Lectures On Mathematical Statistical Mechanics
Lectures On Mathematical Statistical Mechanics
Sgrbhinn Institi
uid Ard-L
einn Bhaile Atha
Cliath
Sraith. A. Uimh 30
Communications of the Dublin Institute for Advanced Studies
Series A (Theoretical Physics), No. 30
LECTURES ON
MATHEMATICAL
STATISTICAL MECHANICS
By
S. Adams
DUBLIN
Institi
uid Ard-L
einn Bhaile Atha
Cliath
Dublin Institute for Advanced Studies
2006
Contents
1 Introduction
2 Ergodic theory
2.1 Microscopic dynamics and time averages . . . . . . .
2.2 Boltzmanns heuristics and ergodic hypothesis . . . .
2.3 Formal Response: Birkhoff and von Neumann ergodic
2.4 Microcanonical measure . . . . . . . . . . . . . . . .
2
. . . . . 2
. . . . . 8
theories
9
. . . . . 13
3 Entropy
16
3.1 Probabilistic view on Boltzmanns entropy . . . . . . . . . . . 16
3.2 Shannons entropy . . . . . . . . . . . . . . . . . . . . . . . . 17
4 The
4.1
4.2
4.3
4.4
Gibbs ensembles
The canonical Gibbs ensemble
The Gibbs paradox . . . . . .
The grandcanonical ensemble
The orthodicity problem . .
5 The
5.1
5.2
5.3
Thermodynamic limit
33
Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
Thermodynamic function: Free energy . . . . . . . . . . . . . 37
Equivalence of ensembles . . . . . . . . . . . . . . . . . . . . . 42
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
6 Gibbs measures
6.1 Definition . . . . . . . . . . . . . . . . .
6.2 The one-dimensional Ising model . . . .
6.3 Symmetry and symmetry breaking . . .
6.4 The Ising ferromagnet in two dimensions
6.5 Extreme Gibbs measures . . . . . . . . .
6.6 Uniqueness . . . . . . . . . . . . . . . .
6.7 Ergodicity . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
20
20
26
27
31
44
44
47
51
52
57
58
60
62
9 Models
9.1 Lattice Gases . . . . . .
9.2 Magnetic Models . . . .
9.3 Curie-Weiss model . . .
9.4 Continuous Ising model .
.
.
.
.
.
.
.
.
.
.
.
.
ii
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
73
74
75
77
84
Preface
In these notes we give an introduction to mathematical statistical mechanics,
based on the six lectures given at the Max Planck institute for Mathematics in
the Sciences February/March 2006. The material covers more than what has
been said in the lectures, in particular examples and some proofs are worked
out as well the Curie-Weiss model is discussed in section 9.3. The course
partially grew out of lectures given for final year students at the University
College Dublin in spring 2004. Parts of the notes are inspired from notes of
Joe Pule at University College Dublin.
The aim is to motivate the theory of Gibbs measures starting from basic
principles in classical mechanics. The first part covers Sections 1 to 5 and
gives a route from physics to the mathematical concepts of Gibbs ensembles
and the thermodynamic limit. The Sections 6 to 8 develop a mathematical
theory for Gibbs measures. In Subsection 6.4 we give a proof of the existence of phase transitions for the two-dimensional Ising model via Peierls
arguments. Translation invariant Gibbs measures are characterised by a variational principle, which we outline in Section 7. Section 8 gives a quick introduction to the theory of large deviations, and Section 9 covers some models
of statistical mechanics. The part about Gibbs measures is an excerpt of
parts of the book by Georgii ([Geo88]). In these notes we do not discuss
Boltzmanns equation, nor fluctuations theory nor quantum mechanics.
Some comments on the literature. More detailed hints are found throughout the notes. The books [Tho88] and [Tho79] are suitable for people, who
want to learn more about the physics behind the theory. A standard reference in physics is still the book [Hua87]. The route from microphysics to
macrophysics is well written in [Bal91] and [Bal92]. The old book [Kur60] is
nice for starting with classical mechanics developing axiomatics for statistical
mechanics. The following books have a higher level with special emphasis
on the mathematics. The first one is [Khi49], where the setup for the microcanonical measures is given in detail (although not in used modern manner).
The standard reference for mathematical statistical mechanics is the book
[Rue69] by Ruelle. Further developments are in [Rue78] and [Isr79]. The
book [Min00] contains notes for a lecture and presents in detail the twodimensional Ising model and the Pirogov-Sinai theory, the latter we do not
study here. A nice overview of deep questions in statistical mechanics gives
[Gal99], whereas [Ell85] and [Geo79],[Geo88] have their emphasis on probability theory and large deviation theory. The book [EL02] gives a very nice
introduction to the philosophical background as well as the basic skeleton of
statistical mechanics.
iii
I hope these lectures will motivate further reading and perhaps even further research in this interesting field of mathematical physics and stochastics.
Many thanks to Thomas Blesgen for reading the manuscript. In particular I
thank Tony Dorlas, who gave valuable comments and improvements.
Leipzig, Easter 2006
Stefan Adams
iv
Introduction
ness comes into play due to lack of knowledge (macroscopic description). The
large number of subsystems is replaced in a mathematical idealisation by infinitely many subsystems such that the extensive quantities are scaled to stay
finite in that limit. Stochastic limit procedures as the law of large numbers,
central limit theorems and large deviations principles will provide appropriate tools. In these notes we will give a glimpse of the basic concepts. In
the second chapter we are concerned mainly with the Mechanics in the name
Statistical Mechanics. Here we motivate the basic concepts of ensembles via
Hamilton equations of motions as done by Boltzmann and Gibbs.
2
2.1
Ergodic theory
Microscopic dynamics and time averages
or
= Rd )N .
(2.1)
Specify, at a given instant of time, the values of positions and momenta of the
N particles. Hence, one has to specify 2dN coordinate values that determine
a single point in the phase space respectively . Each single point in
the phase space corresponds to a microscopic state of the given system of N
particles. Now the question arises whether the 2dN dimensional continuum
of microscopic states is reasonable. Going back to Boltzmann [Bol84] it seems
that at that time the 2dN dimensional continuum was not really deeply
accepted ([Bol74], p. 169):
Therefore if we wish to get a picture of the continuum in words,
we first have to imagine a large, but finite number of particles
with certain properties and investigate the behaviour of the ensembles of such particles. Certain properties of the ensemble may
approach a definite limit as we allow the number of particles ever
more to increase and their size ever more to decrease. Of these
properties one can then assert that they apply to a continuum,
and in my opinion this is the only non-contradictory definition of
a continuum with certain properties...
and likewise the phase space itself is really thought of as divided into a finite number of very small cells of essentially equal dimensions, each of which
2
for i = 1, . . . , N, j = 1, . . . , d,
(2.2)
H
H
and pi =
,
pi
qi
i = 1, . . . , N,
(2.3)
where the dot denotes as usual differentiation with respect to the time variable. If J denotes the 2n 2n matrix
0 1ln
,
1ln 0
3
H
q1
,...,
H H
H
,
...,
.
qN p1
pN
d
dt
F (t (x)) = F 0 (t (x))
(x)
dt
dt
dt
(x)
dt
dt
= f 0 (H(t (x))) (H(t (x)))T
(x)
dt
dt
= f 0 (H(t (x))) hH(t (x)),
(x)i
dt
= f 0 (H(t (x))) hH(t (x)) , J H(t (x))i.
= f 0 (H(t (x))) H 0 (t (x))
The following theorem is the well known Liouvilles Theorem.
4
Theorem 2.2 (Liouvilles Theorem) The Jacobian | det 0t (x)| of a Hamiltonian flow is constant and equal to 1.
Proof.
Z
trace A(s)ds .
Now
Thus
dt (x)
= (v t )(x).
dt
d0t (x)
= v 0 (t (x))0t (x)
dt
Rt
so that det 0t (x) = exp 0 trace v 0 (s (x))ds because 00 (x) is the identity map
on . Now
v(x) = J H(x) and v 0 (x) = JH 00 (x),
2
H
is the Hessian of H at x. Since H is twice continuwhere H 00 (x) = xi x
j
00
ously differentiable, H (x) is symmetric. Thus
trace (JH 00 (x)) = trace((JH 00 (x))T ) = trace ((H 00 (x))T J T )
= trace (H 00 (x)(J)) = trace (JH 00 (x)).
Therefore trace (JH 00 (x)) = 0 and det(0t (x)) = 1.
From Lemma 2.1 and Theorem 2.2 it follows that a probability measure
on the phase space , i.e., an element of the set P(, B ) of probability
measure on with Borel algebra B , whose Radon-Nikodym density with
respect to the Lebesgue measure is a function of the Hamiltonian H alone is
stationary with respect to the Hamiltonian flow = {t : t R}.
Corollary 2.3 Let P(, B )R with density = F H for some function
F : R R be given, ie., (A) = A (x)dx for any A B . Then
1
t =
for any t R.
5
Proof.
We have
Z
Z
(A) = 1lA (x)(x)dx = 1lA (t (x))(t (x))| det 0t (x)|dx
Z
= 1l1
(x)(x)dx = (1
t A).
t A
For such a stationary probability measure one gets a unitary group of time
evolution operators in an appropriate Hilbert space as follows.
Theorem 2.4 (Koopmans Lemma:) Let 1 be a subset of the phase space
invariant under the flow , i.e., t 1 1 for all t R. Let
P(1 , B1 ) be a probability measure on 1 stationary under the flow that
is 1
= for all t R. Define Ut f = f t for any t R and any
t
function f L2 (1 , ), then {Ut : t R} is a unitary group of operators in
the Hilbert space L2 (1 , ).
