You are on page 1of 186

Dirty tricks for statistical mechanics

Martin Grant
Physics Department, McGill University

°
c MG, August 2004, version 0.91
ii

Preface
These are lecture notes for PHYS 559, Advanced Statistical Mechanics,
which I’ve taught at McGill for many years. I’m intending to tidy this up
into a book, or rather the first half of a book. This half is on equilibrium, the
second half would be on dynamics.
These were handwritten notes which were heroically typed by Ryan Porter
over the summer of 2004, and may someday be properly proof-read by me.
Even better, maybe someday I will revise the reasoning in some of the sections.
Some of it can be argued better, but it was too much trouble to rewrite the
handwritten notes. I am also trying to come up with a good book title, amongst
other things. The two titles I am thinking of are “Dirty tricks for statistical
mechanics”, and “Valhalla, we are coming!”. Clearly, more thinking is necessary,
and suggestions are welcome.
While these lecture notes have gotten longer and longer until they are al-
most self-sufficient, it is always nice to have real books to look over. My favorite
modern text is “Lectures on Phase Transitions and the Renormalisation Group”,
by Nigel Goldenfeld (Addison-Wesley, Reading Mass., 1992). This is referred
to several times in the notes. Other nice texts are “Statistical Physics”, by
L. D. Landau and E. M. Lifshitz (Addison-Wesley, Reading Mass., 1970) par-
ticularly Chaps. 1, 12, and 14; “Statistical Mechanics”, by S.-K. Ma (World
Science, Phila., 1986) particularly Chaps. 12, 13, and 28; and “Hydrodynamic
Fluctuations, Broken Symmetry, and Correlation Functions”, by D. Forster (W.
A. Benjamin, Reading Mass., 1975), particularly Chap. 2. These are all great
books, with Landau’s being the classic. Unfortunately, except for Goldenfeld’s
book, they’re all a little out of date. Some other good books are mentioned in
passing in the notes.
Finally, at the present time while I figure out what I am doing, you can read
these notes as many times you like, make as many copies as you like, and refer
to them as much as you like. But please don’t sell them to anyone, or use these
notes without attribution.
Contents

1 Motivation 1

2 Review of Stat Mech Ideas 5

3 Independence of Parts 7
3.1 Self Averaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2 Central Limit Theorem . . . . . . . . . . . . . . . . . . . . . . . 11
3.3 Correlations in Space . . . . . . . . . . . . . . . . . . . . . . . . . 15

4 Toy Fractals 21
4.1 Von Koch Snowflake . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.2 Sierpinskii Triangle . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.3 Correlations in Self-Similar Objects . . . . . . . . . . . . . . . . . 27

5 Molecular Dynamics 31

6 Monte Carlo and the Master Equation 39

7 Review of Thermodynamics 49

8 Statistical Mechanics 55

9 Fluctuations 59
9.1 Thermodynamic Fluctuations . . . . . . . . . . . . . . . . . . . . 62

10 Fluctuations of Surfaces 67
10.1 Lattice Models of Surfaces . . . . . . . . . . . . . . . . . . . . . . 70
10.2 Continuum Model of Surface . . . . . . . . . . . . . . . . . . . . 72
10.3 Impossibility of Phase Coexistence in d=1 . . . . . . . . . . . . . 88
10.4 Numbers for d=3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

11 Broken Symmetry and Correlation Functions 93


11.1 Proof of Goldstone’s Theorem . . . . . . . . . . . . . . . . . . . . 96

iii
iv CONTENTS

12 Scattering 99
12.1 Scattering from a Flat Interface . . . . . . . . . . . . . . . . . . . 103
12.2 Roughness and Diffuseness . . . . . . . . . . . . . . . . . . . . . . 106
12.3 Scattering From Many Flat Randomly–Oriented Surfaces . . . . 108

13 Fluctuations and Crystalline Order 113

14 Ising Model 121


14.1 Lattice–Gas Model . . . . . . . . . . . . . . . . . . . . . . . . . . 126
14.2 One–dimensional Ising Model . . . . . . . . . . . . . . . . . . . . 127
14.3 Mean–Field Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 130

15 Landan Theory and Ornstien–Zernicke Theory 139


15.1 Landan Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
15.2 Ornstein — Zernicke Theory of Correlations . . . . . . . . . . . . 145

16 Scaling Theory 151


16.1 Scaling With ξ . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

17 Renormalization Group 161


17.1 ǫ–Expansion RG . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
Chapter 1

Motivation

Why is a coin toss random?


Statistical Mechanics: Weighting for averages is
by “equal a priori probability.” So, all many—body
quantum states with the same energy (for a given
fixed volume) have equal weight for calculating prob-
abilities.
This is a mouthful, but at least it gives the right
answer. There are two microscopic states, HEADS
and TAILS, which are energetically equivalent. So
the result is random.
However nice it is to get the right answer, and however satisfying it is to
have a prescription for doing other problems, this is somewhat unsatisfying. It
is worth going a little deeper to understand how microscopic states (HEADS or
TAILS) become macroscopic observations (randomly HEADS or TAILS).
To focus our thinking it is useful to do an experiment. Take a coin and flip
it, always preparing it as heads. If you flip it at least 10–20cm in the air, the
result is random. If you flip it less than this height, the result is not random.
Say you make N attemps, and the number of heads is nheads and the number
of tails is ntails , for different flipping heights h. Then you obtain something like
this:

0 2 4 6 8 10

h
Figure 1.1: Bias of Flips With Height h

1
2 CHAPTER 1. MOTIVATION

There are a lot of other things one could experimentally measure. In fact, my
cartoon should look exponential, for reasons we will see later on in the course.
Now we have something to think about: Why is the coin toss biased for
flips of small height? In statistical mechanics language we would say, “why is
the coin toss correlated with its initial state?” Remember the coin was always
prepared as heads. (Of course, if it was prepared randomly, why flip it?) There
is a length scale over which those correlations persist. For our example, this
scale is ∼ 2–3cm, which we call ξ, the correlation length. The tailing off of the
graph above follows:
Correlations ∼ e−ξ/h
which we need more background to show. But if we did a careful experiment,
we could show this is true for large h. It is sensible that

ξ ≥ coin diameter

as a little thinking will clarify.


This is still unsatisfying though. It leads to the argument that coin tosses are
random because we flip the coin much higher than its correlation length. Also,
we learn that true randomness requires ξ/h → 0, which is analogous to requiring
an infinite system in the thermodynamic limit. [In fact, in passing, note that
for ξ/h = O(′) the outcome of the flip (a preponderance of heads) does not
reflect the microscopic states (HEADS and TAILS equivalent). To stretch a bit,
this is related to broken symmetry in another context. That is, the macroscopic
state has a lower symmetry than the microscopic state. This happens in phase
transitions, where ξ/L = O(′) because ξ ≈ L as L → ∞, where L is the edge
length of the system. It also happens in quantum mesoscopic systems, where
ξ/L = O(′) because L ≈ ξ and L is somewhat small. This latter case is more
like the present examlpe of coin tossing.]
But what about Newton’s laws? Given an initial configuration, and a precise
force, we can calculate the trajectory of the coin if we neglect inhomogeneities in
the air. The reason is that your thumb cannot apply a precise force for flipping to
within one correlation length of height. The analogous physical problem is that
macroscopic operating conditions (a fixed pressure or energy, for example) do
not determine microscopic states. Many such microscopic states are consistent
with one macroscopic state.
These ideas, correlations on a microscopic scale, broken symmetries, and
correlation lengths, will come up throughout the rest of the course. The rest of
the course follows this scenario:

• Fluctuations and correlations: What can we do without statistical me-


chanics?

• What can we do with thermodynamics and statistical mechanics?

• Detailed study of interfaces.

• Detailed study of phase transitions.


3

The topics we will cover are as current as any issue of Physical Review E. Just
take a look.
We’ll begin the proper course with a review of, and a reflection on, some
basic ideas of statistical mechanics.
4 CHAPTER 1. MOTIVATION
Chapter 2

Review of Stat Mech Ideas

In statistical physics, we make a few assumptions, and obtain a distrobution


to do averages. A famous assumption is that of “equal a priori probability.”
By this it is meant that all microscopic states, consistent with a paticular fixed
total energy E and number of particles N , are equally likely. This gives the
microcanonical ensemble. It can then be checked that thermodynamic quanti-
ties have exceedingly small fluctuations, and that one can derive — from the
microcanonical ensemble — the canonical ensemble, where temperature T and
volume V are fixed. It is this latter ensemble which is most commonly used.
Microscopic states, consistent with T and V are not equally likely, they are
weighted with the probability factor

e−Estate /kB T (2.1)

where Estate is the energy of the state, and kB is Boltzmann’s Constant. The
probability of being in a state is
1 −Estate /kB T
ρ= e (2.2)
Z
where the partition function is
X
Z= e−Estate /kB T (2.3)
{states}

The connection to thermodynamics is from the Helmholtz free energe

F = E − TS (2.4)

where S is the entropy, and


e−F/kB T = Z (2.5)
Note that since F is extensive, that is F ∼ O(N ), we have
−f
Z = (e kB T )N

5
6 CHAPTER 2. REVIEW OF STAT MECH IDEAS

where F = f N . This exponential factor of N arises from the sum over states
which is typically X
= eO(N ) (2.6)
{states}

For example, for a magnetic dipole, which can have each constivent up or down,
there are evidently

dipole #1 × dipole #2...dipole #N = 2 × 2 × ...2 = 2N = eN ln 2 (2.7)

total states. The total number of states in a volume V is


VN VN V
≈ N = eN ln N (2.8)
N! N
where the factor N ! accounts for indistinguisability.
These numbers are huge in a way which cannot be imagined. Our normal
scales sit somewhere between microscopic and astronomical. Microscopic sizes
are O(Ȧ) with times O(10−12 sec). Astronomical sizes, like the size and age of
the universe are O(1030 m) and O(1018 s). The ratio of these numbers is

1030 m
∼ 1040 (2.9)
10−10 m
This is in no way comparable to (!)
23
1010 (2.10)

In fact, it seems a little odd (and unfair) that a theorist must average over eN
states to explain what an experimentalist does in one experiment.
Chapter 3

Independence of Parts

It is worth thinking about experiments a bit to make this clear. Say an ex-
perimentalist measures the speed of sound in a sample, as shown. The experi-

Figure 3.1: Experiment 1

mentalist obtains some definite value of c, the speed of sound. Afterwards, the
experimentlalist might cut up the sample into many parts so that his or her
friends can do the same experiment. That is as shown in Fig. 3.2. Each of

Figure 3.2: Experiment 2

7
8 CHAPTER 3. INDEPENDENCE OF PARTS

these new groups also measure the sound speed as shown. It would seem that
the statistics improve because 4 experiments are now done?! In fact one can
continue breaking up the system like this, as shown in Fig. 3.3.

Figure 3.3: Experiment 3

Of course, 1 experiment on the big system must be equivalent to many


experiments on individual parts of a big system, as long as those individual
parts are independent. By this it is meant that a system “self-averages” all its
independent parts.
The number of independent parts is (to use the same notation as the number
of particles)
L3
N = O( 3 ) (3.1)
ξ
for a system of edge length L. The new quantity introduced here is the cor-
relation length ξ, the scale over which things are correlated, so they are not
independent. The correlation length must be at least as large as an atom so

ξ ≥ O(Ȧ) (3.2)

and indeed usually ξ ≈ O(Ȧ). Hence for a macroscopic length, the number of
independent points is
N ∼ O(1023 ) (3.3)
3.1. SELF AVERAGING 9

of course. Now, each independent part has a few degrees of freedom, such as
quantum numbers. Hence, the total number of states in a typical system are
R R R
dq1 dq2 ... dqN
quantum quantum quantum
states states states
of part 1 of part 2 of part N
R
Say dqi = 4 for an example, so that the total number of states is

eO(N ) (3.4)

as it should. Now, however, we have a feeling for the origin of some of the
assumptions entering statistical physics.
We can go a little farther with this, and prove the central–limit theorem of
probability theory.

3.1 Self Averaging


Consider an extensive quantity X. If a system consists of N independent parts
as shown, then

i=1 i=2 i=3

i=4 i=5 i=6

i=7 i=8 i=N

Figure 3.4: N independent parts

N
X
X= Xi (3.5)
i=1

where Xi is the value of X in independent part i. The average value of X is


N
X
hXi = hXi i (3.6)
i=1

But since hXi i =


6 hXi i(N ), we have

hXi = O(N ) (3.7)

It is handy to introduce the intensive quantity


X
x= (3.8)
N
10 CHAPTER 3. INDEPENDENCE OF PARTS

then
hxi = O(1)
This is not very amazing yet. Before continuing, it is worth giving an example
of what we mean by the average. For example, it might be a microscopically
long, but macroscopically short time average
R
dt X(t)
hXi = τ
τ
as drawn. As also indicated, the quantity h(X − hXi)2 i1/2 is a natural one to

X
h(X − hXi)2 i1/2
hXi

τ t

Figure 3.5: Time Average

consider. Let the deviation of X from its average value be

∆X = X − hXi (3.9)

Then clearly
h∆Xi = hXi − hXi = 0
as is clear from the cartoon. The handier quantity is
h(∆X)2 i = h(X − hXi)2 i
= hX 2 − 2XhXi + hXi2 i (3.10)
2 2
= hX i − hXi
Giving the condition, in passing,

hX 2 i ≥ hXi2 (3.11)

since h(∆X)2 i ≥ 0. Splitting the system into independent parts, again, gives

XN N
X
h(∆X)2 i = h ∆Xi ∆Xj i
i=1 j=1

Now using that i may or may not equal j


N
X N
X
h(∆X)2 i = h (∆Xi )2 i + h (∆Xi )(∆Xj )i (3.12)
i=1 i=1
(i=j) j=1
(i6=j)
3.2. CENTRAL LIMIT THEOREM 11

But box i is independent of box j, so,

h∆Xi ∆Xj i = h∆Xi ih∆Xj i


=0×0 (3.13)
=0

and,
N
X
h(∆X)2 i = h(∆Xi )2 i
i=1
or
h(∆X)2 i = O(N ) (3.14)
Also, the intensive x = X/N satisfies
1
h(∆x)2 i = O( ) (3.15)
N
This implies that fluctuations in quantities are very small in systems of many
(N → ∞) independent parts. That is, because

h(∆X)2 i1/2
X ≈ hXi{1 + O( )}
hXi (3.16)
1
≈ hXi{1 + O( 1/2 )}
N
So, in essence, X ≈ hXi. The result

h(∆X)2 i 1
= O( ) (3.17)
hXi2 N

that relative fluctuations are very small, if N is large, is called the self-averaging
lemma. Note that, since N = V /ξ 3 ,

h(∆X)2 i ξ3
2
= O( ) (3.18)
hXi V

so fluctuations measure microscopic correlations, and become appreciable for


ξ3
V ∼ O(1). This occurs in quantum mesoscopic systems (where V is small),
and phase transitions (where ξ is large), for example.

3.2 Central Limit Theorem


We will now prove (to a physicist’s rigor) the central–limit theorem for systems
if many independent parts. In the same manner as above we will calculate all
the moments, such as the nth moment,

h(∆X)n i (3.19)
12 CHAPTER 3. INDEPENDENCE OF PARTS

or at least their order with N .


For simplicity we will assume that all the odd moments vanish

h(∆X)2n+1 i = 0

for intiger n. This means the distrobution ρ is symetric about the mean hXi as
shown. It turns out that one can show that the distrobution function is symetric
ρ ρ ρ

hxi x hxi x hxi x

Figure 3.6: Symetric distribution functions ρ(x)

as (N → ∞), but lets just assume it so the ideas in the algebra are clear. It can
be shown (see below) that we can then reconstruct the distrobution ρ(X) if we
know all the even moments
h(∆X)2n i
for positive integer n. We know n = 1, so let’s do n = 2, as we did h(∆X)2 i
above

XN N
X N
X N
X
h(∆X)4 i = h ∆Xi ∆Xj ∆Xk ∆Xl i
i=1 j=1 k=1 l=1
N N
(∆Xi )2 (∆Xj )2 i + Oh∆Xi:
X X 0 (3.20)
=h (∆Xi )4 i + 3h i
i=1 1=1
(i=j=k=l) j=1
(i6=j)

= O(N ) + O(N 2 )

So as (N → ∞)
N
X N
X
h(∆X)4 i = 3h (∆Xi )2 ih (∆Xj )2 i
i=1 j=1

or
h(∆X)4 i = 3h(∆X)2 i2 = O(N 2 ) (3.21)
Note that the largest term (as (N → ∞)) came from pairing up all the ∆Xi ’s
appropriately. A little thought makes it clear that the largest term in h(∆X)2n i
will result from this sort of pairing giving,

h(∆X)2n i = const. h(∆X)2 in = O(N n ) (3.22)


3.2. CENTRAL LIMIT THEOREM 13

We will now determine the const. in Eq. 3.22. Note that

h(∆X)2n i = h∆X ∆X ∆X...∆X ∆Xi

There are (2n − 1) ways to pair off the first ∆X. After those two ∆X’s are
gone, there are then (2n − 3) ways to pair off the next free ∆X. And so on,
until finally, we get to the last two ∆X’s which have 1 way to pair off. Hence

h(∆X)2n i = (2n − 1)(2n − 3)(2n − 5)...(1)h(∆X)2 in ≡ (2n − 1)!!h(∆X)2 in

or more conventionally,

2n!
h(∆X)2n i = h(∆X)2 in (3.23)
2n n!
Now we know all the even moments in terms of the first two moments hXi and
h(∆X)2 i. We can reconstruct the distribution with a trick. First, let us be
explicit about ρ. By an average we shall mean,
Z ∞
hQ(X)iX = ρ(X)Q(X) (3.24)
−∞

where Q(X) is some quantity. Note that



hδ(X − X )iX ′ = ρ(X) (3.25)

where δ(X) is the Dirac delta function! Recall the Fourrier representation
Z ∞
′ 1 ′
δ(X − X ) = dk e−ik(X−X ) (3.26)
2π −∞

Averaging over X gives
Z ∞
′ 1 ′
ρ(X) = hδ(X − X )iX ′ = dk e−ikX heikX iX ′ (3.27)
2π −∞

In statistics, heikX i is called the characteristic function. But we can now expand
the exponential (for convenience, let hXi = 0 for now, one only has to add and
′ ′
subtract it in e−ik(X−X ) = e−ik(∆X−∆X ) but it makes the algebra messier) as

′ X (ik)n ′ n
e−ikX = (X )
n=0
n!

So that we have,
Z ∞ ∞
1 ikX
X (ik)n n
ρ(X) = dk e hX i (3.28)
2π −∞ n=0
n!
14 CHAPTER 3. INDEPENDENCE OF PARTS

where the prime superscript has been dropped. Note that Eq. 3.28 means that
ρ(X) can be reconstructed from all the hX n i, in general. For our case, all the
odd moments vanish, so
Z ∞ ∞
1 X (−k 2 )n 2n
ρ(X) = dk eikX hX i
2π −∞ n=0
(2n)!

