Intro Particle Systems
Intro Particle Systems
An Introduction
Thomas M. Liggett*
LNS0417001
'tmlmath.ucla.edu
Abstract
Interacting particle systems is a large and growing field of proba-
bility theory that is devoted to the rigorous analysis of certain types of
models that arise in statistical physics, biology, economics, and other
fields. In these notes, we provide an introduction to some of these
models, give some basic results about them, and explain how certain
important tools are used in their study. The first chapter describes con-
tact, voter and exclusion processes, and introduces the tools of coupling
and duality. Chapter 2 is devoted to an analysis of translation invariant
linear voter models, using primarily the duality that is available in that
case. Chapters 3-5 are concerned with the exclusion process, beginning
with the symmetric case, in which one can also use duality, and fol-
lowing with asymmetric systems, which are studied using coupling and
other monotonicity techniques. At the end, we report on some very
recent work on the stationary distributions of one dimensional systems
with positive drift.
Contents
1 Introduction 5
1.1 Examples 6
1.2 Coupling 10
1.3 Duality 15
References 56
Interacting Particle Systems - An Introduction 5
1 Introduction
These notes are intended to provide a first exposure to the area of interacting
particle systems. More comprehensive treatments of this field can be found
in the author's 1985 and 1999 books, as well as in the references in the latter
book. The main prerequisite for reading these notes is a good background
in measure theoretic probability theory.
The processes we will discuss here are continuous time Feller processes
t]t on the compact configuration space {0,1} S , where S is a countable set.
(Recall that a Feller process is a strong Markov process whose transition
measures are weakly continuous in the initial state.) A very simple such
process is that in which the coordinates r]t(x) evolve according to indepen-
dent two state Markov chains with transitions from 0 to 1 and from 1 to 0
at rate 1. Even this simple example illustrates some of the important differ-
ences between particle systems and more classical Feller processes such as
Brownian motion on Rd. The distribution of Brownian motion at any pos-
itive time is equivalent to Lebesgue measure, and hence these distributions
at different positive times are equivalent to each other. In our example of
independent two state Markov chains, the situation is entirely different. For
example, if r/o = 1, then the distribution at every time is a homogeneous
product measure with time dependent density, so the distributions of the
process at different times are mutually singular with respect to one another.
The main issues to be considered here are:
(a) a description of the class X of stationary distributions for the process,
and
(b) limit theorems for the distribution of r]t as t —> oo.
Settling these issues is naturally very easy in the example of independent
two state Markov chains. The only stationary distribution is the product
measure with density \, and there is convergence to this limit for every initial
distribution.
In general, all that can be said is that since {0,1} S is compact and
the process is Feller, X is always nonempty. Typically, there will be some
stationary distributions that can be written down explicitly, and the task is
to determine whether or not these exhaust all of X. If not, these additional
stationary distributions must usually be constructed by some type of limiting
argument. The set X is convex, and we will denote its extreme points by Xe.
The process r]t is usually described by specifying the rates at which tran-
sitions occur. If S is finite, saying that the transition 77 —> £ (for 77 ^ £)
6 T.M. Liggett
where c(r], () is the rate at which transitions occur from 77 to (. Our choices of
rates will guarantee that the series in (2) converges uniformly for cylinder / ' s .
The stationary distributions of the process are determined by the generator
as follows:
Theorem 1. A probability measure [/, on {0,1} 5 is stationary for the process
if and only if
Qfdfi = 0
{ j](u)
r](y)
r](x)
iiu^x,y,
iiu — x,
ifu = y.
Interacting Particle Systems - An Introduction 7
Thus f)x is obtained from 77 by changing its value at x, while f)x<Sj is obtained
by interchanging the values at x and y. In the latter case, if r](x) ^ f](y),
the transition 77 —> r]Xjy can be interpreted as moving a particle from x to y
or vice versa.
Example 1. The contact process. Here S is a graph whose vertices have
bounded degree, and A is a positive parameter. Use the notation x ~ y to
mean that the vertices x and y are connected by an edge. Then for each
x G S,
fl
f] —> Vx a t r a t e <
\\\{y~x:V(y) = l}\
Here |yl| denotes the cardinality of the set A. The interpretation is that
sites with r](x) — 1 are infected, while sites with r](x) — 0 are healthy.
Infected sites recover from the infection after an exponential time of rate 1,
while healthy sites become infected at a rate proportional to the number of
infected neighbors. The only "trivial" (in the sense that it can be found by
inspection) stationary distribution is the pointmass 5o on the configuration
77 = 0 .
While we have described the contact process as a model for the spread
of infection, it arises in other contexts as well. For example, it is related to
Reggeon Field Theory in high energy physics, and it is a building block for
more complex models in biology.
Here are some of the answers to our basic questions in the case of the
contact process. They illustrate some of the variety of behavior that even
relatively simple systems can exhibit.
If S is finite, then I — {do}, and r]t is eventually = 0 for any initial
configuration. This follows from finite state Markov chain theory.
