
. By the corollary of Russo’s formula, and since
{
0
B
n
}
only
depends on the edges in B
n
, which are 18n
2
, we get that
P
1
2
(0 B
n
)
1
2p
18n
2
P
p
(0 B
n
).
So
θ(p) (2p)
18n
2
P
1
2
(0 B
n
) A
1
(2p)
18n
2
n
α
1
.
Now take n = b(log 2p)
1/2
c. Then as p &
1
2
, we have
n
1
(2p 1)
1
2
.
Substituting this in, we get
θ(p) C
p
1
2
α
1
/2
.
By similar methods, we can prove that
Theorem.
When
d
= 2 and
p > p
c
, there exists a positive constant
c
such that
P
p
(0 B(n), |C(0)| < ) e
cn
.
It is natural to ask if we have a similar result in higher dimensions. In higher
dimensions, all the techniques in Z
2
don’t work.
Higher dimensions
In d 3, define the slab
S
k
= Z
2
× {0, 1, . . . , k}
d2
.
Then S
k
S
k+1
Z
d
.
In general, for a graph
G
, we let
p
c
(
G
) be the critical probability of bond
percolation in G. Then we have
p
c
(S
k
) p
c
(S
k+1
) p
c
.
So the sequence (p
c
(S
k
)) must converge to a limit. Call this limit
p
slab
c
= lim
k→∞
p
c
(S
k
).
We know that p
slab
c
p
c
.
A lot of the results we can prove for
Z
2
can be proven for
p
slab
c
instead of
p
c
.
So the natural question is how
p
slab
c
is related to
p
c
, and in particular, if they
are equal. This has been an open question for a long time, until Grimmett and
Marstrand proved it.
Theorem
(Grimmett–Marstrand)
.
Let
F
be an infinite-connected subset of
Z
d
with p
c
(F ) < 1. Then for all η > 0, there exists k N such that
p
c
(2kF + B
k
) p
c
+ η.
In particular, for all d 3, p
slab
c
= p
c
.
We shall not prove the theorem, but we shall indicate how the “in particular”
part works.
We take F = Z
2
× {0}
d2
. Then
2kF + B
k
= Z
2
× ([k, k]
d2
Z
d2
),
a translate of S
2k
. So p
c
(S
2k
) = p
c
(2kF + B
k
) p
c
as k .
A consequence of Grimmett-Marstrand is that
Theorem. If d 3 and p > p
c
, then there exists c > 0 such that
P
p
(0 B(n), |C(0)| < ) e
cn
.
1.4 Conformal invariance and SLE in d = 2
Instead of working on
Z
2
, we will now work on the triangular lattice
T
. We
consider site percolation. So every vertex is open (black) with probability
p
and
closed (white) with probability 1 p, independently for different vertices.
Like for Z
2
, we can show that p = p
c
(T) =
1
2
.
Let
D
be an open simply connected domain in
R
2
with
D
a Jordan curve.
Let
a, b, c D
be 3 points labeled in anti-clockwise order. Consider the triangle
T
with vertices
A
= 0,
B
= 1 and
C
=
e
/3
. By the Riemann mapping theorem,
there exists a conformal map ϕ : D T that maps a 7→ A, b 7→ B, c 7→ C.
Moreover, this can be extended to D such that ϕ :
¯
D
¯
T is a homeomor-
phism. If
x
is in the arc
bc
, then it will be mapped to
X
=
ϕ
(
x
) on the edge
BC of T .
Focus on the case
p
=
1
2
. We can again put a triangular lattice inside
D
,
with mesh size
δ
. We let
ac bx
to mean the event that there is an open path
in D joining the arc ac to bx.
Then by RSW (for T), we get that
P
δ
(ac bx) c > 0,
independent of
δ
. We might ask what happens when
δ
0. In particular, does
it converge, and what does it converge to?
Cardy, a physicist studying conformal field theories, conjectured that
lim
δ0
P
δ
(ac bx) = |BX|.
He didn’t really write it this way, but expressed it in terms of hypergeometric
functions. It was Carlesson who expressed it in this form.
In 2001, Smirnov proved this conjecture.
Theorem
(Smirnov, 2001)
.
Suppose (Ω
, a, b, c, d
) and (Ω
0
, a
0
, b
0
, c
0
, d
0
) are con-
formally equivalent. Then
P(ac bd in Ω) = P(a
0
c
0
b
0
d
0
in
0
).
This says percolation at criticality on the triangular lattice is conformally
invariant.
We may also take the dual of the triangular lattice, which is the hexagonal
lattice. Colour a hexagon black if the center is black, and similarly for whites.
Suppose we impose the boundary condition on the upper half plane that the
hexagons are black when x > 0, y = 0, and white when x < 0, y = 0.
Starting at (
x, y
) = (0
,
0), we explore the interface between the black and
white by always keeping a black to our right and white on our right. We can
again take the limit
δ
0. What is this exploration path going to look like in
the limit δ 0? It turns out this is an SLE(6) curve.
To prove this, we use the fact that the exploration curve satisfies the locality
property, namely that if we want to know where we will be in the next step, we
only need to know where we currently are.
2 Random walks
2.1 Random walks in finite graphs
We shall consider weighted random walks on finite graphs. Let
G
= (
V, E
) be a
finite connected graph. We assign a conductances/weights to the edges (
c
(
e
))
eE
.
Here c is a function of the edges, so we require c(xy) = c(yx).
In the weighted random walk on
G
, the probability of going from
x
to
y
(assuming x y) is
P(x, y) =
c(x, y)
c(x)
, c(x) =
X
zx
c(x, z).