(2n)!
But hX 2n i = 2 n
2n n! hX i from above, so,
Z ∞ ∞
1 X 1 −k 2 2 n
ρ(X) = dk eikX ( hX i)
2π −∞ n=0
n! 2

which is the exponential again, so,


Z ∞
1 −k2 2
ρ(X) = dk eikX e 2 hX i (3.29)
2π −∞
−k2 2
(In passing, it is worth remembering that heikX i = e 2 hX i for a Gaussian
distribution.) The integral can be done by completing the square, adding and
−X 2
subtracting 2hX 2 i , so that

−X 2
Z ∞ khX 2 i1/2
1 2hX −(
21/2
+ iX
2hX 2 i1/2
)2
ρ(X) = e 2i dk e
2π −∞

π
−X 2
s Z ∞ *
1 2hX 2 2
= e 2i du e−u
2π hX 2 i −∞

using an obvious sudstitution. Hence, on putting hXi as non–zero again,


−(X−hXi)2
1 2h(∆X)2 i
ρ(X) = p e (3.30)
2πh(∆X)2 i
which is the Gaussian distribution. This is the central–limit theorem for a
system of many independent parts. Its consequences are most transparent if
one deals with the intensive quantity x = X
N , then recall

hxi = O(1)
(3.31)
h(∆x)2 i = O(1/N )
Letting
a2
h(∆x)2 i ≡ (3.32)
N
where a2 = O(1) we have, up to normalization,
(x−hxi)2
ρ(x) ∝ e−N 2a2 (3.33)
so the width of the Gaussian is exceedingly small and it most represents a spike.
3.3. CORRELATIONS IN SPACE 15

ρ(x) O(1/N 1/2 )

hxi = O(1) x

Figure 3.7: Gaussian (Width Exaggerated)

Which is to say again that fluctuations are small, and average values are well

ρ(x) O(1/N 1/2 )

hxi = O(1) x

Figure 3.8: Gaussian (Width (sort of) to scale)

defined for a system of many (N → ∞) independent parts.

3.3 Correlations in Space


The results so far apply to a large system, and perhaps it is not surprising that
the global fluctuations in a large system are small. These results for the global
fluctuations have implications for the correlations of quantities, though. Say
the intensive variable x has different values at points ~r in space. (Precisely, let
d~r, the volume element, be microscopically large, but macroscopically small).
Consider the correlation function
′ ′
h∆x(~r) ∆x(~r )i ≡ C(~r, ~r ) (3.34)

We can say a few things about C(~r, ~r ) from general considerations.

1. If space is translationally invariant, then the origin of the coordinate sys-


′ ′
tem for ~r and ~r is arbitrary. It can be placed at point ~r or ~r . The
′ ′
physical length of intrest is (~r − ~r ) or (~r − ~r). Hence
′ ′ ′
C(~r, ~r ) = C(~r − ~r ) = C(~r − ~r) (3.35)
16 CHAPTER 3. INDEPENDENCE OF PARTS

∆x(~r)
~r − ~r′
∆x(~r′)
~r

~r′
o

Figure 3.9: Transitional invariance (only ~r − ~r′ is important)

and, of course,
′ ′
h∆x(~r) ∆x(~r )i = h∆x(~r − ~r ) ∆x(0)i (3.36)
etc.

2. If space is isotropic, then the direction of ∆x(~r) from ∆x(~r ) is unim-

portant, and only the magnitude of the distance | ~r − ~r | is physically
∆x(~r′)

|~r − ~r′|
∆x(~r)

Figure 3.10: Isotropy (only |~r − ~r′| is important)

relevant. Hence ′ ′
C(~r, ~r ) = C(| ~r − ~r |) (3.37)

So what was a function of 6 numbers (~r, ~r ) is now a function of only 1,

| ~r − ~r |.
3. Usually correlations fall off if points are arbitrarily far apart. That is,
: 0 ∆x(0)i
h∆x(r → ∞) ∆x(0)i = h∆x(r → ∞)ih :0
=0
or
C(r → ∞) = 0 (3.38)
The scale of the correlation is set by the correlation length.
Now we know ′ ′
C(| ~r − ~r |) = h∆x(~r) ∆x(~r )i (3.39)

Subject to these conditions. A constraint on C(| ~r − ~r |) is provided by our
previous result
1
h(∆x)2 i = O( )
N
3.3. CORRELATIONS IN SPACE 17

C(r) O(ξ)

Figure 3.11: Typical correlation function

for a global fluctuation. Clearly,


R
d~r ∆x(~r)
∆x = R
d~r
Z (3.40)
1
= d~r ∆x(~r)
V
So we have
Z Z
1 ′ ′
d~r d~r h∆x(~r) ∆x(~r )i = h(∆x)2 i
V
(3.41)
1
= O( )
N
or, Z Z
′ ′ V2
d~r d~r C(~r − ~r ) = O( )
N
~ = ~r − ~r′ , R
Let R ~′ = 1 ′

2 (~
r − ~r ). The Jacobian of the transformation is unity,
and we obtain
Z Z 2
~′
dR ~ = O( V )
~ C(R)
dR
N
Z (3.42)
=V ~ C(R)
dR ~

~ → ~r,
So letting R Z
V
d~r C(~r) = O( )
N
V Ld
But N = ξ3 , or in d dimensions, N = ξd
, so
Z
d~r C(~r) = O(ξ 3 )

or
Z
d~r h∆x(~r) ∆x(0)i = O(ξ 3 )
(3.43)
18 CHAPTER 3. INDEPENDENCE OF PARTS

for a system of N → ∞ independant parts. This provides a constraint on


correlations. It is a neat result in Fourier space. Let
Z
~ ~
Ĉ(k) ≡ d~r eik·~r C(~r)

and finally
Z
d~k ~
C(~r) = d
e−ik·~r Ĉ(~k)
(2π)
(3.44)

and, Z
~
h|∆x̂(k)|2 i ≡ d~r eik·~r h∆x(~r) ∆x(0)i

Hence the constraint of Eq. 3.43 is

lim Ĉ(k) = O(ξ 3 )


k→0

or

lim h|∆x̂(k)|2 i = O(ξ 3 )


k→0
(3.45)

Many functions satisfy the constraint of Eq. 3.43 (see Fig. 3.11). It turns out
that usually,
h∆x(~r) ∆x(0)i ∼ e−r/ξ (3.46)
for large r, which satisfies Eq. 3.43. This is obtained from analyticity of (Ĉ(k))−1
near k = 0. It can be argued,
1
≈ (const.) + (const.)k 2 + ...
Ĉ(k)
const.
so that h|∆x̂(k)|2 i = const.+k 2 near k = 0. Fourier transformation gives the

result of Eq. 3.46 above.


(Note: In the previous notation with boxes, as in Fig. 3.4, we had
X
∆X = ∆Xi
i

or
1 X
∆x = ∆Xi
N i
This means Z
1 1 X
d~r ∆x(~r) = ∆Xi (3.47)
V N i
3.3. CORRELATIONS IN SPACE 19

This is satisfied by
N
V X
∆x(r) = δ(~r − ~ri )∆Xi
N i=1

where ~ri is a vector pointing to box i, and ∆Xi = ∆X(~ri ).)


Let me emphasize again that all these results follow from a system having
N → ∞ independent parts. To recapitulate, for x = X/N , where X is extensive,
N = V /ξ 3 is the number of independent parts. V is the volume, and ξ is the
correlation length:

hxi = O(1) (3.7)


1
h(∆x)2 i = O( ) (3.15)
N
(x−hxi)2
ρ(x) ∝ e 2h(∆x)2 i (3.30)
Z
d~r h∆x(r) ∆x(0)i = O(ξ 3 ) (3.43)

The purpose of this course is to examine the cases where V /ξ 3 = O(1) and these
results break down! It is worthwhile to give some toy examples now where, in
particular, Eq. 3.43 breaks down.
20 CHAPTER 3. INDEPENDENCE OF PARTS
Chapter 4

Toy Fractals

Toy Fractals are mathamatical constructs that seem to exist in fractional di-
mensions between 1 and 2, or, 2 and 3, or what have you. They can provide a
useful starting point for thinking about real systems where V /ξ 3 = O(1), or if
V = L3 , ξ/L = O(1).
We’ll briefly discuss self-affine, or surface fractals, and volume fractals. This

Figure 4.1: Self–Affine Fractal

is a self–affine fractal where the surface is very jagged, or as is said in physics,


rough. The self-affine fractal has the property that its perimeter length P (or
surface area in three dimensions) satisfies,

P ∼ Lds , as L → ∞ (4.1)

where ds is the surface self-affine fractal exponent. For a circle of diameter L

Pcircle = πL

while for a square of width L

Psquare = 4L

21
22 CHAPTER 4. TOY FRACTALS

In fact, for run-of-the-mill surfaces in d dimensions

Pnonf ractal ∼ Ld−1 (4.2)

Usually
ds > d − 1 (4.3)

I’ll give an example below. This is a volume fractal, note all the holes with

Figure 4.2: Volume Fractal

missing stuff. The fractal has the propert that its mass M satisfies

M ∼ Ldf as L → ∞ (4.4)

where df is the fractal dimension. For a circle M = π/4 L2 , in units of the


density for a diameter L, while for a square M = L3 . For run-of-the-mill
objects
Mnonf ractal ∼ Ld (4.5)

in d dimensions. Usually
df < d (4.6)

Note that since the mass density is M/Ld , this implies that

1
DEN SIT Y OF A F RACT AL ∼ →0 (4.7)
Ld−df

as L → ∞, implying fractals float (!?). Think about this.


Two toy examples show that ds > d − 1 and df < d are possible.

4.1 Von Koch Snowflake


To draw the von Koch snowflake, follows the steps shown. To find the self-affine
4.1. VON KOCH SNOWFLAKE 23

Figure 4.3: Von Koch Snowflake

fractal exponent for the perimeter, imagine I had (somehow) drawn one of these
fractals out a very very large number of steps. Say the smallest straight link is
length l. We’ll calculate the perimeter as we go through increasing numbers of
links in a large complex fractal. Hence

Figure 4.4:

Pn = 4n l

when

Ln = 3n l
(4.8)

Taking the logarithm

ln P/l = n ln 4

and

ln L/l = n ln 3

Dividing these into each other

ln P/l ln 4
=
ln L/l ln 3
24 CHAPTER 4. TOY FRACTALS

or

P/l = (L/l)ds (4.9)

where

ln 4
ds = ≈ 1.26
ln 3
>d−l =1

It is common in the physics literature to write

P (L) = w(L)Ld−1 (4.10)

where w(L) is the width or thickness of a rough interface, and

w(L) ∼ Lx , as L → ∞ (4.11)

where x is the roughening exponent, and clearly, (x ≈ 0.26 for the snowflake),

ds = x + d − 1 (4.12)

4.2 Sierpinskii Triangle


To draw the Sierpinskii triangle, follow the steps shown. To find the fractal

Figure 4.5: Sierpinskii Triangle

exponent for the mass, imagine I had (somehow) drawn one of these fractals
out to a very large number of steps. Say the smallest filled triangle had mass
m and volume equal to v. Then we’ll calculate the mass as we go through
increasing the numbers of triangles. Hence
4.2. SIERPINSKII TRIANGLE 25

Figure 4.6:

Mn = 3n m

when

Vn = 4n v
(4.13)

or since Vn = 21 L2n and v = 21 l2 we have,

L2n = 4n l2 or Ln = 2n l

Taking logarithms,
M
ln = n ln 3
m

and

L
ln = n ln 2
l

or, taking the ratio

ln M/m ln 3
=
ln L/l ln 2

and
M L
= ( )df (4.14)
m l

where

ln 3
df = ≈ 1.585 (4.15)
ln 2
So df < d where d = 2 here.
26 CHAPTER 4. TOY FRACTALS

This is not the common notation in the physics literature. Below, but near
a second–order phase transition, the excess density satisfies

n ∝ |T − Tc |β

where Tc is the critical temperatire. However the correlation length satisfies

ξ ∝ |T − Tc |−v

Or, since ξ’s largest value is limited by the edge length L of the entire system

ξ ∼ L ∼ |T − Tc |−v

Using this above gives

n ∝ [L−1/v ]β

or, Excess density near Tc

1
n∝
Lβ/v

where β and v are critical exponents. But, for a fractal

1
n∝ d − df
L
so
df = d − β/v (4.16)
where β/v ≈ 0.415 for the Sierpinskii snowflake.
For a real system (in the Ising universality class in two dimensions),

x = 1/2, so ds = 1 1/2 and β/v = 1/8, so df = 1 7/8

Figure 4.7:
4.3. CORRELATIONS IN SELF-SIMILAR OBJECTS 27

4.3 Correlations in Self-Similar Objects

Figure 4.8:

A striking thing about a fractal is that it “looks the same” on different


length scales. This is to say that your eye sees similar (or, for these toy fractals,
identical features) correlations if you look on length scales r or r/b where b > 1.
So far as density correlations go, let

C(r) = h∆n(r) ∆n(0)i

(where the average is, say, over all orientations in space for convenience). Then
the evident self–similarity of fractals implies

C(r) ∝ C(r/b)

or

C(r) = f (b) C(r/b)


(4.17)

It is easy to show that


f (b) = b−p (4.18)
where p is a constant. Take the derivative with respect to b of Eq. 4.17, letting

rk ≡ r/b (4.19)

Then
dC(r) df dC dr∗
= 0 = C(r∗ ) + f ∗
db db dr db
28 CHAPTER 4. TOY FRACTALS

or,

df r∗ dC
C(r∗ ) = + f
db b dr∗
and
d ln f (b) d ln C(r∗ )
= (4.20)
d ln b d ln r∗
Because one side only depends on b, and the other only on r∗ . Hence

d ln f (b)
= const. ≡ −p (4.21)
d ln b
So
f (b) ∝ b−p (4.22)

and the invariance relation is

C(r) = b−p C(r/b) (4.23)

(up to a multiplicative constant).


If one lets
b = 1 + ǫ, ǫ << 1 (4.24)

r
C(r) = (1 + ǫ)−p C( )
1+ǫ
∂C
≈ (1 − pǫ){C(r) − ǫr } + ...
∂r
Terms of order ǫ give
∂C
−p C(r) = r
∂r
The solution of which is
1
C∝ p (4.25)
r
Hence, self similarity implies homogeneous correlation functions

C(r) = b−p C(r/b)

and power–law correlation


1
C(r) ∝ (4.26)
rp
Usually, these apply for large r. The sign of p chosen to make C(r → ∞) →
0. Note that this is quite different from what occurs for a system of many
4.3. CORRELATIONS IN SELF-SIMILAR OBJECTS 29

C(r)

1/rp

Figure 4.9: Decay of Correlations in a Self–Similar Object

C(r)

e−r/ξ
ξ

Figure 4.10: Decay of Correlations in a System of Many Independent Parts

independent parts where (we said)

C(r) ∼ e−r/ξ

The restriction Z
d~rC(r) = O(ξ d )

in d–dimensions is evidently unsatisfied for


1
C(r) ∼
rp
if d > p, for then the integral diverges!
Usually,
p=2−η (4.27)
and η ≥ 0, so d > p in d = 2 and 3.
So, fractals, self–affine (rough) surfaces, and second–order phase transition
fluctuations do not satisfy the central limit theorem.
In fact, rough surfaces and phase transitions will be the focus for the rest of
the course.
30 CHAPTER 4. TOY FRACTALS
Chapter 5

Molecular Dynamics

At this point, one might say, computers are so big and fierce, why not simlpy
solve the microscopic problem completely! This is the approach of molecular
dynamics. It is a very important numerical technique, which is very very con-
servative. It is used by physicists, chemists, and biologists.
Imagine a classical solid, liquid or gas composed of many atoms or molecules.
Each atom has a position ~ri and a velocity ~vi (or momentum m~vi ). The Hamil-

particle i ~vi
~vj

|~ri − ~rj | particle j


~ri

~rj
o

Figure 5.1:

tonian is (where m is the atom’s mass)


N
X m~v 2 i
H= + V (~r1 , ...~rN )
i=1
2

Usually, the total potential can be well approximated as the sum of pairwise
forces, i.e.
N
X
V (r1 , ...rN ) = V (|~ri − ~rj |)
i=1, j=1
(i>j)

which only depends on the distance between atoms. This potential usually has

31
32 CHAPTER 5. MOLECULAR DYNAMICS

a form like that drawn. Where the depth of the well must be

ǫ
r0 r

Figure 5.2:

ǫ ∼ (100′ s K)kB

and

ro ∼ (a f ew Ȧ′ s)

This is because, in a vapor phase, the density is small and the well is unimpor-
tant; when the well becomes important, the atoms condense into a liquid phase.
Hence ǫ ∼ boiling point of a liquid, and ro is the separation of atoms in a liquid.
Of course, say the mass of an atom is

m ∼ 10 × 10−27 kg
ro ∼ 2Ȧ

then the liquid density is

m 10 × 10−27 kg
ρ∼ 3
=
ro 10Ȧ3
−27
10 × (103 gm)
=
(10−8 cm)3
ρ ∼ 1gm/cm3

of course. To convince oneself that quantum effects are unimportant, simlpy


estimate the de Broglie wavelength

λ∼ √
mkB T

∼ 1Ȧ/ T

where T is measured in Kelvin. So this effect is unimportant for most liq-


uids. The counter example is Helium. In any case, quantum effects are usually
33

marginal, and lead to no quantitative differences of importance. This is sim-


ply because the potentials are usually parameterized fits such as the Lennard -
Jones potential
σ σ
V (r) = 4ǫ[( )12 − ( )6 ]
r r

where

ro = 21/6 σ ≈ 1.1σ

and

V (ro ) = −ǫ

by construction. For Argon, a good “fit” is

ro = 3.4Ȧ
ǫ = 120K

Nummerically it is convenient that V (2ro ) << ǫ, and is practically zero. Hence


one usually considers

(
Lennard – Jones , r < 2ro
V (r) =
0 , r > 2ro

where ro = 21/6 σ. Lennard-Jones has a run–of–the–mill phase diagram which


is representative of Argon or Neon, as well as most simple run–of–the–mill pure
substances. Simpler potentials like the square well also have the same properties.

P T
s l and l
v s

v Coexistance

T ρ

Figure 5.3: Leonnard Jousium (The solid phase is fcc as I recall)


34 CHAPTER 5. MOLECULAR DYNAMICS

But the hard sphere potential has a more simple–minded phase diagram. This

V 
∞
 , r<σ
V = −ǫ , σ < r < σ′


0 , r > σ′

Figure 5.4:

(
V ∞ , r<σ
V =
0 , r>σ

r
where,

P T
s l
and s
l coex.

T ρ

Figure 5.5:

is because of the lack of an attractive interaction.