If S — Zd, the d-dimensional integer lattice, then there is a critical value
(i) A < \(d) implies that I — {5Q} and r\t —> So weakly for any initial
configuration,
and
(ii) A > \(d) implies that Ie — {5Q, U} for some u ^ 5Q, and r\t —> v weakly
for any initial configuration with infinitely many infected sites.
Remark. \i d — 1, A = A(l), and 770 = 1, then 77^ —> 5o weakly, yet with
probability 1, for every x, r]t(x) — 1 for arbitrarily large i's - see Theorem
8 T.M. Liggett
3.10 of Chapter VI of Liggett (1985) for the latter statement. This illustrates
the fact that weak convergence results often miss interesting a.s. behavior.
If S — Td (d > 2), the tree in which every vertex has d + 1 neighbors,
then there are two critical values satisfying
so that
(i) A < \i(d) implies that I — {5o} and r]t —> So weakly for any initial
configuration,
(ii) Ai (d) < A < A2 (d) implies that Ie is infinite, and
(iii) A > \2(d) implies that le — {60, v} for some v ^ 60, and rjt -> v weakly
for any initial configuration with infinitely many infected sites.
In case (ii), if the initial configuration 77 has finitely many infected sites, then
p(x,y)>0 and ^ p ( s , y ) = 1.
y
The interpretation is that sites are individuals who at any time can have one
of two opinions (denoted by 0 and 1) on an issue. At exponential times of
rate 1, the individual at x chooses a y with probability p(x,y), and adopts
Interacting Particle Systems - An Introduction 9
in order to guarantee that r\t is well defined as a Feller process. The transition
rates are given by
«/(»/)= E p(x,v)[f(vx,y)-f(v)]-
r,(x)=l,r)(y)=0
Note that (3) implies that this series converges uniformly, and hence defines
a continuous function on {0,1} S .
Again, 60 and 61 are stationary, and it is often possible to produce other
stationary distributions explicitly. To see that one should expect to have
more stationary distributions for the exclusion process than for the linear
voter model, consider the case in which S is finite. Assuming irreducibility
of the Markov chain with transition probabilities p(x, y), it is easy to see
that Ie — {5o, S\} in the case of the linear voter model, while for the exclu-
sion process, Ie has one element for each 0 < n < \S\ — it is the stationary
10 T.M. Liggett
\f(x) - f(y)\ = \Ef(Xt) - Ef(Yt)\ < E\f(Xt) - f(Yt)\ < 2||/||0OP(r > t).
To complete the proof, let t —> oo.
R e m a r k s , (a) The assumption of Theorem 2 is satisfied, for example, by
any irreducible recurrent random walk on Zl or Z2. To see this, let pt(x,y)
be the transition probabilities for Xf, let X\ and Y% be two independent
copies of the random walk, and write
By recurrence, the integral of the left side above is infinite. Therefore, the
integral of the right side is infinite, and hence the random walk X\ — Y"/ is
recurrent.
(b) We will use repeatedly the fact that irreducible random walks on Zd
have no nonconstant bounded harmonic functions. For many random walks,
this fact follows from Theorem 2. A coupling proof for general random walks
can be found on pages 69-70 of Liggett (1985).
12 T.M. Liggett
//*</ fdu
at rate 1,
!
at rate \\{y ~ x : 77(2/) = 1}|
1
< 1
at rate \\{y ~ a; : C(y) = 1, »?(l/) = 0}|.
0
exists. (For this, we use the fact that every cylinder function can be written
as a finite linear combination of increasing cylinder functions.) This v is
the distribution that is referred to in our discussion of Example 1 above. It
is often called the upper invariant measure, since it is the largest possible
stationary distribution.
Of course, it is possible that v — So- In fact, that is what happens for
small values of A, as we will see below in Theorem 3. If that is the case,
another application of coupling gives the following limit theorem:
14 T.M. Liggett
as t f oo for all initial distributions /i. To see this, use (4) to write
£o < /i < Si =4- 6Q = $oS(t) < fiS(t) < 6iS(t),
and pass to the limit as t f oo.
We would like to know that the set of A's for which v — So is an interval, so
that we can define the critical value Ac as the right endpoint of that interval.
But this again an easy consequence of coupling. To see this, write v\ for
the limit in (5). We would like to know that v\ is increasing in A. So, take
Ai < A2, and couple copies r]t and Ct of the processes with these parameter
values respectively, and initial distribution S\ so that r]t < Ct for all t > 0.