Proof. Since Ut Ut = U0 = I, Ut is invertible and thus we have to prove
only that Ut preserves inner products.
Z
Z
hUt f, Ut gi = (f t )(x)(g t (x)(dx) = (fg)(t (x))(dx)
Z
Z
1
f (x) = lim
dt(f t )(x).
T 2T T
Suppose that is a probability measure on 1 invariant with respect to t
then
Z T Z
Z
1
dt
(f t )(x)(dx)
f(x)(dx) = lim
T 2T T
1
1
Z T Z
1
= lim
dt
f (x)( 1
t )(dx)
T 2T T
1
Z T Z
1
= lim
dt
f (x)(dx)
T 2T T
1
Z
=
f (x)(dx).
1
Assume now that the observed value is independent of where the system is
at t = 0 in 1 , i.e. if f(x) = f for a constant f, then
Z
Z
f(dx) = f.
f(x)(dx) =
1
Therefore
f =
Z
f (x)(dx).
1
We have made
Z Ttwo assumptions in this argument:
1
(1) lim
dt(f t )(x) exists.
T 2T T
(2) f(x) is constant on 1 .
Statement (1) has been proved by Birkhoff (Birkhoffs pointwise ergodic theorem (see section 2.3 below)): f(x) exists almost everywhere. We shall prove
a weaker version of this, Von Neumanns ergodic theorem.
Statement (2) is more difficult and we shall discuss it later.
The continuity of the Hamiltonian flow entails that for each f Cb (Rd ) and
each x the function
fx : R R, 7 fx (t) = f (x(t))
(2.5)
2.2
Boltzmanns argument (1896-1898) for the introduction of ensembles in statistical mechanics can be broken down into steps which were unfortunately
entangled.
1. Step: Find a set of time dependent functions which admit an invariant mean hi like (2.6).
2. Step: Find a reasonable set of observables, which ensure that the time
average of each observable over an orbit in the phase space is independent of the orbit.
Let Cb (R) be the set of all bounded continuous functions on the real line R,
equip it with the supremum norm |||| = suptR |(t)|, and define
Z T
o
n
1
dt(t) exists .
(2.7)
M = Cb (R) : hi = lim
T 2T T
Lemma 2.6 There exist positive linear functionals : Cb (R) R, normalised to 1, and invariant under time translations, i.e., such that
(i) () 0 for all M,
(ii) linear,
(iii) (1) = 1,
(iv) (s ) = () for all s R, where s (t) = (t s), t, s R,
RT
1
such that () = limT 2T
dt(t) for all Cb (R) where this limit
T
exists.
Our time evolution
(t, x) R 7 t (x) = x(t)
is continuous, and thus we can substitute fx in (2.5) for in the above result.
This allows to define averages on observables as follows.
8
Lemma 2.7 For every f Cb () and every x , there exists a timeinvariant mean x given by
x : Cb () R, f 7 x (f ) = (fx ),
with x depending only on the orbit {x(t) : t R, x(0) = x}.
For any E R+ let
E = {(q, p) : : H(q, p) = E}
denote the energy surface for the energy value E for a given Hamiltonian
H for the system of N particles.
The strict ergodicity hypothesis
The energy surface contains exactly one orbit, i.e. for every x
and E 0
{x(t) : t R, x(0) = x} = E .
There is a more realistic mathematical version of this conjecture.
The ergodicity hypothesis
Each orbit in the phase space is dense on its energy surface, i.e.
{x(t) : t R, x(0) = x} is a dense subset of E .
2.3
We present in this section briefly the important results in the field of ergodic
theory initiated by the ergodic hypothesis. For that we introduce the notion
of a classical dynamical system.
Notation 2.8 (Classical dynamical system) A classical dynamical system is a quadruple (, F, ; ) consisting of a probability space (, F, ),
where F is a algebra on , and a one-parameter (additive) group T (R
or Z) and a group of actions, : T , (t, x) 7 t (x), of the group
T on the phase space , such that the following holds.
(a) fT : T R, (t, x) 7 f (t (x)) is measurable for any measurable
f : R,
(b) t s = t+s for all s, t T ,
9
Z
(dx)x (f ) =
Proof.
(dx)f (x).
Note that in Birkhoffs Theorem one has convergence almost surely. There
exists a weaker version, the following ergodic theorem of von Neumann. We
restrict in the following to classical dynamical system with T = R, the real
time. Let H = L2 (1 , F, ) and define Ut f = f t for any f H. Then by
Koopmans lemma Ut is unitary for any t R.
Theorem 2.10 (Von Neumanns Mean Ergodic Theorem) Let
M = {f H : Ut f = f t R},
then for any g H,
Z
1 T
dtUt g
gT :=
T 0
converges to P g as T , where P is the orthogonal projection onto M.
For the proof of this theorem we need the following discrete version.
Theorem 2.11 Let H be a Hilbert space and let U : H H be an unitary
operator. Let N = {f H : U f = f } = ker(U I) then
N 1
1 X n
U g = Qg
N N
n=0
lim
Let U = U1 and g =
Z
and thus
N
1
X
dtUt f , then
n+1
dtUn+t f =
U g=
R1
dtUt f
n
U g=
dtUt f.
0
n=0
RN
Therefore N1 0 dtUt f converges as N . For T R+ , by writing T =
RT
N + r where 0 r < 1 and N N, we deduce that T1 0 dtUt f converges as
T . Define the operator P by
Z
1 T
P f = lim
dtUt f.
T T 0
Note that P f M and
1
P f = lim
T T
dtUt f M.
0
[ker(I U )] = Range(I U ).
If f ker(I U ) and g Range(I U ), then for some h,
hf, gi = hf, (I U )hi = h(I U )f, hi
= h(I U )f, U hi = 0.
Thus Range(I U ) [ker(I U )] . Since [ker(I U )] is closed,
Range(I U ) [ker (I U )] .
If f [Range(I U )] , then for all g
0 = hf, (I U )U gi = h(I U )f, U gi = h(I U )f, gi.
Therefore (I U )f = 0, that is f ker(I U ). Thus
[Range (I U )] ker (I U ).
11
Then
[ker (I U )] [Range (I U )] = Range (I U ).
If g Range(I U ), then g = (I U )h for some h. Therefore
N 1
1 X n
1
U g = {h U h + U h U 2 h + U 2 h U 3 h + . . . + U N 1 h U N h}
N n=0
N
=
Thus
1
{h U N h}.
N
1
1 N
X n
2khk
U g
0 as N .
N n=0
N
Approximating
of Range(I U ) by elements of Range(I U ), we
PNelements
1 n
1
have that N n=0 U g 0 = P g for all
g [ker (I U )] = Range(I U ).
If g ker (I U ), then
N 1
1 X n
U g = g = P g.
N n=0
Definition 2.12 (Ergodicity) Let = (t )tR be a flow on 1 and a
probability measure on 1 which is stationary with respect to . is said to
be ergodic if for every measurable set F 1 such that t (F ) = F for all
t R, we have (F ) = 0 or (F ) = 1.
Theorem 2.13 (Ergodic flows) = (t )tR is ergodic if and only if the
only functions in L2 (1 , ) which satisfy f t = f are the constant functions.
Proof. Below a.s.(almost surely) means that the statement is true except
on a set of zero measure. Suppose that the only invariant functions are the
constant functions. If t (F ) = F for all t then 1lF is an invariant function
and so 1lF is constant a.s. which means that 1lF (x) = 0 a.s. or 1lF (x) = 1 a.s.
Therefore (F ) = 0 or (F ) = 1l.
Conversely suppose t is ergodic and f t = f . Let F = {x| f (x) < a}.
Then t F = F since
t (F ) = {t (x)| f (x) < a} = {t (x)| f (t (x) < a} = F.
12
Therefore
1
lim
T 2T
Z
dtUt g =
g(x)(dx).
1
Remark 2.14 However proving ergodicity has turned out to be the most difficult part of the programme. There is only one example for which ergodicity
has been claimed to be proved and that is for a system of hard rods (Sinai).
This concerns finite systems. In the thermodynamic limit (see Chapter 5)
ergodicity should hold, but we do not discuss this problem.
2.4
Microcanonical measure
Theorem 2.15 (Riesz-Markov) If l : A() R is linear and for any positive f A() it holds l(f ) 0, then there is a unique Borel measure on
such that
Z
l(f ) = f (x)(dx).
(2.8)
dt1 = Q
Then
dt1
dt .
k1 k 1
We apply this to the Microcanonical Measure.
Choose : R2n A R2n such that 1 = H and so t1 is an energy surface.
Then
Z
Z E+
Z
f ( 1 (t))
f (x)dx =
dt1
dt1 .
1 (t))k
E
t1 kH(
[E,E+]
dx =
Therefore
1
lim
0
Z
f (x)dx =
[E,E+]
Thus
0E (dx) =
In particular
Z
(E) =
E
15
dE
.
kHk
t1
f ( 1 (t))
dt .
kH( 1 (t))k 1
dE
.
||H(N ) ||
3
3.1
Entropy
Probabilistic view on Boltzmanns entropy
We discuss briefly the famous Boltzmann formula S = k log W for the entropy
and give here an elementary probabilistic interpretation. For that let be
a finite set (the state space) and let there be given a probability measure
P() on , where P() denotes the set of probability measures on
with the -algebra being the set of all subsets of . In the picture of Maxwell
and Boltzmann, the set is the set of all possible energy levels for a system of
particles, and the probability measure corresponds to a specific histogram
of energies describing some macrostate of the system. Assume that (x) is
a multiple of n1 for any x , n N, i.e. is a histogram for n trials or a
macrostate for a system of n particles. A microscopic state for the system of
n particles is any configuration n .