If we put c(e) = 1 for all e, then this is just a simple random walk, with
P(x, y) =
(
1
deg x
y x
0 otherwise
.
The weighted random walk is reversible with respect to
π
(
x
) =
c(x)
c
G
, where
c
G
=
P
x
c(x), because
π(x)P (x, y) =
c(x)
c
G
·
c(x, y)
c(x)
= π(y)P(y, x).
Conversely, every reversible Markov chain can be represented as a random walk
on a weighted graph place an edge
{x, y}
if
P
(
x, y
)
>
0. By reversibility,
P(x, y) > 0 iff P(y, x) = 0. Define weights
c(x, y) = π(x)P(x, y)
for all x y.
One way to understand random walks on weighted graphs is to think of them
as currents in electrical networks. We set the resistance on the edges by
r(e) =
1
c(e)
.
Naturally, we would like to talk about currents flowing through the circuit.
Definition
(Flow)
.
A flow
θ
on
G
is a function defined on oriented edges which
is anti-symmetric, i.e. θ(x, y) = θ(y, x).
Not every flow can be thought of as a current. We fix two distinguished
vertices
a
and
z
, called the source and the sink respectively. We think of current
as entering the network through
a
and exiting through
z
. Through any other
vertex, the amount of current going in should equal the amount of current going
out.
Definition (Divergence). The divergence of a flow θ is
div θ(x) =
X
yx
θ(x, y).
By the antisymmetry property, we have
X
x
div θ(x) =
X
x
X
yx
θ(x, y) =
X
xy
θ(x, y) = 0.
Definition (Flow from a to z). A flow θ from a to z is a flow such that
(i) div θ(x) = 0 for all x 6∈ {a, z}. (Kirchhoff’s node law)
(ii) div θ(a) 0.
The strength of the flow from
a
to
z
is
kθk
=
div θ
(
a
).We say this is a unit flow
if kθk = 1.
Since
P
div θ(x) = 0, a flow from a to z satisfies div θ(z) = div θ(a).
Now if we believe in Ohm’s law, then the voltage difference across each edge
is I(x, y)r(x, y). So we have
W (x) = W (y) + r(x, y)I(x, y)
for all
y
. We know that if we sum
I
(
x, y
) over all
y
, then the result vanishes. So
we rewrite this as
c(x, y)W (x) = c(x, y)W (y) + I(x, y).
Summing over all y, we find that the voltage satisfies
W (x) =
X
xy
c(x, y)
c(x)
W (y) =
X
xy
P(x, y)W (y).
Definition
(Harmonic function)
.
Let
P
be a transition matrix on Ω. We call
h
harmonic for P at the vertex x if
h(x) =
X
y
P(x, y)h(y).
We now take this as a definition of what a voltage is.
Definition
(Voltage)
.
A voltage
W
is a function on that is harmonic on
\ {a, z}.
The first theorem to prove is that given any two values
W
(
a
) and
W
(
z
), we
can find a unique voltage with the prescribed values at a and z.
More generally, if
B
Ω, and
X
is a Markov chain with matrix
P
, then we
write
τ
B
= min{t 0 : X
t
B},
Proposition.
Let
P
be an irreducible matrix on and
B
Ω,
f
:
B R
a
function. Then
h(x) = E
x
[f(X
τ
B
)]
is the unique extension of f which is harmonic on \ B.
Proof. It is obvious that h(x) = f(x) for x B. Let x 6∈ B. Then
h(x) = E
x
[f(X
τ
B
)]
=
X
y
P(x, y)E
x
[f(X
τ
B
) | X
1
= y]
=
X
y
P(x, y)E
y
[f(X
τ
B
)]
=
X
Y
P(x, y)h(y)
So h is harmonic.
To show uniqueness, suppose
h
0
be another harmonic extension. Take
g
=
h h
0
. Then g = 0 on B. Set
A =
x : g(x) = max
y\B
g(y)
,
and let
x A \B
. Now since
g
(
x
) is the weighted average of its neighbours, and
g
(
y
)
g
(
x
) for all neighbours
y
of
x
, it must be the case that
g
(
x
) =
g
(
y
) for
all neighbours of y.
Since we assumed
G
is connected, we can construct a path from
x
to the
boundary, where
g
vanishes. So
g
(
x
) = 0. In other words,
max g
= 0. Similarly,
min g = 0. So g = 0.
Given a voltage, Ohm’s law defines a natural flow, called the current flow.
Definition (Current flow). The current flow associated to the voltage W is
I(x, y) =
W (x) W (y)
r(x, y)
= c(x, y)(W (x) W (y)).
The current flow has an extra property, namely that it satisfies the cycle law.
For any cycle e
1
, e
2
, . . . , e
n
, we have
n
X
i=1
r(e
i
)I(e
i
) = 0.
Proposition.
Let
θ
be a flow from
a
to
z
satisfying the cycle law for any cycle.
Let I the current flow associated to a voltage W . If kθk = kIk, then θ = I.
Proof.
Take
f
=
θ I
. Then
f
is a flow which satisfies Kirchhoff’s node law
at all vertices and the cycle law for any cycle. We want to show that
f
= 0.
Suppose not. The we can find some e
1
such that f(e
i
) > 0. But
X
yx
f(x, y) = 0
for all
x
, so there must exist an edge
e
2
that
e
1
leads to such that
f
(
e
2
)
>
0.
Continuing this way, we will get a cycle of ordered edges where
f >
0, which
violates the cycle law.
Let
W
0
be a voltage with
W
0
(
a
) = 1 and
W
0
(
z
) = 0, and let
I
0
be the current
associated to W
0
. Then any other voltage W can be written as
W (x) = (W (a) W (z))W
0
(x) + W (z).