(This is possible because the processes are attractive, and the transition
rates are monotone functions of A.) It follows that the distributions of the
two processes at time t are ordered, and hence so are their limits as t f 00:
Proof. Let r]t be the contact process on S, and let Ct be the branching
random walk on S that is defined as follows. The value Ct(x) is a nonnegative
integer that represents the number of particles at x at time t. The possible
transitions are
C(x) -> C(x) - 1 at rate C(%),
and
C(x) -> C(x) + 1 at rate
In other words, each particle dies at rate 1, and for neighboring vertices x
and y, each particle at y gives birth to a new particle at x at rate A. Now
couple these processes with initial configurations 770 = 1 and Co = 1) so that
Vt(x) < Ct(x) for allt > 0 and x £ 5 . Then
Interacting Particle Systems - An Introduction 15
which, as we will see, implies that E(t(x) < e^" 1 )* for all x G S and
t > 0. (Note that this is true with equality on a homogeneous graph of
constant degree d, since then f(t) — E(t(x) is independent of a; and satisfies
/'(*) = (\d- l)/(t).) So, if Ad < 1, it follows that E(t(x), and hence Erjt(x)
tends to 0 as t f oo.
To check E(t(x) < e ^ " 1 ^ , one can proceed as follows. Rewrite the
expression for the derivative above as
1
We need to show that M(t) < e^" )* for all t > 0. Elementary estimates
imply that M(t) is bounded on bounded t sets. Suppose that C > 1 and
T > 0 satisfy
M(t) < Ce^'1^
for 0 < t < T. (By local boundedness, such a C exists for every choice of
T.) Using the above integral inequality, it follows that
M(t) < c V A d - ^
for 0 < t < T, with C = C - (C - l)e~ A d r . Iterating this argument gives
M(t) < e^- 1 )*
for 0 < t < T as required.
1.3 Duality
Suppose H(r], () is a nonnegative continuous function of two variables. The
Markov processes f)t and (t are said to be dual with respect to H if
t) (6)
for all 77, ( and all t > 0. For reasonable choices of if, this relation means
that probabilities related to one of the processes can be expressed in terms
16 T.M. Liggett
and
n —> n — 1 at rate 6(n)
for n > 1, where /3(n) > 0 and 6(n) > 0 for all n > 1. Letting H(rj,() =
' w e c a n c o m P u t e the left side of (7) as follows:
where Q2 is the generator of the birth and death process with transitions
and
n —> n + 1 at rate 5(n + l).
Note that this chain is irreducible, and the roles of the /3's and £'s have been
reversed. Writing (6) as
and passing to the limit as t f 00, we see that f)t has positive probability of
escaping to 00 if and only if (t is positive recurrent, and in this case,
is absorbed at 0) =
Interacting Particle Systems - An Introduction 17
1 J1 if 77 = 0 on A,
I0 otherwise.
Then
{ if (77, A\{x})
-if(77, A)
0
if x G A and 77(3:) = 1,
iixcA and rj(x) = 0,
so
[H(r,x,A)-H(r,,A)]
+A Y [H(r}x,A)-H(r],A)]
t)(a!)=O,t)(»)=l
H(n,A\{x})-\
xeA,r)(x)=l
xEA
Note that the right side of (9) is simply the generator of the process with
transitions
A -> A\{x} at rate 1 if x G A,
(10)
A -> A U {y} at rate \\{x G A : x ~ y}| if y
applied to the function H(r], •). But this process, At, can be thought of as
the contact process itself on configurations with finitely many infected sites,
18 T.M. Liggett
. . fl if 77 = I o n A . .
(x) = { ' . (12)
IU otherwise,
p(x,y)[H(Vx,A)-H(r,,A)]
p(x,y)H(r,,A\{x})[l-2r,(x)]{r,(x)[l-r,(y)]+r,(y)[l-r,(x)]}
p(x,y)[H(r,,(A\{x})U{y})-H(r,,A)].
xeA,yes
The right side above can be seen as the result of applying to H(r],A), as
a function of A, the generator of the process in which points in At move
according to the Markov chain with transition probabilities p(x, y), with the
proviso that a point that moves to an occupied site coalesces with the point
at that site. Thus At is called the coalescing Markov chain process. The
duality relation (6) becomes
This will play an essential role in our analysis of the linear voter model in
the next chapter.
3d. Duality for the symmetric exclusion process. Again take H(r], A)
as in (12), and let fi be the generator of the exclusion process. Let
{ A
A\{x}U{y}
if x,y € A or x,y ^ A,
iix£A,y£A,
A\{y}U{x} ifyeA,x£A,
and write
QH(;A)(r})= £ ) ri(x)[l-Ti(yMx,v)[H(Tlx,y,A)-H(Ti,A)]
x,yeS
x,yeS
£) p(x,y)[l -r,(y)]H(r,,A)
xeA,y<£A
+\p(y, x) - p(x, y)]H(v, A U {y})}.
In order to get a duality statement, we need to recognize the right side above
as the generator of a Markov process, applied to H(r], A) as a function of
A. This is easy to do if p(x,y) — p(y,x) for all x, y, and in this case, we
conclude that
p(x,y)[H(r,,AXty)-H(r,,A)].