Boltzmanns idea: The entropy of a macrostate P() corresponds to
the degree of uncertainty about the actual microstate n when only
is known and thus can be measured by log Nn (), the logarithmic number of
microstates leading to .
16
Proof.
The entropy H() counts the number of possibilities to obtain the macrostate
or histogram , and thus it describes the hidden multiplicity of the true
microstates consistent with the observed . It is therefore a measure of the
complexity inherent in .
3.2
Shannons entropy
We give a brief view on the basic facts on Shannons entropy, which was
established by Shannon 1949 ([Sha48] and [SW49]). We base the specific form
of the Shannon entropy functional on probability measures just on a couple of
clear intuitive arguments. For that we start with a sequence of four axioms on
a functional S that formalises the intuitive idea that entropy should measure
the lack of information (or uncertainty) pertaining to a probability measure.
For didactic reasons we limit ourselves to probability measures on a finite
set = {1 , . . . , n } of elementary events. Let P P()
P be the probability
measure with P ({i }) = pi [0, 1], i = 1, . . . , n, and ni=1 pi = 1. Now we
formulate four axioms for a functional S acting on the set P() of probability
measures.
Axiom 1: To express the property that S is a function of P P() alone
and not of the order of the single entries, one imposes:
17
(a) For every permutation Sn , where Sn is the group of permutations of n elements, and any P P() let P P() be defined as
P ({i }) = p(i) for any i = 1, . . . , n. Then
S(P ) = S(P ).
(b) S(P ) is continuous in each of the entries pi = P ({i }), i = 1, . . . , n.
The next axiom expresses the intuitive fact that the outcome is most random
for the uniform distribution.
Axiom 2: Let P (uniform) ({i }) =
1
n
for i = 1, . . . , n. Then
The next axiom states that the entropy remains constant, whenever we extend our space of outcomes with vanishing probability.
Axiom 3: Let P 0 P(0 ) where 0 = {n+1 } and assume that
P 0 ({n+1 }) = 0. Then
S(P 0 ) = S(P )
for P P() with P ({i }) = P 0 ({i }) for all i = 1, . . . , n.
Finally we consider compositions.
0
Axiom 4: Let P P() and Q P(0 ) for some set 0 = {10 , . . . , m
}
0
with m N. Define the probability measure P Q P( ) as
m
X
l=1
n
X
pi log pi
i=1
for P()
19
4.1
,N (x)
eH (x)
=
Z (, N )
, x ,
20
(4.10)
(4.11)
Z
and
2 (x2 ) =
0 (x1 , x2 )dx1 .
0
with i = 0, 1, 2,
log Z (, N )
with =
1
.
kT
22
S(,N ) S() k
((1 log Z (, N ) H(N ) (x)),N (x)dx
Z
(N )
(1 log Z (, N ) H (x))(x))dx
S(,N
) S().
Z
= k log Z (, N ) + k
H(N ) (x),N (x)dx.
(N )
2
log Z =
,N (x) H (x)
2
,N (x)H(N ) (x)dx dx 0.
Thermodynamic functions
For the canonical ensemble the relevant thermodynamical variables are the
temperature T (or = (kT )1 ) and the volume V of the region R. We
have already defined the entropy S of the canonical ensemble by
S (, N ) = k log Z (, N ) +
where U = E (H(N ) ) =
,N
1
E (H (N ) ),
T ,N
1
= log Z (, N ).
23
1
E (H (N ) ))
T ,N
1 X 2
H (x) =
p
2m i=1 i
, x = (q, p) .
h2
2m
24
12
P (, N ) =
=
=
.
V
V
V
T
Let aN (, v) be the free energy per particle considered as a function of the
specific density v, that is,
aN (, v) =
1
A (, N ),
N N
since
lim
1
(d log log v 1) ,
1
log N ! log N
N
= 1.
a(, v)
v
,
T
1
and thus p(, v) = v
. We can also define the free energy density as a
function of the particle density , i.e.,
fl (, ) =
1
A (, Vl ),
Vl l
25
T
Clearly f (, ) = a(, 1 ). For the ideal gas we get
f (, ) =
(d log + log 1)
and therefore
Finally we want to check the relative dispersion of the energy in the canonical
ensemble. Let hH(N ) i = E (H(N ) ). Then
p(, ) =
,N
2 log Z (, N )
h(H(N ) hH(N ) i)2 i
.
=
( log Z (, N ))2
hH(N ) i2
This gives for the ideal gas
q
h(H(N ) hH(N ) i)2 i
(N )
hH i
4.2
1
1
1
= ( dN ) 2 = O(N 2 ).
2
(4.15)
vessels are in equilibrium having the same temperature and pressure. Now
imagine that the wall between the two vessels is gently removed. The aggregate vessel is now filled with a gas that is still in equilibrium at the same
temperature and pressure. Denote by S1 and S2 the entropy on each side of
the wall. Since the corresponding canonical Gibbs ensembles are independent of one another, the entropy S12 of the aggregate vessel - before the wall
is removed - is exactly S1 + S2 . However an easy calculation gives us
S12 (S1 + S2 ) = k((N1 + N2 ) log(V1 + V2 ) N1 log V1 N2 log V2 )
(4.16)
V2
V1
+ N2 log
> 0.
= k N1 log
V1 + V2
V1 + V2
This shows that the informational (Shannon) entropy has increased, while
we expected the thermodynamic entropy to remain constant, since the wall
between the two vessels is immaterial from a thermodynamical point of view.
This is the Gibbs paradox.
We have indeed lost information in the course of removing the wall. Imagine
the gas before removing the wall consists of yellow molecules in one vessel
and of blue molecules in the other. After removal of the wall we get a
uniform greenish mixture throughout the aggregate vessel. Before we knew
with probability 1 that a blue molecule was initially in the vessel where we
had put it, after removal of the wall we only know that it is in that part of
the aggregate vessel with probability NN1 N1 2 .
The Gibbs paradox is resolved in classical statistical mechanics with an ad
hoc ansatz. Namely, instead of the canonical partition function Z (, N )
one takes N1 ! Z (, N ) and instead of the microcanonical partition function
(E, N ) one takes N1 ! (E, N ). This is called the correct Boltzmann
counting. The appearance of the factorial can be justified in quantum
mechanics. It has something to do with the in-distinguishability of identical particles. A state describing a system of identical particles should be
invariant under any permutation of the labels identifying the single particle variables. However, this very interesting issue goes beyond the scope of
this lecture, and we will therefore assume it from now on. In Subsection 5.1
we give another justification by computing the partition function and the
entropy in the microcanonical ensemble.
4.3
but we deal with systems with a given temperature. Similarly we like not to
specify the number of particles but the average number of particles. In the
grandcanonical ensemble the system can have any number of particles with
the average number determined by external sources. The grandcanonical
Gibbs ensemble is obtained if the canonical ensemble is put in a particlebath, meaning that the particle number is no longer fixed, only the mean of
the particle number is determined by a parameter. This was similarly done
in the canonical ensemble for the energy, where one considers a heat-bath.
The phase space for exactly N particles in box Rd can be written as
,N = { ( Rd ) : = {(q, pq ) : q
b }, Card (b
) = N },
(4.17)
where
b , the set of positions occupied by the particles, is a locally finite
subset of , and pq is the momentum of the particle at positions q. If the
number of the particles is not fixed, then the phase space is
= { ( Rd ) : = {(q, pq ) : q
b }, Card (b
) finite}.
(4.18)
phase space in for N particles, and equip it with the -algebra B generated
,
(N )
(N )
,
(x) = Z (, )1 e(H
(x)N )
, N N,
,N
X
1
N
e
Z (, N )
fN (x),N (dx).
(4.20)
E , (f ) =
Z (, ) N =0
,N
28
1
log Z (, ).
For the grandcanonical measure we have a Principle of Maximum Entropy very similar to those for the other two ensembles. We maximise the
entropy subject to the constraint that the mean energy E , (H ) and the
Z
X
N =1
,N
Then the grandcanonical ensemble/measure , , where and are determined by E , (H ) = E and E , (N ) = N0 , N0 N, maximises the entropy
among the absolutely continuous probability measures on with mean energy E and mean particle number N0 .
Proof. As in the two previous cases we use a log a b log b (a b)(1 +
log a) and so
a log ta b log tb (a b)(1 + log a + log t).
eN eHN (x)
)
and put a = (N
, (x), b = N (x) and t = N !.
N !Z (, )
Then, writing Z for Z (, ),
Let ,
N (x) =
(N )
(N )
,
(x) log(N !,
(x)) N (x) log(N !N (x))
(N )
(,
(x) N (x))(1 log Z H(N ) (x) + N ).
29
) S() k
Z
nX
+ (1 log Z)(0)
,
(1 log Z)0
)
(1 log Z H(N ) (x) + N )(N
, (x)dx
N =1 ,N
XZ
o N =1
,N
Thermodynamic Functions:
We shall write Z for Z (, ) and we suppress for a while some obvious
sub-indices and arguments. We have already defined the entropy S by
S = k log Z +
1
(E , (H ) E , (N )),
T
1
=
log Z E , (N ) .
30
V
V
V
1
It is argued that log Z should be independent of V for large V and therefore
V
we can write
1
1
1
1
V
1
P =
log Z =
V log Z =
log Z +
log Z
V
V
V
V
V V
1
log Z.
V
Therefore we define the pressure by the equation
P =
1
log Z.
V
S=
A
T
is also satisfied.
All the thermodynamic functions can be calculated from Z = Z (, ).
Therefore all calculations in the grandcanonical ensemble begin with the
calculation of the partition function Z = Z (, ).