Finally, lets get to the molecular dynamics method. Newton’s equations (or
Hamilton’s) are,
d~ri
= ~vi
dt
~i
F
and ~ai = m
N
d~vi 1 X ∂
= V (|~ri − ~rj |)
dt m j=1 ∂~ri
(i6=j)

Clearly, we can simply solve these numerically given all the positions and veloc-
ities initially, i.e. {~ri (t = 0), ~vi (t = 0)}. A very simple (in fact too simple) way
35

to do this is by
d~ri ri (t + ∆t) − ri
=
dt ∆t
for small ∆t. This gives

~ri (t + ∆t) = ~ri (t) + ∆t ~vi (t)

and

1 ~
~ri (t + ∆t) = ~vi (t) + ∆t Fi
m

where F~i (t) is determined by {~ri (t)} only. A better (but still simple) method
is due to Verlet. It should be remembered that the discretization method has a
little bit of black magic in it.
Valet notes that,

(∆t)2
~ri (t ± ∆t) = ~ri (t) ± ∆t~vi + ~ai (t) + ...
2
d2 ~
where ~ai = F~i /m = ri
dt2 . Add this to itself

1 ~
~ri (t + ∆t) = 2~ri (t) − ~ri (t − ∆t) + (∆t)2 Fi + O(∆t)
m
And us the simple rule

ri (t + ∆t) − ri (t − ∆t)
~vi (t + ∆t) =
2∆t
Thats all. You have to keep one extra set of ri ’s, but it is the common method,
and works well. To do statistical mechanics, you do time averages, after the
system equilibrates. Or ensemble averages over different initial conditions. The
ensemble is microcanonical sine energy is fixed. Usually one solves it for a fixed
volume, using periodic boundary conditions. A common practice is to do a

Figure 5.6: Periodic Boundary Conditions


36 CHAPTER 5. MOLECULAR DYNAMICS

“cononical” ensemble by fixing temperature through the sum rule

1 1
mhv 2 i = kB T
2 2
or,
hv 2 i
T =
mkB
This is done by constantly rescaling the magnitude of the velocities to be
p
|v| ≡ hv 2 i1/2 ≡ mkB T

What about microscopic reversibility and the arrow of time?


Recall the original equations

∂~ri
= ~vi
∂t

∂~vi 1 X ∂
= V (|ri − rj |)
∂t m j ∂~ri
(j6=i)

The equations are invariant under time reversal t → −t if

~r → ~r
~v → −~v

How then can entropy increase? This is a deep and uninteresting question.
Entropy does increase; its a law of physics. The real question is how to calculate
how how fast entropy increases; e.g. calculate a viscosity or a diffusion constant,
which set the time scale over which entropy increases.
Nevertheless, here’s a neat way to fool around with this a bit. Take a gas
of N atoms in a box of size Ld where energy is fixed such that the average
equilibrium separation between atoms

seq = O(3 or 4)ro

where ro is a parameter giving the rough size of the atoms. Now initialize the
atoms in a corner of the box with

s(t = 0) = O(ro )

or smaller (see fig. 5.7). Now, at time t = later, let all ~v → −~v . By microscopic
reversibility you should get fig. 5.8. Try it! You get fig. 5.9. It is round off
37

t=0 t = later t = equilibrium

“Entropy”

later t

Figure 5.7:

t=0 t = later t = 2later


“Entropy”
1

later t

Figure 5.8:

“Entropy”
1

later t

Figure 5.9:

error, but try to improve your precision. You will always get the same thing.
Think about it. The second law is a law, not someone’s opinion.
38 CHAPTER 5. MOLECULAR DYNAMICS
Chapter 6

Monte Carlo and the


Master Equation

An improtant description for nonequilibrium statistical mechanics is provided by


the master equation. This equation also provides the motivation for the Monte
Carlo method, the most important numerical method in many–body physics.
The master equation, despite the pretentious name, is phenomenology fol-
lowing from a few reasonable assumptions. Say a system is defined by a state
{S}. This state might, for example, be all the spins ±1 that an Ising model has
at each site. That is {S} are all the states appearing in the
X
Z= e−E{S}/kB T (6.1)
{S}

partition function above. We wish to know the probability of being in some


state {S} at a time t,
p({S}, t) =? (6.2)
from knowledge of probabilities of states at earlier times. In the canonical
ensemble (fixed T , fixed number of particles)

Peq ({S}) ∝ e−E{S}/kB T (6.3)

of course. For simplicity, lets drop the { } brackets and let

{S} ≡ S (6.4)

As time goes on, the probability of being in state S increases because one makes
transitions into this state from other (presumably nearby in some sence) states
S ′ . Say the probability of going from S → S ′ , in a unit time is

W (S, S ′ ) ≡ W (S ← S ′ ) (6.5)
(Read As)

This is a transition rate as shown in fig. 6.1. Hence

39
40 CHAPTER 6. MONTE CARLO AND THE MASTER EQUATION
W (S, S ′ )

S′ S

W (S ′ , S)

Figure 6.1:

W (S, S ′ ) P (S ′ ) (6.6)

are contributions increasing P (S).


However, P (S) decreases because one makes transitions out of S to the states
S ′ with probability
W (S ′ , S) ≡ W (S ′ ← S) (6.7)
(Read As)

So that
W (S ′ , S) P (S) (6.8)
are contributions decreasing P (S) as time goes on.
The master equation incorporates both these processes as

∂P (S, t) X
= [W (S, S ′ ) P (S ′ ) − W (S ′ , S) P (S)] (6.9)
∂t ′ S

which can be written as


∂P (S, t) X
= L(S, S ′ ) P (S ′ ) (6.10)
∂t ′ S

where the inappropriately named Liouvillian is


X
L(S, S ′ ) = W (S, S ′ ) − ( W (S ′′ , S))δS, S ′ (6.11)
S ′′

Note that the master equation

1. has no memory of events further back in time than the differential element
“dt”. This is called a Markov process

2. is linear.

3. is not time reversal invariant as t → −t (since it is first order in time).

In equilibrium,
∂Peq (S)
=0 (6.12)
∂t
41

since we know this to be the definition of equilibrium. Hence in Eq. 6.9,

W (S, S ′ ) Peq (S ′ ) = W (S ′ , S) Peq (S) (6.13)

which is called detailed balance of transitions or fluctuations. For the canonical


ensemble therefore,

W (S, S ′ ) ′


= e−(E(S)−E(S ))/kB T (6.14)
W (S , S)

This gives a surprisingly simple result for the transitions W , it is necessary (but
not sufficient) that they satisfy Eq. 6.14. In fact, if we are only concerned with
equilibrium properties, any W which satisfies detailed balance is allowable. We
need only pick a convenient W !
The two most common W ’s used in numerical work are the metropolis rule
(
e−∆E/kB T , ∆E > 0
WM etropolis (S, S ′ ) = (6.15)
1 , ∆E ≤ 0

and the Glauber rule


1 ∆E
WGlauber (S, S ′ ) = (1 − tanh ) (6.16)
2 2kB T
where
∆E = E(S) − E(S ′ ) (6.17)
Let us quickly check that WM etropolis satisfies detailed balance. If E(S) > E(S ′ )

W (S, S ′ ) e−(E(S)−E(S ))/kB T


= = e−∆E/kB T
W (S , S) 1

Now if E(S) < E(S ′ )

W (S, S ′ ) 1
= −(E(S)−E(S ′ ))/k T = e−∆E/kB T
W (S ′ , S) e B

where ∆E = E(S) − E(S ′ ) as above. So Detailed balance works. You can


check yourself that the Glauber rule works. The form of the rule is shown in
figures 6.2 and 6.3. To do numerical work via Monte Carlo, we simply make
transitions from state {S ′ } → {S} using the probability W ({S}, {S ′ }). A last
bit of black magic is that one usually wants the states {S ′ } and {S} to be
“close”. For something like the Ising model, a state {S} is close to {S ′ } if the
two states differ by no more than one spin flip, as drawn in fig. 6.4. Let us do
this explicitly for the 2 dimensional Ising model in zero field, the same thing
that Onsager solved! X
EST AT E = −J Si Sj (6.18)
hiji
42 CHAPTER 6. MONTE CARLO AND THE MASTER EQUATION

Metropolis
1 1 1

∆E ∆E ∆E
Low T Moderate T High T

Figure 6.2:

Glauber
1 1 1
1/2 1/2

∆E ∆E ∆E
Low T Moderate T High T

Figure 6.3:

Source State {S ′ } Target State {S}

? or

Outcome with
probability
W ({S}, {S ′ })

Figure 6.4:
43

?
Case 1
prob = e−8J/kB T

?
Case 2
prob = e−4J/kB T

?
Case 3
prob = 1

?
Case 4
prob = 1

?
Case 5
prob = 1

Figure 6.5:
44 CHAPTER 6. MONTE CARLO AND THE MASTER EQUATION

where Si = ±1 are the spins on sites i = 1, 2, ... N spins (not to be confused


with our notation for a state {S}!), and hiji means only nearest neighbours
are summed over. The positive constant J is the coupling interaction. If only
one spin can possibly flip, note there are only five cases to consider as shown in
fig. 6.5.
To get the probabilities (Metropolis)
For Case 1:

EBef ore = −4J , EAf ter = +4J


∆E = 8J
W (S, S ′ ) = Prob. of flip = e−8J/kB T

For Case 2:

EBef ore = −2J , EAf ter = +2J


∆E = 4J
W (S, S ′ ) = Prob. of flip = e−4J/kB T

For Case 3:

EBef ore = 0 , EAf ter = 0


∆E = 0
W (S, S ′ ) = Prob. of flip = 1

For Case 4:

EBef ore = +2J , EAf ter = −2J


∆E = −4J
W (S, S ′ ) = Prob. of flip = 1

For Case 5:

EBef ore = +4J , EAf ter = −4J


∆E = −8J
W (S, S ′ ) = Prob. of flip = 1

This is easy if W = 1, if W = e−4J/kB T , then you have to know the temper-


ature in units of J/kB . Let’s say, for example,

W = 0.1

You know the spin flips 10% of the time. You can enforce this by comparing
W = 0.1 to a random number uniformly between 0 and 1. If the random number
is less that or equal to W , then you flip the spin.
Here is an outline of a computer code. If you start at a large temperature and
45

Initialize System
- How many spins, How long to average

time t = 0 - How hot, (give T = (...)J/kB )?

- How much field H?


- Give all N spins some initial value (such as Si = 1).
use periodic boundary conditions

Do time = 0 to maximum time

Grinding Do i = 1 to N

any 1mcs - Pick a spin at random Do this N times


time step - Check if it flips (each spin is flipped on average)
as N updates - Update the system

Calculate magnetization per spin


P
hM i = N1 h i Si i
Output Calculate energy per spin
P
hE −1
Ji= N h hiji Si Sj i

averaging over time (after throwing away some early time).

Figure 6.6:
46 CHAPTER 6. MONTE CARLO AND THE MASTER EQUATION

M
1
high T > Tc

t in Monte Carlo step 1, 1 unit of time


corresponds to attempting N spin flips

Figure 6.7:

M transient
time average to get hM i low T < Tc
1

Figure 6.8:

hM i

M ∼ |T − Tc |−beta , β = 1/8

T
Tc ≈ 2.269J/kB

Figure 6.9:
47

hE/Ji

T
Tc

−2

Figure 6.10:

initially Si = 1, the magnetization quickly relaxes to zero (see fig. 6.7). Below Tc ,
the magnetization quickly relaxes to its equilibrium value (see fig. 6.8). Doing
this for many temperatures, for a large enough latice gives fig. 6.9. Doing this
for the energy per spin gives fig. 6.10. which is pretty boring until one plots the
heat capacity per spin
1 ∂hEi
C≡
J ∂T
Try it. You’ll get very close to Ousager’s results with fairly small systems

C
ln |T − Tc |

T c ≈ 2.269 J/kB

Figure 6.11:

N = 10 × 10, or 100 × 100 over a thousand or so time steps, and you’ll do much
better than the mean field theory we did in class.
As an exercise, think how would you do this for the solid–on–solid model
X
Z= e−EState /kB T
{States}

where
N
X
EState = J |hi − hi+1 |
i=1
where hi are heights at sites i.
48 CHAPTER 6. MONTE CARLO AND THE MASTER EQUATION
Chapter 7

Review of Thermodynamics

There are a lot of good thermodynamics books, and a lot of bad ones too. The
ones I like are “Heat and Thermodynamics” by Zemansky and Dittman (but
beware their mediocre discussion of statistical mechanics, and their confusing
discussion of phase transitions), and Callin’s book, “Thermostatics”, (which is
overly formal).
A thermodynamic state of a system is described by a finite number of vari-
ables, usually at least two. One needs at least two because of conservation
laws for energy E and number of particles N . So one needs at least (E, N ) to
specify a system, or two independent variables which could give (E, N ). Some
systems need more than two, and for some systems, it is a useful approach to
only consider one variable, but for now, we will consider that two are enough.
There is a deep question lurking here. Thermodynamics describes systems
in the thermodynamic limit: N → ∞, volume V → ∞, such that number
density n = N/V = const. Each classical particle has, in three dimensions, 6
independent degrees of freedom describing its position and momentum. So there
are not 2, but ∼ 1023 degrees of freedom! This is a very deep question. Many
people have worried about it. But let’s take the physics view. We need at least
two variables, so let’s use two. Experimentally, this describes thermodynamic
systems on large length and time scales, far from, for example, phase transitions.

Consider a pure system which can be in a gas, liq-


2 out of uid, or simple solid phase. It can be described by tem-
perature T , pressure P , entropy S, in addition to E,
T , P , S, E, V ,
V , and N . Temperature is a ”good” thermodynamic
N , ... variable by the zeroth law of thermodynamics (two
systems in thermal equilibrium with a third system
are independent
are in thermal equilibrium with each other). Entropy
is a ”good” thermodynamic variable by the second
law of thermodynamics (the entropy of an isolated
Figure 7.1: Our System system never decreases).
Variables can be classified by whether they are

49
50 CHAPTER 7. REVIEW OF THERMODYNAMICS

extencive or intensive. Say one considers a system of two independent parts as


drawn in fig. 7.2.

1 2 1 2

Figure 7.2: A system with two independent parts.

An extensive variable satisfies

Vtotal = V1 + V2
Etotal = E1 + E2
Stotal = S1 + S2

while an intensive variable satisfies

T = T1 = T2
P = P1 = P2

From mechanics, the work done on a system to displace it by dx is

−F dx

where F is the force. Acting on an area, we have


¯
dW = −P dV
|{z} |{z} |{z}
Small Pressure Differential
amount volume
of work change
Two new concepts arise in thermodynamics. Temperature, which is something
systems in thermal equilibrium share; and heat, which is a form of energy ex-
changed between systems at different temperatures.
The law of energy concervation (fist law of thermodynamics) is
dE = ¯
dW + ¯
dQ
|{z} |{z} |{z}
change in work heat
internal done on transfered
energy system
The second law of thermodynamics is
¯ = T dS
dQ

where S is entropy. Heat always flows from the hot system to the cold system.
In fig. 7.3, both sides change due to transfer of heat Q. For the hot side
51

THOT TCOLD

Figure 7.3: System with hot and cold parts exchanging heat Q.

−Q
(∆S)HOT =
THOT
for the cold side
−Q
(∆S)COLD =
TCOLD
Hence
1 1
(∆S)T OT AL = Q [ − ]
TCOLD THOT
So
(∆S)T OT AL ≥ 0
This is the usual way the second law is presented. Note that this is the only
law of physics that distinguishes past and future.
As an aside, it is interesting to note that S must be a “good” thermodynam-
ics variable, or else the second law is violated. If we assume the opposite (that
two distinct states have the same S) we can create a perpetual motion machine,
as drawn in fig. 7.4. Blobbing up the 1st and 2nd laws gives

(T, S)

P (T, S) Q=W

isotherm

adiabat

Figure 7.4: Perpetual Motion Machine

dE = −P dV + T dS

where only 2 out of E, P, V, T and S are independent. This is the most impor-
tant and useful result of thermodynamics.
To get further, we need equations of state, which relate the variables. Whithin
thermodynamics, these are determined experimentally. For example, the ideal
52 CHAPTER 7. REVIEW OF THERMODYNAMICS

gas equation of state is

P = P (V, T ) = N kB T /V

and
E = E(T )
where usually
E =CT
and C is a constant determined experimentally. Then
N kB T
C dT = − dV + T dS
V
So, for an adiabatic process (wherein no heat is exchanged, so dS = 0), a change
in volume causes a change in temperature.
In general, equations of state are not known. Instead one obtains differential
equations of state
∂P ∂P
dP = ( )T dV + ( )V dT
∂V ∂T

and

∂E ∂E
dE = ( )T dV + ( )V dT
∂V ∂T
Clearly, if we know all the (T, V ) dependence of these derivatives, such as ( ∂E
∂T )V ,
we know the equations of state. In fact, most work focuses on the derivatives.
Some of these are so important they are given their own names.
1 ∂V
Isothermal Compressibility kT = −
( )T
V ∂P
1 ∂V
Adiabatic Compressibility ks = − ( )s
V ∂P
1 ∂V
Volume Expansivity α= ( )P
V ∂T
∂S
Heat Capacity at Constant Volume CV = T ( )V
∂T
∂S
Heat Capacity at Constant Pressure CP = T ( )P
∂T
Of course, kT = kT (T, V ), etc.
These results rest on such firm foundations, and follow from such simple
principles, that any microscopic theory must be consistent with them. Further-
more, a microscopic theory should be able to address the unanswered questions
of thermodynamics, e.g., calculate kT or CV from first principles.
These quantities can have very striking behavior. Consider the phase dia-
gram of a pure substance (fig. 7.5). The lines denote 1st order phase transi-
53

P
s l

Figure 7.5: Solid, liquid, vapor phases of pure substance

s l
Pc

T1 T2 T

Figure 7.6:

tions. The dot denotes a continuous phase transition, often called second order.
Imagine preparing a system at the pressure corresponding to the continuous
transition, and increasing temperature from zero through that transition (see
fig. 7.6).
Figure 7.7 shows the behavior of kT and CV : At the first-order transition,
kT has a delta-function divergence while CV has a cusp. At the continuous
transition, both CV and kT have power-law divergences. We will spend a lot of
time on continuous, so-called second order transitions in the course.
54 CHAPTER 7. REVIEW OF THERMODYNAMICS

kT P ≡ Pc

T1 T2

Cv P ≡ Pc

T1 T2

Figure 7.7:
Chapter 8

Statistical Mechanics

We will develop statisitical machanics with “one eye” on thermodynamics. We


know that matter is made up of a large number of atoms, and so we expect
thermodynamic quantities to be avarages of these atomic properties. To do the
averages, we need a probability weight.
We will assume that, in a isolated system, the probability of being in a state
is a function of the entropy alone. This is because, in thermodynamics, the
equilibrium state is the one with the maximum entropy. Hence, the probability
weight
ρ = ρ(S)
To get the form of S we need to use
1. S is extensive.
2. S is maximized in equilibrium.
3. S is defined up to a constant.
Since S is extensive, two independent systems of respective entropies S1 and S2
have total entropy (S1 + S2 ) (see fig. 8.1). But

1 2

Figure 8.1:

ρ(Stotal ) = ρ(S1 )ρ(S2 )


Since they are independent (and note I will not worry about normalization for
now). Since S is extensive we have
ρ(S1 + S2 ) = ρ(S1 )ρ(S2 )

55
56 CHAPTER 8. STATISTICAL MECHANICS

or
ln ρ(S1 + S2 ) = ln ρ(S1 ) + ln ρ(S2 )
This is of the form
f (S1 + S2 ) = f (S1 ) + f (S2 )
where f is some function. The solution is clearly

f (S) = AS

where A is a constant, so
ρ = eAS
But S is defined up to a constant so

ρ = BeS/kB , A = 1/kB

where the constant B is for normalization, and our earlier carelessness with
normalization has no consequence. This is the probability weight for the micro-
cannonical ensemble since S = S(E, V ) in thermodynamics from

T dS = dE + P dV

That is, it is the ensemble for given (E, V). The more handy ensemble, for given
(T, V) is the canonical one. Split the system into a reservoir and a system. They
are still independent, but the resevoir is much bigger than the system (fig. 8.2).