Thus we see that in the symmetric case, the exclusion process is self-dual,
and hence
PV (r]t = 1 on A) = PA(r] = Ion At), (14)
where At is the finite version of the exclusion process. We will use (14)
extensively in Chapter 3. The lack of duality in the asymmetric case makes
the process more challenging to analyze, but we are rewarded with more
interesting and diverse behavior. We will address the asymmetric exclusion
process in Chapters 4 and 5.
20 T.M. Liggett
Considering the nature of the duals in the three cases, we see that for the
symmetric exclusion process, the size of the dual does not change with time.
By contrast, the dual of the linear voter model has a cardinality that can
decrease with time, while in the case of the contact process, the cardinality
of the dual can both increase and decrease. These differences have a large
effect on how the duality relation is exploited.
Remark. It might appear that this would show that for each x and y,
r]t(x) — r]t(y) from some time on. This is not necessarily the case, since
changing the t in the argument changes the random walks X(s) and Y(s).
An easy example is the nearest neighbor symmetric linear voter model in 1
dimension. If the initial configuration is • • • 1 1 1 0 0 0 • • •, for example,
then the configuration at later times has the same form, with the boundary
between l's and 0's moving like a simple symmetric random walk. Therefore,
every site changes opinion infinitely often.
We will now formalize the above argument using duality directly. Let
At be the dual process of coalescing random walks. It will be convenient to
define
g(A) = PA(\At\ < \A\ for some t > 0).
We will regard g(A) as a measure of how spread apart the points in A are
— g(A) small corresponds to A being spread out.
Lemma 1. (a) IfAcB, then g(A) < g(B).
(b) For any nonempty A,
g(A) <
BCA,\B\=2
Exercise. Write down the transition rates for the Markov chain (At, Bt) on
{(A, B) : A C B} used in the proof of part (a) of Lemma 1.
22 T.M. Liggett
Proof. For part (a), let X(t) and Y(t) be coalescing random walks starting at
x and y respectively, and let r be the coalescing time. By duality (equation
(13) of Chapter 1),
=J
Interacting Particle Systems - An Introduction 23
Therefore,
\nS(t){v : V = 1 on A} - \\ < 2PA(\At\ > 1),
which tends to 0 as t -> 00 by Lemma 2.
Exercise. Is it the case that the weak limit of /J,S(t) exists for every initial
distribution /i?
2.2 The transient case
In this section, we assume that for the independent random walks X(t) and
Y(t), X(t) — Y(t) is transient. In this case, the behavior of the process f)t is
quite different. We begin by showing how to construct a nontrivial stationary
distribution fj,p for each density 0 < p < 1. Let vp be the homogeneous
product measure on {0,1} S with density p:
vp\r\ : 77 = 1 on A} — p\A\
for each finite A C S.
By duality,
(4)
G(x,y)=/ Px(Z(t) =
Jo
be the Green function for the random walk Z(t). By the strong Markov
property applied at the hitting time of y,
G(x,y) = Px(Z(t) = y for some t > 0)G(y,y).
Similarly, for T > 0,
For part (b), use part (a) and Lemma l(b), together with the fact
that Xi(t) — Xj(t) —> oo a.s. as t —> oo for i ^ j by the transience as-
sumption. Part (c) follows from part (b) and Lemma l(a), since At and
(Xi(t),..., Xn{t)) can be coupled together in such a way that the contain-
ment
^ C { X i (*),...,*„(*)}
is maintained.
We are now in a position to develop some properties of the stationary
distributions [ip we have constructed. For the statement of part (b) below,
recall that a shift invariant probability measure on {0, \}z is said to be
ergodic if it assigns probability 0 or 1 to every shift invariant subset of
{0,1} Z . These turn out to be the extremal shift invariant distributions.
Theorem 5. For each 0 < p < 1, the measure [ip defined in (3) has the
following properties:
(a) The coordinates are asymptotically independent in the sense that
for all A,
(b) Up is translation invariant and spatially ergodic,
(c) np{n : r](x) = 1} = p,
and
(d)
(Of course, this is an abuse of notation since the limit of At does not exist;
by \Aool, we mean the limit of \At\, which does exist by monotonicity.)
Therefore, part (a) follows from
and A2 be copies of the coalescing random walks process that are coupled
so that
(a) A\ and A2 are independent,
and
(b) At = A\ U A\ for t < r = inf{s > 0 : A] n A2 ^ 0}.
Given disjoint sets .A1 and .A2, start the coupled chains with initial states
A\ = A1, Al = A2, and Ao = A1 U A2. Since |4x>| = | ^ | + 1^1 on the
event {r = oo}, (5) implies that
e?(o,o)
Exercise. Write down the transition rates for the Markov chain (A], A2, At)
used in the proof of part (b) of Theorem 5.
Next we check that we have captured all the extremal stationary distri-
butions via the construction in Theorem 5.
Theorem 6. (a) I is the closed convex hull of {/ip : 0 < p < 1}.
(b) le = {HP : 0 < p < 1}.