4.4
E (Tkin )
=
N
, Tkin
31
1 X 2
=
p,
2m i=1 i
dU + P dV
Tkin
is an exact differential at least in the thermodynamic limit. This will then
provide the second law of thermodynamics. Let us provide a heuristic check
for the canonical ensemble. Here,
Z
(N )
1
E,N (Tkin ) =
Tkin (x)eH (x) dx,
Z (, N ) ,N
and U = Z (, N ). The pressure in the canonical ensemble can be
calculated as
Z
X
(N )
1 2 a dq2 dqN dp1 dpN
N
p
,
P (,N ) =
eH (x)
Z (, N ) p>0
2m A
N!
Q
where the sum goes over all small cubes Q adjacent to thePboundary of the
box with volume V by a side with area a while A = Q a is the total
area of the container surface and q1 is the centre of Q. Let n(Q, v)dv, where
1
v = 2m
p is the velocity, be the density of particles with normal velocity v
that are about to collide with the external walls of Q. Particles will cede a
momentum 2mv = p in normal direction to the wall at the moment of their
collision (mv mv due to elastic boundary conditions). Then
XZ
va
dvn(Q, v)(2mv)
A
v>0
Q
is the momentum transferred per unit time and surface area to the wall.
Gaussian calculation gives then after a couple of steps that due to F (, N ) =
1 log Z (, N ) and S (E; N ) = (U F (, N )) we have that T =
2 Tkin
, and that
(k)1 = dk
N
T dS = d(F + T S ) + pdV = dU + pdV,
with p = 1 V
log Z (, N ). Details can be found in [Gal99], where also
references are provided for rigorous proofs for the orthodicity in the canonical
ensemble. The orthodicity problem is more difficult in the microcanonical
ensemble. The heuristic approach goes similar. However, for a rigorous proof
of the orthodicity one needs here a proof that the expectation of the kinetic
energy in the microcanonical ensemble satisfies
E(Tkin
) = E(Tkin ) (1 + N ) , > 0,
32
5.1
Definition
limit when this systems becomes large ( Rd ), and this limit is identified
with a thermodynamic function. Any singularities for these thermodynamic
functions in the thermodynamic limit may correspond to phase transitions
(see [Min00],[Gal99] and [EL02] for details on theses singularities).
Taking the thermodynamic limit thus involves letting tend to infinity, i.e., approaching Rd or Zd respectively. We have to specify how tends
to infinity. Roughly speaking we consider the following notion.
Notation 5.1 (Thermodynamic limit) A sequence (n )nN of boxes n
Rd is a cofinal sequence approaching Rd if the following holds,
(i) n Rd as n ,
(ii) If hn = {x Rd : dist(x, ) h} denotes the set of points with
distance less or equal h to the boundary of , the limit
|hn |
=0
n ||
lim
exists.
The thermodynamic limit consists thus in letting n for a cofinal sequence (n )nN of boxes with the following additional requirements for the
microcanonical and the canonical ensemble:
Microcanonical ensemble: There are energy densities n (0, ), given
En
with En as n , and particle densities n (0, ),
as n = |
n|
n
given as n = N
with Nn as n , such that n and
n
n (0, ) as n .
Canonical ensemble: There are particle densities n (0, ), given as
n
with Nn as n , such that n (0, ) as n .
n = N
n
In some models one needs more assumptions on the cofinal sequence of
boxes, for details see [Rue69] and [Isr79].
We check the thermodynamic limit of the following simple model in the
microcanonical ensemble, which will give also another justification of the
correct Boltzmann counting.
Ideal gas in the microcanonical ensemble
Consider a non-interacting gas of N identical particles of mass m in d dimensions, contained in the box of volume V = ||. The gradient of the
34
1 X 2
1
H (x) =
pi = (0, . . . , 0, p1 , . . . , pn ) , x ,
2m i=1
m
where as usual n = N d. We have
n
1 X 2
2
|H(x)| = 2
pi = H(x) , x ,
m i=1
m
2
(2mE) 2 N d1 =
exp(S/k)V N
.
mcdN
2N
V (N d2)
1 exp
U (S, V ) = E =
,
2
2m
(mcN d ) (N d2)
2S
k(N d2)
and the temperature as the partial derivative of the internal energy with
respect to the entropy is
U
2
2U
2U
T =
=
U=
2
S V
k(N d 2)
kN d
kN d(1 N d )
for large N . This gives for large N the following relations
d
U N kT,
2
35
U
d
N k.
T V
2
U
2
2U
N kT
U
P =
=
.
2
V S d(1 N d ) V
dV
V
CV =
The previous relation is the empirical ideal gas law in which k is Boltzmanns
constant. We can therefore identify k in the definition of the entropy with
Boltzmanns constant.
We need to calculate c . We have via a standard trick
Z
Z
Z
2
x2
|x|2
2
=
e dx =
S (r)er dr
e
dx =
0
R
Z
Z
c
c
1 t
1 r2
2
t e dt =
.
= c
r e dr =
2 0
2
2
0
This gives c =
2 2
,
( 2 )
tx1 et dt.
(x) =
0
Note that if n N then (n) = (n 1)!. The behaviour of (x) for large
positive x is given by Stirlings formula
1
(x) 2xx 2 ex .
This gives limx ( x1 log (x) log x) = 1. We now have for the entropy of
the non-interacting gas in a box of volume V
Nd
1
2 2
S (E, N ) = k log mV N N d (2mE) 2 N d1 .
( 2 )
Let v be the specific volume and the energy per particle, that is
v=
V
N
and
E
.
N
1
S (N, N ),
N N
36
S (E, N ) = k log
= k log
,
(5.21)
(n/d)!
N!
where we put Plancks constant h = 1. Then
d
d d
d
sN (, v) k log v + log N + log(4m) + log( ) d log h
2
2 2
2
1
log(N !)
N
h d
d
d
d
d+2
k
+ log v + log log
log( )
2
2
2
2
2
4m
as N .
5.2
We shall prove that the canonical free energy density exists in the thermodynamic limit for very general interactions. We consider a general interacting
gas of N identical particles of mass m in d dimensions, contained in the box
of volume of V with elastic boundary conditions. The Hamiltonian for this
system is
N
1 X 2
(N )
H =
p + U (r1 , . . . , rN ).
2m i=1 i
37
12 N d Z
Z
1 1
1 2m
U (q)
e
dq
=
eU (q) dq.
=
dN
N ! h2
N
!
N
N
Theorem 5.3 (Fisher-Ruelle) Let be a stable and tempered pair potential. Let R be as above in Definition 5.2 and for (0, ) let L0 be such that
(L0 + R)d N. Let Ln = 2n (L0 + R) R and let n be the cube centred at
the origin with side Ln and volume Vn = |n | = Ldn . Let Nn = (L0 + R)d 2dn .
If
1
fn (, ) =
A (, Nn ),
Vn n
then limn fn (, ) exists.
38
39
R
(1)
n-1
(2)
n-1
(4)
n-1
(3)
n-1
0
n
n = {(r1 , . . . rNn ) N
n | each n1 contains Nn1 rk s}.
Note that for (q1 , . . . qNn ) n there are no rk s in the corridor between the
(i)
n1 .
Let
(2)
n = {(q1 , . . . qNn ) Nn : q1 , . . . , qNn1 (1)
n1 , qNn1 +1 , . . . , q2Nn1 n1 ,
n
(2d )
n
Since n N
n
Z
1
dq1 . . . dqNn eUNn (q1 ,...,qNn )
Zn
Nn !dNn n
Z
Nn !
1 1
=
dq1 . . . dqNn eUNn (q1 ,...qNn ) .
(Nn1 !)2d Nn ! dNn n
n
Since is tempered, we get for q1 , . . . , qNn
UNn (q1 , . . . , qNn ) UNn1 (q1 , . . . , qNn1 ) . . .
+ UNn1 (q(2d 1)Nn1 +1 , . . . , q2d Nn1 ).
Thus
Zn
!2d
Nn1
n1
= (Zn1 )2 .
Therefore
gn =
1
1
1
d
log Zn = d
log Zn d
log(Zn1 )2 = gn1 .
Nn
2 Nn1
2 Nn1
Nn !dNn
n
N
n
n
N
n
Vn Nn BNn
e
.
Nn !dNn
Thus
1
gn log Vn
log Nn ! d log + B = log
Nn
1
log Nn ! d log + B
Nn
log + 2 d log + B
Nn
Vn
+ log Nn
5.3
Equivalence of ensembles
p(, ) = lim
f (, ) = lim
s(, ) = lim
42
is called entropy.
{s(, ) }
sup
(cp) ,>()
f (, ) = inf { 1 s(, )}
>()
this set generated by counting variables (see [Geo95] for details). Then each
Gibbs ensemble can be extended trivially to a probability on (, B) just by
putting the whole mass on a subset. Therefore it makes sense to consider
all weak limit points in the thermodynamic limit. If the limit points are not
unique, i.e., there are several accumulation points, one considers the whole
set of accumulation points closed appropriately as the set of equilibrium
states/measure or Gibbs measures.
Equivalence at the level of states/measures is given if all accumulation points
of the different Gibbs ensembles belong to the same set of equilibrium points
or the same set of Gibbs measure ([Geo93],[Geo95],[Ada01]).
In the next section we develop the mathematical theory for Gibbs measures without any limiting procedure.
Gibbs measures
6.1
Definition
Let Zd the square lattice for dimensions d 1 and let E be any finite set.
d
Define := E Z = { = (i )iZd : i E} the set of configurations with
values in the state space E. Let E be the power set of E, and define the
d
algebra F = E Z such that (, F) is a measurable space. Denote the set
of all probability measures on (, F) by P(, F).
