getdoc9f87. 358KB Jun 04 2011 12:04:46 AM

J
i
on
r
t
c
e
El

o

u

a
rn l

o

f
P


c

r

ob
abil
ity

Vol. 10 (2005), Paper no. 15, pages 525-576.
Journal URL
http://www.math.washington.edu/∼ejpecp/
Logarithmic Sobolev Inequality for Zero–Range Dynamics:
Independence of the Number of Particles

Paolo Dai Pra
Dipartimento di Matematica Pura e Applicata, Universit`a di Padova
Via Belzoni 7, 35131 Padova, Italy
e-mail: daipra@math.unipd.it
and
Gustavo Posta

Dipartimento di Matematica, Politecnico di Milano
Piazza L. Da Vinci 32, 20133 Milano, Italy
e-mail: guspos@mate.polimi.it
Abstract
We prove that the logarithmic-Sobolev constant for Zero-Range Processes in a box
of diameter L may depend on L but not on the number of particles. This is a first, but
relevant and quite technical step, in the proof that this logarithmic-Sobolev constant
grows as L2 , that will be presented in a forthcoming paper ([3]).

Keywords and phrases: Logarithmic Sobolev Inequality, Zero-Range Processes.
AMS 2000 subject classification numbers: 60K35, 82C22.
Abbreviated title: log-Sobolev inequality for Zero-Range.
Submitted to EJP on April 26, 2004. Final version accepted on March 5, 2005.

525

1

Introduction


The zero-range process is a system of interacting particles moving in a discrete lattice Λ,
that here we will assume to be a subset of Zd . The interaction is “zero range”, i.e. the
motion of a particle may be only affected by particles located at the same lattice site. Let
c : N → [0, +∞) be a function such that c(0) = 0 and c(n) > 0 for every n > 0. In
the zero-range process associated to c(·) particles evolve according to the following rule:
for each site x ∈ Λ, containing ηx particles, with probability rate c(ηx ) one particle jumps
from x to one of its nearest neighbors at random. Waiting jump times of different sites are
independent. If c(n) = λn then particles are independent, and evolve as simple random
walks; nonlinearity of c(·) is responsible for the interaction. Note that particles are neither
created nor destroyed. When Λ is a finite lattice, for each N ≥ 1, the zero-range process in
Λ restricted to configurations with N particles is a finite irreducible Markov chain, whose
unique invariant measure νΛN is proportional to
Y

1
,
c(ηx )!
x∈Λ
where
c(n)! =


½

(1.1)

1
for n = 0
c(n)c(n − 1) · · · c(1) otherwise.

Moreover the process is reversible with respect to νΛN .
If the function n 7→ c(n) does not grow too fast in n, then the zero range process can
be defined in the whole lattice Zd . In this case the extremal invariant measures form a one
parameter family of uniform product measures, with marginals
µρ [ηx = k] =

1
α(ρ)k
,
Z(α(ρ)) c(k)!


(1.2)

where ρ ≥ 0, Z(α(ρ)) is the normalization, andR α(ρ) is uniquely determined by the condition
µρ [ηx ] = ρ (we use here the notation µ[f ] for f dµ).
In this paper we are interested in the rate of relaxation to equilibrium of zero-range
processes. In general, for conservative systems of symmetric interacting random walks in a
spatial region Λ, for which the interaction between particles is not too strong, the relaxation
time to equilibrium is expected to be of the order of the square of the diameter of Λ, as it is
the case for independent random walks. On a rigorous basis, however, this result has been
proved in rather few cases. For dynamics with exclusion rule and finite-range interaction
(Kawasaki dynamics) relaxation to equilibrium with rate diam(Λ)2 has been proved (see
[7, 2]) in the high temperature regime. For models without the exclusion rule, i.e. with a
possibly unbounded particle density, available results are even weaker (see [5] and [9]). Zerorange processes are special models without the exclusion rule. In some respect zero-range
processes may appear simpler than Kawasaki dynamics: the interaction is zero-range rather
than finite-range and, as a consequence, invariant measures have a simpler form. However,
they exhibit various fundamental difficulties, including:

526

• due to unboundedness of the density of particles various arguments used for exclusion

processes fail; in principle the rate of relaxation to equilibrium may depend on the
number of particles, as it actually does in some cases;
• there is no small parameter in the model; one would not like to restrict to “small”
perturbations of a system of independent particles.
In [5] a first result for zero-range processes has been obtained. Let EνΛN be the Dirichlet
form associated to the zero-range process in Λ = [0, L]d ∩ Zd with N particles. Then, under
suitable growth conditions on c(·) (see Section 2), the following Poincar´e inequality holds
νΛN [f, f ] ≤ CL2 EνΛN (f, f ),

(1.3)

where C may depend on the dimension d but not on N or L. Moreover, by suitable test
functions, the L−dependence in (1.3) cannot be improved, i.e. one can find a positive
constant c > 0 and functions f = fN,L so that νΛN [f, f ] ≥ cL2 EνΛN (f, f ) for all L, N . In other
terms, (1.3) says that the spectral gap of EνΛN shrinks proportionally to L12 , independently of
the number of particles N . It is well known that Poincar´e inequality controls convergence
to equilibrium in the L2 (νΛN )-sense: if (StΛ )t≥0 is the Markov semigroup corresponding to the
process, then for every function f

µ


¤
£
¢2 i
2t
N
N
2
N
N
Λ
ν
.
(1.4)
(f

ν
[f
])
≤ exp −

νΛ St f − νΛ [f ]
Λ
Λ
CL2

Poincar´e inequality is however not sufficient to control convergence in stronger norms, e.g.
in total variation, that would follow from the logarithmic-Sobolev inequality
p p
EntνΛN (f ) ≤ s(L, N )EνΛN ( f , f ),
(1.5)
where EntνΛN (f ) = νΛN [f log f ] − νΛN [f ] log νΛN [f ]. The constant s(L, N ) in (1.5) is intended
to be the smallest possible, and, in principle, may depend on both L and N .
Our main aim is to prove that s(L, N ) ≤ CL2 for some C > 0, i.e. the logarithmicSobolev constant scales as the inverse of the spectral gap. This is the first conservative
system with unbounded particle density for which this scaling is established (see comments
on page 423 of [4]). It turns out that the proof of this result is very long and technical, and
it roughly consists in two parts. In the first part one needs to show that
s(L) := sup s(L, N ) < +∞,

(1.6)


N ≥1

i.e. that s(L, N ) has an upper bound independent of N , while in the second part a sharp
induction in L is set up to prove the actual L2 dependence. For this induction to work one
has to choose L sufficiently large as a starting point, and for this L an upper bound for
s(L, N ) independent of N has to be known in advance. Note that for models with bounded
particle density inequality (1.6) is trivial.
This paper is devoted to the proof of (1.6), while the induction leading to the L 2 growth
is included in [3]. The proof of (1.6) is indeed very long, and relies on quite sharp estimate
on the measure νΛN . After introducing the model and stating the main result in Section 2, we
devote Section 3 to the presentation of the essential steps of the proof, leaving the (many)
technical details for the remaining sections.
527

2

Notations and Main result

Throughout this paper, for a given probability space (Ω, F, µ) and f : Ω → R measurable,
we use the following notations for mean value and covariance:

Z
µ[f ] := f dµ,
µ[f, g] := µ [(f − µ[f ])(g − µ[g])]
and, for f ≥ 0,

Entµ (f ) := µ[f log f ] − µ[f ] log µ[f ],

where, by convention, 0 log 0 = 0. Similarly, for G a sub-σ-field of F, we let µ[f |G] to denote
the conditional expectation,
µ[f, g|G] := µ[(f − µ[f |G])(g − µ[g|G])|G]
the conditional covariance, and
Entµ (f |G) := µ[f log f |G] − µ[f |G] log µ[f |G]
the conditional entropy.
If A ⊂ Ω, we denote by 1(x ∈ A) the indicator function of A. If B ⊂ A is finite we
will write B ⊂⊂ A. For any x ∈ R we will write ⌊x⌋ := sup{n ∈ Z : n ≤ x} and
⌈x⌉ := inf{n ∈ Z : n ≥ x}.
Let Λ be a possibly infinite subset of Zd , and ΩΛ = NΛ be the corresponding configuration
space, where N = {0, 1, 2, . . .} is the set of natural numbers. Given a configuration η ∈ Ω Λ
and x ∈ Λ, the natural number ηx will be referred to as the number of particles at x.
Moreover if Λ′ ⊂ Λ ηΛ′ will denote the restriction of η to Λ′ . For two elements σ, ξ ∈ ΩΛ ,

the operations σ ± ξ are defined componentwise (for the difference whenever it returns an
element of ΩΛ ). In what follows, given x ∈ Λ, we make use of the special configuration δ x ,
having one particle at x and no other particle. For f : ΩΛ → R and x, y ∈ Λ, we let
∂xy f (η) := f (η − δ x + δ y ) − f (η).
Consider, at a formal level, the operator
LΛ f (η) :=