So, noting that W
0
(a) = 1, we have
W (a) = W (x) = (W (a) W (z))(W
0
(a) W
0
(x)).
Thus, if I is the current associated to W , then
kIk =
X
xa
I(a, x) =
X
xa
W (a) W (x)
r(a, x)
= (W (a) W (z))kI
0
k.
So we know that
W (a) W (z)
kIk
=
1
kI
0
k
,
and in particular does not depend on W .
Definition
(Effective resistance)
.
The effective resistance
R
eff
(
a, z
) of an electric
network is defined to be the ratio
R
eff
(a, z) =
W (a) W (z)
kIk
for any voltage
W
with associated current
I
. The effective conductance is
C
eff
(a, z) = R
eff
(a, z)
1
.
Proposition. Take a weighted random walk on G. Then
P
a
(τ
z
< τ
+
a
) =
1
c(a)R
eff
(a, z)
,
where τ
+
a
= min{t 1 : X
t
= a}.
Proof. Let
f(x) = P
x
(τ
z
< τ
a
).
Then
f
(
a
) = 0 and
f
(
z
) = 1. Moreover,
f
is harmonic on
\ {a, z}
. Let
W
be
a voltage. By uniqueness, we know
f(x) =
W (a) W (x)
W (a) W (z)
.
So we know
P
a
(τ
z
< τ
+
a
) =
X
xa
P(a, x)f (x)
=
X
xa
c(a, x)
c(a)
W (a) W (x)
W (a) W (z)
=
1
c(a)(W (a) W (z))
X
xa
I(a, x)
=
kIk
c(a)(W (a) W (z))
=
1
c(a)R
eff
(a, z)
.
Definition
(Green kernel)
.
Let
τ
be a stopping time. We define the Green
kernel to be
G
τ
(a, x) = E
a
"
X
t=0
1(X
t
= x, t < τ )
#
.
This is the expected number of times we hit x before τ occurs.
Corollary. For any reversible chain and all a, z, we have
G
τ
z
(a, a) = c(a)R
eff
(a, z).
Proof.
By the Markov property, the number of visits to
a
starting from
a
until
τ
z
is the geometric distribution Geo(P
a
(τ
z
< τ
+
a
)). So
G
τ
z
(a, a) =
1
P
a
(τ
z
< τ
+
a
)
= c(a)R
eff
(a, z).
Practically, to compute the effective resistance, it is useful to know some
high school physics. For example, we know that conductances in parallel add:
c
1
c
2
=
c
1
+ c
2
Thus, if
e
1
and
e
2
are two edges with the same endvertices, then we we can
replace them by a single edge of conductance the sum of the conductances. We
can prove that this actually works by observing that the voltages and currents
remain unchanged outside of
e
1
, e
2
, and we check that Kirchhoff’s and Ohm’s
law are satisfied with
I(e) = I(e
1
) + I(e
2
).
Similarly, resistances in series add. Let
v
be a node of degree 2 and let
v
1
and
v
2
be its 2 neighbours. Then we can replace these 2 edges by a single edge of
resistance
r
1
+
r
2
. Again, to justify this, we check Kirchoff’s and Ohm’s law with
I(v
1
v
2
) = I(v
1
v) + I(vv
2
).
r
1
r
2
v
1
v v
2
=
r
1
+ r
2
We can also glue 2 vertices of the same potential, keeping all other edges.
Current and potential are unchanged since current doesn’t flow if there is the
same potential.
Exercise.
Let
T
be a finite connected tree, and
a, z
two vertices. Then
R
eff
(
a, z
)
is the graph distance between a and z.
It turns out it is not convenient to work with our definition of effective
resistance. Instead, we can use the well known result from high school physics
P = IV = I
2
R. For any flow, we define
Definition
(Energy)
.
Let
θ
be a flow on
G
with conductances (
c
(
e
)). Then the
energy is
E(θ) =
X
e
(θ(e))
2
r(e).
Here we sum over unoriented edges.
Note that given any flow, we can always increase the energy by pumping
more and more current along a cycle. Thus, we might expect the lowest energy
configuration to be given by the flows satisfying the cycle law, and if we focus
on unit flows, the following should not be surprising:
Theorem
(Thomson’s principle)
.
Let
G
be a finite connected graph with
conductances (c(e)). Then for any a, z, we have
R
eff
(a, z) = inf{E(θ) : θ is a unit flow from a to z}.
Moreover, the unit current flow from a to z is the unique minimizer.
Proof.
Let
i
be the unit current flow associated to potential
ϕ
. Let
j
be another
unit flow a z. We need to show that E(j) E(i) with equality iff j = i.
Let k = j i. Then k is a flow of 0 strength. We have
E(j) = E(i + k) =
X
e
(i(e) + k(e))
2
r(e) = E(i) + E(k) + 2
X
e
i(e)k(e)r(e).
It suffices to show that the last sum is zero. By definition, we have
i(x, y) =
ϕ(x) ϕ(y)
r(x, y)
.
So we know
X
e
i(e)k(e)r(e) =
1
2
X
x
X
yx
(ϕ(x) ϕ(y))k(x, y)
=
1
2
X
x
X
yx
ϕ(x)k(x, y) +
1
2
X
x
X
yx
ϕ(x)k(x, y).
Now note that
P
yx
k
(
x, y
) = 0 for any
x
since
k
has zero strength. So this
vanishes.
Thus, we get
E(j) = E(i) + E(k),
and so E(j) E(i) with equality iff E(k) = 0, i.e. k(e) = 0 for all e.
Using this characterization, it is then immediate that
Theorem
(Rayleigh’s monotonicity principle)
.