Definition 6.1 (Random field) Let P(, F). Any family (i )iZd of
random variables which is defined on the probability space (, F, ) and which
takes values in (E, E) is called a random field.
If one considers the canonical setup, where i : E are the projections
for any i Zd , a random field is synonymous with a probability measure
P(, F). Let S = { Zd : || < } be the set of finite volume
subsets of the square lattice Zd . Cylinder events are defined as { A}
for any A E and any projection : E for S. Then F is the
smallest - algebra containing all cylinder events. If S the algebra
F on contains all cylinder events {00 A} for all A E and 0 .
44
for any S.
(6.22)
(6.23)
AS,A6=
exists. The term exp(H ()) is called the Boltzmann factor for
some parameter > 0, where is the inverse temperature.
Example 6.3 (Pair potential) Let A = 0 whenever |A| > 2 and let
J : Zd Zd R, : E E R and : E R symmetric and measurable.
Then a general pair interaction potential is given by
0 if |A| > 2
We combine configurations outside and inside of any finite set of random
variable as follows. Let and , S, be given. Then Zd \
with (Zd \ ) = and Zd \ (Zd \ ) = Zd \ . With this notation we can
define a nearest-neighbour Hamiltonian with given boundary condition.
45
1
2
J(i, j)(i , j ) +
X
i,jc ,
hiji=1
i,j,hiji=1
(i , j ) +
J(i, i)(i )
denotes a Hamiltonian in with nearest-neighbour interaction and with configurational boundary condition , where hx, yi = maxi{1,...,d} |xi yi |
for x, y Zd . Instead of a given configurational boundary condition one can
model the free boundary condition and the periodic boundary condition as
well.
In the following we fix a probability measure P(E, E) on the state space
and call it the reference or a priori measure. Later we may also consider the Lebesgue measure as reference measure. Choosing a probability
measure as a reference measure for finite sets gives just a constant from
normalisation.
Definition 6.5 (Gibbs measure)
(i) Let , S, > 0 the inverse temperature and be an interaction potential. Define for any event A F
Z
1
(A|) = Z ()
(d)1lA (Zd \ ) exp H (Zd \ )
(6.24)
with normalisation or partition function
Z
Z () =
(d) exp H (Zd \ ) .
(6.25)
where is the Gibbs distribution for the parameter (6.24). The set
of Gibbs measures for inverse temperature with interaction potential
is denoted by G(, ).
46
(iii) An interaction potential is said to exhibit a first-order phase transition if |G(, )| > 1 for some > 0.
If the interaction potential is known we may skip the explicit appearance
of the interaction potential and write instead G() for the set of Gibbs measure with inverse temperature . However, the parameter can ever be
incorporated in the interaction potential .
Remark 6.6
(i) (A|) is T -measurable for any event A F.
(ii) The equation (6.25) is called the DLR-equation or DLR-condition
in honour of R. Dobrushin, O. Lanford and D. Ruelle.
6.2
47
0 , else
for .
(6.26)
H () = J
||
X
i i+1 h
||
X
i .
(6.27)
i=1
i=1
The partition function depends on the inverse temperature > 0, the coupling constant J and the external field h R, and is given by
X
X
(per)
Z (, J, h) =
exp H () .
(6.28)
1 =1
|| =1
V1 ,2 V2 ,3 V||1,|| V|| ,1
1 =1
with
|| =1
1
1
Vi i+1 = exp hi + Ji i+1 + hi+1
2
2
for any and i = 1, . . . , ||. Hence, Z (, J, h) = Trace V|| with the
symmetric matrix
(J+h)
e
eJ
V=
eJ e(Jh)
that has the eigenvalues
J
= e
cosh(h) e
2J
sinh (h) + e
2J
12
(6.29)
||
Z (, J, h) = + + .
48
(6.30)
This is a smooth expression in the external field parameter h and the inverse
temperature ; it rules out the appearance of a discontinuous isothermal
magnetisation: so far, no phase transition. The thermodynamic limit of the
free energy per volume is
1
1
log Z (, J, h) = log + ,
(6.31)
Z ||
1
because ||1 log Z (, J, h) = log + ||
1+( + )|| . The magnetisation
in the canonical ensemble is given as the partial derivative of the specific free
energy per volume,
f (, J, h) = lim
sinh(h)
m(, J, h) = h f (, J, h) = q
.
sinh2 (h) + e4J
This is symmetric, m(, J, 0) = 0 and limh m(, J, h) = 1 and for all
h 6= 0 we have |m(, J, h)| > |m(, 0, h)| saying that the absolute value of the
magnetisation is increased by the non-vanishing coupling constant J. The
set G(, ) of Gibbs measures contains only one element, called J,h , see
[Geo88] for the explicit construction of this measure as the corresponding
Markov chain distribution, here we outline only the main steps.
1.) The nearest-neighbour interaction J,h in (6.26) defines in the usual way
the Gibbs distributions J,h (|) for any finite Z and any boundary
condition . Define the function g : E 3 (0, ) by
J,h
{i}
(i = y|) = g(i1 , y, i+1 ),
y E, i Z, .
(6.32)
We compute
g(x, y, z) = ey(h+Jx+Jz) /2 cosh((h + Jx + Jz))
Fix any a E. Then the matrix
g(a, x, y)
Q=
g(a, a, y) x,yE
for x, y, z E.
(6.33)
(6.34)
n
Y
i=1
6.3
Before we study the two-dimensional Ising model, we briefly discuss the role
of symmetries for Gibbs measures and their connections with phase transitions. As is seen by the spontaneous magnetisation below the Curie temperature, the spin system takes one of several possible equilibrium states
each of which is characterised by a well-defined direction of magnetisation.
In particular, these equilibrium states fail to be preserved by the spin reversal (spin-flip) transformation. Thus breaking of symmetries has some
connection with the occurrence of phase transitions.
Let T denote the set of transformations
: , 7 (i 1 i )iZd ,
where : Zd Zd is any bijection of the lattice Zd , and the i : E E, i
Zd , are invertible measurable transformations of E with measurable inverses.
Each T is a composition of a spatial transformation and the spin
transformations i , i Zd , which act separately at distinct sites of the square
lattice Zd .
Example 6.7 (Spatial shifts) Denote by = (i )iZd the group of all
spatial transformations or spatial shifts or shift transformations
j : , (i )iZd 7 (ij )iZd .
Example 6.8 (Spin-flip transformation) Let the state space E be a symmetric Borel set of R and define the spin-flip transformation
: , ()iZd 7 (i )iZd .
Notation 6.9 The set of all translation invariant probability measures on
is denoted by P (, F) = { P(, F) : = i1 for any i Zd }.
The set of all translation invariant Gibbs measures for the interaction potential and inverse temperature is denoted by G (, ) = { G(, ) : =
i1 for any i Zd }.
Definition 6.10 (Symmetry breaking) A symmetry T is said to be
broken if there exists some G(, ) such that () 6= for some .
A direct consequence of symmetry breaking is that |G(, )| > 1, i.e.,
when there is a symmetry breaking the interaction potential exhibit a phase
transition. There are models where all possible symmetries are broken as
well as models where only a subset of symmetries is broken. A first example is the one-dimensional inhomogeneous Ising model, which is probably the
51
simplest model showing symmetry breaking. The one-dimensional inhomogeneous Ising model on the lattice N has the inhomogeneous nearest-neighbour
interaction potential = (A )AS defined
a sequence (Jn )nN of real
P for2J
numbers Jn > 0 for all n N such that nN e n < , as follows
Jn n n+1 , if A = {n, n + 1},
A =
.
0 otherwise ,
This model is spatial inhomogeneous, the potential is invariant under the
spin-flip transformation , but some Gibbs measures are not invariant under this spin-flip transformation (for details see [Geo88]) for > 0. The
simplest spatial shift invariant model which exhibits a phase transition is
the two-dimensional Ising model, which we will study in the next subsection 6.4. This model breaks the spin-flip symmetry while the shift-invariance
is preserved. Another example of symmetry breaking is the discrete twodimensional Gaussian model by Shlosman ([Shl83]). Here the spatial shift
invariance is broken. More information can be found in [Geo88] or [GHM00].
6.4
Let E = {1, 1} be the state space and define the nearest-neighbour interaction potential = (A )AS as
i j , if A = {i, j}, |i j| = 1
A =
.
0 , otherwise
The interaction potential is invariant under the spin flip transformation
and the shift-transformations i , i Zd . Let + , be the Dirac measures for
the constant configurations + and . The interaction potential
takes its minimum at + and , hence + and are ground states for
the system. The ground state generacy implies a phase transition if + ,
are stable in the sense that the set of Gibbs measure G(, ) is attracted by
each of the measures + and for . Let d denote the Levy metric
compatible with weak convergence in the sense of probability measures.
Theorem 6.11 (Phase transition) Under the above assumptions it holds
lim d(G (, ), + ) = lim d(G (, ), ) = 0.
52
Remark 6.12
(i) + (0 ) is the mean magnetisation. Thus: The two-dimensional Ising
ferromagnet admits an equilibrium state/measure of positive magnetisation although there is no action of an external field. This phenomenon
is called spontaneous magnetisation.
(ii) G (, ) > 1 + (0 ) > 0 goes back to [LL72]. Moreover, the Griffiths inequality implies that the magnetisation + (0 ) is a non-negative
non-decreasing function
there is a critical
of . Moreover
inversetemperature c such that G (, ) = 1 when < c and G (, ) > 1
when > c . The value of c is
1
1
c = sinh1 1 = log 1 + 2)
2
2
and the magnetisation for c is
+ (0 ) = 1 (sinh 2)4
81
(iii) For the same model in three dimensions one has again + , G (, ),
but there also exist non-shift-invariant Gibbs measures ([Dob73]).