XX

c(ηx )∂xy f (η),

(2.1)

x∈Λ y∼x

where y ∼ x means |x − y| = 1, and c : N → R+ is a function such that c(0) = 0 and
inf{c(n) : n > 0} > 0. In the case of Λ finite, for each N ∈ N \ {0}, LΛ is the infinitesimal
generator of a irreducible Markov chain on the finite state space {η ∈ ΩΛ : η Λ = N }, where
X
η Λ :=
ηx
x∈Λ

is the total number of particles in Λ. The unique stationary measure for this Markov chain
is denoted by νΛN and is given by
1 Y 1
νΛN [{η}] := N
,
(2.2)
ZΛ x∈Λ c(ηx )!
528

where c(0)! := 1, c(k)! := c(1) · · · · · c(k), for k > 0, and Z ΛN is the normalization factor. The
measure νΛN will be referred to as the canonical measure. Note that the system is reversible
for νΛN , i.e. LΛ is self-adjoint in L2 (νΛN ) or, equivalently, the detailed balance condition
c(ηx )νΛN [{η}] = c(ηy + 1)νΛN [{η − δx + δy }]

(2.3)

holds for every x ∈ Λ and η ∈ ΩΛ such that ηx > 0.
Our main result, that is stated next, will be proved under the following conditions.
Condition 2.1 (LG)
sup |c(k + 1) − c(k)| := a1 < +∞.
k∈N

As remarked in [5] for the spectral gap, N -independence of the logarithmic-Sobolev constant
requires extra-conditions; in particular, our main result would not hold true in the case
c(k) = c1(k ∈ N \ {0}). The following condition, that is the same assumed in [5], is a
monotonicity requirement on c(·) that rules out the case above.
Condition 2.2 (M) There exists k0 > 0 and a2 > 0 such that c(k) − c(j) ≥ a2 for any
j ∈ N and k ≥ j + k0 .
A simple but key consequence of conditions above is that there exists A0 > 0 such that
A−1
0 k ≤ c(k) ≤ A0 k for any k ∈ N.

(2.4)

In what follows, we choose Λ = [0, L]d ∩ Zd . In order to state our main result, we define
the Dirichlet form corresponding to LΛ and νΛN :
EνΛN (f, g) = −νΛN [f LΛ g] =

1 XX N
ν [c(ηx )∂xy f (η)∂xy g(η)] .
2 x∈Λ y∼x Λ

(2.5)

Theorem 2.1 Assume that conditions (LG) and (M) hold. Then there exists a constant
C(L) > 0, that may only depend on a1 , a2 , the dimension d and L, such that for every choice
of N ≥ 1, L ≥ 2 and f : ΩΛ → R, f > 0, we have
³p p ´
f, f .
(2.6)
EntνΛN (f ) ≤ C(L)EνΛN

3

Outline of the proof

For simplicity, the proof will be outlined in one dimension. The essential steps for the
extension to any dimension are analogous to the ones for the spectral gap, that can be found
in [4], Appendix 3.3. However, in the most technical and original estimates contained in this
work (see Sections 7 and 8), proofs are given in a general dimension d ≥ 1.

529

3.1

Step 1: duplication

The idea is to prove Theorem 2.1 by induction on |Λ|. Suppose |Λ| = 2L, so that Λ = Λ1 ∪Λ2 ,
|Λ1 | = |Λ2 | = L, where Λ1 , Λ2 are two disjoint adjacent segments in Z. By a basic identity
on the entropy, we have
h
i
EntνΛN (f ) = νΛN EntνΛN [·|ηΛ ] (f ) + EntνΛN (νΛN [f |η Λ1 ]).
(3.1)
1

η

N −η

Note that νΛN [·|η Λ1 ] = νΛΛ1 1 ⊗ νΛ2 Λ1 . Thus, by the tensor property of the entropy (see [1],
Th. 3.2.2):
·
¸
i
h
N
N
νΛ EntνΛN [·|ηΛ ] (f ) ≤ νΛ Ent ηΛ1 (f ) + Ent N −ηΛ1 (f ) .
(3.2)
νΛ

1

1

νΛ

2

Now, let s(L, N ) be the maximum of the logarithmic-Sobolev constant for the zero-range
process in volumes Λ with |Λ| ≤ L and less that N particles, i.e. s(L, N ) is the smallest
constant such that
p p
EntνΛn (f ) ≤ s(L, N )EνΛn ( f , f ).
for all f > 0, |Λ| ≤ L and n ≤ N . Then, by (3.2),
·
¸
i
h
p p
p p
N
N
νΛ EntνΛN [·|ηΛ ] (f ) ≤ s(L, N )νΛ E ηΛ1 ( f , f ) + E N −ηΛ1 ( f , f )
νΛ
νΛ
1
2
p1 p
= s(L, N )EνΛN ( f , f ).

(3.3)

Identity (3.1) and inequality (3.3) suggest to estimate s(L, N ) by induction on L. The
hardest thing is to make appropriate estimates on the term EntνΛN (νΛN [f |η Λ1 ]). Note that
this term is the entropy of a function depending only on the number of particles in Λ 1 .

3.2

Step 2: logarithmic Sobolev inequality for the distribution of
the number of particles in Λ1

Let
γΛN (n) = γ(n) := νΛN [η Λ1 = n].
γ(·) is a probability measure on {0, 1, . . . , N } that is reversible for the birth and death process
with generator
¸
·
¸
·
γ(n − 1)
γ(n + 1)
∧ 1 (ϕ(n + 1) − ϕ(n)) +
∧ 1 (ϕ(n − 1) − ϕ(n))
(3.4)
Aϕ(n) =
γ(n)
γ(n)
and Dirichlet form
D(ϕ, ϕ) = −hϕ, AϕiL2 (γ) =

N
X
n=1

[γ(n) ∧ γ(n − 1)](ϕ(n) − ϕ(n − 1))2 .

Logarithmic Sobolev inequalities for birth and death processes are studied in [8]. The nontrivial proof that conditions in [8] are satisfied by γ(n), leads to the following result.
530

Proposition 3.1 The Markov chain with generator (3.4) has a logarithmic Sobolev constant
proportional to N , i.e. there exists a constant C > 0 such that for all ϕ ≥ 0
√ √
Entγ (ϕ) ≤ CN D( ϕ, ϕ).
We now apply Proposition 3.1 to the second summand of the r.h.s. of (3.1), and we
obtain
¸2
·q
N
q
X
N
νΛN [f |η Λ1 = n] − νΛN [f |η Λ1 = n − 1]
[γ(n) ∧ γ(n − 1)]
EntνΛN (νΛ [f |η Λ1 ]) ≤ CN
n=1

N
X

¡ N
¢2
γ(n) ∧ γ(n − 1)
νΛ [f |η Λ1 = n] − νΛN [f |η Λ1 = n − 1] , (3.5)
N
= n] ∨ νΛ [f |η Λ1 = n − 1]

2

, x, y > 0.
where we have used the inequality ( x − y)2 ≤ (x−y)
x∨y
≤ CN

3.3

ν N [f |η Λ1
n=1 Λ

Step 3: study of the term νΛN [f |η Λ1 = n] − νΛN [f |η Λ1 = n − 1]

One of the key points in the proof of Theorem 2.1 consists in finding the “right” representation for the discrete gradient νΛN [f |η Λ1 = n] − νΛN [f |η Λ1 = n − 1], that appears in the r.h.s.
of (3.5).
Proposition 3.2 For every f and every n = 1, 2, . . . , N we have

 
¯
¯
γ(n − 1) 1  N  X

¯
h(η
)c(η
)∂
f
ν
νΛN [f |η Λ1 = n] − νΛN [f |η Λ1 = n − 1] =
 Λ 
x
y yx ¯η Λ1 = n − 1
¯
γ(n) nL
x∈Λ
1
y∈Λ2




¯
¯
 X
¯

+νΛN f,
h(ηx )c(ηy )¯η Λ1 = n − 1 ,
¯
x∈Λ

(3.6)

1
y∈Λ2

where

h(n) :=

n+1
.
c(n + 1)

Moreover, by exchanging the roles of Λ1 and Λ2 , the r.h.s. of (3.6) can be equivalently written
as, for every n − 0, 1, . . . , N − 1,

 
¯
¯
γ(N − n)
1

 N X
¯

h(ηy )c(ηx )∂xy f ¯η Λ1 = n
 νΛ 
¯
γ(N − n + 1) (N − n + 1)L
x∈Λ
1
y∈Λ2


¯
¯

 X
¯
+ νΛN f,
h(ηy )c(ηx )¯η Λ1 = n . (3.7)
¯
x∈Λ


1
y∈Λ2

531

The representations (3.6) and (3.7) will be used for n ≥
convenience, we rewrite (3.6) and (3.7) as

N
2

and n <

N
2

respectively. For

νΛN [f |η Λ1 = n] − νΛN [f |η Λ1 = n − 1] =: A(n) + B(n),

(3.8)

where

A(n) :=

and

B(n) :=
























¯
#
¯
¯
γ(n−1) 1 N
ν
for n ≥
x∈Λ1 h(ηx )c(ηy )∂yx f ¯η Λ1 = n − 1
γ(n) nL Λ
y∈Λ2
¯
¯
#
"
¯
P
¯
γ(N −n)
1
for n <
− γ(N
νN
x∈Λ1 h(ηy )c(ηx )∂xy f ¯η Λ1 = n
−n+1) (N −n+1)L Λ
y∈Λ2
¯
"

P

¯
#
¯
¯
γ(n−1) 1 N
ν f, x∈Λ1 h(ηx )c(ηy )¯η Λ1 = n − 1
for n ≥
γ(n) nL Λ
y∈Λ2
¯
¯
#
"
¯
P
¯
γ(N −n)
1
− γ(N
ν N f, x∈Λ1 h(ηy )c(ηx )¯η Λ1 = n for n <
−n+1) (N −n+1)L Λ
y∈Λ2
¯
"

P

N
2

(3.9)
N
,
2

N
2

(3.10)
N
.
2

Thus, our next aim is to get estimates on the two terms in the r.h.s. of (3.8). It is useful
to stress that the two terms are qualitatively different. Estimates on A(n) are essentially
insensitive to the precise form of c(·). Indeed, the dependence of A(n) on L and N is of the
same order as in the case c(n) = λn, i.e. the case of independent particles. Quite differently,
the term B(n) vanishes in the case of independent particles,
since,
¤ in that case, the term
£
P
N
=
n

1
. Thus, B(n) somewhat
h(η
)c(η
)
is
a.s.
constant
with
respect
to
ν
·|η
x∈Λ1
x
y
Λ1
Λ
y∈Λ2

depends on interaction between particles. Note that our model is not necessarily a “small
perturbation” of a system of independent particles; there is no small parameter in the model
that guarantees that B(n) is small enough. Essentially all technical results of this paper are
concerned with estimating B(n).