System
fixed T , V

Resevoir

Figure 8.2:

Then

S(Eresevoir ) = S(Etotal − E)
∂S
≈ S(Dtotal ) − ( )total E
∂E

and

E
S(Eresevoir ) = const −
T
57

So, for the canonical ensemble


Estate

ρstate = const e kB T

The normalization factor is


X
Z= e−Estate /kB T
states

The connection to thermodynamics is from the Helmholtz free energy F =


E − T S, where
Z = e−F/kB T
We haven’t shown this, but it is done in many books.
58 CHAPTER 8. STATISTICAL MECHANICS
Chapter 9

Fluctuations

Landan and Lifshitz are responsible for “organizing” the theory of fluctuations.
They attribute the theory to Einstein in their book on statistical physics. I
am following their treatment. First, lets do toy systems, where entropy only
depends on one variable, x.
The probability weight
ρ ∼ eS(x)/kB
so the probability of being in x → x + dx is

const eS(x)/kB dx

where the constant gives normalization. At equilibrium, S is maximized, and x


is at its mean value of hxi as drawn in fig. 9.1. Near the maximum we have

hxi x

Figure 9.1:

¯ ¯
∂S ¯¯ 1 ∂ 2 S ¯¯
S(x) = S(hxi) + (x − hxi) + (x + hxi)2 + ...
∂x ¯x=hxi 2 ∂x2 ¯x=hxi

But S(hxi) is just a constant, ∂S/∂hxi = 0, and ∂ 2 S/∂hxi2 is negative definite,


since S is a maximum. Hence,
S(x) 1
= (x − hxi)2
kB 2σ 2

59
60 CHAPTER 9. FLUCTUATIONS

where
σ 2 ≡ −kB /(∂ 2 S/∂hxi2 )
So the probability of being in x → x + dx is
2
/2σ 2
const. e−(x−hxi)

which is simply a Gaussian distribution. Since it is sharply peaked at x = hxi,


we will let integrals vary from x = −∞ to x = ∞. Then

h(x − hxi)2 i = σ 2

i.e.,
−kB
h(x − hxi)2 i =
∂ S/∂hxi2
2

This is called a fluctuation-dissipation relation of the first kind. Fluctuations in


x are related to changes in entropy. They are also called correlation–response
relations. As an aside, the “second kind” of fluctuation–dissipation relations
turn up in nonequilibrium theory.
Since S is extensive, we have

h(∆x)2 i = O(1/N )

again, if x is intensive. Note that, from before we said that

h(∆x)2 i = O(ξ 3 /L3 )

So that fluctuations involved microscopic correlation lengths. Now we have

−kB
= O(ξ 3 /L3 )
∂ 2 S/∂hxi2

So thermodynamic derivatives are determined by microscopic correlations. This


is why we need statistical mechanics to calculate thermodynamic derivatives.
We need more degrees of freedom than x to do anything all that interesting,
but we have enough to do something.
Consider isothermal toys. These are things like balls, pencils, ropes, and
pendulums, in thermal equilibrium. They have a small number of degrees of
freedom. For constant temperature, using thermodynamics

T ∆S = ∆E −
|{z} W
|{z}
≡0 work done
for simple to system
systems

Hence ∆S = −W/T . Since entropy is only defined up to a constant, this is all


we need.
This is how it works. Consider a simple pendulum (fig. 9.2). The amount of
61

Figure 9.2: Simple pendulum of mass m, length l, displaced an angle θ.

work done by the thermal background to raise the pendulum a height x is

mgx

where g is the gravitational acceleration. From the picture

x = l − l cos θ
θ2
≈l
2

so

mglθ2
W =
2

2
and S = − mglθ
2 + const. so, since hθi = 0, we have

kB T
hθ2 i =
mgl

This is a very small fluctuation for a pendulum like in the physics labs. When
the mass is quite small, the fluctuations are appriciable. This is the origin, for
example, of fluctuations of small particles in water, called Brownian motion,
which was figured out by Einstein. There are many other isothermal toys.
62 CHAPTER 9. FLUCTUATIONS

Time to
Fall
5s

Figure 9.3:

Landan and Lifshitz give several examples with solutions. My favorite, not given
in Landan’s book though, is the pencil balanced on its tip. It is interesting
to estimate how long it takes to fall as a function of temperature (at zero
temperature, one needs to use the uncertainty principle).

9.1 Thermodynamic Fluctuations


We can use this approach to calculate the corrections to thermodynamics. Again
imagine we have two weakly interacting systems making up a larger system. Due
to a fluctuation, there is a change in entropy of the total system due to each
subsystem: “1”, the system we are interested in, and “2”, the reservoir

2 “1” The system


“2” The resevior
1 N fixed, common T , P

Figure 9.4:

∆S = ∆S1 + ∆S2
dE2 + P dV2
= ∆S1 +
| T
{z }
thermodynamics for
the reservoir
But when the reservoir increases its energy, the system decreases its energy.
The same thing for the volume. However, they have the same P (for mechanical
equilibrium) and T (for thermal equilibrium). So we have
∆E1 P
∆S = ∆S1 − − ∆V1
T T
9.1. THERMODYNAMIC FLUCTUATIONS 63

The probability weight is e∆S/kB , so we have


− k 1 T (∆E−T ∆S+P ∆V )
P robability ∝ e B

where I have dropped the “1” subscript. There are still only 2 relevant variables,
so, for example:

∆E =∆E(S, V )
∂E ∂E
= ∆S + ∆V +
∂S ∂V
1 ∂2E ∂2E ∂2E
+ [ 2 (∆S)2 + 2 (∆S∆V ) + (∆V )2 ]
2 ∂S ∂S∂V ∂V 2
to second order. By ∂E/∂V , I mean (∂E/∂V )S , but it is already pretty messy
notation. Rearranging:
∂E ∂E 1 ∂E ∂E
∆E = ∆S + ∆V + [∆ ∆S + ∆ ∆V ]
∂S ∂V 2 ∂S ∂V
But
∂E
( )V = T
∂S
and
∂E
( )S = −P
∂V
so we have
1
∆E =T ∆S − P ∆V + (∆T ∆S − ∆P ∆V )
2
The term in brackets gives the first–order correlation to thermodynamics. In
the probability weight

P robability ∝ e−(∆T ∆S−∆P ∆V )/2kB T

There are still only two independent variables, so, for example
∂S ∂S
∆S = ( )V ∆T + ( )T ∆V
∂T ∂V
CV ∂P
= ∆T + ( )T ∆V
T ∂T
Likewise
∂P ∂P
∆P = ( )V ∆T + ( )T ∆V
∂T ∂V
∂P 1
=( )V ∆T − ∆V
∂T κT V
64 CHAPTER 9. FLUCTUATIONS

Putting these expressions into the probability weight, and simplifying, gives
−CV
[ 2k 2 (∆T )2 − 2k 1
(∆V )2 ]
P robability ∝ e BT B T κT V

Evidently the probabilities for fluctuations in T and V are independent since

P robability = P rob(∆T ) · P rob(∆V )

Also, all the averages are gaussians, so by inspection:

kB T 2
h(∆T )2 i =
CV
h(∆V )2 i = kB T κT V

and

h∆T ∆V i = 0

If we had chosen ∆S and ∆P as our independent variables, we would have


obtained κ V
[ −1 (∆S)2 − 2kS T (∆P )2 ]
P robability ∝ e 2kB CP B

so that

h(∆S)2 i = kB Cp
kB T
h(∆P )2 i =
κS V
h∆S ∆P i = 0

Higher–order moments are determined by the Gaussian distributions. So, for


example,

h(∆S)4 i = 3h(∆S)2 i2
= 3(kB CP )2

Note that any thermodynamic quantity is a function of only two variables.


Hence, its fluctuations can be completly determined from, for example h(∆T )2 i,
h(∆V )2 i, and h∆T ∆V i. For example

∆Q = ∆Q(T, V )
∂Q ∂Q
=( )V ∆T + ( )T ∆V
∂T ∂V
so
∂Q 2 : 0 ∂Q 2
h(∆Q)2 i =( ) h(∆T )2 i + ... h∆T ∆V i + ( ) h(∆V )2 i
∂T V ∂V T
9.1. THERMODYNAMIC FLUCTUATIONS 65

and

∂Q 2 kB T 2 ∂Q 2
h(∆Q)2 i =( ) +( ) kB T κT V
∂T V CV ∂V T
Check this, for example, with Q = S or Q = P . Note the form of these relations
is
h(∆x)2 i = O(1/N )
for intensive variables, and

h(∆X)2 i = O(N )

for extensive variables. Also note the explicit relationship between corrections
to thermodynamics (h(∆T )2 i), thermodynamic derivatives (CV = T (∂S/∂T )V ),
and microscopic correlations (1/N = ξ 3 /L3 ).
The most important relationship is

h(∆V )2 i = kB T κT V

although not in this form. Consider the number density n = N/V . Since N is
fixed
N N
∆n = ∆ = − 2 ∆V
V V
Hence
V4
h(∆V )2 i = h(∆n)2 i
N2
so,

N2
h(∆n)2 i = kB T κT V
V4
or

n2 kB T κT
h(∆n)2 i =
V
for the fluctuations in the number density. This corresponds to the total fluc-
tuation of number density in our system. That is, if ∆n(~r) is the fluctuation of
number density at a point in space ~r,
Z
1
∆n = d~r∆n(~r)
V

(Sorry for the confusing tooking notation.) Hence


Z Z
2 1
h(∆n) i = 2 d~r d~rh∆n(~r)∆n(~r′)i
V
66 CHAPTER 9. FLUCTUATIONS

Again, for translational invariance of space, and isotropy of space

h∆n(~r)∆n(~r′)i = C(|~r − ~r′|)

In the same way as done earlier we have


Z
2 1
h(∆n) i = d~rC(r)
V

where the dummy integral “~r” is “~r −~r′” of course. But h(∆n)2 i = n2 kB T κT /V ,
so we have the thermodynamic “sum rule”:
Z Z
d~r C(r) ≡ d~rh∆n(r)∆n(0)i = n2 kB T κT

If Z
~
Ĉ(k) ≡ d~r eik·~r C(~r)

we have
lim Ĉ(k) = n2 kB T κT
k→0

The quantity Ĉ(k) is called the structure factor. It is directly observable by x–


ray, neutron, light, etc. scattering. For a liquid, one typically obtains something
like this The point at k = 0 is fixed by the result above.

Ĉ(k) (peaks correspond to short scale order)

Ĉ(k)

Figure 9.5: Typical structure factor for a liquid.


Chapter 10

Fluctuations of Surfaces

A striking fact is that, although most natural systems are inhomogeneous, with
complicated form and structure, almost all systems in thermodynamic equilib-
rium must have no structure, by Gibbs’ phase rule.
As an example, consider a pure system which can be in the homogeneous
phases of solid, liquid, or vapour. Each phase can be described by a Gibbs
energy gs (T, P ), gl (T, P ), or gv (T, P ). In a phase diagram, the system chooses
its lowest Gibbs free energy at each point (T, P ). Hence, the only possibility
for inhomogeneous phases are along lines and points:
gs (T, P ) = gl (T, P )
gives the solidification line
gl (T, P ) = gv (T, P )
gives the vaporization line, while
gs (T, P ) = gl (T, P ) = gv (T, P )
gives the triple point of solid–liquid–vapour coexistance.

P
s l

Figure 10.1:

Inhomogeneous states exist on lines in the phase diagram, a set of measure


zero compared to the area taken up by the P –T plane.

67
68 CHAPTER 10. FLUCTUATIONS OF SURFACES

Much modern work in condensed–matter shysics concerns the study of in-


homogeneous states. To get a basis for this, we shall consider the most simple
such states, equilibrium coexisting phases, separated by a surface.

Phase Diagram Cartoon of System


P

Pc Vapor
l

v Liquid

Tc T

Figure 10.2:

As every one knows, in a finite volume droplet, droplets are spheres while
much larger droplets are locally flat. This is the same for liquid–vapour systems,
oil–water systems, magnetic domain walls, as well as many other systems. The
reason for this is simple thermodynamics. A system with a surface has extra
free energy proportional to the area of that surface. Using the Helmholtz free
energy now, for convenience,
Z
F = (Bulk Energy) + σ d~x (10.1)
| {z }
Surface Area

where the positive constant σ is the surface tension. Note that the extra free

R
Bulk Energy σ d~x
y

~x

Figure 10.3:

energy due to the surface is minimized by minimizing the surface area.


The coordinate system is introduced in the cartoon above. Note that y is
perpendicular to the surface, while ~x is a (d − 1)–dimensional vector parallel to
the surface. Furthermore, the edge length of the system is L, so its total volume

V = Ld
69

while the total surface area


A = Ld−1
Equation 10.1, ∆F = σLd−1 , is enough to get us started at describing fluc-
tuations of a surface at a non–zero temperature. Of course, Eq. 10.1 is ther-
modynamics, while in statistical mechanics, we need a probability function, or
equivalently, the partition function.
X
Z= e−Estate /kB T (10.2)
{states}

to calculate such fluctuations strangely enough, using Eq. 10.1 we can construct
models for Estate in Eq. 10.2! This sounds suspicious since Estate should be
microscopic, but it is a common practice.
We shall make the decision to limit our investigations to long-length-scale
phenomena. So it is natural to expect that phenomena on small length scales,
like the mass of a quark or the number of gluons, do not affect our calculations.
In particle physics they call this renormalizability: if we consider length scales

r > a few Angstroms

all the physics on scales below this only gives rise to coupling constants entering
our theory on those larger length sacles. The main such coupling constant we
shall consider is simply the radius of an atom

ro ≈ 5Ȧ (10.3)

so that we consider thength scales

r ≥ ro

Usually this appears as a restriction, and introduction of coupling constant, in


wavenumber k, so that we consider

k≤Λ

where Λ ≡ 2π/ro . In particle physics this is called an ultraviolet cutoff. I


emphasize that for us Λ or ro is a physically real cutoff giving the smallest
length scale in which our theory is well defined.
This is not too dramatic so far. An important realization in statistical
mechanics is that many microscopic systems can give rise to the same behaviour
on long length scales. In the most simple sense, this means that large–length–
scale physics is independent of the mass of the gluon. We can choose any
convenient mass we want, and get the same large–length–scale behaviour. In
fact, it is commonplace to go much much further than this, and simlpy invent
microscopic models — which are completely unrelated to the true microscopics
in question! — which are consistent with thermodynamics. The rule of thumb
is to construct models which are easy to solve, but still retain the correct long–
length–scale physics.
70 CHAPTER 10. FLUCTUATIONS OF SURFACES

This insight, which has dramatically changed condensed–matter physics over


the last twenty years, is called “universality”. Many microscopic models have the
same long–length–scale physics, and so are said to be in the same “universality
class”. The notion of a universality class can be made precise in the context of
the theory of phase transitions. Even without precision, however, it is commonly
applied throughout modern condenced–matter physics!

10.1 Lattice Models of Surfaces


To get a sense of this, let us reconsider our system of intrest, an interface. We
will construct a “microscopic” model, as shown. We have split the ~x and y axes

~x

Physical Surface Microscopic modeled abstraction

Figure 10.4:

into discrete elements. Instead of a continuous variable where

−L/2 ≤ x ≤ L/2 and − L/2 ≤ y ≤ L/2 (10.4)

Say (in d = 2), we have a discrete sum along the x direction


L L
i = − , ..., −3, −2, −1, 0, 1, 2, 3, ...,
2 2
(where L is measured in units of ro , the small length scale cutoff), and a similar
discrete sum along the y direction
L L
j = − , ..., −3, −2, −1, 0, 1, 2, 3, ..., (10.5)
2 2
Actually, for greater convenience, let the index along the x axis vary as

i = 1, 2, 3, ..., L (10.6)

All the possible states of the system then correspond to the height of the in-
terface at each point i, called hi , where hi is an integer −L/2 ≤ hi ≥ L/2 as
shown in fig. 10.5. Hence the sum in the partition function is
L L/2
X X Y X
= = (10.7)
{states} {hi } i=1 hi =−L/2
10.1. LATTICE MODELS OF SURFACES 71

i=1 i=2 i=3 i=4 i=L

h5 = 4

h1 = 0
h2 = 1 h3 = 0 h4 = 1

Figure 10.5:

This is a striking simplification of the microscopics of a true liquid–vapour sys-


tem, but as L → ∞, the artificial constant imposed by the i, j discrete grid will
become unimportant.
Now let us invent a “microscopic” energy Estate = E{hi } which seems phys-
ically reasonable, and consistent with the free energy ∆f = σLd−1 . Physically,
we want to discourage surfaces with too much area, since the free energy is
minimized by a flat surface. It is natural for a microscopic energy to act with a

Discouraged by Estate Encouraged by Estate

Figure 10.6:

microscopic range. So we enforce flatness, by enforcing local flatness


local Estate = +(hi − hi+1 )2 (10.8)
So the energy increases if a system does not have a locally flat interface. In
total
L
X
Estate = J (hi − hi+1 )2 (10.9)
i=1
where J is a positive constant. This is called the discrete Gaussian solid–on–
solid model. Equally good models(!) are the generalized solid–on–solid models
L
X
n
Estate =J |hi − hi+1 |n (10.10)
i=1
72 CHAPTER 10. FLUCTUATIONS OF SURFACES

where n = 1 is the solid–on–solid model, and n = 2 is the discrete Gaussian


model. With Eq. 10.9, the partition function is
L
Y X − kBJ T P |hi′ −hi′ +1 |2
Z=( )e i′ =1 (10.11)
i hi

This is for d = 2, but the generalization to d=3 is easy. Note that

kB T
Z = Z( , L) (10.12)
J
One can straight forwardly solve Z numerically. Analytically it is a little tough.
It turns out to be easier to solve for the continuum version of Eq. 10.11 which
we shall consider next.