Proof of Theorem 6.11.
B = {{i, j} Z2 : |i j| = 1, {i, j} 6= }
the set of all nearest-neighbour bonds which emanate from sites in . Each
bond b = {i, j} B should be visualised as a line segment between i
and j. This line segment crosses a unique dual line segment between two
nearest-neighbour sites u, v in the dual cube (shift by 12 in the canonical
directions). The associate set b = {u, v} is called the dual bond of b, and we
write
B = {b : b B } = {{u, v} : |u v| = 1}
for the set of all dual bonds. Note
1
b = {u + : |u (i + j)/2| = }.
2
A set c B is called a circuit of length l if c = {{u(k1) , u(k) } : 1 k l} for
some (u(0) , . . . , u(l) ) with u(l) = u(0) , |{u(1) , . . . , u(l) }| = l and {u(k1) , u(k) }
B , 1 k l. A circuit c surrounds a site a if for all paths (i(0) , . . . , i(n) )
in Z2 with i(0) = a and i(n)
/ and {i(m1) , i(m) } B for all 1 m n there
(m1)
exits a m N with {i
, i(m) } c. We denote the set of circuits which
surround a by Ca . We need a first lemma.
53
k = 1, . . . , l,
which cross the horizontal half-axis from a to the right for example. The
remaining l 1 dual bonds are successively added, at each step there are at
most 3 possible choices.
The ingenious idea of Peierls ([Pei36]) was to look at circuits which occur in
a configuration. For each we let
B () = {b : b = {i, j} B , i 6= j }
denote the set of all dual bonds in B which cross a bond between spins of
opposite sign. A circuit c with c B () is called a contour for . We let
outside of be constant. As in Figure 3 we put outside + -spins. If a site
a is occupied by a minus spin then a is surrounded by a contour for .
The idea for the proof of Theorem 6.11 is as follows. Fix + boundary
condition outside of . Then the minus spins in form (with high probability) small islands in an ocean of plus spins. Then in the limit Z2 one
obtains a + G() which is close to the delta measure + for sufficiently
large. As
+ and are distinct, so are + and when is large. Hence
G() > 1 when is large. We turn to the details. The next lemma just
ensures the existence of a contour for positive boundary conditions and one
minus spin in . We just cite it without any proof.
Lemma 6.14 Let with i = +1 for all i c and a = 1 for some
a . Then there exists a contour for which surrounds a.
Now we are at the heart of the Peierls argument, which is formulated in the
next lemma.
Lemma 6.15 Suppose c B is a circuit. Then
(c B ()|) e2|c|
for all > 0 and for all .
54
+ +
+ +
-
+ +
+
+ +
+
+
+ +
+
+ + +
+ + +
+
+
+ +
+
+
+
- -
55
+ +
+
+
+
+
- - +
- +
+
+ + +
+ - - - +
- - +
- - +
+ + +
+ +
-
- + +
+ +
-
+ + +
- - +
- +
+
- + + - +
+ - +
+ - +
- - +
- +
- - - + - + + +
-
+ +
+ +
+ +
+
+ +
+
- - +
+ +
- -
+
+
+
- +
+ + + + +
+ +
+ + +
+ + +
+ + + +
+
+
+
+
+
+
+
+
+
+
+
+
+
+ + +
+
+
- - + +
+ + +
- - +
- - +
+ + - - -
Proof.
{i,j}B
= |B | 2|{{i, j} B : 6= j }| = |B | 2|B |.
Now we define two disjoint sets of configurations which we need later for an
estimation.
A1 = { : Zd \ = Zd \ , c B ()}
A2 = { : Zd \ = Zd \ , c B () = }.
There is a mapping c : with
, if i is surrounded by c,
(c )i =
,
, otherwise
which flips all spins in the interior of the circuit c. Moreover, for all {i, j}
B we have
i j , if {i, j}
/c
(c )i (c )j =
,
i j , if {i, j} c
resulting in B (c )4B () = c (this was the motivation behind the definition of mappings), where 4 denotes the symmetric difference of sets. In
particular we get that c is a bijection from A2 to A1 , and we have
H () H (c ) = 2|B ()| 2|B (c )| = 2|c|.
Now we can estimate with the help of the set of events A1 , A2
P
P
exp(H
())
A exp(H (c ))
A
1
(c B ()|) P
= P 2
A2 exp(H ())
A2 exp(H ())
= exp(2|c|).
Now we finish our proof of Theorem 6.11. For > 0 define
X
r() = 1
l(3e2 )l ,
l1
cCa
56
l1
1 X
N +i (| + ).
|N | i
N
As P(, F) is compact, the sequence (N+ )N N has a cluster point + and one
can even show that + G (, ) ([Geo88]). Our estimation above gives
then + (a = 1) r(), and in particular one can show that
lim + = +
and
Note = (+ ). Hence,
+ (0 = 1| + ) = (0 = +1| ).
If is so large that + (0 = 1) r() < 31 , then
1
+ (0 = 1) = (0 = +1) < .
3
But {0 = 1} {0 = +1} = , and hence
2
+ (0 = +1) = 1 + (0 = 1) .
3
6.5
The set G(, ) of Gibbs measures for some interaction potential and
inverse temperature > 0 is a convex set, i.e., if , G(, ) and 0 < s < 1
then s + (1 s) G(, ). An extreme Gibbs measure (or in physics: a
pure state) is an extreme element of the convex set G(, ). The set
of all extreme Gibbs measures is denoted by ex G(, ). Below we give a
characterisation of extreme Gibbs measures. But first we briefly discuss
microscopic and macroscopic quantities. A real function f : R is
said to describe a macroscopic observable if f is measurable with respect
to the tail--algebra T . The T -measurability of a function f means that the
value of f is not affected by the behaviour of any finite set of spins. For
example, the event
n
o
1 X
i exists and belongs to B
B BR ,
lim
n |n |
i
n
57
6.6
Uniqueness
In this subsection we give a short intermezzo about the question of uniqueness of Gibbs measures, i.e., the situation when there is a most one Gibbs
measure possible for the given interaction potential. One might guess that
this question has something to due with the dependence structure introduced
from the interaction potential. One is therefore led to check the dependence
structure of the conditional Gibbs distributions at one given lattice site. For
58
sup
,,
d
= d
Z \{j}
Z \{j}
||{i}
(|) {i}
(|)||,
jZd
(ii) If
sup
iZd A3i
Example 6.19 (Lattice gas) Let E = {0, 1} and let the reference measure
be the counting measure. Let K : S R be any function on the set of all
finite subsets of Zd and define for the interaction potential by
Q
K(A) , if A = iA i = 1
A () =
0 , otherwise
for any A S. Note that (A ) = |K(A)|. Thus uniqueness is given whenever
X
sup
(|A| 1)|K(A)| < 4.
iZd A3i
Example 6.20 (One-dimensional systems) Let be a shift-invariant interaction potential and d = 1. Then there is at most one Gibbs measure
whenever
X
diam (A)(A ) < .
AS,
min A=0
6.7
Ergodicity
(6.38)
of all shift-invariant events. A F-measurable function f : R is Imeasurable if and only if f is invariant, in that f i = f for all i Zd . A
standard result in ergodic theory is the following theorem.
Theorem 6.21 (i) A probability measure P (, F) is extreme in
P (, F) if and only if is trivial on the invariant -algebra.
(ii) Each P (, F) is uniquely determined (within P (, F)) by its
restriction to I.
60
One can show that each extreme measure is a limit of finite volume Gibbs
distributions with suitable boundary conditions. Now, what about ergodic
Gibbs measures? The ergodic Theorem 6.24 below gives an answer: If
ex P (, F) and (N )N N a sequence of cubes with N Zd as N one
gets
1 X
f (i )
(f ) = lim
N |N |
i
N
= lim
for a.a.
(6.39)
The multidimensional ergodic theorem says something about the limiting behaviour of RN f as N . Let (N )N N be a cofinal sequence of boxes with
N Zd as N .
Theorem 6.24 (Multidimensional Ergodic Theorem) Let a probability measure P (, F) be given. For any measurable f : R with
(|f |) < ,
lim RN f = (f |I) a.s.
N
() log ().
Now (H)H() = F () is called the free energy of , and for any P()
we have
F () log Z
and F () = log Z
if and only if = .
To see this, apply Jensens inequality for the convex function (x) = x log x
and conclude by simple calculation
()
() X
X
=
()
(H) H() + log Z =
() log
()
()
X
()
()
= (1) = 0,
()
62
for the entropy. We wish to show that the thermodynamic limit of the entropy
exists, i.e.wish to show that
h() = lim
1
H ()
|n | n
exists for any cofinal sequence (n )nN of finite volume boxes in S. Essential
device for the proof of the existence of this limit is the following sub-additivity
property.
Proposition 7.4 (Strong Sub-additivity) Let , S and P(, F)
be given. Then
H () + H () H () + H ().
Proof.
(7.40)
exists in [, ).
64
(7.41)
Proof.
Choose
c > := inf
Sr.B.
1
a()
||
1
and let Sr.B. be such that ||
a() < c. Denote by Nn the number of
disjoint translates of contained in n . Then n is split into Nn translates
of and a remainder in the boundary layer. Choose Nn as large as possible.
n|
Then limn N|n ||
= 1. The sub-additivity gives
lim sup
Now, both Proposition 7.4 and Lemma 7.5 provide the main steps of the
proof of the following theorem.