Let
G
be a finite connected
graph and (
r
(
e
))
e
and (
r
0
(
e
))
e
two sets of resistances on the edges such that
r(e) r
0
(e) for all e. Then
R
eff
(a, z; r) R
eff
(a, z; r
0
).
for all a, z G.
Proof. By definition of energy, for any current i, we have
E(i; r) E(i; r
0
).
Then take the infimum and conclude by Thompson.
Corollary.
Suppose we add an edge to
G
which is not adjacent to
a
. This
increases the escape probability
P
a
(τ
z
< τ
+
a
).
Proof. We have
P
a
(τ
z
< τ
+
z
) =
1
c(a)R
eff
(a, z)
.
We can think of the new edge as an edge having infinite resistance in the old
graph. By decreasing it to a finite number, we decrease R
eff
.
Definition
(Edge cutset)
.
A set of edges Π is an edge-cutset separating
a
from
z if every path from a to z uses an edge of Π.
Theorem
(Nash–Williams inequality)
.
Let
k
) be disjoint edge-cutsets sepa-
rating a from z. Then
R
eff
(a, z)
X
k
X
eΠ
k
c(e)
!
1
.
The idea is that if we combine all paths in each Π
k
to a single path with
conductance
P
eΠ
k
c
(
e
), then every path has to use all of these edges, so we
can bound the effective resistance.
Proof.
By Thompson, it suffices to prove that for any unit flow
θ
from
a
to
z
,
we have
X
e
(θ(e))
2
r(e)
X
k
X
eΠ
k
c(e)
!
1
.
Certainly, we have
X
e
(θ(e))
2
r(e)
X
k
X
eΠ
k
(θ(e))
2
r(e).
We now use Cauchy–Schwarz to say
X
eΠ
k
(θ(e))
2
r(e)
!
X
eΠ
k
c(e)
!
X
eΠ
k
|θ(e)|
p
r(e)c(e)
!
2
=
X
eΠ
k
|θ(e)|
!
2
.
The result follows if we show that the final sum is 1.
Let
F
be the component of
G \
Π
k
containing
a
. Let
F
0
be the set of all
edges that start in F and land outside F . Then we have
1 =
X
xF
div θ(x)
X
eF
0
|θ(e)|
X
eΠ
k
|θ(e)|.
Corollary. Consider B
n
= [1, n]
2
Z
2
. Then
R
eff
(a, z)
1
2
log(n 1).
There is also an upper bound of the form
R
eff
(
a, z
)
c log n
. So this is
indeed the correct order. The proof involves constructing an explicit flow with
this energy.
Proof. Take
Π
k
= {(v, u) B
n
: kvk
= k 1, kuk
= k}.
Then
|Π
k
| = 2(k 1),
and so
R
eff
(a, z)
1
X
k=1
1
|Π
k
|
1
2
log(n 1).
We shall need the following result:
Proposition.
Let
X
be an irreducible Markov chain on a finite state space. Let
τ
be a stopping time such that
P
a
(
X
τ
=
a
) = 1 and
E
a
[
τ
]
<
for some
a
in
the state space. Then
G
τ
(a, x) = π(x)E
a
[τ].
A special case of this result was proven in IB Markov Chain, where
τ
was
taken to be the first return time to
a
. The proof of this is very similar, and we
shall not repeat it.
Using this, we can prove the commute-time identity.
Theorem
(Commute time identity)
.
Let
X
be a reversible Markov chain on a
finite state space. Then for all a, b, we have
E
a
[τ
b
] + E
b
[τ
a
] = c(G)R
eff
(a, b),
where
c(G) = 2
X
e
c(e).
Proof.
Let
τ
a,b
be the first time we hit
a
after we hit
b
. Then
P
a
(
X
τ
a,b
=
a
) = 1.
By the proposition, we have
G
τ
a,b
(a, a) = π(a)E
a
[τ
a,b
].
Since starting from
a
, the number of times we come back to
A
by
τ
a,b
is the
same as up to time τ
b
, we get
G
τ
a,b
(a, a) = G
τ
b
(a, a) = c(a)R
eff
(a, b).
Combining the two, we get
π(a)E
a
[τ
a,b
] = c(a)R
eff
(a, b).
Thus, the result follows.
In particular, from the commute-time identity, we see that the effective
resistance satisfies the triangle inequality, and hence defines a metric on any
graph.
2.2 Infinite graphs
Finally, let’s think a bit about how we can develop the same theory for infinite
graphs.
Let
G
be an infinite connected weighted graph with conductances (
c
(
e
))
e
.
Let 0 be a distinguished vertex. Let (
G
k
= (
V
k
, E
k
))
k
be an exhaustion of
G
by
finite graphs, i.e.
V
k
V
k+1
for all k;
0 V
k
for all k;
S
k
V
k
= V ;
E
k
is the set of edges of E with both end points in V
k
.
Let
G
n
be the graph obtained by gluing all the vertices of
V \V
n
into a single
point z
n
and setting
c(x, z
n
) =
X
zV \V
n
c(x, z)
for all
x V
n
. Now observe that
R
eff
(0
, z
n
;
G
n
) is a non-decreasing function of
n (Rayleigh). So the limit as n exists. We set
R
eff
(0, ) = lim
n→∞
R
eff
(0, z
n
; G
n
).
This is independent of the choice of exhaustion by Rayleigh.
Now we have
P
0
(τ
+
0
= ) = lim
n→∞
P
0
(τ
z
n
< τ
+
0
) =
1
c(0)R
eff
(0, z
n
; G
n
)
=
1
c(0)R
eff
(0, )
.
Definition
(Flow)
.