10.2 Continuum Model of Surface


Instead of using a discrete mesh, let’s do everything in the continuum limit.
Let h(~x) be the height of the interface at any point ~x as shown below. This

y y = h(~x)

is the interface
h(~x)
~x

Figure 10.7:

is a useful description if there are no overhangs or bubbles, which cannot be


described by y = h(~x). Of course, although it was not mentioned above, the
These give a multivalued h(~x), as shown

Overhang Bubble

Figure 10.8:

“lattice model” (fig. 10.5) also could not have overhangs or bubbles. The states
10.2. CONTINUUM MODEL OF SURFACE 73

the system can have are any value of h(~x) at any point ~x. The continuum
analogue of Eq. 10.7 is
X X YZ
= = dh(~x)
{states} {h~
x} ~
x
Z −Z (10.13)
≡ Dh(·)

Q R
where the last step just shows some common notation for ~x dh(~x). Things
are a little subtle because ~x is a continuous variable, but we can work it through.
One
Q R other weaselly point is that Q the sum over states is dimensionless, while
x dh(~
~ x) has dimensions of ( x L). This constant factor of dimensionality
does not affect our results, so we shall ignore it.
For the energy of a state, we shall simply make it proportional to the amount
of area in a state h(~x), as shown in fig. 10.7.

Estate {h(~x)} ∝ (Area of surface y = h(~x))

or
Estate = σ(L′ )d−1 (10.14)
′ d−1
where (L ) is the area of the surface y = h(~x), and σ is a positive constant.
It is worth stopping for a second to realize that this is consistent with out
definition of excess free energy ∆F = σLd−1 . In fact, it almost seems a little
too consistent, being essentially equal to the excess free energy!
This blurring of the distinction between microscopic energies and macro-
scopic free energies is common in modern work. It means we are considering a
“coarse–grained” or mesoscopic description. Consider a simple example
X
e−F/kB T ≡ Z = e−Eh /kB T
{h}

where h labels a state. Now say one has a large number of states which have
the same energy Eh . Call these states, with the same Eh , H. Hence we can
rewrite X X
e−Eh /kB T = gH e−EH /kB T
{h} {H}

where
gH = # of states h for state H
Now let the local entropy
S = kB ln gH
So we have X
e−F/kB T = e−FH /kB T
{H}

where FH = EH −T SH . This example is partially simlpe since we only partially


summed states with the same energy. Nevertheless, the same logic applies if one
74 CHAPTER 10. FLUCTUATIONS OF SURFACES

does a partial sum over states which might have different energies. this long
digression is essentially to rationalize the common notation (instead of Eq. 10.14)
of
Estate = Fstate = σ(L′ )d−1 (10.15)
Let us now calculate (L′ )d−1 for a given surface y = h(~x)

p
(dh)2 + (dx)2

dh
y dx

Figure 10.9:

As shown above, each element (in d = 2) of the surface is


p
dL′ = (dh)2 + (dx)2
r
dh (10.16)
= dx 1 + ( )2
dx
In d–dimensions one has
r
d−1 ′ d−1 ∂h 2
d L =d x 1+( ) (10.17)
∂~x
Integrating over the entire surface gives
Z r
′ d−1 d−1 ∂h
(L ) = d x 1 + ( )2
∂~x
so that Z r
d−1 ∂h 2
Estate = σ d ~x 1 + ( ) (10.18)
∂~x
Of course, we expect
∂h 2
( ) << 1 (10.19)
∂~x
(that is, that the surface is fairly flat), so expanding to lowest order gives
Z
d−1 σ ∂h
Estate = σL + dd−1 ~x( )2 (10.20)
2 ∂x
| {z }
Extra surface area due to fluctuation

The partition function is therefore,


X
Z= e−Estate /kB T
{states}
10.2. CONTINUUM MODEL OF SURFACE 75

or,
− σL
d−1 YZ −σ
Z
∂h
Z=e k
BT ( dh(x)) exp dd−1 ~x( ′ )2 (10.21)
2kB T ∂~x
~
x

This looks quite formidable now, but it is straightforward to evaluate. In par-


ticle physics it is called a free–field theory. Note the similarity to the discrete
partition function of Eq. 10.11. The reason Eq. 10.21 is easy to evaluate is that
it simply involves a large number of Gaussian integrals
Z
2
dh e−ah

As we know, Gaussian integrals only have two important moments,

hh(x)i

and
hh(~x) h(~x′ )i
All properties can be determined from knowledge of these two moments! How-
ever, space is translationally invariant and isotropic in the ~x plane (the plane
of the interface), so properties involving two points do not depend upon their

|x1 − x2 |

x~1
x~2
o

Figure 10.10: ~x plane

exact positions, only the distance between them. That is

hh(x)i = const.

and
hh(~x1 ) h(~x2 )i = G(|~x1 − ~x2 |) (10.22)
For convenience, we can choose

hhi = 0 (10.23)

then our only task is to evaluate the correlation function

G(x) = hh(~x) h(0)i


(10.24)
= hh(~x + x′ ) h(~x′ )i
76 CHAPTER 10. FLUCTUATIONS OF SURFACES

To get G(x), we use the partition function from Eq. 10.21, from which the
probability of a state is
−σ
dd−1 ~
x( ∂h 2
R
x)
ρstate ∝ e 2kB T ∂~
(10.25)

The subtle thing here is the gradients which mix the probability of h(~x) with that
of a very near by h(~x′ ). Eventually this means we will use fourier transforms.
To see this mixing, consider
Z Z Z
∂h ∂
d~x( )2 = d~x[ d~x′ h(~x′ )δ(~x − ~x′ )]2
∂~x ∂~x
Z Z Z
′ ∂ ∂
= d~x d~x d~x′′ h(~x)h(~x′′ )( δ(~x − ~x′ )) · ( δ(~x − ~x′′ ))
∂~x ∂~x

Using simple properties of the Dirac delta function. Now let

x → x′′
x′ → x
x′′ → x′

Since they are just dummy indices, giving


Z Z Z
∂ ∂
= d~x d~x′ d~x′′ h(~x)h(~x′ )( ′′ δ(~x′′ − ~x)) · ( ′′ δ(~x′′ − ~x′ ))
∂~x ∂~x
Z Z Z
∂ ∂
= d~x d~x′ d~x′′ h(~x)h(~x′ ) · ′ (δ(~x′′ − ~x) δ(~x′′ − ~x′ ))
∂~x ∂~x
Z Z
′ ′ ∂ ∂
= d~x d~x h(~x)h(~x ) · δ(~x − ~x′ )
∂~x ∂~x′
Z Z
∂2
= d~x d~x′ h(~x)h(~x′ )[− 2 δ(~x] − ~x′ )
∂~x

(you can do this by parts integration, and there will be terms involving surfaces
at ∞, which all vanish.) Hence,
Z Z Z
∂h
dd−1 ~x( )2 = dd−1 ~x dd−1 ~x′ h(~x)M (~x − ~x′ )h(~x′ ) (10.26)
∂~x

where the interaction matrix,

∂2
M (~x) = − δ(~x) (10.27)
∂~x2
which has the form shown in fig. 10.11. This “Mexican hat” interaction matrix
appears in many different contexts. It causes h(~x) to interact with h(~x′ ).
In fourier space, gradients ∂/∂~x become numbers, or rather wave numbers ~k.
We will use this to simplify our project. It is convenient to introduce continuous
10.2. CONTINUUM MODEL OF SURFACE 77
2

M (~x) = − ∂~
x2 δ(~
x)

~x

Figure 10.11:

and discrete fourier transforms. One or another is more convenient at a given


time. We have,
R d−1~ ~
d k ik·~ x
 (2π) d−1 e ĥ(k) (Continuous)
h(~x) = 1
P i~
k·~
x (10.28)
 Ld−1 e ĥk (Discrete)
~
k

where
Z
~
ĥ(k) = dd−1 ~xe−ik·~x h(~x)
(10.29)
= ĥk

Closure requires R
dd−1~
k i~
 (2π)d−1
e k·~x
δ(~x) = 1
P i~
k·~
e x (10.30)
 Ld−1
~
k

and Z
dd−1 ~x i~k·~x
δ(~k) = e
(2π)d−1
but Z
1 ~
δ~k, 0 = dd−1 ~xeik·~x (10.31)
Ld−1
is the Kronecker delta (
1 , if ~k = 0
δ~k, 0 = (10.32)
0 , otherwise
not the Dirac delta function.
We have used the density os states in ~k space as
L d−1
( )

so that the discrete → continuum limit is
78 CHAPTER 10. FLUCTUATIONS OF SURFACES

Z
X L d−1
→( ) dd−1~k

~
k

and
L d−1 ~
δ~k, 0 → ( ) δ(k) (10.33)

Using the Fourier transforms we can simplify the form of Estate in Eq. 10.20.
Consider
Z
∂h
dd−1 ~x( )2 =
∂x
Z Z
∂ d~k ~
= d~x( eik·~x ĥ(~k))2
∂~x (2π)d−1
Z Z Z
d~k d~k ′ ~ ~′
= d~x d−1 d−1
(−~k · ~k ′ )ĥ(~k)ĥ(~k ′ )ei(k+k )·~x
(2π) (2π)
Z ~ Z ~ ′ Z
dk dk ~k · ~k ′ )ĥ(~k)ĥ(~k ′ )[ d~xei(~k+~k′ )·x ]
= (−
(2π)d−1 (2π)d−1
Z
d~k
= k 2 ĥ(k)ĥ(−k)
(2π)d−1

after doing ~k ′ integral. But h(~x) is real, so ĥ∗ (~k) = ĥ(−k) (from Eq. 10.29),
and we have
Z Z
d−1 ∂h 2 dd−1~k 2 ~ 2
d ~x( ) = k |ĥ(k)| (Continuum)
∂~x (2π)d−1
or Z
∂h 2 1 X 2
dd−1 x( ) = d−1 k |ĥ~k |2 (Discrete) (10.34)
∂~x L
~
k

Hence modes of ĥ~k are uncoupled in fourier space; the matrix M (~x) has been
replaced by the number k 2 , (see Eqs. 10.26 and 10.27). Therefore we have
σ X 2
Estate = σLd−1 + k |ĥ~k |2 (10.35)
2Ld−1
~
k
P
For the partition function we need to sum out all the states {states} using the
fourier modes ĥk rather than h(~x). This is a little subtle because h(~x) is real
while ĥ~k is complex. Of course not all the components are independent since
ĥk = ĥ∗−k , so we double count each mode. To do the sum properly we have
X YZ Y Z
= dh(~x) = ′ d2 ĥ~k (10.36)
{states} ~
x ~
k
10.2. CONTINUUM MODEL OF SURFACE 79

where d2 ĥk = d ℜĥk d ℑĥk


Y
and ′ ← Restriction to half of ~k space to avoid double counting
~
k

For the most part, this prime plays no role in the algebra below. Recall that
ky
Q′ Integrating in kg > 0, only, for example
k

kx

Figure 10.12:

the only nontrivial moment is


G(x) = hh(~x)h(0)i
= hh(~x + ~x′ )h(~x′ )i
from Eq. 10.23 above. This has a simple consequence for the Fourier mode
correlation function
hĥ(~k) ĥ(~k ′ )i
Note
Z Z
~ ~′ ′
hĥ(~k) ĥ(~k ′ )i = dd−1 ~x dd−1 ~x′ e−ik·~x−ik ·~x hh(~x) h(~x′ )i
Z Z
~ ~′ ′
= d~x d~x′ e−ik·~x−ik ·~x G(~x − ~x′ )

Let,
1
~y = ~x − ~x′ ~x = ~y ′ + ~y
2
1 1
~y = (~x + ~x′ ) ′ ′
~x = ~y + ~y
2 2
so that d~xd~x′ = d~y d~y ′ , and
Z Z
~ ~ ′ −i(~
k−~
k′ )· y
~ ~ ~′ ′
hĥ(k) ĥ(k )i = d~y e 2 G(y) × d~y ′ e−i(k+k )·~y

so,
Z
~
hĥ(~k) ĥ(~k ′ )i = d~y e−ik·~y G(y)(2π)d−1 δ(~k + ~k ′ )
Z
~
= [ d~x e−ik·~x G(~x)](2π)d−1 δ(~k + ~k ′ ) , letting ~y → ~x
80 CHAPTER 10. FLUCTUATIONS OF SURFACES

as (
Ĝ(~k) (2π)d−1 δ(~k + ~k ′ ) (Continuum)
hĥ(~k) ĥ(~k ′ )i = (10.37)
Ĝ~k Ld−1 δ~k+~k′ , 0 (Discrete)
Hence, in Fourier space, the only non–zero Fourier cos is

hĥ~k ĥ−~k i = h|ĥ~k |2 i

which is directly related to Ĝ~k . Using the probability


σ
k2 |ĥk |2
P
− ~
2kB T Ld−1 k
ρstate ∝ e

we have,
′′2
σ
|ĥk′′ |2
P
Q R − ~ k
~
k′ ′ d2 ĥk′ |ĥK |2 e 2kB T Ld−1 k′′
h|ĥ~k |i = σ
(10.38)
k′′2 |ĥk′′ |2
P
Q′ R 2 − k~′′
2kB T Ld−1
~
k′ d ĥ k ′ e

Except for the integrals involving ĥ~k , everything on the numerator and denomi-
nator cancels out. We have to be careful with real and imaginary parts though.
Consider X X
k 2 |ĥk |2 = k 2 ĥk ĥ−k
k k

For the particular term involving ĥ15 , for example

(15)2 ĥ15 ĥ−15 + (−15)2 ĥ−15 ĥ−15 = 2(15)2 ĥ15 ĥ−15

giving a factor of 2. In any case,


R − 2σ
k2 |ĥk |2
d2 ĥ~k |ĥk |2 e 2kB T Ld−1
h|ĥ~k |2 i = 2σ
R − k2 |ĥk |2
d2 hk e 2kB T Ld−1

(These steps are given in Goldenfeld’s book in Section 6.3, p.174.) Now let

ĥ~k = Reiθ

so that R∞ 2
2 R dR R2 e−R /a
h|ĥk | i = 0R ∞
0
R dR e−R2 /a
where
kB T 1
a= (10.39)
σk 2 Ld−1
A little fiddling gives
kB T 1
h|ĥk |2 i = a =
σk 2 Ld−1
10.2. CONTINUUM MODEL OF SURFACE 81

Hence we obtain, from Eq. 10.37


hĥ(~k) ĥ(~k ′ )i = Ĝ(~k)(2π)d−1 δ(~k + ~k ′ )
where
kB T 1
Ĝ(~k) =
σ k2
and
G(x) = hh(~x)h(0)i (10.40)
From this correlation function we will obtain the width of the surface. We will

W W , width of surface
y

~x

Figure 10.13:

define the width to be the root–mean–square of the height fluctuations averaged


over the ~x plane. That is
Z
1
w2 = d−1 dd−1 ~xh(h(~x) − hh(~x)i)2 i (10.41)
L
which is overcomplicated for our case, since hh(~x)i = 0, and hh(~x)h(~x)i = G(0),
from above. So we have the simpler form
w2 = G(0) (10.42)
From the definition of fourier transform,
Z
dd−1~k kB T
G(0) = (10.43)
(2π)d−1 σk 2

which has a potential divergence as ~k → 0, or ~k → ∞. We shall always restrict


our integrations so that

≤ |~k| ≤ Λ (10.44)
L
|{z}
|{z}
Lattic constant
System size

Then it is easy to do the integral of Eq. 10.43. First in d = 2,


Z Λ
kB T 1 1
G(0) = dk 2
σ 2π −Λ k
Z
kB T Λ 1
= dk 2
πσ 2π L
k
82 CHAPTER 10. FLUCTUATIONS OF SURFACES

and
kB T
G(0) = L,d=2
2π 2 σ
so r
kB T 1/2
w= L ,d=2 (10.45)
2π 2 σ
Similarly, in d = 3,
Z Λ
kB T 1 k dk
G(0) = 2π
σ 4π 2 2π
L
k2
and
kB T
G(0) = ln (LΛ/2π) , d = 3
2πσ
so, r
kB T LΛ 1/2
w= (ln ) ,d=3 (10.46)
2πσ 2π
It is easy to obtain L dependences as dimension of space varies. One obtaines,


undef ined , d ≤ 1

LX , 1<d<3
w= 1/2
(10.47)


(lnL) , d=3

const. , d>3

where
3−d
X= (10.48)
2
is called the roughening exponent. The fact that w is undefined for d ≤ 1 is
explained below.
Eq. 10.47 implies the interface is rough and increases with L for d ≤ 3.
However, note that the width (for d > 1) is not large compared to the system

W ∼ L1/2 Width in d = 2 remains small


compared to system as L → ∞
L

Figure 10.14:

itself, as L → ∞.
w 1 1
= 1−X = (d−1)/2 (10.49)
L L L
Hence the width is small for d > 1. In d = 1, it appears the width is comparable
to the system size! This means the interface — and indeed phase coexistance
10.2. CONTINUUM MODEL OF SURFACE 83

itself — does not exist in d = 1. This is one signature of a lower critical


dimension, at and below which, phase coexistance cannot occur at any nonzero
temperature.
In general, one defines the roughening exponent X by
w ∼ LX (10.50)
as L → ∞. This is equivalent to the definition of the surface correlation expo-
nent ηs
ˆ ∼ 1
G(k) (10.51)
k 2−ηs
as k → 0. (Here ηs = 0, of course.) The two exponents are related by the
formula
3 − d ηs
X= − (10.52)
2 2
The surface turns out to be self-affine, in the way we discussed earlier for the
Von Koch snowflake, since correlation functions like Ĝ(~k) are power–law like.
The self–affine dimension of the rough surface is
ds = X + d − 1 (10.53)
To obtain the correlation function in real space is a little awkward because of
the divergence of G(X → 0) as L → 0. For convenience, consider
g(x) ≡ h(h(~x) − h(0))2 i (10.54)
Multiplying this out gives
g(x) = 2(hh2 i − hh(~x)h(0)i
This is simply 
2 (G(0) − G(x))

g(x) = or (10.55)

2 (w2 − G(x))

The interpretation of g(x) from Eq. 10.54 is simple, it is the (squared) difference
in the heights between two points a distance x apart, as shown in fig. 10.15. Not
surprisingly, we expect (for large L)
g(x → L) = 2w2 (10.56)