Interacting Particle Systems - An Introduction 27
\h(A)-Uth(A)\<g(A). (8)
Therefore,
\U,h(A) -Ut+.h(A)\ <U,g(A).
The right side tends to zero as s —> 00 by Lemma 4(b), so lim^oo Ush
exists, and is harmonic for the irreducible random walk (Xi(t),..., Xn(t)) on
Sn. Such harmonic functions are constant (see the second remark following
the proof of Theorem 2 of Chapter 1), so we conclude that there are constants
cn so that
lim Ush(A) — C\A\
for every A. Of course, cn depends on h, and therefore on the stationary
measure /i. Passing to the limit in (8), we conclude that
\h(A)-clAl\<g(A). (9)
Note that by Lemmas l(b) and 4(a), g(A) is small if A is spread out.
Therefore, (9) says that if A is spread out and has cardinality n, then h(A)
28 T.M. Liggett
C = / pn7(dp).
Jo
A necessary and sufficient condition for this is that
0 (10)
= fpm{l-p)n1{dp).
\h*(A)-h(A)\<2g(A).
Applying Vt to this and using the harmonicity for Vt of both h and h* gives
\h*(A)-h(A)\<2Vtg(A).
Letting t —> oo and using Lemma 4(c) gives h* = h, and hence /i* = /i. Thus
H = / Vplidp), (11)
./o
which completes the proof of part (a).
To prove part (b), note first that fip G Xe for each p by Theorem 5(b),
since the ergodic measures are extremal in the class of all translation invari-
ant measures. So,
W : 0 < p < 1} C Xe.
For the reverse containment, take | i E l e . By part (a), /i is a mixture of
/ip's. Since /i is extremal, it must be one of the /ip's.
Exercise. Use the fact that (10) is a necessary and sufficient condition for cn
to be a moment sequence to prove de Finetti's Theorem, which states that
every exchangeable measure /i on {0,1} S , where S is infinite, can be written
as a mixture of the homogeneous product measures vp. (See Section VII.4 of
Feller (1966). An exchangeable measure is one for which /i{rj : rj = 1 on A}
depends on A only through its cardinality.) We will use de Finetti's Theorem
in Chapters 3 and 4.
Next is a convergence theorem for translation invariant initial distribu-
tions.
Theorem 7. Suppose /i is translation invariant and spatially ergodic. Then
lim
t—>00
for each x G Sn. To see that (13) implies (12), use (7) to write
: r,(x) = 1, r,(y) = 1} - p2 = J e
be the characteristic function of the random walk jumps. Since the random
walk is irreducible, \4>{0)\ — 1 if and only if 9 — 0. Note that
Exei<Xt,9> _ ei<x,0>e-t[
Now define
Interacting Particle Systems - An Introduction 31
Then
2
J[Wt(x,r,)-Wt(x,r,)]2fi(dr,) = 7(d0),
= v)Wt(y,T}),
Since all bounded harmonic functions for the random walk are constant, we
conclude that
W(x,rj) = W(0,rj) a.s.(/i) (15)
where r u is the shift: Tur](y) — r](y + u). Therefore, W(x + u, rf) — W(x,Tur])
a.s., so by (15), W(0,77) is an a.s. invariant random variable. Since /i is
ergodic, W(0,77) is a.s. constant, and since it has mean p, we conclude that
Wt(x,rj) -^ p
Uth{{xu...,xn})= f f[Wt{xi,r,)dn ^ (T
r / = l on B}. (1)
Interacting Particle Systems - An Introduction 33
Since the cardinality of At does not change with t, this shows that every
exchangeable measure is stationary. Conversely, suppose /i is stationary.
Then fiS(t) — /i for each t, so (1) implies that the function /(-A) = /i{r/ : 77 =
1 on A} is harmonic for At. By Theorem 2, / depends on A only through
its cardinality, so /i is exchangeable.
The proof of Theorem 2 depends very much on whether the random walk
X(t) with transition probabilities p(x,y) is transient or recurrent. We will
explain the proof in these two cases in the next two sections.
3.1 The recurrent case
To prove Theorem 2 in this case, it suffices to construct a coupling of two
copies At and Bt of the finite exclusion process for any initial states AQ, BQ
of cardinality n so that \A0 n Bo\ — n — 1 with the following property:
Note that the marginal processes have the following rates: The pair (X(t), Y(t))
has the transition rates for two independent random walks with transition
probabilities p(-, •) and Ct evolves like an exclusion process, until the first
time r that X(t) — Y(t). This again uses the symmetry of ]?(•,•)• To see
this, write a configuration at some time as
* * * * *
Bt: * * * * *
ct X(t) Y(t)
Then we can regard the exclusion process with n particles as a Markov chain
on T. Let Vt be its semigroup, and let Ut be the semigroup for the system
of n independent random walks X(t) — (Xi(t), ...,Xn(t)). We need to prove
that Vtf — f implies / is constant if / is a bounded symmetric function
Interacting Particle Systems - An Introduction 35
(Px(X(t) G T) —> 1 is needed here because Lemma 4(b) refers to the set
By (4), Usf has a limit as s —> oo. Since this limit is harmonic for Ut, it is
a constant; call it C. We conclude that
\f(x)-C\<g(x), xeSn.