3.4

Step 4: estimates on A(n)

The following proposition gives the key estimate on A(n).
Proposition 3.3 There is a constant C > 0 such that
A2 (n) ≤

¢
CL2 ¡ N
νΛ [f |η Λ1 = n] ∨ νΛN [f |η Λ1 = n − 1]
N
·
¸
p p
p p
γ(n − 1)
×
EνΛN [·|ηΛ =n−1] ( f , f ) + EνΛN [·|ηΛ =n] ( f , f ) .
1
1
γ(n)

Remark 3.4 Let us try so see where we are now. Let us ignore, for the moment the term
B(n), i.e. let us pretend that B(n) ≡ 0. Thus, by (3.8) and Proposition 3.3 we would have

532

¡

¢2 CL2 ¡ N
¢
νΛN [f |η Λ1 = n] − νΛN [f |η Λ1 = n − 1] ≤
νΛ [f |η Λ1 = n] ∨ νΛN [f |η Λ1 = n − 1]
N
·
¸
p p
p p
γ(n − 1)
×
EνΛN [·|ηΛ =n−1] ( f , f ) + EνΛN [·|ηΛ =n] ( f , f ) (3.11)
1
1
γ(n)
Inserting (3.11) into (3.5) we get, for some possibly different constant C,
p p
EntνΛN (νΛN (f |η Λ1 )) ≤ CL2 EνΛN ( f , f ),

where we have used the obvious identity:
h
p p
p p i
νΛN EνΛN [·|G] ( f , f ) = EνΛN ( f , f ),

(3.12)

(3.13)

that holds for any σ-field G. Inequality (3.12), together with (3.1) and (3.3) yields
s(2L, N ) ≤ s(L, N ) + CL2 .

(3.14)

sup s(2, N ) < +∞,

(3.15)

Thus, if we can show that
N

(see Proposition 3.7 next) then we would have
sup s(L, N ) ≤ CL2
N

for some C > 0, i.e. we get the exact order of growth of s(L, N ). In all this, however, we
have totally ignored the contribution of B(n).

3.5

Step 5: preliminary analysis of B(n)

We confine ourselves to the analysis of B(n) for n ≥ N2 , since the case n < N2 is identical.
Consider the covariance that appears in the r.h.s. of the first formula of (3.10). By elementary
η
N −η
properties of the covariance and the fact that νΛN [·|η Λ1 ] = νΛΛ1 1 ⊗ νΛ2 Λ1 , we get
¯

¯
"
"
##
¯
X
X
X


¯
h(ηx )νΛN2−n+1 f,
c(ηy )
νΛN  f,
h(ηx )c(ηy )¯ η Λ1 = n − 1 = νΛn−1
1
¯
x∈Λ1
x∈Λ1
y∈Λ2
¯
y∈Λ2
"
#
#
"
X
X
n−1
N −n+1
N −n+1
+ ν Λ 1 νΛ 2
[f ],
h(ηx ) νΛ2
c(ηy ) . (3.16)


x∈Λ1

y∈Λ2

It follows by Conditions (LG) and (M) (see (2.4)) that, for some constant C > 0, h(η x ) ≤ C
and c(ηy ) ≤ Cηy . Thus, a simple estimate on the two summands in (3.16), yields, for some
533

C>0


"
#2
X
C
γ
(n

1)
 νΛN −n+1 νΛn−1 [f ],
B 2 (n) ≤
c(ηy ) +
1
γ 2 (n)
n2 2
y∈Λ
2

2

2

+

"

C(N − n + 1) n−1 N −n+1
νΛ 1 νΛ 2
[f ],
n2 L2
x∈Λ

Thus, our next aim is to estimate the two covariances in (3.17).

3.6

X

1

#2 
h(ηx )  . (3.17)

Estimates on B(n): entropy inequality and estimates on moment generating functions

By (3.17), estimating B 2 (n) consists in estimating two covariances. In general, covariances
can be estimated by the following entropy inequality, that holds for every probability measure
ν (see [1], Section 1.2.2):
ν[f, g] = ν[f (g − ν[g])] ≤

£
¤ 1
ν[f ]
log ν et(g−ν[g]) + Entν (f ),
t
t

(3.18)

where f ≥ 0 and t > 0 is arbitrary. Since in (3.17) we need to estimate the square of a
covariance, we write (3.18) with −g in place of g, and obtain

¡ £
¤
£
¤¢ 1
ν[f ]
log ν et(g−ν[g]) ∨ ν e−t(g−ν[g]) + Entν (f ).
(3.19)
t
t
£
¤
Therefore, we first get estimates on the moment generating functions ν e±t(g−ν[g]) , and then
optimize (3.19) over t > 0.
Note that the covariances in (3.17) involve functions of ηΛ1 or ηΛ2 . In next two propositions we write Λ for Λ1 and Λ2 , and denote by N the number of particles in Λ. Their proof
can be found in [3].
|ν[f, g]| ≤

Proposition 3.5 Let x ∈ Λ. Then there is a constant AL depending on L = |Λ| such that
for every N > 0 and t ∈ [−1, 1]
i
h
N
2
(3.20)
νΛN et(c(ηx )−νΛ [c(ηx )]) ≤ eAL N t ,

and

i
h

N
2
νΛN etN (h(ηx )−νΛ [h(ηx )]) ≤ AL eAL (N t + N |t|) ,

(3.21)

Using (3.18), (3.19) and (3.21) and optimizing over t > 0, we get estimates on the covariances
appearing in (3.17).
Proposition 3.6 There exists a constant CL depending on L = |Λ| such that the following
inequalities hold:
"
#2
X
νΛN f,
(3.22)
c(ηx ) ≤ CL N νΛN [f ] EntνΛN (f ),
x∈Λ

534

νΛN

"

f,

X

#2

h(ηx )

x∈Λ



CL N −1 νΛN [f ]

h

νΛN [f ]

i

+ EntνΛN (f ) .

(3.23)

Inserting these new estimates in (3.17) we obtain, for some possibly different CL ,
Ã
h
i
2
(n−1) N
n−1
N
−n+1
B 2 (n) ≤ CLγγ2 (n)
Ent
(f
)
νΛ [f |η Λ1 = n − 1] N −n+1
ν
Λ1
νΛ
n2
2
Ã
!!
(3.24)
h
i
2
(N −n+1)
N −n+1
N
νΛ [f |η Λ1 = n − 1] + νΛ2
+ n3
.
EntνΛn (f )
1

In order to simplify (3.24), we use Proposition 7.5, which gives
γ(n − 1)
Cn
n


C(N − n + 1)
γ(n)
N −n+1

(3.25)

for some C > 0. It follows that
γ 2 (n − 1) N − n + 1
γ(n − 1) C 3
C2

,

γ 2 (n)
n2
N −n+1
γ(n) n
and

C2
γ 2 (n − 1) (N − n + 1)2

.
γ 2 (n)
n3
n

Thus, (3.24) implies, for some CL > 0 depending on L, recalling also that n ≥
N

νΛN [f |η Λ1

N
,
2

³
γ(n) ∧ γ(n − 1)
2
B (n) ≤ CL γ(n − 1) νΛN [f |η Λ1 = n − 1]
= n] ∨ νΛN [f |η Λ1 = n − 1]
h

h
i
N −n+1
n
N −n+1 (f ) + ν
Ent
(f
)
. (3.26)
Ent
+ νΛn−1
νΛ
Λ2
ν
1
Λ2

1

Now, we bound the two terms

Entν n−1 (f ) and Entν N −n+1 (f )
Λ1

Λ2

by, respectively,
p p
p p
s(N, L)Eν n−1 ( f , f ) and s(N, L)Eν N −n+1 ( f , f ),
Λ1

Λ2

and insert these estimates in (3.6). What comes out is then used to estimate (3.5), after
having obtained the corresponding estimates for n < N2 . Recalling the estimates for A(n),
straightforward computations yield
p p
p p
(3.27)
EntνΛN (νΛN [f |η Λ1 ]) ≤ CL EνΛN ( f , f ) + CL νΛN [f ] + CL s(N, L)EνΛN ( f , f ),

for some CL > 0 depending on L. Inequality (3.27), together with (3.3), gives, for a possibly
different constant CL > 0
p p
p p
EntνΛN (f ) ≤ CL EνΛN ( f , f ) + CL νΛN [f ] + CL s(N, L)EνΛN ( f , f ).
(3.28)
535

To deal with the term νΛN [f ] in (3.28) we use the following well known argument. Set
¡√
√ ¢2
f=
f − νΛN [ f ] . By Rothaus inequality (see [1], Lemma 4.3.8)
p p
EntνΛN (f ) ≤ EntνΛN (f ) + 2νΛN [ f , f ].