Theorem 7.6 (Specific entropy) Fix a finite signed reference measure
on the measurable state space (E, E). Let P (, F) be a translation
invariant probability measure and (n )nN a cofinal sequence of boxes with
n Zd as n . Then,
(a)
1
Hn ()
n |n |
h() = lim
exists in [, (E)].
Now the following question arises. What happens if we take instead of the
reference measure any Gibbs measure and evaluate the relative entropy? We
analyse this question in the following. To define the specific energy of a
translation invariant probability measure it proves useful to introduce the
following function. Let = (A )AS be a translation invariant interaction
potential. Define the function f : R as
X
f =
|A|1 A .
(7.42)
A30
E (f ) = (f ) = lim
(7.43)
exists.
Notation 7.9 (Specific free energy) E (f ) or (f ) is called the specific (internal) energy per site of relative to . The quantity f () =
(f ) h() is called the specific free energy of for .
Proof of Theorem 7.8. For the proof see any of the books [Geo88],[Rue69]
or [Isr79]. The proof goes back to Dobrushin [Dob68b] and Ruelle [Rue69].
We continue our investigations with the previously occurred question of the
relative entropy with respect to a given Gibbs measure.
Theorem 7.10 (Pressure) Let P (, F) and G (, ), > 0,
be a translation invariant interaction potential, (n )nN be a cofinal sequence
of boxes with n Zd as n and (n )nN be a sequence of configurations
n . Then,
(a) P () = limn
1
|n |
log Zn (n ) exists.
66
1
H (|)
|n | n
(7.44)
(d) log
()
d
Z
= H () +
(d)H (Zd \ ) + log Z ().
We can draw an easy corollary, which is the first part of the variational
principle for Gibbs measures.
Corollary 7.12 (First part variational principle) For a translation invariant interaction potential and P (, F) we have h(|) 0. If
moreover G (, ) then h(|) = 0.
Proof. The assertions are due to Dobrushin ([Dob68a]) and Lanford and
Ruelle ([LR69]).
The next theorem gives the reversed direction and a summary of the whole
variational principle.
Theorem 7.13 (Variational principle) Let be a translation invariant
interaction potential, (n )nN a cofinal sequence of boxes with n Zd as
n and P (, F). Then,
(a) Let P (, F) be such that lim inf n
G (, ).
1
H (|)
|n | n
= 0. Then
(c) h(|) : P (, F) [0, ] is an affine lower semi continuous functional which attains its minimum 0 on the set G (, ). Equivalently,
G () is the set on which the specific free energy functional
f : P (, F) [0, ]
attains its minimum P ().
Proof. This variational principle is due to Lanford and Ruelle ([LR69]).
A transparent proof which reveals the significance of the relative entropy is
due to Follmer ([Fol73]).
8.1
Motivation
Consider the coin tossing experiment. The microstates are elements of the
configuration space = {0, 1}N equipped with the product measure P ,
where P({0, 1}) is given as = 0 0 +1 1 with 0 +1 = 1. If 0 = 1 = 21
we have a fair coin. Recall the projections j : {0, 1}, j N, and
consider the mean
N
1 X
SN () =
j ()
N j=1
for .
If m denotes the mean (m = 12 for a fair coin), the weak law of large
numbers (WLLN) tells us that for > 0
P (SN (m , m + )) 1
as N ,
as N .
In particular one can even prove exponential decay of the latter probability,
which we sketch briefly. The problem of decay of probabilities of rare events
68
is the main task of large deviations theory. For simplicity we assume now
that m = 21 . Then
1
log P (SN (z , z + ))
N N
inf
I(x),
F (z, ) = lim
=
x(z,z+)
N
1 X
() P({0, 1})
N j=1 j
for any ;
where T 0 = id and (T )j = j+1 is the shift and (N ) is the periodic continuation of the restriction of to N .
The latter example can be connected to our experience with Gibbs measures
and distributions as follows. Let N = [N, N ]d Zd , N N, and define the
periodic empirical field as
(per)
RN
() =
1 X
(N ) P (, F)
|N | k k
for all ,
The WLLN implies for all of these probabilities exponential decay of the rare
events given by a function I as the rate in N . This will be generalised in the
next subsection, where such functions I are called rate functions.
8.2
Definition
In the following we consider the general setup, i.e. we let X denote a Polish
space and equip it with the corresponding Borel--algebra BX .
Definition 8.1 (Rate function) A function I : X [0, ] is called a
rate function if
(1) I 6= ,
(2) I is lower semi continuous,
(3) I has compact level sets.
Definition 8.2 (Large deviations principle) A sequence (PN )N N of probability measures PN P(X, BX ) on X is said to satisfy the large deviations
principle with rate (speed) N and rate function I if the following upper and
lower bound hold,
1
log PN (C) inf I(x)
xC
N N
1
lim inf log PN (O) inf I(x)
N N
xO
lim sup
for C X closed,
(8.45)
for O X open.
Let us consider the following situation. Let (Xi )iN be i.i.d. real-valued
random variables, i.e., there is a probability space (, F, P) such that each
random variable has the distribution = P X11 P(R, BR ). Denote
1
the distribution of the mean SN by N = PN SN
P(R, BR ). For this
situation there is the following theorem about a large deviations principle
70
lim sup
Proof.
(8.46)
An important tool in proving large deviations principle is the following alternative version of the well-known Varadhan Lemma ([DZ98]).
Theorem 8.4 (Tilted LDP) Let the sequence (PN )N N of probability measures PN P(X, BX ) satisfy a large deviations principle with rate (speed) N
and rate function I. Let F : X R be a continuous function that is bounded
from above. Define
Z
eN F (x) PN (dx) , S BX .
JN (S) =
S
71
Proof. See [dH00] or [DZ98] for the original version of Varadhans Lemma
and [Ell85] or [Dor99] for a version as in the theorem.
8.3
(8.47)
where H(per)
is the Hamiltonian in N with interaction potential and periN
odic boundary conditions. Recall that ,
denotes the Gibbs distribution in
N
N with configurational boundary condition and ,per
the Gibbs disN
tribution in N with periodic boundary condition. Further, if G (, )
is a Gibbs measure, h(|) = h(|) denotes the specific relative entropy with
respect to the Gibbs measure with respect to the given interaction potential . Denote by e() the evaluation -algebra for the probability measures
on . Note that the mean energy h, i can be identified as a linear form on
a vector space of finite range interaction potentials. In particular we define
(per)
N () = hRN
(), i
for any interaction potential with finite range. In the limit N one
gets a linear functional on the vector space V of all interaction potentials
with finite range (see [Isr79] and[Geo88] for details on this vector space).
Theorem 8.5 (LDP for Gibbs measures) Let N = [N, N ]d Zd , >
0 and be an interaction potential with finite range. Then the following
assertions hold.
(per) 1
(a) Let G (, ) be given. Then the sequence ( (RN
) )N N of
(per) 1
) P(P(, F), e()) satisfies a large
probability measures (RN
deviations principle with rate (speed) |N | and rate function h(|).
(b) Let ,
be the Gibbs distribution in N with boundary condition .
N
Then for any closed set F P(, F) and any open set G P(, F),
1
(per)
log sup ,
(RN
F ) inf {h() + h, i + P ()},
N
F
N |N |
1
(per)
lim inf
log sup ,
(RN
G) inf {h() + h, i + P ()}.
N
N |N |
G
(8.48)
lim sup
72
1
lim inf
log sup ,
(N K) inf {JV ( )},
N
N |N |
K
lim sup
(8.49)
Models
73
9.1
Lattice Gases
Note that there is no need here for N ! since the particles are indistinguishable.
The thermal wavelength is put equal to 1. The grandcanonical partition
function is then
Z (, ) =
V
X
N =0
eN
eH () =
N =0
P ,
i i =N
e(H ()
V
X
i )
e(H ()
i )
P ,
i i =N
The thermodynamic functions are defined in the usual way. The probability
for a configuration {0, 1} is
P
e(H () i i )
.
Z (, )
The Hamiltonian is of the form
H () =
i j (qi qj ),
i,j, i6=j
where qi is the position vector of the site i . However this is too difficult
to solve in general. We consider two simplifications of the potential energy:
Mean-field Models: is taken to be a constant. Therefore
X
H () =
i j .
i,j, i6=j
Note that for Mean-field models the lattice structure is not important since
the interaction does not depend on the location of the lattice sites and therefore we can take = {1, 2, . . . , V }.
Nearest-neighbour Models: In these models we take
J , if |qi qj | = 1
(qi qj ) =
,
0 , if |qi qj | > 1
that is the interaction is only between nearest neighbours and is then equal
to J, J R. If we denote a pair of neighbouring sites i and j by hi, ji we
have
X
H () = J
i j .
hi,ji
9.2
Magnetic Models
In magnetic models at each site of there is a dipole or spin. This spin could
be pointing upwards or downwards, that is, along the direction of the external
magnetic field or in the opposite direction. For i we set i = 1 if the spin
at the site i is pointing upwards and i = 1 if it is pointing downwards.
The term {1, 1} is called a configuration. For a configuration we
have an energyP
E() and an interaction with an external magnetic field of
strength h, h i i . The partition function is then
X
Z (, h) =
e(E()h
{1,1}
1
log Z (, h).
V
e(E()h i i )
.
Z (, h)
The total magnetic moment is the random variable
X
M () =
i
i
75
i )
and therefore
P
E(M ) =
{1,1}
P
P
( i i )e(E()h i i )
Z (, h)
1
log Z (, h).
h
E(M )
= f (, h).
V
h
Note that
f (, h).
h
1
1 X
i (b + )V,
i = E() (a + )
2 i
2
i
1
1
(, ) = (1 + m (, a + )).