Let
G
be an infinite graph. A flow
θ
from 0 to
is an
anti-symmetric function on the edges such that div θ(x) = 0 for all x 6= 0.
Theorem.
Let
G
be an infinite connected graph with conductances (
c
(
e
))
e
.
Then
(i) Random walk on G is recurrent iff R
eff
(0, ) = .
(ii)
The random walk is transient iff there exists a unit flow
i
from 0 to
of
finite energy
E(i) =
X
e
(i(e))
2
r(e).
Proof. The first part is immediate since
P
0
(τ
+
0
= ) =
1
c(0)R
eff
(0, )
.
To prove the second part, let
θ
be a unit flow from 0 to
of finite energy. We
want to show that the effective resistance is finite, and it suffices to extend
Thompson’s principle to infinite graphs.
Let
i
n
be the unit current flow from 0 to
z
n
in
G
n
. Let
v
n
(
x
) be the associated
potential. Then by Thompson,
R
eff
(0, z
n
; G
n
) = E(i
n
).
Let
θ
n
be the restriction of
θ
to
G
n
. Then this is a unit flow on
G
n
from 0 to
z
n
. By Thompson, we have
E(i
n
) E(θ
n
) E(θ) < .
So
R
eff
(0, ) = lim
n→∞
E(i
n
).
So by the first part, the flow is transient.
Conversely, if the random walk is transient, we want to construct a unit flow
from 0 to
of finite energy. The idea is to define it to be the limit of (
i
n
)(
x, y
).
By the first part, we know
R
eff
(0
,
) =
lim E
(
i
n
)
<
. So there exists
M >
0 such that
E
(
i
n
)
M
for all
n
. Starting from 0, let
Y
n
(
x
) be the number
of visits to
x
up to time
τ
z
n
. Let
Y
(
x
) be the total number of visits to
x
. Then
Y
n
(x) % Y (x) as n almost surely.
By monotone convergence, we get that
E
0
[Y
n
(x)] % E
0
[Y (x)] < ,
since the walk is transient. On the other hand, we know
E
0
[Y
n
(x)] = G
τ
z
n
(0, x).
It is then easy to check that
G
τ
z
n
(0,x)
c(x)
is a harmonic function outside of 0 and
z
n
with value 0 at z
n
. So it has to be equal to the voltage v
n
(x). So
v
n
(x) =
G
τ
z
n
(0, x)
c(x)
.
Therefore there exists a function v such that
lim
n→∞
c(x)v
n
(x) = c(x)v(x).
We define
i(x, y) = c(x, y)(v(x) v(y)) = lim
n→∞
c(x, y)(v
n
(x) v
n
(y)) = lim
n→∞
i
n
(x, y).
Then by dominated convergence, we know
E
(
i
)
M
, and also
i
is a flow from 0
to .
Note that this connection with electrical networks only works for reversible
Markov chains.
Corollary. Let G
0
G be connected graphs.
(i) If a random walk on G is recurrent, then so is random walk on G
0
.
(ii) If random walk on G
0
is transient, so is random walk on G.
Theorem
(Polya’s theorem)
.
Random walk on
Z
2
is recurrent and transient on
Z
d
for d 3.
Proof sketch. For d = 2, if we glue all vertices at distance n from 0, then
R
eff
(0, )
n1
X
i=1
1
8i 4
c · log n
using the parallel and series laws. So R
eff
(0, ) = . So we are recurrent.
For
d
= 3, we can construct a flow as follows let
S
be the sphere of radius
1
4
centered at the origin. Given any edge
e
, take the unit square centered at
the midpoint
m
e
of
e
and has normal
e
. We define
|θ
(
e
)
|
to be the area of the
radial projection on the sphere, with positive sign iff hm
e
, ei > 0.
One checks that
θ
satisfies Kirchoff’s node law outside of 0. Then we find
that
E(θ) C
X
n
n
2
·
1
n
2
2
< ,
since there are
n
2
edges at distance
n
, and the flow has magnitude
1
n
2
.
Another proof sketch.
We consider a binary tree
T
ρ
with edges joining generation
to n + 1 having resistance ρ
n
for ρ to be determined. Then
R
eff
(T
ρ
) =
X
n=1
ρ
2
n
,
and this is finite when ρ < 2.
Now we want to embed this in
Z
3
in such a way that neighbours in
T
ρ
of generation
n
and
n
+ 1 are separated by a path of length of order
ρ
n
. In
generation
n
of
T
ρ
, there are 2
n
vertices. On the other hand, the number of
vertices at distance
n
in
Z
3
will be of order (
ρ
n
)
2
. So we need (
ρ
n
)
2
2
n
. So
we need ρ >
2. We can then check that
R
eff
(0, ; Z
3
) c · R
eff
(T
ρ
) < .
Note that this method is rather robust. We only need to be able to embed
our
T
ρ
into
Z
3
. Using a similar method, Grimmett, Kesten and Zhang (1993)
showed that simple random walk on a supercritical bond percolation is transient
for d 3.
3 Uniform spanning trees
3.1 Finite uniform spanning trees
The relation between uniform spanning trees and electrical networks was discov-
ered by Kirchoff in 1850’s. Of course, he was not interested in uniform spanning
tress, but in electrical networks themselves. One of the theorems we will prove is
Theorem
(Foster’s theorem)
.
Let
G
= (
V, E
) be a finite weighted graph on
n
vertices. Then
X
eE
R
eff
(e) = n 1.
We can prove this using the commute-time formula, but there is a one-line
proof using this connection between electrical networks and uniform spanning
trees.
Definition
(Spanning tree)
.