Using this inequality and replacing f by f in (3.28) we get, for a different C L ,
p p
p p
p p
EntνΛN (f ) ≤ CL EνΛN ( f , f ) + CL νΛN [ f , f ] + CL s(N, L)EνΛN ( f , f )
p p
p p
≤ DL EνΛN ( f , f ) + DL s(N, L)EνΛN ( f , f )
(3.29)
where, in the last line, we have used the Poincar´e inequality (1.3). Therefore
s(2L, N ) ≤ DL [s(L, N ) + 1],

(3.30)

that implies (2.6), provided we prove the following “basis step” for the induction.
Proposition 3.7
sup s(2, N ) < +∞
N

The proof of Proposition 3.7 is also given in Section 8. As we pointed out above, (3.30) gives
no indication on how s(L) = supN s(N, L) grows with L. The proof of the actual L2 -growth
is given in [3].

4

Representation of νΛN [f |η Λ1 = n] − νΛN [f |η Λ1 = n − 1]:
proof of Proposition 3.2

We begin with a simple consequence of reversibility
Lemma 4.1 Let Λ = Λ1 ∪ Λ2 be a partition of Λ, N ∈ N. Define γ(n) := νΛN [η Λ1 = n] and
τxy f := ∂xy f + f . Then:
¯
i
h

¯
c(ηy )
γ(n−1) N

η
=
n

1
for any x ∈ Λ1 and y ∈ Λ2
ν
τ
f
¯

Λ1
 γ(n) Λ c(ηx +1) yx
νΛN [f 1(ηx > 0)|η Λ1 = n] =
¯
h
i


ν N c(ηy ) τ f ¯¯ η = n
for any x, y ∈ Λ
Λ

c(ηx +1) yx

Λ1

for any f ∈ L1 (νΛN ) and n ∈ {1, . . . , N }.

536

2

Proof. Assume that y ∈ Λ2 . We use first the detailed balance condition (2.3), then the
change of variables ξ 7→ ξ + δy − δx to obtain
P
N
£
¤
ξ:ξx >0 νΛ [η = ξ]f (ξ)1(ξ Λ1 = n)
N
νΛ f 1(ηx > 0)|η Λ1 = n =
νΛN [¯
ηΛ1 = n]
P
c(ξy +1)
N
ξ:ξx >0 νΛ [η = ξ − δx + δy ] c(ξx ) f (ξ)1(ξ Λ1 = n)
=
νΛN [η Λ1 = n]
¯
i
h

¯
c(ηy )
γ(n−1) N

η
=
n

1
if x ∈ Λ1
ν
τ
f
¯

Λ1
 γ(n) Λ c(ηx +1) yx
=
¯
h
i


ν N c(ηy ) τ f ¯¯ η = n
if x ∈ Λ
Λ

c(ηx +1) yx

2

Λ1

Lemma 4.2 Let Λ = Λ1 ∪ Λ2 be a partition of Λ, N ∈ N. Define γ(n) := νΛN [η Λ1 = n] and
h : N → R by h(n) := (n + 1)/c(n + 1). Then:
νΛN [f |η Λ1 = n] − νΛN [f |η Λ1 = n − 1]
¯
¯
"
#
#!
à "
¯
¯
X
X
γ(n − 1)
¯
¯
h(ηx )¯η Λ1 = n − 1
=
,
h(ηx )∂yx f ¯η Λ1 = n − 1 + νΛN f, c(ηy )
νΛN c(ηy )
¯
¯
nγ(n)
x∈Λ
x∈Λ
1

1

(4.1)

for any f ∈ L1 (νΛN ), y ∈ Λ2 and n ∈ {1, . . . , N }.

Proof. We begin by writing
¯
¯
¤ 1 X N£
¤
1 X N£
νΛN [f |η Λ1 = n] = −
νΛ ηx ∂xy f ¯η Λ1 = n +
νΛ ηx τxy f ¯η Λ1 = n .
n x∈Λ
n x∈Λ
1

1

By Lemma 4.1 we get

¯
#
"
¯
¯
¤
£
γ(n

1)
c(η
)
¯
y
N
N
− νΛ ηx ∂xy f ¯η Λ1 = n =
ν
(ηx + 1)∂yx f ¯η Λ1 = n − 1
¯
γ(n) Λ c(ηx + 1)
¯
¤
γ(n − 1) N £
=
νΛ c(ηy )h(ηx )∂yx f ¯η Λ1 = n − 1 ,
γ(n)

and similarly

Thus

¯
¯
¤ γ(n − 1) N £
¤
£
νΛN ηx τxy f ¯η Λ1 = n =
νΛ c(ηy )h(ηx )f ¯η Λ1 = n − 1 .
γ(n)

νΛN [f |η Λ1 = n]
¯
¯
à "
"
#
#!
¯
¯
X
X
γ(n − 1)
¯
¯
N
N
h(ηx )∂yx f ¯η Λ1 = n − 1 + νΛ c(ηy )f
νΛ c(ηy )
h(ηx )¯η Λ1 = n − 1
=
.
¯
¯
nγ(n)
x∈Λ
x∈Λ
1

1

537

Letting f ≡ 1 in this last formula we obtain
¯
#
"
¯
X
γ(n − 1) N
¯
ν c(ηy )
h(ηx )¯η Λ1 = n − 1 = 1
¯
nγ(n) Λ
x∈Λ
1

that implies

¯
#
"
¯
X
γ(n − 1) N
¯
h(ηx )¯η Λ1 = n − 1
ν c(ηy )f
¯
nγ(n) Λ
x∈Λ
1

¯
#
"
¯
X
¤
£
γ(n

1)
¯
= νΛN f |η Λ1 = n − 1 +
h(ηx )¯η Λ1 = n − 1
ν N f, c(ηy )
¯
nγ(n) Λ
x∈Λ
1

and, therefore,

¤
£
νΛN [f |η Λ1 = n] − νΛN f |η Λ1 = n − 1
¯
¯
"
#
#!
à "
¯
¯
X
X
γ(n − 1)
¯
¯
h(ηx )¯η Λ1 = n − 1
.
=
h(ηx )∂yx f ¯η Λ1 = n − 1 + νΛN f, c(ηy )
νΛN c(ηy )
¯
¯
nγ(n)
x∈Λ
x∈Λ
1

1

Proof of Proposition 3.2. Equation (3.6) is obtained by (4.1) by averaging over y ∈ Λ 2 .

5

Bounds on A(n): proof of Proposition 3.3

Proof of Proposition 3.3. The Proposition is proved if we can show that for 1 ≤ n < N/2
A2 (n) ≤

¢
CL2 ¡ N
νΛ [f |η Λ1 = n] ∨ νΛN [f |η Λ1 = n − 1]
N −n
¸
·
p p
p p
γ(n − 1)
×
EνΛN [·|ηΛ =n−1] ( f , f ) + EνΛN [·|ηΛ =n] ( f , f ) ,
1
1
γ(n)

while for n ≥ N/2
A2 (n) ≤

¢
CL2 ¡ N
νΛ [f |η Λ1 = n] ∨ νΛN [f |η Λ1 = n − 1]
n
·
¸
p p
p p
γ(n − 1)
×
EνΛN [·|ηΛ =n−1] ( f , f ) + EνΛN [·|ηΛ =n] ( f , f ) . (5.1)
1
1
γ(n)

538

We will prove only this last bound being
√ the
√ the previous one identical. So assume
√ proof of
n ≥ N/2 and notice that ∂yx f = (∂yx f )( f + τyx f ). By Cauchy-Schwartz inequality
(

¯
#)2
¯
¯
c(ηy )h(ηx )∂yx f ¯η Λ1 = n − 1
¯
x∈Λ1 , y∈Λ2
¯
#
"
·
¸2
³ p ´2 ¯
X
γ(n − 1)
¯
c(ηy )h(ηx ) ∂yx f ¯η Λ1 = n − 1

νΛN
¯
nLγ(n)
x∈Λ1 , y∈Λ2
¯
"
#
³p
X
p ´2 ¯¯
N
× νΛ
c(ηy )h(ηx )
f + τyx f ¯η Λ1 = n − 1
¯
x∈Λ1 , y∈Λ2
¯
"
#
¸2
·
³ p ´2 ¯
X
γ(n

1)
¯
νΛN
c(ηy ) ∂yx f ¯η Λ1 = n − 1
≤ 2a2
¯
nLγ(n)
x∈Λ1 , y∈Λ2
¯
"
#
¯
X
¯
× νΛN
c(ηy ) (f + τyx f ) ¯η Λ1 = n − 1 (5.2)
¯

γ(n − 1) N
ν
nLγ(n) Λ

"