36 T.M. Liggett
The intuition behind this inequality is the following: g(x) is small if the
coordinates of x are widely separated. Since the difference between the in-
dependent system and the interacting system is that the latter evolution
keeps the coordinates of the process apart, one might expect that the co-
ordinates are generally further apart in the interacting system than in the
independent system. This relation of being "further apart" cannot be main-
tained in a pathwise sense, but (7) would say that the intuition is correct in
an expected value sense.
It is easier to consider this issue for another function on Sn defined by
where
A(u) = Pu(Z(t) = 0 for some t > 0).
Here Z(t) is the underlying random walk on S. Since
where U and V are the generators of Ut and Vt respectively. (To check this,
integrate the following identity from 0 to t:
(U -V)Utg*{x) —
-Utg*(x)]
1 "
= g ^2p(xi,Xj)[Utg*(xi,...,Xi-i,Xj,Xi+i,...,xn)
i,...,xn) -2Utg*(x)],
where we have used the symmetry of p(-, •) again. To check that each sum-
maud on the right above is nonnegative, it is therefore enough to check that
{ +1
-1
0
ifu = xi,
ifu = a;2,
otherwise.
Then
Utg*(x)=
Pu(Xl(ti) = w)Pv(X2(U) = w)
L
j=l w u
Xi(tj - U) # X2{tj - U) V j > i) > 0.
For a € 7i, let va be the product measure on {0,1} 5 with marginals given
by a:
va{n • v(x) = 1} = a(x).
Interacting Particle Systems - An Introduction 39
Finally, let pt(x,y) be the transition probabilities for the continuous time
Markov chain with jump probabilities p(x, y) and exponential holding times
with mean 1.
Theorem 3. (a) For every a € %,
Ha - lim uaS(t)
tHX>
exists.
(b) Halv '• vix) — 1} — &{x) for all x £ 5 , and
= 1} < a(x)a(y)
for all x ^ y €E S.
(c) na is a product measure if and only if a is a constant.
(d) le = {tia:a£ %}.
(e) If the probability measure fi on {0,1} 5 satisfies
lim
2/1,2/2
lim
Note that Theorem 1 is a special case of Theorem 3. For part (a), this
uses the fact that random walks on Zd have only constant bounded harmonic
functions, so H consists only of the constants in [0,1] in this case. To deduce
part (b) of Theorem 1 from Theorem 3, use the argument at the end of the
proof of Theorem 7 of Chapter 2.
Here is an example in which H is very large. Let S — T2, the tree in
which each vertex has 3 neighbors. Let p(x, y) — 1/3 if a; and y are neighbors,
and p(x,y) — 0 otherwise. To construct a nonconstant a £ ? f , proceed as
follows: Fix two adjacent vertices a, b, and let a(a) — 1/3 and a(b) — 2/3. If
£ is a distance n from a and n + 1 from 6, let a(x) — (l/3)2~", while if a; is a
40 T.M. Liggett
distance n from b and n+1 from a, let a(x) — 1 —(l/3)2~™. Let T* be the half
of the tree made up of sites that are closer to b than to a. Then the harmonic
function we have defined is a(x) — Px(X(t) €E T* from some time on). By
Theorem 3, there is an extremal stationary distribution for the corresponding
exclusion process with these marginals. It is not known explicitly what this
measure looks like. There are of course many other elements of %.
Remark. If p(x,y) is symmetric, then it satisfies both (a) and (b) (with
IT ^constant). In this case, we checked that the homogeneous product mea-
sures are stationary using duality in Chapter 3.
Interacting Particle Systems - An Introduction 41
fnfdva = o
for all cylinder functions / . In the interest of simplicity, we will carry out
the computation in the case in which S is finite, so that all the sums below
are finite. The general case follows by approximation, or by a more careful
handling of the sums involved. (See Theorem 2.1 of Chapter VIII of Liggett
(1985).) Then
P(x,y)[f(r]x,y) - f(ri)]dva
x,y:r)(x)=l,T)(y)=O
,V) I _ _ [f(rix,y)-f(ri)]dva.
It suffices to assume that 0 < a(x) < 1 for all x. Making the change of
variables 77 —> r]x>y in the integral, write
2p(x,y) /
,y Jrr.ri(x)=l,
(2)
x,y
(Recall that these sums are finite, since we are taking S finite in this com-
putation.)
Turning to case (b) and interchanging the roles of x and y, (2) becomes
f :fj(»)=O,fj(*)=l
42 T.M. Liggett
x,y J T]-.T](y)=0,T](x)=l
so that / £lfdva — 0.