* 0h(0)i
since, G(L) = hh(L) h(0)i = hh(L)ih * 0because the points 0 and L are so far
appart so as to be uncorrelated, so G(L) = 0.
In any case, we have
Z i~
k·~
x
kB T 1 d−1~ 1 − e
g(x) = 2 d k( )
σ (2π)d−1 k2
Z (10.57)
kB T 1 ~ x
d−1~ 1 − cos k · ~
=2 d k( )
σ (2π)d−1 k2
84 CHAPTER 10. FLUCTUATIONS OF SURFACES

p
g(~x − ~x′ )

h(~x) h(~x′ )

Figure 10.15:

Consider d = 2. Let
Z Λ
kB T 1 1 − cos kx
g(x) = 2 dk( )
σ 2π −Λ k2
Z
kB T Λ 1 − cos kx
=2 dk
πσ 2π/L k2
let u = kx
Z Λx
kB T 1 − cos u
g(x) = 2 x{ du }
πσ 2πx/L u2
let Λ → ∞ and L → ∞ in the integral (for convenience)
Z ∞
kB T 1 − cos u
g(x) = 2 x( du )
πσ 0 u2
| {z }
π/2

so,
kB T
g(x) = x, d = 2 (10.58)
σ
In d = 3,
Z 2π Z Λ
kB T 1 1 − cos kx cos θ
g(x) = 2 dθ kdk( )
σ 4π 2 0 2π/L k2

This is a little more involved, and I will just quote the answer
kB T
g(x) = ln xΛ , d = 3 (10.59)
4πσ
In general, one obtians


undef ined , d=1

x2X , 1<d<3
g(x) = (10.60)


ln x , d=3

const. , d>3
10.2. CONTINUUM MODEL OF SURFACE 85

where, again X = (3 − d)/2. This confirms our earlier claim that correlations
are power — law like.
(There is a subtlety here involving constraints which deserves a minute of
time. The limit g(x → L) ∼ L2X , but the constraint is not the same as 2w2 as
derived above. Reconsider the integral of Eq. 10.58:
Z Λx
kB T 1 − cos u
g(x) = 2 x{ du }
πσ 2πx
L
u2

Now just let Λ = ∞. Rewriting gives


Z ∞ Z 2πx
kB T 1 − cos u L 1 − cos u
g(x) = 2 x{ du 2
− du }
πσ 0 u 0 u2
Z 2πx
kB T π L 1 − cos u
=2 x{ − du }
πσ 2 0 u2

or, for d = 2
kB T
g(x, L) = xf (x/L) (10.61)
σ
where Z 2πx
2 1 − cos u
f (x) = 1 − du (10.62)
π 0 u2
Note that
f (x → ∞) = 0
but
f (x → 0) = 1
This is called a scaling function. End of digression.)
f (x/L)

0
x/L

Figure 10.16:

We are pretty well done now, except we shall calculate the “renormalized”
surface tension from the partition function. Recall from Eq. 10.21
Z Z
− σL
d−1 Y
−σ ∂h
Z = e kB T ( dh(x)) exp dd−1 ~x( ′ )2
2kB T ∂~x
~
x
86 CHAPTER 10. FLUCTUATIONS OF SURFACES

Switching to Fourier space using Eqs. 10.34 and 10.36 gives


− σL
d−1 Y
−σ 1 X 2
Z = e kB T ( ′d2 ĥk ) exp d−1
k |ĥk |2
2kB T L (10.63)
~
k k
−F/kB T
≡e
So calculating Z gives the thermodynamic free energy F . We will call the
“renormolized”, or more properly the thermodynamic surface tension σ ∗
∂F
σ∗ ≡ ( )T,V,N
∂A
or,
∂F
σ∗ = ( )T,V,N (10.64)
∂Ld−1
going back to Eq. 10.63
− σL
d−1 Y Z − σ k2
|ĥk |2
Z=e k BT ′ d2 ĥ~k e 2kB T Ld−1

~
k

Since all modes are independent. But


Z Z ∞ Z ∞
2
d ĥ~k = dℜĥ~k dℑĥ~k
−∞ −∞

or if ĥ~k = Reiθ
Z Z 2π Z ∞
d2 ĥk = dθ RdR
0 0
(10.65)
Hence
− σL
d−1 Y Z 2π Z ∞
−( 2kσ T k2
)R2
Z=e kBT ′ dθ R dR e B Ld−1

~ 0 0
k
σ k2
let u = ( )R2
2kB T Ld−1
Z
− σL
d−1 Y 1 ∞
:0
Z=e kBT ′2π 2 du e−u
~
2( 2kσB T Lkd−1 ) 0
k

− σL
d−1 Y 2πkB T Ld−1
=e kB T

σk 2
~
k
d−1 2πkB T Ld−1
− σL
Q′
=e kB T
eln ~
k σk2

σLd−1 X 2πkB T Ld−1


= exp(− + ′ ln )
kB T σk 2
~
k
10.2. CONTINUUM MODEL OF SURFACE 87

σLd−1 1 X 2πkB T Ld−1


Z = exp(− + ln ) (10.66)
kB T 2 σk 2
~
k

But, Z = exp(−F/kB T ), up to a constant, so

kB T X 2πkB T Ld−1
F = σLd−1 − ln (10.67)
2 σk 2
~
k

is the free energy due to the surface. The second term gives the contribution due
to fluctuations. To take the derivative with respect to ∂/∂Ld−1 , it is important
to note that the ~k’s get shifted with L’s, and that it is the (kL) combination
which is not a function of L. So to take the derivative, we rewrite

2π/k 2π/k ′ note k shifts to k ′


as L → L + dL

L L + dL

Figure 10.17:

kB T X 2πkB T Ld−1
F = σLd−1 − ln
2 σ(kL)2
~
k

or,
kB T X kB T X 2πkB T
F = σLd−1 − ln(Ld−1 )(d+1)/(d−1) − ln
2 2 σ(kL)2
~
k ~
k
or,
kB T d + 1 X kB T X 2πkB T
F = σLd−1 − ( )(ln Ld−1 ) − ln (10.68)
2 d−1 2 σ(kL)2
~
k ~
k

The last term is independent of Ld−1 , so taking the derivative ( ∂L∂F


d−1 )T,V,N = σ

gives,
kB T d + 1 1 X
σ∗ = σ − ( )(Discrete)
2 d − 1 Ld−1
~
k
or,
Z
kB T d + 1 dd−1~k
σ∗ = σ − ( )(Continuum) (10.69)
2 d−1 (2π)d−1
88 CHAPTER 10. FLUCTUATIONS OF SURFACES

is the renormalized surface tension. Note that fluctuations decrease the value
of σ to the lower σ ∗ .
Z Z Λ
dd−1~k 1 Λ
In d = 2, d−1
= dk = (10.70)
(2π) 2π −Λ π
Z Z 2π Z Λ
dd−1~k 1 1 2 Λ2
In d = 3, = dθ kdk = πΛ = (10.71)
(2π)d−1 4π 2 0 0 4π 2 4π
Z
dd−1~k
In d = 1, =1 (10.72)
(2π)d−1

Hence we have,

3kB T
σ∗ = σ − Λ ,d = 2

kB T 2
σ∗ = σ − Λ ,d = 3

and since d = 1 is odd, let d = 1 + ǫ, for very small ǫ, and

kB T
σ∗ = σ − , d = 1 + ǫ , ǫ << 1
ǫ
(10.73)

(It is worth nothing that this is the first term in the d = 1 + ǫ expansion, where
R d−1 q ∂h 2
we have only considered the leading term in EState = σ d ~x 1 + ( ∂~x ) ≈
R
σLd−1 + σ2 dd−1 ~x( ∂h 2
∂x ) + ...)

10.3 Impossibility of Phase Coexistence in d=1


Note that, again, the renormalized surface tension vanishes as d → 1 (as ǫ → 0).
In our simple calculation σ ∗ actually becomes negative infinity at d = 1.
In fact, the lack of phase coexistance in d = 1 is a theorm due to Landan
and Pierls, which is worth quickly proving. (See Goldenfeld’s book, p.46.)
A homogeneous phase has an energy

EBulk ∝ +Ld−1 (10.74)

Putting in a surface gives an extra energy

ESurf ace ∝ +Ld−1 (10.75)

So the extra energy one gets due to adding a surface is

∆E1Surf ace ∝ +Ld−1


(10.76)
In d = 1, ∆E1Surf ace ∝ +O(1)
10.3. IMPOSSIBILITY OF PHASE COEXISTENCE IN D=1 89

Ebulk ∝ Ld
Esurf ace ∝ Ld−1

~x

Figure 10.18:

Basically, this looks bad for surfaces.


Since the energe increases, it looks like one wants to have the fewest possible
surfaces, so the picture on the previous page is that one: two coexisting phases
with one well–defined surface.
However, the system’s state is chosen by free energy minimization not energy
minimization, where
F = E − TS
Surfaces are good for entropy because they can occur anywhere along the y axis,
as shown. The total number of such ways is L along the y axis. Of course there

or or or etc.

Figure 10.19:

are all rotations and upside down interfaces, so that in the end there are Ld
ways. The entropy is

∆S = kB ln(numberof states)

so we obtain, for a surface,

∆S = +kB ln Ld

or,
∆S = +kB d(ln L) (10.77)
is the increase in entropy due to having a surface. Hence, the total change in
the free energy due to a surface is

∆F = +O(Ld−1 ) − kB T O(ln L) (10.78)


90 CHAPTER 10. FLUCTUATIONS OF SURFACES

So, for d > 1, L → ∞, the free energy increases if a surface is added, but for
d = 1 (at T > 0) the free energy decreases if a surface is added!
This means coexisting phases are stable for d > 1, but in d = 1 they are
unstable. Since one interface decreases F , we can add another and another and
another, and keep on decreasing F until there are no longer coexisting phases
(of bulk energy E ∼ Ld as L → ∞), just a mishmash of tiny regions of phase
fluctuations. Hence, for systems with short–ranged forces, at T > 0, there can
be no phase coexistance in d = 1. (Short–ranged forces were assumed when we
said the surface energy ESurf ace ∼ Ld−1 ).

10.4 Numbers for d=3


Some of these things may seem a little academic, since we have constantly done
everything in arbitrary d. Our results for d = 2, apply to films absorbed on
liquid surfaces and to ledges in crystal surfaces, for example. Experiments have

Figure 10.20:

been done on these systems, confirming our d = 2 results, such as w ∼ L1/2 .


In d = 3, it is worth looking again at two of our main results, Eq. 10.47
r
kB T LΛ 1/2
w= (ln ) (10.79)
2πσ 2π
kB T 2
σ∗ = σ − Λ

or
4πkB T Λ 2
σ ∗ = σ(1 − ( ) )
σ 2π
10.4. NUMBERS FOR D=3 91

or
σ ∗ = σ(1 − T /T ∗ ) (10.80)
where
σ
T∗ ≡ (10.81)
4πkB (Λ/2π)2
Now, reconsider the phase diagram of a pure substance. Our analysis describes,

P
s l
Pc Critical point

triple point v
T
Tc

Figure 10.21: Phase Diagram

for example, the behavior along the liquid-vapour coexistance line. Near the
triple point of a typical fluid, the surface tension is such that
r
kB Tt
≈ 1Ȧ (10.82)
2πσ
where Tt is the triple point temperature. Similarly, one roughly has

Λ
≈ 10Ȧ (10.83)

Putting these in Eq. 10.79 above implies

L 1/2
w = 1Ȧ(ln ) (10.84)
10Ȧ
Putting in numbers is amusing. The width divergence is very weak.


2.6Ȧ, L = 10000Ȧ (light wavelength)

4.3Ȧ, L = 10cm (coffee cup)
w=


5.9Ȧ, L = 1000km (ocean)

9.5Ȧ, L = 1030 m (universe)
¡
Of course, in a liquid–vapour system, gravity plays a role. (Certainly it plays
a role if one had a liquid–vapour system the size of the universe.) The extra
energy due to gravity is
Z
1
EGravity = dd−1 ~x(ρl − ρv )gh2 (x)
2
92 CHAPTER 10. FLUCTUATIONS OF SURFACES

where ρ is the mass density of the liquid or vapour, and g ¢is gravity. This
changes our results in a way you should work out for yourself.
Putting in numbers for the renormalized surface tension gives the result

σ ∗ (T = Tt ) ≈ σ(0.3)

which is a much bigger result!


Note that σ ∗ vanishes (assume σ 6= σ(T )) at T = Tt , as T is increased. It is
tempting, although a little too rough to interpret

T ∗ = Tc

putting in numbers gives T ∗ ≈ 1.4Tt , which is Tc to the right order. More


precisely, it is known that

σ∗ = Const.(1 − T /Tc )µ

for T → Tc , where µ ≈ 1.24 in d = 3. So our theory, and (over)–interpretation


gives µ = 1, instead. RThispis not too bad, considering we only went to lowest
order in the EState = d~x 1 + (∂h/∂~x)2 expansion, and did not consider bulk
properties at all.
Chapter 11

Broken Symmetry and


Correlation Functions

A major concept in modern work is symmetry. These results for surfaces are
an example of more general properties of correlations in systems with broken
symmetries.
By a broken symmetry, we mean the following. Note that in the partition
function X
Z= e−EStates /kB T
{States}

the energy of a state has a “symmetry”

EState (h + Const.) = EState (h) (11.1)

Since Z
σ dh
EState = σLd−1 + dd−1 ~x( )2
2 ∂~x
P
Naively, since the sum {States} does not break this symmetry, we would expect
it to be respected by any quantity obtained from Z. The two main quantities
we have obtianed are the moments hhi and hhhi. As we found before,

hhi = 0 (11.2)

In fact we could have the average value be any result, without changing our
algebra. Note that h’s expectation value is a definite number. In particular

hh + Const.i =
6 hhi

This is a common signature of a broken symmetry variable: Its correlations and


averages do not respect that full symmetry of Z or EState .
It is also instructive to think of the symmetry of Eq. 11.1 in a more abstract
way. Note that homogeneous systems, whether liquid or vapour, are transla-
tionally invariant and rotationally invariant. However, a system with a surface,

93
94CHAPTER 11. BROKEN SYMMETRY AND CORRELATION FUNCTIONS

Figure 11.1:

as drawn in fig. 11.1, only has translational invariance and isotropy in the ~x
plane of the interface; There is no invariance in the y direction. Hence the
system with a surface has a lower symmetry (a broken symmetry) compared to
the liquid or vapour system.
This perfect symmetry of a homogeneous system is lost precisely because h,
the broken symmetry variable, has a definite value, like hhi = 0.
It is instructive to see that the fact Z and EState have this symmetry implies
the form of the correlation functions of the broken symmetry variable h. Con-
sider d = 2 only. Since a surface can be anywhere for a given system (it only
has a definite value for a given system), it can be different places for different
systems. In a very large system, part of hhi might have one value, while far far

Figure 11.2:

away on the x axis it might want to have another value. The free energy cost of
such a fluctuation, if it involves very large length scales (equivalent to k → 0)
is infinitesimal.
σ
∆EState (k) = k 2 |ĥk |2
2
as k → 0. Indeed, if one has a very very large system encompassing the four
shown in fig. 11.2, the four subsystems behave like independent systems which

Figure 11.3: Surface of a very big system


95

each have a definite value for hhi. Since the energy cost of such fluctuations
is infinitesimal as k → 0, we can set up long–wavelength sloshing modes with
infinitesimal kB T . Roughly,
kB T σ
= k 2 |ĥk |2
2 2
and
kB T 1
h|ĥk |2 i ∼
σ k2
as we recall. These sloshing modes are capillary waves driven by thermal noise.
But we need not know the form of EState to obtain h|ĥk |2 i ∼ 1/k 2 . We
only need to know that it breaks a symmetry of Z and EState . As shown in
fig. 11.3, fluctuations occur, say ±∆y, over a long distance ∆x. The values of
∆y (which could be ∼ Ȧ) and ∆x (which could be ∼ meters) are not important
for the argument. Let us invent a model where, after one goes a distance ∆x
to the right, one goes up or down ∆y, at random, as shown in fig. 11.4. Of

step 2 vapor
y
step 1 liquid
x

Figure 11.4: Interface (in units of ∆x and ∆y)

course, this is a one–dimensional random walk along the y axis in “time” x.


The root–mean–square deviation of y from its initial value is

yRM S ∼ x1/2

for large “time” x. Of course yRM S is related to the correlation function


(g(x))1/2 from Eq. 10.54, so we have

g(x) ∼ x

Fourier transforming, and fiddling via Eqs. 10.55 and 10.56, gives Ĝ(k) ∼ 1/k 2
for small k so that,
h|ĥk |2 i ∼ 1/k 2
as expected. Our argument was only in d = 2, but it can be generalized to other
dimensions.
Note that it was the fact that h was a broken symmetry variable which gave
rise to the 1/k 2 ; no special details of the interaction were necessary.
Before obtaining the general result, note that the system reacts to a sym-
metry breaking variable by “shaking it” in such a way as to try and restore the
96CHAPTER 11. BROKEN SYMMETRY AND CORRELATION FUNCTIONS
A symmetry breaking variable ends up
having its fluctuations try to restore
the symmetry.

w ∼ Lx

Figure 11.5:

broken symmetry. The rough width of a surface is this attempt to restore the
full translational invariance in the y direction, as shown in fig. 11.5. In fact, in
d = 1, the fluctuations are so strong that the full symmetry is restored, and one
has translational invariance in the y drection (since coexisting phases cannot
exist). In, 1 < d ≤ 3, the fluctuations do not restore the symmetry, but they
make the surface rough, or diffuse, as L → ∞. In d > 3, fluctuations do not
affect the flat interface.
Finally, it should be noted that the cartoon on p. 64 is self–similar and self
affine, in the same sense as the von–koch snowflake we discussed below. Note
the important new ingredient: Broken symmetry ⇔ power–law correlations ⇔
self–similar structure.