X

x∈Λ1 , y∈Λ2

where a := khk+∞ ((2.4) implies boundedness of h). In order to bound the last factor in
(5.2) we observe that by Lemma 4.1

so that
"
νΛN

¯
¤
£
νΛN c(ηy )τyx f ¯η Λ1 = n − 1 =

¯
£
¤
γ(n)
νΛN c(ηx )f ¯η Λ1 = n ,
γ(n − 1)

¯
#
¯
¯
c(ηy ) (f + τyx f ) ¯η Λ1 = n − 1
¯
x∈Λ1 , y∈Λ2
X µ £
¯
¤
νΛN c(ηy )f ¯η Λ1 = n − 1 +
=
X

x∈Λ1 , y∈Λ2

¯
£
¤
γ(n)
νΛN c(ηx )f ¯η Λ1 = n
γ(n − 1)



¾
½
£ ¯
¤
¤
£ ¯
γ(n)nL
N
N
ν f ¯η Λ 1 = n ,
≤ a (N − n + 1)LνΛ f ¯η Λ1 = n − 1 +
γ(n − 1) Λ

where the fact c(k) ≤ ak (see (2.4)) has been used to obtain the last line. This implies that
γ(n − 1) N
ν
γ(n)nL Λ

"

X

x∈Λ1 , y∈Λ2

≤a

½

¯
#
¯
¯
c(ηy ) (f + τyx f ) ¯η Λ1 = n − 1
¯

¾
¤
¤
£ ¯
γ(n − 1)(N − n + 1)L N £ ¯¯
N
¯
νΛ f η Λ 1 = n − 1 + ν Λ f η Λ 1 = n
γ(n)nL
¡ £ ¯
¤
¤¢
£ ¯
≤ B1 νΛN f ¯η Λ1 = n − 1 ∨ νΛN f ¯η Λ1 = n

539

where in last step we used Proposition 7.5. By plugging this bound in (5.2) we get
(

γ(n − 1) N
ν
nLγ(n) Λ

"

¯
#)2
¯
¯
c(ηy )h(ηx )∂yx f ¯η Λ1 = n − 1
¯
x∈Λ1 , y∈Λ2
¯
¡ £
¤
¤¢
£ ¯
≤ B2 νΛN f ¯η Λ1 = n − 1 ∨ νΛN f ¯η Λ1 = n
¯
"
#
³ p ´2 ¯
X
γ(n − 1) N
¯
ν
c(ηy )h(ηx ) ∂yx f ¯η Λ1 = n − 1 . (5.3)
×
¯
nLγ(n) Λ x∈Λ , y∈Λ
X

1

2

Now we have to bound the last factor in the right hand side of (5.3). For x, y ∈ Z the path
between x and y will be the sequence of nearest-neighbor integers γ(x, y) = {z0 , z1 , . . . , zr }
of Z such that for i = 1, . . . , r |zi−1 − zi | = 1, for i 6= j zi 6= j, z0 = x and zr = y. Obviously,
for x, y ∈ Λ, we have r = |γ(x, y)| ≤ 2L. Now let such a γ(x, y) be given. Notice that if
ηy ≥ 1, we can write
r−1 ³
X
p
p ´
(∂yx f )(η) =
∂zk+1 zk f (η − δzr + δzk+1 ).
k=0

By Jensen inequality, we obtain
r−1 ³
X
p ´2
p ´2
c(ηy ) ∂yx f (η) ≤ 2Lc(ηy )
∂zk+1 zk f (η − δzr + δzk+1 ),

³

so

k=0

¯
#
¯
´
³
p
2
γ(n − 1) N
¯
νΛ
c(ηy )h(ηx ) ∂yx f ¯η Λ1 = n − 1
¯
γ(n)nL
x∈Λ1 ,y∈Λ2
¯
#
"
r−1
³
p ´2 ¯¯
2γ(n − 1) X X N

ν c(ηy )h(ηx ) τzr zk+1 ∂zk+1 zk f ¯η Λ1 = n − 1
¯
γ(n)n x∈Λ ,y∈Λ k=0 Λ
1
2
¯
#
"
r−1
¯
´
³
X
X
p
2
2aγ(n − 1)
¯
N

νΛ c(ηy ) τyzk+1 ∂zk+1 zk f ¯η Λ1 = n − 1 ,
¯
γ(n)n
"

X

x∈Λ1 ,y∈Λ2 k=0

where again a = khk+∞ . By Lemma 4.1
νΛN

"

¯
#
³
p ´2 ¯¯
c(ηy ) τyzk+1 ∂zk+1 zk f ¯η Λ1 = n − 1
¯
i
h

¡
√ ¢2 ¯¯
γ(n)
N

f
η
=
n
if zk+1 ∈ Λ1
ν
c(η
)

¯ Λ1
zk+1
zk+1 zk

 γ(n−1) Λ
=
i
h

¡
√ ¢2 ¯¯

ν N c(η
)

f
η
=
n

1
if zk+1 ∈ Λ2
¯
zk+1
zk+1 zk
Λ1
Λ
540

which implies
¯
#
³ p ´2 ¯
¯
c(ηy )h(ηx ) ∂yx f ¯η Λ1 = n − 1
¯
x∈Λ1 ,y∈Λ2
¯
"
(
#
³
´2 ¯
X
p
2a X
¯
ν N c(ηzk+1 ) ∂zk+1 zk f ¯η Λ1 = n − 1 +

¯
n x∈Λ ,y∈Λ k:z ∈Λ Λ
1
2
1
k+1
¯
#)
"
¯
´
³
X
p
2
γ(n − 1)
¯
νΛN c(ηzk+1 ) ∂zk+1 zk f ¯η Λ1 = n − 1
¯
γ(n) k:z ∈Λ
2
k+1
¯
( "
#
³ p ´2 ¯
2aL2 X
¯

νΛN c(ηx ) ∂xy f ¯η Λ1 = n +
¯
n

γ(n − 1) N
ν
nLγ(n) Λ

"

X

x,y∈Λ
|x−y|=1

¯
#)
"
³ p ´2 ¯
γ(n − 1) N
¯
.
ν c(ηx ) ∂xy f ¯η Λ1 = n − 1
¯
γ(n) Λ

By plugging this bound in (5.3) we get (5.1).

6

Rough estimates: proofs of Propositions 3.5 and 3.6

Proof of Proposition 3.5. We begin with the proof of (3.20). Denote by ΩN
Λ the set of
configurations having N particles in Λ. Then
νΛN [{η}] =

Y 1
1(η ∈ ΩN
Λ)
,
c(η
ZΛN
x )!
x∈Λ

where ZΛN is a normalization factor. It is therefore easily checked that
c(ηx )νΛN [{η}]

ZΛN −1 N −1
=
ν
[{η − δx }]
ZΛN Λ

(6.1)

for any x ∈ Λ. Thus, for every function f ,
νΛN [c(ηx )f ] =

ZΛN −1 N −1
ν
[σx f ],
ZΛN Λ

(6.2)
N −1

N


where σx f (η) = f (η + δx ). Letting f ≡ 1 in (6.2) we have
rewrite (6.2) as
νΛN [c(ηx )f ] = νΛN [c(ηx )]νΛN −1 [σx f ].
Now, let t > 0, and define

¤
£
ϕ(t) = νΛN etc(ηx ) .
541

= νΛN [c(ηx )], and so we
(6.3)

We now estimate ϕ(t) by means of the so-called Herbst argument (see [1], Section 7.4.1). By
direct computation, Jensen’s inequality and (6.3), we have
¤
£
¤
£
¤
£
tϕ′ (t) − ϕ(t) log ϕ(t) = tνΛN c(ηx )etc(ηx ) − νΛN etc(ηx ) log νΛN etc(ηx )
£
¤
£
¤
≤ tνΛN c(ηx )etc(ηx ) − tνΛN [c(ηx )]νΛN etc(ηx )
X
£
¤
£
¤¢
¡
(6.4)
= t
νΛN [c(ηx )] νΛN −1 etc(ηx +1) − νΛN etc(ηx ) .
x∈Λ

We now claim that, for every x ∈ Λ and 0 ≤ t ≤ 1
¯ N −1 £ tc(η +1) ¤
£ tc(ηx ) ¤¯
£
¤
N
x
¯ν
¯ ≤ CL tνΛN etc(ηx ) ,
e

ν
e
Λ
Λ

(6.5)

for some constant CL depending on L but not on N . For the moment, let us accept (6.5), and
show how it is used to complete the proof. By (6.4), (6.5) and the fact that ν ΛN [c(ηx )] ≤ CN/L
we get
tϕ′ (t) − ϕ(t) log ϕ(t) ≤ CL N t2 ϕ(t),
for 0 ≤ t ≤ 1 and some possibly different CL . Equivalently, letting ψ(t) =

log ϕ(t)
,
t

ψ ′ (t) ≤ CL N.