Let
G
= (
V, E
) be a finite connected graph. A
spanning tree
T
of
G
is a connected subgraph of
G
which is a tree (i.e. there are
no cycles) and contains all the vertices in G.
Let
T
be the set of all spanning trees of
G
. Pick
T T
uniformly at random.
We call it the uniform spanning tree. We shall prove the following theorem:
Theorem. Let e 6= f E. Then
P(e T | f T ) P(e T ).
It is tempting to ask a very similar question if we instead considered all
spanning forests, do we get the same result? Intuitively, we should, but it turns
out this is an open question. So let’s think about trees instead.
In order to prove this, we need the following result:
Theorem (Kirchoff). Let T be a uniform spanning tree, e an edge. Then
P(e T ) = R
eff
(e)
Of course, this immediately implies Foster’s theorem, since every tree has
n 1 edges.
Notation.
Fix two vertices
s, t
of
G
. For all every edge
e
= (
a, b
), define
N
(
s, a, b, t
) to be the set of spanning tress of
G
whose unique path from
s
to
t
passes along the edge (a, b) in the direction from a to b. Write
N(s, a, b, t) = |N(s, a, b, t)|,
and N the total number of spanning trees.
Theorem. Define, for every edge e = (a, b),
i(a, b) =
N(s, a, b, t) N(s, b, a, t)
N
.
Then
i
is a unit flow from
s
to
t
satisfying Kirchoff’s node law and the cycle law.
Note that from the expression for i, we know that
i(a, b) = P(T N(s, a, b, t)) P(T N(s, b, a, t)).
Proof.
We first show that
i
is a flow from
s
to
t
. It is clear that
i
is anti-
symmetric. To check Kirchoff’s node law, pick
a 6∈ {s, t}
. We need to show
that
X
xa
i(a, x) = 0.
To show this, we count how much each spanning tree contributes to the sum. In
each spanning tree, the unique path from
s
to
t
may or may not contain
a
. If
it does not contain
a
, it contributes 0. If it contains
A
, then there is one edge
entering
a
and one edge leaving
a
, and it also contributes 0. So every spanning
tree contributes exactly zero to the sum.
So we have to prove that
i
satisfies the cycle law. Let
C
= (
v
1
, v
2
, . . . , v
n+1
=
v
1
) be a cycle. We need to show that
n
X
i=1
i(v
i
, v
i+1
) = 0.
To prove this, it is easier to work with “bushes” instead of trees. An
s/t
bush is
a forest that consists of exactly 2 trees,
T
s
and
T
t
, such that
s T
s
and
t T
t
.
Let
e
= (
a, b
) be an edge. Define
B
(
s, a, b, t
) to be the set of
s/t
bushes such
that a T
s
and b T
t
.
We claim that
|B
(
s, a, b, t
)
|
=
N
(
s, a, b, t
). Indeed, given a bush in
B
(
s, a, b, t
),
we can add the edge
e
= (
a, b
) to get a spanning tree whose unique path from
s
to t passes through e, and vice versa.
Instead of considering the contribution of each tree to the sum, we count the
contribution of each bush to the set. Then this is easy. Let
F
+
= |{(v
j
, v
j+1
) : B B(s, v
j
, v
j+1
, t)}
F
= |{(v
j
, v
j+1
) : B B(s, v
j+1
, v
j
, t)}
Then the contribution of B is
F
+
F
N
.
By staring it at long enough, since we have a cycle, we realize that we must have
F
+
= F
. So we are done.
Finally, we need to show that i is a unit flow. In other words,
X
xs
i(s, x) = 1.
But this is clear, since each spanning tree contributes
1
N
to
i
(
s, x
), and there are
N spanning trees.
We can now prove the theorem we wanted to prove.
Theorem. Let e 6= f E. Then
P(e T | f T ) P(e T ).
Proof.
Define the new graph
G.f
to be the graph obtained by gluing both
endpoints of
f
to a single vertex (keeping multiple edges if necessary). This
gives a correspondence between spanning trees of
G
containing
f
and spanning
trees of G.f. But
P(e T | f T ) =
number of spanning trees of G.f containing e
number of spanning trees of G.f
.
But this is just
P
(
e UST of G.f
), and this is just
R
eff
(
e
;
G.f
). So it suffices
to show that
R
eff
(e; G.f ) R
eff
(e; G).
But this is clear by Rayleigh’s monotone principle, since contracting
f
is the
same as setting the resistance of the edge to 0.
In practice, how can we generate a uniform spanning tree? One way to do so
is Wilson’s method.
Definition
(Loop erasure)
.
Let
x
=
hx
1
, . . . x
n
i
be a finite path in the graph
G
. We define the loop erasure as follows: for any pair
i < j
such that
x
i
=
x
j
,
remove x
i+1
, x
i+2
, . . . , x
j
, and keep repeating until no such pairs exist.
To implement Wilson’s algorithm, distinguish a root vertex of
G
, called
r
,
and take an ordering of
V
. Set
T
0
=
{r}
. Define
T
i
inductively as follows: take
the first vertex in the ordering is not in
T
i
. We start a simple random walk from
this vertex, and run until it hits
T
i
, and erase the loops. Then set
T
i+1
to be
the union of T
i
with this (loop-erased) path. Repeat until all vertices are used.
Theorem (Wilson). The resulting tree is a uniform spanning tree.
Note that in particular, the choice of ordering is irrelevant to the resulting
distribution.
To prove this theorem, it is convenient to have a good model of how the
simple random walk on
G
is generated. Under any vertex
G
, place an infinite
number of cards that contain iid instructions, i.e. each card tells us to which
neighbour we jump. Different vertices have independent stacks. Whenever we
are at a vertex
x
, look at the top card underneath it and go to the neighbour it
instructs. Then throw away the top card and continue.