2
2
76
9.3
Curie-Weiss model
i j =
ii<jV
V
X
2V
!2
i
i=1
, {1, 1} ,
where > 0 and is any finite set with || = V . We sketch here only
some explicit calculations, more on the model can be found in the books
[Ell85],[Dor99], [Rei98], and [TKS92]. The partition function is given by
X
PV
Z (, h) =
e(E()h i=1 i ) .
{1,1}V
Z (, h) = e 2
exp
{1,1}V
2V
V
X
!2
i
+ h
V
X
i .
i=1
i=1
1 2
e 2 y dy = 2,
1 2
+ax)
e( 2 x
dx =
1 2
2e 2 a
or
Z
1 2
1
e
=
e( 2 x +ax) dx.
2
p PV
Using this identity with a = V
i=1 i we get
1 2
a
2
#
r
V
X
1 2
Z (, h) = e
exp x + x
+ h
i dx
2
V
i=1
{1,1}V
" r
#
Z
V
X
X
1 2
1
exp x
= e 2
e 2 x
+ h
i dx.
V
2
V
i=1
2
"
{1,1}
77
Now
X
exp(
V
X
i ) = (2 cosh )V .
i=1
{1,1}V
Therefore
Z (, h) = e
Putting =
x
V
"
21 x2
#
V
r2
+ h
2 cosh x
dx.
V
, we get
2 V
Z (, h) = e
= e 2 2V
V
2
12 Z
V
2
12 Z
V
2
exp(
) cosh( + h) d
2
eV G(h,) d,
where
1
G(h, ) = 2 + log cosh( + h).
2
The free energy per lattice site is
1
1
1
f (, h) =
log Z (, h) =
log 2
log
V
2V
2V
Z
1
log
eV G(h,) d.
V
2
Therefore by Laplaces Theorem (see for example [Ell85] or [Dor99]), the free
energy per lattice site in the thermodynamic limit is
Z
1
1
f (, h) = log 2 lim
log
eV G(h,) d
V V
1
1
= log 2 sup G(h, ).
R
Suppose that the supremum of G(h, ) is attained at (h). Then
1
1
f (, h) = log 2 G(h, (h))
and
G
(h, (h)) = (h) + tanh((h, (h)) + h) = 0,
78
y
y= tanh (
+h
y=
y= tanh (
h/
m0
(h )
1
f (, h) =
G(h, (h))
h
h
G
= tanh((h) + h) +
(h, (h)) (h)
h
= (h),
m(, h) =
since
G
(h, (h))
= 0.
Since
f (, h) = f (, h)
and
it is sufficient to consider the case h 0 (see Figure 7 and 8). The expression
m0 = limh0 m(h) is called the spontaneous magnetisation, this is the
mean magnetisation as the magnetic field is decreased to zero,
m0 = lim m(h) = lim (h).
h0
m0
h0
= 0 , if 1
.
> 0 , if > 1
79
y= tanh (
+h
y=
)
y= tanh (
h/
(h )
Figure 6: h > 0, 1
-f ( ,
h
em
p
o
l
s
1_ ln!2
h
Figure 7: > 1
80
-f ( ,
h
_1 ln!2
h
Figure 8: 1
We can consider this model from the point of view of a lattice gas. Consider
a lattice gas with potential energy
H () =
V
Let ti =
(i +1)
.
2
X
ii<jV
i j =
2V
V
X
i=1
Then
V
X
i=1
V
1 X
i = (
i + V ).
2 i=1
81
!2
ti
V
X
i
+
2V i=1
Therefore
H () =
8V
V
X
!2
i
i=1
V
V
X
V
X
i
+
i + .
4 i=1
8
4V i=1
4
We can neglect the last two terms, because is small and the expectation of
a single spin is zero, and we take
H () =
8V
V
X
!2
i
i=1
V
X
V
i
.
4 i=1
8
Then
H ()
V
X
i=1
i =
8V
V
X
!2
i
i=1
X
( + )
i ( + )V
4 2 i=1
8 2
V
X
= E() ( + )
i ( + )V.
4 2 i=1
8 2
with =
and
and
1
1
(, ) = ( + ) f (, + )
8 2
4 2
1
1
(, ) = (1 + m(, + )).
2
4 2
Let 0 = 2 , then
1
1
(, ) = ( + ) f (, ( 0 ))
8 2
2
and
1
1
(, ) = (1 + m(, ( 0 ))).
2
2
4
If > , then (, ) has a discontinuity in its derivative at 0 and (, )
has a discontinuity at 0 (see Figure 9).
82
( , )
0
( , )
Figure 9: >
83
9.4
In the continuous Ising model the state space E = {1, +1} is replaced by
d
the real numbers R. Let = RZ denote the space of configurations. Due
to the non-compactness of the state space severe mathematical difficulties
arise. We note that the continuous Ising model can be seen as an effective
model describing the height of an interface, here the functions give the
height of an interface for some reference height; and any collection (x )xZd or
probability measure P P(, F) is called random field of heights. Details
about this model can be found in [Gia00] and [Fun05]. One first considers
the so-called massive model, where there is a mass m > 0 implying a selfinteraction. Let S, and m > 0. We write synonymously x = (x)
for . Nearest neighbour heights do interact with an elastic interaction
potential V : R R, which we assume to be strictly convex with quadratic
growth, and which depends only on the difference in the heights of the nearest
2
neighbours. In the simplest case V (r) = r2 one gets the Hamiltonian
H ()
X m2
x
2x +
1 X
(x y )2 ,
4d x,y
|xy|=1
1
eH () (d),
Z ()
where
(d) =
dx
x (dx )
x
/
is the product of the Lebesgue measure at each single site in and the Dirac
measure at x for x c . The term is called reference measure in with
boundary . The thermodynamic limit exists for the model with m > 0
in any dimension. However, for the most interesting case m = 0 this exists
only for d 3. These models are called massless models or harmonic
crystals. The interesting feature of these models is that there are infinitely
many Gibbs measures due to the continuous symmetry. Hence we are in
a regime of phase transitions (see [BD93] for some rigorous results for this
regime). The massless models have been studied intensively during the last
fifteen years (see [Gia00] for an overview). The main technique applied is
the random walk representation. This can be achieved when one employs
summation by parts to obtain a discrete elliptic problem.
84
85
References
[AA68]
[Ada01]
S. Adams. Complete Equivalence of the Gibbs Ensembles for onedimensional Markov Systems. Journal Stat. Phys., 105(5/6), 2001.
[AGL78] M. Aizenman, S. Goldstein, and J.L. Lebowitz. Conditional Equilibrium and the Equivalence of Microcanonical and Grandcanonical Ensembles in the Thermodynamic limit. Commun. Math.
Phys., 62:279302, 1978.
[Aiz80]
M. Aizenman. Instability of phase coexistence and translation invariance in two dimensions. Number 116 in Lecture Notes in
Physics. Springer, 1980.
[AL06]
[Bal91]
[Bal92]
[BD93]
[Bir31]
G.D. Birkhoff. Proof of the ergodic theorem. Proc. Nat. Acad. Sci.
USA, 17:656600, 1931.
[BL99a]
G.M. Bell and D.A. Lavis. Statistical Mechanics of Lattice Systems, volume I. Springer-Verlag, 2nd edition, 1999.
[BL99b]
G.M. Bell and D.A. Lavis. Statistical Mechanics of Lattice Systems, volume II. Springer-Verlag, 1999.
[Bol84]
L. Boltzmann. Uber
die Eigenschaften monozyklischer und anderer damit verwandter Systeme, volume III of Wissenschaftliche
Abhandlungen. Chelsea, New York, 1884. reprint 1968.
[Bol74]
[Bre57]
[CK81]
[dH00]
R.L. Dobrushin. Investigation of Gibbsian states for three dimensional lattice systems. Theor. Prob. Appl., 18:253271, 1973.
[Dor99]
T.C. Dorlas. Statistical Mechanics, Fundamentals and Model Solutions. IOP, 1999.
[DZ98]
A. Dembo and O. Zeitouni. Large Deviations Techniques and Applications. Springer Verlag, 1998.
[EL02]
[Ell85]
[FO88]
[Fol73]
[FS97]
[Fun05]
[Gal99]
[Geo79]
[Geo88]
[Geo93]
[Geo95]
[Gib02]
J.W. Gibbs. Elementary principles of statistical mechanics, developed with special reference to the rational foundations of thermodynamics. Scribner, New York, 1902.
[GM67]
[Hua87]
[Isi24]
E. Ising. Beitrag zur theorie des ferro- und paramagnetismus. Dissertation, Mathematisch-Naturwissenschaftliche Fakultat der Universitat Hamburg, 1924.
[Isr79]
88
[Jay89]
[Khi49]
[Khi57]
[Kur60]
[KW41]
[Len20]
[LL72]
[LPS95]
J.T. Lewis, C.E. Pfister, and W.G. Sullivan. Entropy, concentration of probability and conditional limit theorems. Markov Process.
Related Fields, 1(3):319386, 1995.
[LR69]
[Oll88]
[Pei36]
[Rei98]
89
[RR67]
[Rue69]
[Rue78]
D. Ruelle. Thermodynamic formalism: The Mathematical Structures of Classical Equilibrium. Addison-Wesley, 1978.
[Sha48]
[Shl83]
[SW49]
C.E. Shannon and W. Weaver. The mathematical theory of information. University of Illinois Press, Budapest, 1949.
[Tho74]
R.L. Thompson. Equilibrium States on Thin Energy Shells. Memoirs of the American Mathematical Society. AMS, 1974.
[Tho79]
[Tho88]
C. J. Thompson.
Clarendon, 1988.
Addison-
90