(6.6)

Observing that limt↓0 ψ(t) = νΛN [c(ηx )], by (6.6) and Gronwall lemma we have
ψ(t) ≤ νΛN [c(ηx )] + CL N t,
from which (3.20) easily follows, for t ≥ 0.
We now prove (6.5). We shall use repeatedly the inequality
|ex − ey | ≤ |x − y|e|x−y| ex ,

(6.7)

which holds for x, y ∈ R. Using (6.7) and the fact that, by condition (LG), kc(η x + 1) −
c(ηx )k∞ ≤ a1 , we have
¯ N −1 £ tc(η +1) ¤
£ tc(ηx ) ¤¯
£
¤
N −1
x
¯ν
¯ ≤ a1 tea1 t ν N −1 etc(ηx ) ,
e

ν
e
Λ
Λ
Λ

from which we have that (6.5) follows if we show
¯ N −1 £ tc(η ) ¤
£
¤¯
£
¤
¯ν
e x − νΛN etc(ηx ) ¯ ≤ CL tνΛN etc(ηx ) .
Λ

(6.8)

At this point we use the notion of stochastic order between probability measures.
For
R
R two
probabilities µ and ν on a partially ordered space X, we say that ν ≺ µ if f dν ≤ f dµ
for every integrable, increasing f . This is equivalent to the existence of a monotone coupling
of ν and µ, i.e. a probability P on X × X supported on {(x, y) ∈ X × X : x ≤ y}, having
marginals ν and µ respectively (see e.g. [6], Chapter 2).
As we will see, (6.8) would not be hard to prove if we had νΛN −1 ≺ νΛN . Under assumptions
(LG) and (M) a slightly weaker fact holds, namely that there is a constant B > 0 independent

of N and L such that if N ≥ N ′ + BL then νΛN ≺ νΛN (see [5], Lemma 4.4). In what follows,
542

we may assume that N > BL. Indeed, in the case N ≤ BL there is no real dependence on
N , and (3.20) can be proved by observing that
νΛN

h

N [c(η )])
t(c(ηx )−νΛ
x

e

i

≤1+

t2 N
N
2
2
νΛ [c(ηx ), c(ηx )]e2t sup{c(ηx ):η∈ΩΛ } ≤ 1 + CL t2 ≤ eCL t ≤ eCL N t
2

as 1 ≤ N ≤ BL ((3.20) is trivial for N = 0). For N ≥ BL we use the rough inequality

¯ N −1 £ tc(η ) ¤
¤
£
¤¯
£
¤¯ ¯
£
¯ν
e x − νΛN etc(ηx ) ¯ ≤ ¯νΛN −1 etc(ηx ) − νΛN −1−BL etc(ηx ) ¯
Λ
¯ £
¤
¤¯
£
+ ¯νΛN etc(ηx ) − νΛN −1−BL etc(ηx ) ¯ . (6.9)

Denote by Q the probability measure on NΛ × NΛ that realizes a monotone coupling of νΛN −1
and νΛN −1−BL . In other words, Q has νΛN −1 and νΛN −1−BL as marginals, and
Q[{(η, ξ) : ηx ≥ ξx ∀x ∈ Λ}] = 1.

(6.10)

Using again (6.7)
¯ £
¯ N −1 £ tc(η ) ¤
¤¯
£ tc(ηx ) ¤¯
N −1−BL
x
¯ = ¯Q etc(ηx ) − etc(ξx ) ¯
¯ν
e

ν
e
Λ
Λ
£
¤
≤ tQ |c(ηx ) − c(ξx )|et|c(ηx )−c(ξx )| etc(ηx ) . (6.11)
Since, Q-a.s., ηx ≥ ξx ∀x ∈ Λ and η Λ = ξ Λ + BL, then necessarily ηx ≤ ξx + BL ∀x ∈ Λ.
Thus, it follows that, for some C > 0,
|c(ηx ) − c(ξx )| ≤ CL Q-a.s.
Inserting this inequality in (6.11), we get, for t ≤ 1 and some CL > 0,
¯ N −1 £ tc(η ) ¤
£ tc(ηx ) ¤¯
¤
£
N −1−BL
x
¯ν
¯ ≤ CL tν N −1 etc(ηx ) .

ν
e
e
Λ
Λ
Λ

With the same arguments, it is shown that
¯ N £ tc(η ) ¤
£
¤¯
£
¤
¯νΛ e x − ν N −1−BL etc(ηx ) ¯ ≤ CL tνΛN etc(ηx ) .
Λ

In order to put together (6.12) and (6.13), we need to show that, for 0 ≤ t ≤ 1,
£
¤
£
¤
νΛN −1 etc(ηx ) ≤ CL νΛN etc(ηx ) ,

(6.12)

(6.13)

(6.14)

for some L-dependent CL . By condition (LG) and (6.3) we have
£
¤
£
¤
νΛN −1 etc(ηx ) ≤ eta1 νΛN −1 etc(ηx +1) =

£
¤
£
¤
eta1
νΛN c(ηx )etc(ηx ) ≤ CL νΛN etc(ηx ) ,
N
νΛ [c(ηx )]

where, for the last step, we have used the facts that, for some ǫ > 0, νΛN [c(ηx )] ≥ ǫ NL and
c(ηx ) ≤ ǫ−1 N νΛN -a.s. (for both inequalities we use (2.4)).
The proof (3.20) is now completed for the case t ≥ 0. For t < 0, it is enough to observe
that our argument is insensitive to replacing c(·) with −c(·).
543

We now prove (3.21) The idea is to use the fact that the tails of N (h(ηx ) − νΛN [h(ηx )])
are not thicker than those of c(ηx ) − νΛN [c(ηx )]. Similarly to (6.1), note that
h(ηx )νΛN [{η}]

ZΛN +1
=
(ηx + 1)νΛN +1 [{η + δx }],
N


(6.15)

ZΛN +1 N +1
ν
[ηx f (η − δx )].
ZΛN Λ

(6.16)

which implies
νΛN [h(ηx )f ] =
Letting f ≡ 1 in (6.16) we obtain

L
ZΛN +1
= νΛN [h(ηx )]
N
N +1

that, together with the previously obtained identity
νΛN [h(ηx )] =

N −1

N


N +1
LνΛN +1 [c(ηx )]

= νΛN [c(ηx )], yields
.

(6.17)

Thus
¯
¯
¯ ηx + 1
¯
N
+
1
¯

|h(ηx ) −
= ¯¯
N +1
c(ηx + 1) LνΛ [c(ηx )] ¯
¯
¯¸
·
¯
¯
¯
¯
N
+
1
1
N
+1
¯c(ηx + 1) − ν
¯ + c(ηx + 1) ¯ηx + 1 −
¯

+
1)
[c(η
)]

x
x
Λ
¯
L ¯
c(ηx + 1)νΛN +1 [c(ηx )]
¯
¯¸
·
¯ ¯
C ¯¯
N + 1 ¯¯
N +1
¯
¯

c(ηx + 1) − νΛ [c(ηx )] + ¯ηx + 1 −
, (6.18)
N
L ¯
νΛN [h(ηx )]|

where, in the last step, we use the fact that νΛN +1 [c(ηx )] ≥ ǫN/L for some ǫ > 0. In (6.18) and
in what follows, the L-dependence of constants is omitted. For ρ > 0, let c(ρ) be obtained
by linear interpolation of c(n), n ∈ N. Observe that, by (1.3) with f (η) = η x ,
νΛN [ηx , ηx ] ≤ CL2 νΛN [c(ηx )] ≤ CL N
for some CL > 0 that depends on L. Thus, by Condition (LG)
¯¸
·¯
q

¯
¯ N
¯
¯
N
N
¯νΛ [c(ηx )] − c(N/L)¯ ≤ c1 νΛ ¯ηx − ¯ ≤ c1 ν N [ηx , ηx ] ≤ C N ,
Λ
¯
¯
L

(6.19)

for some C > 0, possibly depending on L. From (6.18) and (6.19) it follows that, for some
C>0
¯
¯

¯
¯
N
N
N |h(ηx ) − νΛ [h(ηx )]| ≤ C ¯¯ηx − ¯¯ + C N .
(6.20)
L

Moreover, from Condition (M) it follows that there is a constant C > 0 such that
¯
¯
¯
¯
N
¯ηx − ¯ ≤ C (|c(ηx ) − c(N/L)| + 1) .
¯

544

Thus, for every M > 0 there exists C > 0 such that

¯
¯
¯c(ηx ) − νΛN [c(ηx )]¯ ≤ M ⇒ N |h(ηx ) − νΛN [h(ηx )]| ≤ CM + C N .

(6.21)

It follows that, for all M > 0
h
√ i
¯
£¯
¤
νΛN N (h(ηx ) − νΛN [h(ηx )])) > CM + C N ≤ νΛN ¯c(ηx ) − νΛN [c(ηx )]¯ > M

≤ νΛN [c(ηx ) − νΛN [c(ηx )] > M ] + νΛN [νΛN [c(ηx )] − c(ηx ) > M ]. (6.22)

Note that (6.22) is trivially true for M ≤ 0, so it holds for all M ∈ R.
Now, take t ∈ (0, 1]. We have
Z +∞
h
i
N [h(η )])
tN (h(ηx )−νΛ
N
x
= t
etz νΛN [N (h(ηx ) − νΛN [h(ηx )]) > z] dz
νΛ e
−∞
Z +∞
h
√ i
z
N
tz N
≤ t
e νΛ c(ηx ) − νΛ [c(ηx )] > − C N dz +
C
−∞
Z +∞
h
√ i
z
+t
etz νΛN νΛN [c(ηx )] − c(ηx ) > − C N dz
C
−∞
Z +∞

et(CM +C N ) νΛN [c(ηx ) − νΛN [c(ηx )] > M ] dM
= Ct
−∞
Z +∞

et(CM +C N ) νΛN [νΛN [c(ηx )] − c(ηx ) > M ] dM
+ Ct

h
i
h−∞
√ ³
N
N
Ct N
N
= Ce
νΛ eCt(c(ηx )−νΛ [c(ηx )]) + νΛN e−Ct(c(ηx )−νΛ [c(ηx )])
≤ 2CeCt


N AN t2

e

,

where, in the last step, we have used (3.20). This completes the proof for the case t ∈ (0, 1].
For t ∈ [−1, 0) we proceed similarly, after having replaced, in (6.22), N (h(η x ) − νΛN [h(ηx )])
with its opposite.