Remark. Note that in case (b), va is not only stationary, but is reversible as
well. This means that the stationary process f)t obtained by using va as the
initial distribution (when extended to t € (—00,00)) has the property that
{fhi t € (—00, 00)} and {rj-t, t € (—00, 00)} have the same joint distributions.
Analytically, it means that the semigroup of the process is symmetric on
J fS(t)gdva = JgS(t)fdua
for f,g€. L2(va)-
4.2 Coupling
The basic coupling of two copies of the exclusion process is the process (ijt, Ct)
with the following transitions at rate p(x,y):
shift invariant are exactly the exchangeable measures. Since all exchangeable
measures are mixtures of homogeneous product measures by de Finetti's
Theorem (see the exercise preceding Theorem 7 in Chapter 2), the fact that
every exchangeable measure is stationary follows from Theorem l(a). We
begin with a basic fact about the coupled process - in a shift invariant
equilibrium situation, discrepancies of opposite type cannot coexist. This is
a consequence of the fact that discrepancies cannot be created in the coupled
process.
Theorem 3. If v* 6 l * n S * , then
= ((x) =
= CM = 1}
y
The first two terms on the right of (3) correspond to a discrepancy moving
from y to x, the third and fourth terms correspond to a discrepancy moving
from x to y, and the final term corresponds to the destruction of the discrep-
ancy at x. If v* 6 l * , then the left side of (3) is zero. If v* € S*, then the
^-probabilities appearing on the right are functions of y — x. Therefore in
this case, the first and fourth terms cancel, and the second and third terms
cancel. This implies that
for all x, y such that p(x, y) +p(y, x) > 0. Using the irreducibility of p(x, y)
and the stationarity of v* again, it then follows that (4) holds for all x,y as
required. To check this, one shows by induction on n that if
x — £0,2:1, ...,xn — y satisfy p(xi,Xi+i) > 0 for each i, (5)
then (4) holds for that pair x, y. We have proved the basis step n — 1
already. For the induction step, assume this is true for n — 1 (and all x, y
which can be joined by a path of length n — 1 as in (5)), and take x, y joined
by a path of length n as in (5). Use the notation:
>v*
46 T.M. Liggett
Since v* is stationary for the coupled process, the left side above is zero by
the induction hypothesis, and therefore the right side is also zero as required.
It would follow that the net rate of flow of discrepancies to the right across
N, for example, is
P(x, vV{(ri, 0 : V(x) = 0, ((x) = 1, r}(y) = C(y) = 0}
x<N<y
which vanishes by the mean zero assumption. This is not quite the way the
proof goes, but this computation does show how the mean zero assumption
arises.
Remark. We are very far from understanding the structure of Ie in the
general translation invariant case on Zd. In fact, for d > 2, Ie has been
completely described only in the symmetric case covered in Chapter 3. Cer-
tainly it is often not the case that
Ie = {vP,0<p< 1}.
For example, suppose there is a nonzero v € Zd so that
< >
x ) (6)
48 T.M. Liggett
for all x. Then n(x) — ce<x'v> is reversible for p(x,y), so Theorem l(b)
provides examples on stationary distributions that are not shift invariant.
An instructive example is the following. Let {ei,..., e^} be the standard
basis vectors in Zd, and take
with all other p(x, y)'s zero. Assume that all the <Vs and /%'s are strictly
positive, and set
, OLi
t>j = l O g — .
Pi
Then (6) holds with v — (vi,..., vj). The corresponding stationary distribu-
tion v satisfies
^ (7)
Note that this is constant on hyperplanes that are orthogonal to v, while one
might have expected it to be constant on hyperplanes that are orthogonal
to the the mean vector
It would be very interesting to know that all the extremal stationary distri-
butions in this case are the homogeneous product measures and the product
measures with marginals given by (7).
The picture is much more complete in the one dimensional case, however.
That is the topic of the next chapter.
/io, for example, there must be discrepancies of one type on the negative
half line, and of the other type on the positive half line. This is the worst
that can happen, though. It turns out that every element of X* concentrates
on configurations that have all discrepancies of one type to the left of all
discrepancies of the other type. (See Liggett (1976) for details.) Once this
is established, the proof of the following result follows the lines of the proof
of Theorem 4 of Chapter 4.
Theorem 1. If p > \, then Xe — {up, 0 < p < 1} U {/in, -oo < n < oo}.
Thus we see that the asymmetric case allows for the existence of more sta-
tionary distributions than does the symmetric case.
Next, we look at the limiting behavior of the process when the initial
distribution is highly non-translation invariant. For 0 < \,p < 1, let V\JP be
the product measure with
(A ifz<0
:
vxAv V{x) = 1} = <
[p if x > 0.
50 T.M. Liggett
(A ifx<0,
u(x,0) = \
[p if x > 0.