11.1 Proof of Goldstone’s Theorem


We shall now prove Goldstone’s theorem. The theorem is, if B is a broken
symmetry variable, where B(~r) is B’s local variable, (~r is a field point in d
dimensions) then
hB̂(~k) B̂(~k ′ )i = (2π)d δ(~k + ~k ′ )Ĝ(k)
where
lim Ĝ(k) → ∞
k→0
and usually
Ĝ(k) ∝ 1/k 2
We will give not so much a proof of this as an indication of how to work it out
for whatever system one is interested in. To make things easier, say
hBi = 0
If this is not true, redefine B ⇒ B − hBi. Assuming one has identified B as an
important broken symmetry variable, then the free energy’s dependence on B
can be found by a Taylor expansion:
∂F 1 ∂2F 2
F (B) = F (0) + B+ B + ...
∂B 2 ∂B 2
11.1. PROOF OF GOLDSTONE’S THEOREM 97

∂F
But F (0) is just a constant, and ∂B = 0 at equilibrium, so

1 ∂2F 2
F (B) = B + ... (11.3)
2 ∂B 2
Of course, by F we mean EState or FState in the sense of Eq. 10.10. Now, it
turns out that it is important to deal with the local value of B(~r), or the Fourier
component B̂(~k). If we use the local value for the expansion in Eq. 11.3 we have,
Z Z
1 ∂2F
F (B) = d ~r dd~r′
d
B(~r)B(~r′ ) (11.4)
2 ∂B(~r)∂B(~r′ )
2
which looks worse than it is. Since ∂B(~r∂)∂B(~
F
r ′ ) is a thermodynamic derivative of
some kind, in a system which is (we assume) rotationally invariant and transla-
tionally invariant,
∂2F
≡ C(~r, ~r′ ) = C(~r − ~r′ ) (11.5)
∂B(~r)∂B(~r′ )
In fourier space
Z Z
1 dd~k dd~k ∂2F
F (B) = B̂(~k) B̂(~k ′ ) (11.6)
2 (2π)d (2π) ∂ B̂(~k) ∂ B̂(~k ′ )
d

But translational invariance gives


∂2F
= (2π)d δ(~k + ~k ′ )Ĉ(k) (11.7)
∂ B̂(~k) ∂ B̂(~k ′ )

and, if B(~x) is real, so B̂ ∗ (−k) = B̂(k), we obtain (finally)


( R d
1 d k 2
d Ĉ(k)|B̂(k)| (Continuum)
F (B) = 2 1 (2π) P 2
(11.8)
2Ld k Ĉk |B̂k |
~ (Discrete)

We can obtain the form of Ĉ(~k) form expanding close to ~k = 0.

Ĉ(~k) = (Const.) + (Const.)k 2 + (Const.)k 4 + ... (11.9)

where we have assumed everything is analytic near k = 0. Hence terms like


~k · ~k = k 2 and (~k · ~k)2 = k 4 appear, but not nonanalytic terms like (~k · ~k)|~k| = k 3 ,
(which has the absolute value sign).
Finally, we use that B breaks a symmetry of F . Hence F can have no
dependence on B in the thermodynamic limit ~k = 0. Therefore

lim Ĉ(k) = 0 (11.10)


k=0

and, if we have analyticity

Ĉ(k) = (Const.)k 2 (11.11)


98CHAPTER 11. BROKEN SYMMETRY AND CORRELATION FUNCTIONS

for a broken–symmetry variable near ~k = 0. This gives


X
F (B) = (Const.) ~ k |2
k 2 |B (11.12)
~
k

This is the same as Eq. 10.35 so we have

hB̂(~k)B̂(~k ′ )i = (2π)d δ(~k + ~k ′ )Ĝ(k)

where
1 ~k = 0
Ĝ(k) ∝ ,
k2
as claimed. If Ĉ(k) is not analytic as assumed in Eq. 11.9 we get the weaker
condition
lim Ĝ(k) → ∞
k→0

Finally, it is amusing to note that if B is not a broken symmetry variable, from


Eq. 11.9 we have
X
F (B) = (Const.) (k 2 + ξ −2 )|B̂k |2
~
k

where ξ −2 is a constant. This gives hB̂(~k)B̂(~k ′ )i = (2π)d δ(~k + ~k ′ )Ĝ(k) where


1
Ĝ(k) ∝ k2 +ξ −2 as k → 0. Fourier transforming gives

hB(~r)B(0)i = G(r) ∝ e−r/ξ

as r → ∞, as we anticipated earlier, when correlations were discussed.


Chapter 12

Scattering

Figure 12.1:

Scattering experiments evidently involve the change in momentum of some


incident particle or radiation by a sample. The momentum transfered to the
sample (in units of h̄) is
~q = ~ki − ~kf (12.1)
by conservation of momentum. This, along with the intensity of the scattered
radiation
I(~q) (12.2)
is measured at the detector. We shall only consider elastic scattering where
energy, w in units of h̄, is unchanged. Furthermore, we are implicitly considering
the first Born approximation and there by neglecting multiple scattering.
The intensity of the scattered radiation might be due to an electromagnetic
field. That is
I(~q) ∝ |Ê(q)|2 (12.3)

99
100 CHAPTER 12. SCATTERING

where Ê is the electric field. Clearly the electric field variations with ~q are due
to variations in the dielectric properties of the sample, i.e.,

Ê(q) ∝ δǫ̂(q) (12.4)

where δǫ̂ is the dielectric variability. Now, if δǫ̂ is a function of only a small
number of local thermodynamic variables, like density and temperature

∂ǫ ∂ǫ
δǫ̂ = ( )δ ρ̂ + ( )δ T̂ (12.5)
∂ρ ∂T

Usually ∂ǫ/∂T is exceedingly small, so we have

I(q) ∝ |ρ̂(q)|2 (12.6)

A similar argument can be constructed if the scattering is not electromagnetic,


giving the same result. We’ll not focus on |ρ̂(q)|2 letting

S(q) ≡ |ρ̂(q)|2 (12.7)

be the structure factor.


I Should note that, although I’ve skimmed this, the details giving Eq. 12.6
are often experimentally important. But that’s their job, ours is to figure out
S(q). We will deal with two cases for ρ̂(q)

1. ρ corresponds to the order parameter, a local density or local concentra-


tion.

2. The order parameter is not the local density, but instead corresponds to a
sublattice concentration or magnetization, or to a Bragg peak’s intensity.

The first of these is easier. Consider a system of uniform density (we’ll do


interfaces later), then
hρ(x) ρ(0)i = ρ2 (12.8)
of course, so the fourier transform gives

S(q) ∝ δ(q) (12.9)

A cartoon (fig. 12.2) may make this clear. The second case is more subtle. An
example is the sublattice concentration of a binary alloy. Say atoms are A and
B in the alloy, and that locally, A wants to be neighbours with B but not with
another A. This makes the ordered state look like a checker–board as shown in
fig. 12.3.
Assuming the scattering is different from A or B atoms this sort of order can
be seen by looking at S(q). However, it must be realized that the ordering of,
say, the A atoms is on a sublattice, which has twice the lattice spacing of the
original system as shown in fig. 12.4. Hence the density of A atoms is uniform
on a sublattice with twice the lattice constant of the original system, and the
101

y S(q)

Uniform boring system qy , or, qx


hρ(x) ρ(0)i = ρ2 S(q) ∝ δ(q)

Scattering as shaded region

Figure 12.2:

Perfectly ordered
A − B alloy

=A

=B

Figure 12.3:

Sublatice 1 (Shaded)

2a
a

(Sublatice 2 is unshaded)

Figure 12.4:
102 CHAPTER 12. SCATTERING

S(q) qy
q0

or qx
−q0 q0
−q0
−q0 q0 qy

Figure 12.5:

scattering will show peaks not at ~q = 0 but at ~q = q~0 corresponding to that


structure X
S(~q) ∝ δ(~q − q~0 ) (12.10)
q
~0

where q0 = 2π/(2a), and a is the original lattice spacing.


The same thing happens in a crystal, where Bragg peaks form at specific
positions in ~q space. By monitoring the height and width of such peaks, the
degree of order can be determined. A further complication is that usually the
sample will not be so well aligned with the incident radiation that one gets
spots on the qx and qy axes as shown. Instead they appear at some random
orientation (fig. 12.6). This might seem trivial (one could just realign the sam-
qy qy qy

q0 q0
or or q0 etc
qx qx qx

Figure 12.6:

ple), except that often one has many different crystallite orientations scattering
simultaneously, so that all the orientations indecated above are smeared into a
ring of radious q0 .
qy

q0 qx

Figure 12.7: Average of Many Crystallites


12.1. SCATTERING FROM A FLAT INTERFACE 103

12.1 Scattering from a Flat Interface


Consider a flat interface without roughening, as drawn in fig. 12.8. For simplicity
we will usually consider an interface embedded in dimension d = 2. If the density
ρ

y=0
1
y
is interface
y
x

Grey Scale Density

Figure 12.8:

is uniform, it can be written as

ρ(x, y) = A θ(y) + B (12.11)

where θ(y) is the Heaviside step function, and A and B are constants. For
simplicity, we will let A = 1 and B = 0 from now on:

ρ(x, y) = θ(y) (12.12)

To get S(q) we need the fourier transform of θ. Let

θ =H +C

where C is a to–be–determined constant. Taking a derivative gives


dθ dH
= +0
dy dy
or,
dH
= δ(y)
dy
Fourier transform gives
1
Ĥ(ky ) =
iky

so that
1
θ̂(ky ) = + C δ(ky )
iky
To determine C, note that

θ(x) + θ(−x) = 1
104 CHAPTER 12. SCATTERING

so

θ̂(ky ) + θ̂(−ky ) = δ(ky )

Hence C = 1/2 and,


1 1
θ̂(ky ) = + δ(ky ) (12.13)
iky 2
The two–dimensional Fourier transform of ρ(x, y) = θ(y) is therefore
1 1
ρ̂(~k) = δ(kx ) + δ(~k) (12.14)
iky 2

From this we obtain the scattering from a flat interface y = 0 as the modulus,
1
S(k) = |ρ̂(k)|2 ∝ δ(kx ) (12.15)
ky2

where we have neglected the trivial bulk factor delta function. If the unit vector
normal to the interface is n̂ and that along the surface is n̂⊥ , then
1
S(k) ∝ δ(~k · n̂⊥ ) (12.16)
~
(k · n̂)2

As a cartoon (fig. 12.10), the scattering appears as a streak pointed in the

n̂⊥ y

Figure 12.9:

S(k) ky
kx = 0

kx

ky Grey Scale of S(k)

Figure 12.10: Scattering from interface y = 0

direction of n̂ (ŷ in fig. 12.10).


12.1. SCATTERING FROM A FLAT INTERFACE 105

y=0 ρ
is interface

x y

Grey Scale Density

Figure 12.11:

Lets also consider the scattering from a line defect, which is similar, but has
important differences. In this case,
ρ(x, y) = δ(y) (12.17)
Then
ρ̂ = δ(kx ) (12.18)
and
S(k) ∝ δ(kx ) (12.19)
or if the unit vector normal to the surface is n̂, and that along the surface is
n̂⊥ ,
S(k) ∝ δ(~k · n̂⊥ ) (12.20)
for a line defect. Now, if ρ is not the order parameter (like the second case
S(k)
kx = 0 ky

kx

ky
Grey Scale of S(k)

Figure 12.12: Scattering from a line defect at y = 0

discussed above) so that the density doesn’t vary across the interface, but an
order parameter corresponding to sublattice concentration or crystalline orien-
tation does, we have to incorporate the wavenumber k~0 corresponding to that
structure. For simplicity, we will assume that all one has to do is modulate the
structure up to that wavenumber. That is, for an interface
1
S(k) ∝ δ((~k − ~k0 ) · n̂⊥ ) (12.21)
~ ~
|(k − k0 ) · n̂|2
106 CHAPTER 12. SCATTERING

while for a line defect


S(k) ∝ δ((~k − ~k0 ) · n̂⊥ ) (12.22)

12.2 Roughness and Diffuseness


So far this is only geometry. And even as geometry, we have not considered
a diffuse interface. Before doing roughness, lets consider a diffuse interface as
shown in fig. 12.13. One example of this is

y ρ

ξ
x

ξ y

Figure 12.13: A diffuse interface, with no roughness

ρdif f use (x, y) ∼ tanh(y/ξ) (12.23)

So that
δ(kx )
S(k) = |ρ̂(k)dif f use |2 ∼ (1 + O(ξky )2 ) (12.24)
ky2
for an interface at y = 0. So the diffuseness only effects scattering at kx = 0

ky

kx
extra
scattering
due to diffuseness

Figure 12.14:

and large ky . On the other hand, let us now consider a rough interface, with no
diffuseness, where the interface is given by

y = h(x) (12.25)

and
ρRough (x, y) = θ(y − h(x)) (12.26)
12.2. ROUGHNESS AND DIFFUSENESS 107

Figure 12.15: A very rough interface, with no diffuseness

If h(x) involves a small variation, we can accordingly expand to obtain

ρRough ≈ θ(y) + h(x)δ(y) + ... (12.27)

since dθ/dy = δ(y).


Taking the fourier transform and squaring this gives
δ(kx )
S(k) = |ρ̂Rough (k)|2 ∼ + |ĥ(kx )|2 (12.28)
ky2

Averaging over thermal fluctuations using


1
h|ĥ(kx )|2 i ∼ (12.29)
kx2−η
which defines η, gives
δ(kx ) T
SRough (~k) ∼ 2
+ 2−η (12.30)
ky kx
where T is a constant. It would appear, then, that one could deconvolute the
ky

kx

Extra scattering due to roughness

Figure 12.16:

geometry δ(k x) 2
ky2 from the diffuseness O(ξky ) , from the roughness 1/kx
2−η
, since
they have different signatures in k space. For excellent samples this is in fact
true, but if there are many randomly oriented interfaces present, these signatures
are smeared out, and washed away.
108 CHAPTER 12. SCATTERING

12.3 Scattering From Many Flat Randomly–Oriented


Surfaces
Evidently the scattering from many randomly–oriented surfaces gives a pinwheel–
like cluster of streaks as shown in fig. 12.17. Angularly averaging over all orien-
ky ky
1/k p
Angular
Average
kx

Figure 12.17:

tations of the interface gives


R
dn̂|ρ̂k |2
S= R (12.31)
dn̂
where n̂ is the unit vector normal to the surface. We will write

n̂ = − sin θx̂ + cos θŷ (12.32)

and
n̂⊥ = cos θx̂ + sin θŷ (12.33)
with
~k = k cos β x̂ + k sin θŷ (12.34)
δ(~
k·n̂⊥ )
Then, for scattering from a surface where |ρ̂k |2 = |~
k·n̂|2
we have (in d = 2)
Z 2π
1 δ[k cos β cos θ + k sin β sin θ]
S= dθ
2π 0 | − k cos β sin θ + k sin β cos θ|2
Z 2π
1 1 δ[cos (θ − β)]
= dθ
2π k 3 0 | sin (θ − β)|2
Z 2π
1 1 δ[cos θ]
= dθ
2π k 3 0 | sin θ|2

But cos θ = 0 at θ = π/2, 3π/2 where | sin θ|2 = 1, so


1 1
SSurf aces = (12.35)
π k3
Similarly, in arbitrary dimension one has
1
SSurf aces ∝ (12.36)
k d+1
12.3. SCATTERING FROM MANY FLAT RANDOMLY–ORIENTED SURFACES109

for scattering from a surface. For scattering from a line defect which has
|ρ̂(k)|2 ∝ δ(kx ), one similarly obtains

1
Slinedef ects ∝ (12.37)
k d−1
We shall call these results, Porod’s Law. They involve only geometry.
The case where ρ is not the order, and a modulated structure exists is more
subtle. If the wavenumber of that structure is ~k0 then this introduces a new
angle via
~k0 = k0 cos αx̂ + k0 sin αŷ (12.38)
Now there are two possible averages

Figure 12.18:

1. α or β, average over crystallite orientation, or, angles in a non–angularly


resolved detector. These are equivalent (look at the cartoon for a few
seconds), and so we only need to do one of these two averages.

2. θ, averages over surface orientation.

If we average over α or β, for a fixed n̂, it is easy to anticipate the answer as


shown in fig. 12.19, where in the cartoon the regions around (kx = k0 , ky = 0)
ky ky

α or β
average
kx kx

interface n̂ = ŷ

Figure 12.19:

retain the singularity of the original δ(kx − k0 )/(ky − k0 )2 .


110 CHAPTER 12. SCATTERING

ky ky

k0 θ Average k0
kx kx

Figure 12.20:

This is basically a detector problem, so instead we will consider averaging θ


first. In fact it is obvious what the result of such an average is, it must be the
previous result, but shifted from ~k = 0 to ~k = ~k0 , that is
1
Ssurf ace (k) ∝
|~k − ~k0 |d+1
(12.39)
1
Sline (k) ∝
|~k − ~k0 |d−1

If a further average is done over crystallite orientation, we can take these ex-
pressions and angularly average them over the angle between ~k and ~k0 . Clearly
after averaging we will have

|k| − |k0 |
S̄ = S( ) ≡ S(∆k) (12.40)
|k0 |

First, let
φ=β−α (12.41)
so that
(~k − ~k0 )2 = k 2 − 2kk0 cos φ + k02
or, on using
k ≡ k0 (∆k + 1) (12.42)
we have
(~k − ~k0 )2 = k02 {(∆k)2 + 2(1 − cos φ)(∆k + 1)} (12.43)
Hence in two dimensions, for a surface, we have
Z 2π
1 1
S̄(∆k) = dφ (12.44)
2πk03 0 |(∆k)2 + 2(1 − cos φ)(∆k + 1)|3/2

which is some weird integral. We will work out its asymptotic limits ∆k >> 1
and ∆k << 1.
Firstly, if ∆k >> 1, then
Z 2π
1 1
S̄(∆k) = dφ + ...
2πk03 (∆k)3 0
12.3. SCATTERING FROM MANY FLAT RANDOMLY–ORIENTED SURFACES111

or (in arbitrary d)
1
S̄(∆k) ∼ , ∆k >> 1 (12.45)
(∆k)d+1

If ∆k << 1, consider

Z 2π
? 1 1
S̄(∆k → 0) = dφ
2πk03 0 |2(1 − cos φ)|3/2

which diverges due to the φ = 0 (and φ = π) contribution. Oops. So we had


better keep a little ∆k. For convenience, we will also expand φ around 0.

Z 2π
1 1
S̄(∆k) ≈ dφ + ...
2πk03 0 |(∆k)2 + φ2 |3/2

Now let u = φ/∆k

Z :
2π/∆k

1 1 1
S̄(∆k) = du
2πk03 (∆k)2 0 |1 + u2 |3/2

so,
1
S̄(∆k) ∼ as ∆k → 0 (12.46)
(∆k)2

for all dimensions d.


If one does this in d dimensions with

1
S= (12.47)
|k − ~k0 |γ
~

then
(
1 k−k0
(k−k0 )γ−(d−1)
, k0 << 1
S̄ = 1 k−k0
(12.48)
(k−k0 )γ , k0 >> 1

If γ − (d − 1) > 0. For the case γ = d − 1, one obtains


(
k−k0
− ln (k − k0 ), k0 << 1
S̄ ∼ 1 k−k0 (12.49)
(k−k0 )d−1
, k0 >> 1

It should be noted that the actual scattering is given by a weird boring integral
112 CHAPTER 12. SCATTERING

ky 1
, near tails
|k−k0 |d+1

1
|k−k0 |2 , near center

k0
kx

Figure 12.21:

like Eq. 12.44, which is pretty close to an elliptical integral.