Proof of Proposition 3.6. We first prove inequality (3.22). Clearly, it is enough to show
that, for all x ∈ Λ,
νΛN [f, c(ηx )]2 ≤ CL N νΛN [f ] EntνΛN (f ),

for some CL > 0
By the entropy inequality (3.19) and (3.20) we have, for 0 < t ≤ 1:
νΛN [f, c(ηx )]2 ≤ 2νΛN [f ]2 A2L N 2 t2 +
Set
t2∗ =

EntνΛN (f )
N νΛN [f ]
545

.

2
Ent2ν N (f ).
Λ
t2

(6.23)

(6.24)

Suppose, first, that t∗ ≤ 1. Then if we insert t∗ in (6.23) we get
νΛN [f, c(ηx )]2 ≤ 2νΛN [f ]A2L N EntνΛN (f ) + 2N νΛN [f ] EntνΛN (f ) =: CL N νΛN [f ] EntνΛN (f ). (6.25)
In the case t∗ > 1, we easily have
νΛN [f, c(ηx )]2 ≤ CνΛN [f ]2 N 2 ≤ CνΛN [f ]2 N 2 t∗ 2 = CN νΛN [f ] EntνΛN (f ).

(6.26)

By (6.25) and (6.26) the conclusion follows.
Let us now prove (3.23). As before, by (3.19) and (3.21), for t ∈ (0, 1]:
N 2 νΛN [f, h(ηx )]2 ≤
Here we set

¤ 2
2νΛN [f ]2 £ 2
2 4
2 2
2
+ 2 Ent2ν N (f ).
N
t
log
A
+
A
t
N
+
A
L
L
L
Λ
t2
t
t2∗ =

EntνΛN (f ) ∨ νΛN [f ]
N νΛN [f ]

(6.27)

,

and proceed as in (6.25) and (6.26).

7

Local limit theorems

The rest of the paper is devoted to the proofs of all technical results that have been used
in previous sections. For the sake of generality, we state and prove all results in dimension
d ≥ 1.
Using the language of statistical mechanics, having defined the canonical measure νΛN ,
for ρ > 0 we consider the corresponding grand canonical measure
η

µρ [{η}] := P

α(ρ)ηΛ νΛΛ [{η}]
ξ

ξΛ Λ
ξ∈ΩΛ α(ρ) νΛ [{ξ}]

=

Y α(ρ)ηx
1
,
Z(α(ρ)) x∈Λ c(ηx )!

(7.1)

where α(ρ) is chosen so that µρ (ηx ) = ρ, x ∈ Λ, and Z(α(ρ)) is the corresponding normalization. Clearly µρ is a product measure with marginals given by (1.2). Monotonicity and the
Inverse Function Theorem for analytic functions, guarantee that α(ρ) is well defined and it
is a analytic function of ρ ∈ [0, +∞). We state here without proof some direct consequences
of Conditions 2.1 and 2.2. The proofs of some of them can be found in [5].
Proposition 7.1
1. Let σ 2 (ρ) := µρ [ηx , ηx ], then
σ 2 (ρ)
σ 2 (ρ)
≤ sup
< +∞
ρ>0
ρ
ρ
ρ>0

0 < inf

(7.2)

2. Let α(ρ) be the function appearing in (7.1); then
0 < inf

ρ>0

α(ρ)
α(ρ)
≤ sup
< +∞
ρ
ρ
ρ>0
546

(7.3)

7.1

Local limit theorems for the grand canonical measure

The next two results are a form of local limit theorems for the density of η Λ under µρ .
Define pρΛ (n) := µρ [η Λ = n] for ρ > 0, n ∈ N and Λ ⊂⊂ Zd . The idea is to get a Poisson
N/|Λ|
approximation of pΛ (n) for very small values of N/|Λ| and to use the uniform local limit
theorem (see Theorem 6.1 in [5]) for the other cases.
Lemma 7.2 For every N0 ∈ N \ {0} there exist A0 > 0 such that
¯
¯
¯ N/|Λ|
N n −N ¯¯ A0
¯
sup
¯pΛ (n) − n! e ¯ ≤ |Λ|
N ∈N\{0}, n∈N
n≤N ≤N0

for any Λ ⊂⊂ Zd .

Proof. Let ρ := N/|Λ| and assume, without loss of generality, that 0 ∈ Λ. Notice that
µρ [η Λ = n]
¯
¯
i h
i h
i
i
h
h
¯
¯
= µρ η Λ = n¯ max ηx ≤ 1 µρ max ηx ≤ 1 + µρ η Λ = n¯ max ηx > 1 µρ max ηx > 1 .
x∈Λ

x∈Λ

x∈Λ

x∈Λ

We begin by proving that

h
i
µρ max ηx > 1 = O(|Λ|−1 ),
x∈Λ

(7.4)

uniformly in 0 < N ≤ N0 .
Indeed
h
i
µρ max ηx > 1 = 1 − (1 − µρ [η0 > 1])|Λ| ,
x∈Λ

and

+∞
+∞
+∞
1 X α(ρ)k
α(ρ)2 X α(ρ)k
α(ρ)2 X c(k)! α(ρ)k
µρ [η0 > 1] =
=
=
.
Z(ρ) k=2 c(k)!
Z(ρ) k=0 c(k + 2)!
Z(ρ) k=0 c(k + 2)! c(k)!

Since c(k)!/c(k + 2)! is uniformly bounded, we have µρ [η0 > 1] ≤ B1 α(ρ)2 = O(|Λ|−2 ),
uniformly in 0 < N ≤ N0 . Thus
¡
¢|Λ|
(1 − µρ [η0 > 1])|Λ| = 1 − O(|Λ|−2 )
= O(|Λ|−1 ),

which establishes (7.4).
Now let

ρe :=

A trivial calculation shows that
ρe =

+∞
X
k=0

¯
i
¯
kµρ η0 = k ¯η0 ≤ 1 .
h

µρ [η0 ] − µρ [η0 1(η0 > 1)]
ρ − µρ [η0 1(η0 > 1)]
µρ [η0 1(η0 ≤ 1)]
=
=
.
µρ [η0 ≤ 1]
µρ [η0 ≤ 1]
µρ [η0 ≤ 1]
547

Moreover
+∞
+∞
1 X kα(ρ)k
α(ρ)2 X (k + 2)α(ρ)k
µρ [η0 1(η0 > 1)] =
=
Z(ρ) k=2 c(k)!
Z(ρ) k=0 c(k + 2)!

+∞
B2 α(ρ)2 X (k + 2)α(ρ)k
= B2 α(ρ)2 (ρ + 2) = O(|Λ|−2 ),

Z(ρ) k=0
c(k)!

and finally
ρe =

ρ + O(|Λ|−2 )
= ρ + O(|Λ|−2 ).
1 + O(|Λ|−2 )

Observe that for any n ∈ {0, . . . , |Λ|}, we have
¯
i µ|Λ|¶
h
¯
µρ η Λ = n¯ max ηx ≤ 1 =
ρe(1 − ρe)|Λ|−n .
x∈Λ
n

(7.5)

(7.6)

This comes from the fact that the random variables {ηx : x ∈ Λ}, under the probability
measure µρ [·| maxx∈Λ ηx ≤ 1], are Bernoulli independent random variables with mean ρe.
The remaining part of the proof follows the classical argument of approximation of the
binomial distribution with the Poisson distribution. Using (7.5) and (7.6), after some simple
calculations we get
¯
i µ|Λ|¶
¯
µρ η Λ = n¯ max ηx ≤ 1 =
ρe(1 − ρe)|Λ|−n
x∈Λ
n
¸n ·
¸|Λ|−n
·
N
N
|Λ|!
−2
−2
+ O(|Λ| )
+ O(|Λ| )
1−
=
n!(|Λ| − n)! |Λ|
|Λ|
·
¸|Λ|
¤
N
1 £
−1 n
−2
1−
N + O(|Λ| )
+ O(|Λ| )
=
n!
|Λ|
·
¸−n
|Λ|(|Λ| − 1) · · · · · (|Λ| − n + 1)
N
N n −N
−2
×
1

+
O(|Λ|
)
e + O(|Λ|−1 )
=
|Λ|n
|Λ|
n!
h

uniformly in 0 < N ≤ N0 and 0 ≤ n < N . This proves that there exist positive constants v1
and B3 such that if Λ ⊂⊂ Zd is such that |Λ| > v1 then
¯
¯
¯ N/|Λ|
N n −N ¯¯ B3
¯p
¯ Λ (n) − n! e ¯ ≤ |Λ|

uniformly in N ∈ N \ {0} with N ≤ N0 and n ∈ N with n ≤ N . The general case follows
easily because the set of n ∈ N, N ∈ N \ {0} and Λ ⊂⊂ Zd such that n ≤ N ≤ N0 , |Λ| ≤ v1
and 0 ∈ Λ is finite.