Given stacks of cards under the vertices of the graph, by revealing the top
card of each vertex, we get a directed graph, where directed edges are of the
form (x, y), where y is the instruction on the top card under x.
If there is no directed cycle, then we stop, and we have obtained a spanning
tree. If there is a directed cycle, then we remove the top cards from all vertices
on the cycle. We call this procedure cycle popping. Then we look at the top
cards again and remove cycles and cards used.
We first prove a deterministic lemma about cycle popping.
Lemma.
The order in which cycles are popped is irrelevant, in the sense that
either the popping will never stop, or the same set of cycles will be popped, thus
leaving the same spanning tree lying underneath.
Proof.
Given an edge (
x, S
i
x
), where
S
i
x
is the
i
th instruction under
x
, colour
(
x, S
i
x
) with colour
i
. A colour is now coloured, but not necessarily with the
same colour.
Suppose
C
is a cycle that can be popped in the order
C
1
, C
2
, . . . , C
n
, with
C
n
=
C
. Let
C
0
be any cycle in the original directed graph. We claim that
either
C
0
does not intersect
C
1
, . . . , C
n
, or
C
0
=
C
k
for some
k
, and
C
0
does not
intersect
C
1
, . . . C
k1
. Indeed, if they intersect, let
x C
0
C
k
, where
k
is the
smallest index where they intersect. Then the edge coming out of
x
will have
colour 1. Then
S
1
x
is also in the intersection of
C
0
and
C
k
, and so the same is
true for the edge coming out of
S
1
x
. Continuing this, we see that we must have
C
k
=
C
0
. So popping
C
k
, C
1
, . . . , C
k1
, C
k+1
, . . . , C
n
gives the same result as
popping C
1
, . . . , C
n
.
Thus, by induction, if
C
is a cycle that is popped, then after performing a
finite number of pops,
C
is still a cycle that can be popped. So either there is
an infinite number of cycles that can be popped, so popping can never stop, or
every cycle that can be popped will be popped, thus in this case giving the same
spanning tree.
Using this, we can prove that Wilson’s method gives the correct distribution.
In Wilson’s algorithm, we pop cycles by erasing loops in the order of cycle
creation. This procedure will stop will probability 1. With this procedure, we
will reveal a finite set of coloured cycles
O
lying over a spanning tree. Let
X
be the set of all (
O, T
), where
O
is the set of coloured cycles lying on top of
T
,
which arise from running the random walks.
Now if we can get to (
O, T
), then we could also get to (
O, T
0
) for any other
spanning spanning tree
T
0
, since there is no restriction on what could be under
the stacks. So we can write
X = X
1
× X
2
,
where
X
1
is a certain set of finite sets of coloured cycles, and
X
2
is the set of all
spanning trees.
Define a function Ψ on the set of cycles by
Ψ(C) =
Y
eC
p(e), p(e) =
c(e)
c(e
)
,
where
e
= (
e
, e
+
), and
c
(
e
) is the sum of the conductances of all edges with
vertex e
.
More generally, if O is a set of cycles, then we write
Ψ(O) =
Y
CO
Ψ(C).
Then
P(get (O, T ) using Wilson’s algorithm) =
Y
e
S
CO
CT )
p(e) = Ψ(O) ·Ψ(T ).
Since this probability factorizes, it follows that the distribution of
O
is indepen-
dent of
T
, and the probability of getting
T
is proportional to Ψ(
T
). So
T
has
the distribution of a uniform spanning tree.
Corollary
(Cayley’s formula)
.
The number of labeled unrooted trees on
n
-
vertices is equal to n
n2
.
Proof. Exercise.
How can we generalize Wilson’s algorithm to infinite graphs? If the graph is
recurrent, then we can still use the same algorithm, and with probability one,
the algorithm terminates. If we have a transient graph, then we can still perform
the same algorithm, but it is not clear what the resulting object will be.
The Aldous–Broder algorithm is another way of generating a uniform spanning
tree. Run a random walk that visits all vertices at least once. For every vertex
other than the root, take the edge the walk crossed the first time it visited the
vertex, and keep the edge in the spanning tree. It is an exercise to prove that
this does give a uniform spanning tree. Again, this algorithm works only on a
recurrent graph. A generalization to transient graphs was found by Hutchcroft.
3.2 Infinite uniform spanning trees and forests
Since recurrent graphs are boring, let G be an infinite transient graph. Then a
“naive” way to define a uniform spanning tree is by exhaustions.
Let
G
n
be an exhaustion of
G
by finite graphs. Define
G
W
n
(where
W
stands
for “wired”) to be the graph
G
n
, where all of the vertices of
G \ G
n
are glued to
a single vertex
z
n
. The “free”
G
n
is just
G
n
considered as a subgraph. It turns
out the wired one is much easier to study.
Let
µ
n
be the uniform spanning tree measure on
G
W
n
. Let
B
be a finite set
of edges B = {e
1
, . . . , e
n
}. Then for T a UST of G
W
n
, we have
µ
n
(B T ) =
m
Y
k=1
µ(e
k
T | e
j
T for all j < k)
=
n
Y
k=1
R
eff
(e
k
; G
W
n
/{e
j
: j < k}),
where
G
W
n
/{e
j
:
j < k}
is the graph obtained from
G
W
n
by contracting the edges
{e
j
: j < k}.