Then the (entropic) solution is given by
ifz<(l-2A)t,
if (1 - 2A)t < x < (1 - 2p)t,
ifx>(l-2p)t,
Interacting Particle Systems - An Introduction 51
lim u(x, t)
(l + e\x\)4p(0,x) <
x<0 x>0
This is possible by the fifth moment assumption on the negative tails, and
the fact the drift is strictly positive. Define the product measure va as in
Section 3.3, with a(x) chosen to be
J 1 , „ !,I_I^ if x < 0
a(x) —
x<u
is nonincreasing in t for every u. The limiting measure as t —> oo is then a
stationary blocking measure.
An important step in checking this is the following symmetry: For 77 €
{ 0 , l } z \ define 77* €E {0,1 } z ' by
rf{x) = l-rj(-x).
Then 77^ also evolves like the exclusion process. To see this, note that if a
particle in rjt moves from it to w, then a nonparticle (a site at which r]t(x) — 0)
moves from v to u. Since va is invariant under the transformation 77 —>• 77*,
so is vaS(t).
Outline of part 2 of the proof. To remove the moment assumption that
we imposed in part 1, we will need to take the p that we are interested
in, whose negative tails have only a second moment, and truncate it to
obtain an approximating sequence pk whose negative tails have a finite fifth
moment. We will need to have some sort of a priori bound that will allow
us to maintain control of the corresponding sequence of stationary blocking
measures for pk as k —> 00. To do so, suppose v is a stationary measure that
satisfies
{q : T](x) = 1} < 00, ^ v{q : r]{x) = 0} < 00, (1)
x<0 x>0
(which implies it is a blocking measure) and Ylx<o x2p{^ix) < °°- While
(1) is stronger than being blocking, note that the stationary measures con-
structed in part 1 have this stronger property. The result of this part will
be applied to p^ for each k.
Let
00
M(n) — Y^ u{r] : r](x) — l,r](x+ n)— 0}.
x=—00
l
for n € Z . For n — 1, this is the expected number of 10's in the configuration
77. Note that M(n) < 00 for all n by (1). For n > 1,
M(-ri) - M(n) = ^2 [v{n(x) = 0,77(3: + n) = 1} - v{q(x) = 1,77(3: + n) = 0}]
X
{x) = 1}] = n,
54 T.M. Liggett
Next, note that for a stationary measure, the net rate at which particles
go from the left of x to the right of x is independent of x. If the stationary
measure is also a profile measure, this rate must be zero (since it tends to
zero at ±00). Therefore, for fixed x,
)M(n) = ^ n p ( 0 , - n ) M ( - n ) .
n=l
Informally speaking, there cannot be many 10's in /i&. The stationary block-
ing measure /i we are trying to produce will be a limit point of /i& as k —> 00.
It is easy to check that such a limit point is stationary (Proposition 2.14 of
Chapter I of Liggett (1985).) The problem is to show that it is a blocking
measure. In particular, we need to rule out the possibility that /i = 5
Interacting Particle Systems - An Introduction 55
So, suppose that /i& converges weakly to ^6Q + ^6\. Then for large k,
relative to /i&, there will be a large number of l's to the left of the origin
with probability close to \. Since /i& concentrates on So, there must also
be a large number of O's to the right of the origin. These O's may occur in
a large clump, or in many small clumps. The latter possibility is ruled out
by (4). So, there must be something close to a long interval of l's followed
by a long interval of O's in the configuration. By Theorem 2, the limiting
distribution of the exclusion process beginning with the configuration
•••1 1 1 1 0 0 0 0 •••
is v\/2- (Recall that the first three parts of Theorem 2 have been proved in
the nonnearest neighbor context.) But /i& is stationary, and fi/2 has many
10's, so this again contradicts (4). Thus n ^ \b~o + \b~\.
Acknowledgments
(a) The author would like to thank J. Steif for his very helpful corrections
and comments on the first draft of these notes.
(b) Preparation of these notes was supported in part by NSF Grant
DMS-00-70465.
56 T.M. Liggett
References
[1] E.D. Andjel, M. D. Bramson, and T. M. Liggett, Shocks in the asym-
metric exclusion process, Probab. Th. Rel. Fields 78 (1988), 231-247.
[2] M.D. Bramson and T. Mountford, Blocking measures for one dimen-
sional nonzero mean exclusion processes, Ann. Probab. (2002).
[5] P.A. Ferrari, J.L. Lebowitz, and E. Speer, Blocking measures for asym-
metric exclusion processes via coupling, Bernoulli (2001).
[6] R. Holley and T.M. Liggett, Ergodic theorems for weakly interacting
infinite systems and the voter model, Ann. Probab. 3 (1975), 643-663.
[7] T.M. Liggett, Coupling the simple exclusion process, Ann. Probab. 4
(1976), 339-356.
[8] T.M. Liggett, Ergodic theorems for the asymmetric simple exclusion
process II, Ann. Probab. 5 (1977), 795-801.
[9] T.M. Liggett, Interacting Particle Systems, Springer, New York, 1985.
[10] T.M. Liggett, Stochastic Interacting Systems: Contact, Voter and Ex-
clusion Processes, Springer, Berlin, 1999.