Proposition 7.3 For any ρ0 > 0, there exist finite positive constants A0 , n0 and v0 such
that:
548

1.

¯
¯
(n−ρ|Λ|)2 ¯
¯p
A0
1

ρ
2
2
sup ¯¯ σ (ρ)|Λ|pΛ (n) − √ e 2σ (ρ)|Λ| ¯¯ ≤ p

n∈N
σ 2 (ρ)|Λ|

(7.7)

for any ρ ≤ ρ0 and any Λ ⊂⊂ Zd such that σ 2 (ρ)|Λ| ≥ n0 ;
2.

¯
¯
(n−ρ|Λ|)2 ¯
¯p
1
A0

ρ
sup ¯¯ σ 2 (ρ)|Λ|pΛ (n) − √ e 2σ2 (ρ)|Λ| ¯¯ ≤ p
ρ>ρ0

|Λ|

(7.8)

n∈N

d

for any Λ ⊂⊂ Z such that |Λ| ≥ v0 .

Proof. This is a special case of the Local Limit Theorem for µρ (see Theorem 6.1 in [5]).
We conclude this section with a bound on the tail of pρΛ which will be used in the regimes
not covered by Lemma 7.2 or Proposition 7.3.
Lemma 7.4 There exists a positive constant A0 such that
pρΛ (n + 1)
A0 ρ|Λ|
ρ|Λ|


ρ
A0 (n + 1)
pΛ (n)
n+1
for any n ∈ N \ {0}, Λ ⊂⊂ Zd and ρ > 0.
Proof. Notice that
pρΛ (n + 1) = µρ [η Λ = n + 1] =

1 X
µρ [ηx , 1(η Λ = n + 1)]
n + 1 x∈Λ

and, by the change of variable ξ := σ − δ x :
X
X
µρ [ηΛ = σ]σx 1(σ Λ = n+1) =
µρ [ηΛ = σ]σx 1(σ Λ = n+1, σx > 0)
µρ [ηx 1(η Λ = n+1)] =
σ∈ΩΛ

=

X

µρ [ηΛ = ξ + δ x ]
(ξx + 1)1(ξ Λ = n)
µ
ρ [ηΛ = ξ]
ξ∈ΩΛ
·
¸
X
ηx + 1
α(ρ)(ξx + 1)
µρ [ηΛ = ξ]
=
1(ξ Λ = n) = α(ρ)µρ
,η = n .
c(ξx + 1)
c(ηx + 1) Λ
ξ∈Ω

µρ [ηΛ = ξ + δ x ](ξx + 1)1(ξ Λ = n) =

ξ∈ΩΛ

σ∈ΩΛ

X

µρ [ηΛ = ξ]

Λ

This means that
pρΛ (n

·
¸
α(ρ) X
ηx + 1
+ 1) =
µρ
1(η Λ = n) .
n + 1 x∈Λ
c(ηx + 1)

By (2.4) we know that there exists a positive constant B0 such that
c(k)
≤ B0
for any k ∈ N \ {0}.
k
Thus, by plugging these bounds in (7.9), we get
B0−1 ≤

α(ρ)|Λ|pρΛ (n)
B0 α(ρ)|Λ|pρΛ (n)
ρ
≤ pΛ (n + 1) ≤
,
B0 (n + 1)
n+1
from which the conclusion follows.
549

(7.9)

7.2

Gaussian estimates for the canonical measure

In this section we will prove some Gaussian bounds on νΛN [η Λ′ = ·], when the volumes |Λ|
and |Λ′ | are of the same order (typically it will be |Λ′ |/|Λ| = 1/2). These bounds are volume
dependent (see Proposition 7.8 below) and so are of limited utility. However they will be
used to prove Proposition 7.9 which, in turn, is used in Section 8 to regularize ν ΛN [η Λ′ = ·]
and prove Gaussian uniform estimates on it.
Assume Λ′ ⊂ Λ ⊂⊂ Zd , N ∈ N\{0} and consider the probability measure on {0, 1, . . . , N }
νΛN [η Λ′

pρΛ′ (n)pρΛ\Λ′ (N − n)

= n] =

pρΛ (N )

.

Notice that νΛN [η Λ′ = ·] does not depend on ρ > 0 (and depends on Λ and Λ′ only through
|Λ| and |Λ′ |). Indeed a simple computation shows that
P
1(ξ Λ\Λ′ =N −n)
1(σ Λ′ =n) P
σ∈ΩΛ′

νΛN [η Λ = n] = P
N P
n=0

Q

x∈Λ′

σ∈ΩΛ′

c(σx )!

1(σ =n)
Q Λ′
x∈Λ′ c(σx )!

ξ∈ΩΛ\Λ′

P

Q

y∈Λ\Λ′

ξ∈ΩΛ\Λ′

c(ξy )!

1(ξ Λ\Λ′ =N −n)
Q
y∈Λ\Λ′ c(ξy )!

,

for any n ∈ {0, 1, . . . , N }. A particular case is when Λ′ ⊂ Λ ⊂⊂ Zd is such that |Λ′ | = |Λ\Λ′ |.
In this case we define
γΛN (n)

:=

νΛN [η Λ′

= n] =

pρΛ′ (n)pρΛ\Λ′ (N − n)
pρΛ (N )

=

pρΛ′ (n)pρΛ′ (N − n)
.
pρΛ (N )

(7.10)

We begin by proving a simple result on the decay of the tails of γΛN .
Proposition 7.5 There exist a positive constant A0 such that for every Λ ⊂⊂ Zd such that
|Λ| ≥ 2 and every N ∈ N \ {0}
N −n
γ N (n + 1)
A0 (N − n)

≤ Λ N
,
A0 (n + 1)
(n + 1)
γΛ (n)

(7.11)

for any n ∈ {0, . . . , N − 1}.
Proof. Let Λ′ ⊂ Λ be a non empty lattice set. Then by (7.10)

pρΛ′ (n + 1)pρΛ\Λ′ (N − n − 1)
γΛN (n + 1)
,
=
pρΛ′ (n)pρΛ\Λ′ (N − n)
γΛN (n)

and (7.11) follows from Lemma 7.4.
Next we prove Gaussian bounds on γΛN . This is a very technical argument. We begin
with the case |Λ| = 2.
¯ := ⌈N/2⌉ for N ∈ N \ {0}. There exist a
Lemma 7.6 A

Dokumen yang terkait

AN ALIS IS YU RID IS PUT USAN BE B AS DAL AM P E RKAR A TIND AK P IDA NA P E NY E RTA AN M E L AK U K A N P R AK T IK K E DO K T E RA N YA NG M E N G A K IB ATK AN M ATINYA P AS IE N ( PUT USA N N O MOR: 9 0/PID.B /2011/ PN.MD O)

0 82 16

ANALISIS FAKTOR YANGMEMPENGARUHI FERTILITAS PASANGAN USIA SUBUR DI DESA SEMBORO KECAMATAN SEMBORO KABUPATEN JEMBER TAHUN 2011

2 53 20

EFEKTIVITAS PENDIDIKAN KESEHATAN TENTANG PERTOLONGAN PERTAMA PADA KECELAKAAN (P3K) TERHADAP SIKAP MASYARAKAT DALAM PENANGANAN KORBAN KECELAKAAN LALU LINTAS (Studi Di Wilayah RT 05 RW 04 Kelurahan Sukun Kota Malang)

45 393 31

FAKTOR – FAKTOR YANG MEMPENGARUHI PENYERAPAN TENAGA KERJA INDUSTRI PENGOLAHAN BESAR DAN MENENGAH PADA TINGKAT KABUPATEN / KOTA DI JAWA TIMUR TAHUN 2006 - 2011

1 35 26

A DISCOURSE ANALYSIS ON “SPA: REGAIN BALANCE OF YOUR INNER AND OUTER BEAUTY” IN THE JAKARTA POST ON 4 MARCH 2011

9 161 13

Pengaruh kualitas aktiva produktif dan non performing financing terhadap return on asset perbankan syariah (Studi Pada 3 Bank Umum Syariah Tahun 2011 – 2014)

6 101 0

Pengaruh pemahaman fiqh muamalat mahasiswa terhadap keputusan membeli produk fashion palsu (study pada mahasiswa angkatan 2011 & 2012 prodi muamalat fakultas syariah dan hukum UIN Syarif Hidayatullah Jakarta)

0 22 0

Pendidikan Agama Islam Untuk Kelas 3 SD Kelas 3 Suyanto Suyoto 2011

4 108 178

ANALISIS NOTA KESEPAHAMAN ANTARA BANK INDONESIA, POLRI, DAN KEJAKSAAN REPUBLIK INDONESIA TAHUN 2011 SEBAGAI MEKANISME PERCEPATAN PENANGANAN TINDAK PIDANA PERBANKAN KHUSUSNYA BANK INDONESIA SEBAGAI PIHAK PELAPOR

1 17 40

KOORDINASI OTORITAS JASA KEUANGAN (OJK) DENGAN LEMBAGA PENJAMIN SIMPANAN (LPS) DAN BANK INDONESIA (BI) DALAM UPAYA PENANGANAN BANK BERMASALAH BERDASARKAN UNDANG-UNDANG RI NOMOR 21 TAHUN 2011 TENTANG OTORITAS JASA KEUANGAN

3 32 52