We want to show that
µ
n
as a sequence of measures converges. So we want
to understand how
R
eff
(
e
k
;
G
W
n
/{e
j
:
j < k}
) changes when we replace
n
with
n
+ 1. By Rayleigh’s monotonicity principle, since
G
W
n
can be obtained from
G
W
n+1
by contracting more edges, we know the effective resistance increases when
we replace n with n + 1. So we know that
µ
n
(B T ) µ
n+1
(B T ).
So (µ
n
(B T )) is an increasing sequence. So we can define
µ(B F) = lim
n→∞
µ
n
(B T ),
and
F
is our uniform forest. We can then extend the definition of
µ
to cylinder
events
{F : B
1
F, B
2
F = ∅}
with B
1
, B
2
finite sets of edges, using inclusion-exclusion:
µ(B
1
F, B
2
F = ) =
X
SB
2
µ(B
1
S F)(1)
|S|
.
Then by commuting limits, we find that in fact
µ(B
1
F, B
2
F = ) = lim
n→∞
µ
n
(B
1
T, B
2
T = ).
If A is a finite union or intersection of cylinder sets of this form, then we have
µ(A) = lim
n→∞
µ
n
(A).
So this gives a consistent family of measures.
By Kolmogorov’s theorem, we get a unique probability measure
µ
supported
on the forests of
G
.
µ
is called the wired uniform spanning forest. One can
also consider the free uniform spanning forest, where we do the same thing but
without gluing, however, this is more complicated and we will not study it.
We can adapt Wilson’s algorithm to generate uniform spanning forests. This
is called Wilson’s method rooted at infinity (BKPS, 2001).
Let
G
be a transient graph, and start with
F
0
=
. Pick an ordering of the
vertices of
V
=
{x
1
, x
2
, . . .}
. Inductively, from
x
n
, we start a simple random
walk until the first time it hits
F
n1
if it does. If it doesn’t, then run indefinitely.
Call this (possibly infinite) path
P
n
. Since
G
is transient,
P
n
will visit every
vertex finitely many times with probability 1. So it makes sense to define the
loop erasure of P
n
. Call it LE(P
n
). We set
F
n
= F
n1
LE(P
n
),
and take
F =
[
n
F
n
.
As before, the order of V that we take does not affect the final distribution.
Proposition.
Let
G
be a transient graph. The wired uniform spanning forest
is the same as the spanning forest generated using Wilson’s method rooted at
infinity.
Proof.
Let
e
1
, . . . , e
M
be a finite set of edges. Let
T
(
n
) be the uniform spanning
tree on
G
W
n
. Let
F
be the limiting law of
T
(
n
). Look at
G
W
n
and generate
T
(
n
)
using Wilson’s method rooted at
z
n
. Start the random walks from
u
1
, u
2
, . . . , u
L
in this order, where (
u
i
) are all the end points of the
e
1
, . . . , e
M
(
L
could be less
than 2M if the edges share end points).
Start the first walk from
u
1
and wait until it hits
z
n
; Then start from the
next vertex and wait until it hits the previous path. We use the same infinite
path to generate all the walks, i.e. for all
i
, let (
X
k
(
u
i
))
k0
be a simple random
walk on
G
started from
u
i
. When considering
G
W
n
, we stop these walks when
they exit
G
n
, .e. if they hit
z
n
. In this way, we couple all walks together and all
the spanning trees.
Let
τ
n
i
be the first time the
i
th walk hits the tree the previous
i
1 walks
have generated in G
W
n
. Now
P(e
1
, . . . , e
M
T (n)) = P
e
1
, . . . , e
M
L
[
j=1
LE(X
k
(u
j
)) : k τ
j
n
.
Let
τ
j
be the stopping times corresponding to Wilson’s method rooted at infinity.
By induction on
j
, we have
τ
n
j
τ
j
as
n
, and by transience, we have
LE(X
k
(u
j
) : k τ
k
j
) LE(X
k
(u
j
) : k τ
j
). So we are done.
Theorem
(Pemantle, 1991)
.
The uniform spanning forest on
Z
d
is a single tree
almost surely if and only if d 4.
Note that the above proposition tells us
Proposition
(Pemantle)
.
The uniform spanning forest is a single tree iff starting
from every vertex, a simple random walk intersects an independent loop erased
random walk infinitely many times with probability 1. Moreover, the probability
that
x
and
y
are in the same tree of the uniform spanning forest is equal to the
probability that simple random walk started from
x
intersects an independent
loop-erased random walk started from y.
To further simplify the theorem, we use the following theorem of Lyons, Peres
and Schramm:
Theorem
(Lyons, Peres, Schramm)
.
Two independent simple random walks
intersect infinitely often with probability 1 if one walks intersects the loop erasure
of the other one infinitely often with probability 1.
Using these, we can now prove the theorem we wanted. Note that proving
things about intersection, rather than collisions, tends to be higher, because any
attempts to, say, define the first intersection time would inherently introduce an
asymmetry.
We prove half of Pemantle’s theorem.
Theorem.
The uniform spanning forest is not a tree for
d
5 with probability
1.
Proof of Pemantle’s theorem.
Let
X, Y
be two independent simple random walks
in Z
d
. Write
I =
X
t=0
X
s=0
1(X
t
= Y
s
).
Then we have
E
x,y
[I] =
X
t
X
s
P
xy
(X
t+s
= 0)
X
t=kxyk
tP
xy
(X
t
= 0).
It is an elementary exercise to show that
P
x
(X
t
= 0)
c
t
d/2
,
so
P
x
(X
t
= 0)
X
t=kxyk
1
t
d/21
.
For d 5, for all ε > 0, we take x, y such that
E
x,y
[I] < ε.
Then
P(USF is connected) P
x,y
(I > 0) E
x,y
[I] < ε.
Since this is true for every ε, it follows that P(USF is connected) = 0.