Finally, note the big mess of power laws, and we have not considered dif-
fuseness or roughness! Everything is just geometry.
Chapter 13

Fluctuations and
Crystalline Order

Usually it is safer to apply Goldstone’s theorem by “re–deriving” it for what-


ever system one is interested in. However, to give an interesting and concrete
example, consider crystalline order in a solid. The density ρ clearly depends on

Density
ρ = ρ(~r)

Atoms on a lattice

Figure 13.1:

positions ~r, as shown, in a regular crystal. A crystal, although it looks sym-


metric in a common sense way, actually has very little symmetry compared to
a liquid or a vapour.
A liquid or a vapour are isotropic in all directions of space and traditionally
invariant in all directions. A crystalline solid has a very restricted symmetry.
In figure 13.1, the system is invariant under (e.g.)
~r → ~r + r0 (nx̂ + mŷ) (13.1)
where n and m are integers (positive or negative), and r0 is the distance between
atoms. Clearly then a crystal has very low symmetry, and–indeed–breaks the
translational and rotational invariance of space!
This means there should be a broken symmetry variable, associated with
this, which satisfies Goldstone’s theorem. Clearly this variable is associated
with the ~r dependence of
ρ(~r) (13.2)

113
114 CHAPTER 13. FLUCTUATIONS AND CRYSTALLINE ORDER

invariance only for

~r ~r → ~r′ , with
~r′
~r′ = r0 (+1x̂ − 1ŷ)

r0 shown here

Figure 13.2:

Let us then introduce a variable ~u(~r) which discribes the local variations in the
density. That is, due to some fluctuation,

~r → ~r + ~u(~r)

and
ρ(~r) → ρ(~r + ~u(~r)) (13.3)
where ~u(~r) is the local displacement vector as shown in fig. 13.3.

Equilibrium averaged position

Actual position due to fluctuation

~u(~r)

Figure 13.3: Jiggled atoms due to thermal fluctuations

Aside: (What does it mean to be translationally invariant

ρ(~r) → ρ(~r + ~u)

Local variable corresponding to broken symmetry

Figure 13.4:

~u(~r)
115

A definite value for hu(~r)i = 0, say, breaks space symmetry (translational and
rotational invariance).)
By construction
h~u(~r)i = 0 (13.4)
In Fourier space, if all directions of fluctuation are equally likely,

h~u ˆ(~k ′ )i ∝ (2π)d δ(~k + ~k ′ )( 1 )


ˆ(~k)~u (13.5)
k2
It is a little hard to see what this does to a density, since ~u appears inside
ρ(~r + ~u(~r)). The thing to do is to expand in a Fourier series, but to get an idea,
consider d = 1. Atoms don’t look like cosines, but this gives the sense of it.
ρ(x) ∝ (cos Gx + 1)
ρ

x
2π/G

Figure 13.5: Rough depiction of a one–dimensional variation of density via


cosine

More precisely, X
ρ(x) = eiGx âG (13.6)
G

where there might be several wavenumbers G giving the density, and âG is a
Fourier coefficient giving the atomic form. In arbitrary dimension we write
P
ρ(x) = G eiGx âG

x
2π/G

Figure 13.6: Better depiction of a one–dimensional variation. But the cosine is


essentially the right idea.

X ~
ρ(~r) = eiG·~r âG
~ (13.7)
~
G

where G~ are called the reciprocal lattice vectors. Then we have a fluctuation
represented by
X ~
ρ(~r + ~u(~r)) = eiG·(~r+~u(~r)) âG
~ (13.8)
~
G
116 CHAPTER 13. FLUCTUATIONS AND CRYSTALLINE ORDER
X ~ ~
hρ(~r + ~u(~r))i = eiG·~r âG
~ he
iG·~
u(~
r)
i (13.9)
~
G

To calculate the average we use the fact that ~u(~r) is a Gaussian variable. There-
fore
~ 1 ~ 2
heiG·~u i = e− 2 hG·~ui
1 ~~
(13.10)
= e− 2 GG:h~u~ui

But ~u is only correlated to ~u if they are in the same direction, we expect. That
is
huα uβ i ∝ δαβ
so
~ 1 2
hu2 i
he−iG·~u i ≈ e 2 G
But, from Eq. 13.5 we have,
Z
dd k 1
hu2 i =
(2π)d k 2
or 
Const.
 d=3
hu2 i = ln L d=2 (13.11)


L d=1
So in d = 3 X ~
hρi = Const. eiG·~r âG
~ (13.12)
~
G

where the Const. = exp − 12 G2 (Const.) is called the deBye–Waller factor. In


d = 2 we have
1
− G2 ln L
2
| {z }
X ~ 1 2
hρi = eiG·~r âG e L const.G /2 (13.13)
~
G

which is a marginal case. In d = 1 we have


X ~ 1 2
hρi = eiG·~r âG e− 2 G L (13.14)
~
G

which is a large effect, much greater than the periodicity as L → ∞. This is the
content of the original Landan–Peierls theorem, which states that a one dimen-
sional variation in density, ρ(x) (whether crystal structure or phase coexistance)
is impossible: The only G which can survive in Eq. 13.14 is G = 0 as L → ∞
which gives ρ(x) → Constant, due to ~u(~r) fluctuations.
It is more revealing to do these things for the scattering intensity I(k), where

I(k) ∝ |ρ̂k |2 (13.15)


117

(Scattering is from coherent density fluctuations). From Eq. 13.7,


Z X ~
~
ρ̂k = dd r e−ik·~r { eiG·~r âG
~}
~
G
X Z
~ ~
= âG dd r e−i(k−G)·~r
~
G

or X
ρ̂k = âG δ(~k − G)
~ (13.16)
~
G
so, X
ρ̂∗k = â∗G δ(~k − G)
~
~
G

and X
|ρ̂k |2 = âG â∗G′ δ(~k − G)δ(~k − G
~ ′)
~ G
G, ~′

or, X
|ρ̂k |2 ∝ |âG |2 δ(~k − G)
~ (13.17)
~
G

where |âG |2 is an atomic form factor. This gives well defined peaks where ~k = G
~
as shown in fig. 13.7. When thermal fluctuations are present

Ik = |ρ̂k |2 Scattering gives sharp


peaks, if thermal fluctuations

are absent.

k
~
various G’s

Figure 13.7:

Z X ~
~ ~
ρ̂k = dd~r e−ik·~r ( eiG·~r âG eiG·u(~r) )
G

or, Z
X ~ ~ ~
ρ̂k = âG dd~r e−i(k−G)·r eiG·u(~r) (13.18)
~
G
so Z
X ~ ~′ ′ ~′ ′
ρ̂∗k = â∗G dd~r′ ei(k−G )·~r e−iG ·u(~r )
~′
G
118 CHAPTER 13. FLUCTUATIONS AND CRYSTALLINE ORDER

Hence
X Z Z
~ ~ ~ ~′ ′ ~ ~′ ′
h|ρ̂k |2 i = âG â∗G dd~r dd~r′ e−i(k−G)·~r ei(k−G )·~r heiG·u(~r)−iG ·u(~r ) i
~ G
G, ~′
(13.19)
This is the same integral as on page so we can simply use that
~ ~′ ′
heiG·u(~r)−iG ·u(~r ) i = f (~r − ~r′ ) (13.20)

and then,
X Z
~ ~~ −G
~ ′ )f (~r)
h|ρ̂k |2 i = âG â∗G dd~r e−i(k−G)·~r (2π)d δ(G (13.21)
~ G
G, ~′

where
~ ~
f (~r) = heiG·~u(r)−iG·u(0) i
G2
r )−u(0))2 i (13.22)
= e− 2 h(u(~

2
= Const. e−G hu(r)u(0)i

so, Z
X ~ ~ 2
h|ρ̂k |2 i = const. ~|
|âG 2
dd r e−i(k−G)·~r e−G hu(~
r )u(0)i
(13.23)
~
G

The correlation function hu(~r)u(0)i can be worked out as before (knowing h|û(k)|2 i ∼
1
k2 ), 
Const.,
 d=3
hu(~r)u(0)i = ln r, d=2 (13.24)


r, d=1
In d = 3, X
h|ρ̂k |2 i = Const. |âG |2 δ(~k − G)
~
~
G

with the same delta function peaks.


In d = 2, the marginal case,
X Z
~ ~
h|ρ̂k |2 i = Const. |âG |2 dd r e−i(k−G)·r e−Const. ln r
~
G
Z ~ ~
X e−i(k−G)·~r
= Const. |âG |2 dd r
rConst.
~
G | {z }
this is the fourier transform of a power
X 1
= Const. |âG |2
~ ~
|k − G|d−Const.
~
G
119

I(k)

d=3

~
G k

Figure 13.8: Delta Function Singulatities

or, letting (d − Const.) ≡ 2 − η, we have


X 1
h|ρ̂k |2 i = Const. |âG |2 (13.25)
~ ~ 2−η
|k − G|
~
G

in d = 2. The thermal fluctuations change the delta function singularities


to power–law singularities! Sometimes they are called η–line shapes (eta–line
shapes).

I(k)

d=2

G k

Figure 13.9: Power Law Singularities

In d = 1, we have
Z
P ~ ~
2
h|ρ̂k | i = Const. ~
G
2
|âG | dd r e−i(k−G)·~r e−(Const.)r
| {z }
this is the Fourier
transform of an
exponential. But
R d i~k′ ·~r −r/ξ
d re e ∝ (k′ )21+ξ−2
for small k ′

So, for d = 1,
X 1
h|ρ̂k |2 i = Const. ~|
|âG 2
(13.26)
~ ~ 2 + ξ −2
(k − G)
~
G
120 CHAPTER 13. FLUCTUATIONS AND CRYSTALLINE ORDER

where ξ −2 is a constant. The thermal fluctuations remove all singularities as


~k = G
~ now! This is another sign that ρ(x) is impossible, and there can be no

I(k)

d=1

G G G

Figure 13.10: No Singularities

1 − d crystalline order (or 1 − d phase coexistance).


There are two other things associated with broken symmetry which I will
not discuss, but you can look up for yourself.

1. Broken symmetries are usually associated with a new (gapless, soft, or


acoustic) mode. For surfaces, these propagating modes are capillary waves.
For crystalline solids, they are shear–wave phonons.

2. Broken symmetries are associated with a “rigidity”. For crystalline solids,


this is simply their rigidity: it takes a lot of atoms to break a symmetry
of nature, and they all have to work together.
Chapter 14

Ising Model

There are many phenomena where a large number of constituents interact to


give some global behavior. A magnet, for example, is composed of many in-
teracting dipoles. A liquid is made up of many atoms. A computer program
has many interacting commands. The economy, the psychology of road rage,
plate techtonics, and many other things all involve phenomena where the global
behavior is due to interactions between a large number of entities. One way
or another, almost all such systems end up being posed as an ”Ising model” of
some kind. It is so common, that it is almost a joke.

The Ising model has a large number of spins which are individually in mi-
croscopic states +1 or −1. The value of a spin at a site is determined by the
spin’s (usually short–range) interaction with the other spins in the model. Since
I have explained it so vaguely, the first, and major import, of the Ising model is
evident: It is the most simple nontrivial model of an interacting system. (The
only simpler model would have only one possible microscopic state for each spin.
I leave it as an exercise for the reader to find the exact solution of this trivial
model.) Because of this, whenever one strips away the complexities of some
interesting problem down to the bone, what is left is usually an Ising model of
some form.

In condensed–matter physics, a second reason has been found to use the


simplified–to–the–bone Ising model. For many problems where one is interested
in large length scale phenomena, such as continuous phase transitions, the addi-
tional complexities can be shown to be unimportant. This is called universality.
This idea has been so successful in the study of critical phenomena, that it has
been tried (or simply assumed to be true) in many other contexts.

For now, we will think of the Ising model as applied to anisotropic magnets.
The magnet is made up of individual dipoles which can orient up, corresponding
to spin +1 or down corresponding to spin −1. We neglect tilting of the spins,

121
122 CHAPTER 14. ISING MODEL

arrow Spin S = +1

arrow Spin S = −1

Figure 14.1:

such as տ, or ր, or ց, and so on. For concreteness, consider dimension d = 2,


putting all the spins on a square lattice, where there are i = 1, 2, 3, ...N spins.
In the partition function, the sum over states is

or or . . .

Figure 14.2: Possible configurations for a N = 42 = 16 system.

X +1
X +1
X +1
X
= ...
{States} s1 =−1 s2 =−1 sN =−1
N X
Y +1
=
i=1 si =−1

where each sum has only two terms, of course. The energy in the exp −
EState /kB T now must be prescribed. Since we want to model a magnet, where
all the spins are +1 or −1, by and large, we make the spin of site i want to
oriented the same as its neighbors. That is, for site i
X
Energy of Site i = (Si − Sneighbors )2
neighbors of i
X
= Si2 − 2 2
Si Sneigh. + Sneigh.
neighbors of i

But Si2 = 1 always, so up to a constant (additive and multiplicative)


X
Energy of Site i = − Si Sneighbors of i
neighbors of i

On a square lattice, the neighbors of “i” are the spin surrounding it (fig. 14.3).
As said before, the idea is that the microscopic constituents -the spins- don’t
know how big the system is, so they interact locally. The choice of the 4 spins
123

Middle Spin is i, 4 Neighbors are


to the North, South, East, and West

Figure 14.3: Four nearest neighbors on a square lattice.

above, below and to the right and left as “nearest neighbors” is a convention.
The spins on the diagonals (NW, NE, SE, SW) are called “next–nearest neigh-
bors”. It doesn’t actually matter how many are included, so long as the inter-
action is local. The total energy for all spins is
X
EState = −J Si Sj
hiji

where J is a positive interaction constant, and hiji is a sum over all i and
j = 1, 2, ...N , provided i and j are nearest neighbors, with no double counting.
Note
N X N
X X 4
= =N×
i=1 j=1
2
hiji
| {z }
neighbors no double counting

If there are “q” nearest neighbors, corresponding to a different interaction, or a


different lattice (say hexagonal), or another dimension of space
X q
= N
2
hiji

So the largest (negative) energy is E = − qJ 2 N . If there is constant external


magnetic field H, which tends to align spins, the total energy is
X X
EState = −J Si Sj − H Si
hiji i

and the partition function is


N
Y +1
X ( k JT ) Sj Sk +( k HT )
P P
Sj
Z= e B hjki B i

i=1 S1 =−1
J H
= Z(N, , )
kB T kB T
As described above, the sum over q nearest neighbors is q = 4 for a two–
dimensional square lattice. For three dimensions, q = 6 on a simple cubic
lattice. The one–dimensional Ising model has q = 2.
124 CHAPTER 14. ISING MODEL

The one–dimensional model was solved by Ising for his doctoral thesis (we
will solve it below). The two–dimensional model was solved by Ousager, for H =
0, in one of the premiere works of mathematical physics. We will not give that
solution. you can read it in Landan and Lifshitz’s “Statistical Physics”. They
give a readable, physical, but still difficult treatment. The three–dimensional
Ising model has not been solved.
The most interesting result of the Ising model is that it exhibits broken
symmetry for H = 0. In the absence of a field, the energy of a state, and the
partition function, have an exact symmetry

Si ↔ −Si

(The group is Z2 .) Hence (?)

hSi i ↔ −hSi i

and
?
hSi i = 0
In fact, for dimension d = 2 and 3 this is not true. The exact result in d = 2 is
surprisingly simple: The magnetization/spin is
(
(1 − cosech2 k2J
BT
)1/8 , T < Tc
m = hSi i =
0, T > Tc

The critical temperature Tc ≈ 2.269J/kB is where (1−cosech2 k2JBT


)1/8 vanishes.
A rough and ready understanding of this transition is easy to get. Consider the
m = hSi i

0
Tc T

Figure 14.4: Two–dimensional square lattice Ising model magnetization per spin

Helmholtz free energy


F = E − TS
where S is now the entropy. At low temperatures, F is minimized by minimizing
E. The largest negative value E has is − 2q JN , when all spins are aligned +1 or
−1. So at low temperatures (and rigorously at T = 0), hSi i → +1 or hSi i → −1.
At high temperatures, entropy maximization minimizes F . Clearly entropy is
maximized by having the spins helter–skelter random: hSi i = 0.
The transition temperature between hSi i = 1 and hSi i = 0 should be when
the temperature is comparable to the energy per spin kB T = O( 2q J) which is
125

in the right ball park. There is no hand–waving explanation for the abrupt
increase of the magnetization per spin m = hSi i at Tc . Note it is not like a
paramagnet which shows a gradual decrease in magnetization with increasing
temperature (in the presence of an external field H). One can think of it (or

Paramagnet No Broken
1
H>0 Symmetry

m = hSi i

Figure 14.5:

at least remember it) this way: At Tc , one breaks a fundamental symmetry of


the system. In some sense, one breaks a symmetry of nature. One can’t break
such a symmetry in a tentative wishy–washy way. It requires a firm, determined
hand. That is, an abrupt change.
The phase diagram for the Ising model (or indeed for any magnet) look like
fig. 14.6. It is very convenient that H is so intimately related with the broken

Paramagnet Only on the shaded line


is there spontaneous
Tc magnetization (and
coexisting phases)

T Tc T
H=0 H>0

forbidden region
for single phase
−1 +1 m +1 m

T
H<0

−1 m

Figure 14.6:

symmetry of the Ising model. Hence it is very easy to understand that broken
symmetry is for H = 0, not H 6= 0. Things are more subtle in the lattice–gas
version of the Ising model.
126 CHAPTER 14. ISING MODEL

14.1 Lattice–Gas Model


In the lattice–gas model, one still has a lattice labeled i = 1, 2...N , but now
there is a concentration variable ci = 0 or 1 at each site. The most simple
model of interaction is the lattice–gas energe
X X
E LG = −ϑ ci cj − µ ci
hiji i

where ϑ is an interaction constant and µ is the constant chemical potential.


One can think of this as a model of a binary alloy (where Ci is the local
concentration of one of the two phases), or a liquid–vapour system (where Ci is
the local density).
This is very similar to the Ising model
X X
E Ising = −J Si Sj − H Si
hiji i

To show they are exactly the same, let


Si = 2Ci − 1
then
X X
E Ising = −J (2Ci − 1)(2Cj − 1) − H (2Ci − 1)
hiji i
X X X
= −4J Ci Cj + 4J Ci + Const. − 2H Ci + Const.
hiji hiji i
X qX X
= −4J Ci Cj + 4J Ci − 2H Ci + Const.
2 i i
hiji

And
X X
E Ising = −4J Ci Cj − 2(H − Jq ) Ci + Const.
hiji i

Since the additive constant is unimportant, this is the same as E LG , provided


ϑ = 4J

and

µ = 2(H − jq )
Note that the plane transition here is not at some convenient point µ = 0, it
is at H = 0, which is the awkward point µ = 2Jq. This is one reason why the
liquid–gas transition is sometimes said — incorrectly — to involve no change
of symmetry (Z2 group) as the Ising model. But the external field must have a
carefully tuned value to get that possibility.
14.2. ONE–DIMENSIONAL ISING MODEL 127

Lattice – gas transition


µ T
hci ≈ 1 µ = 2Jq
Tc
2Jq
hci ≈ 0

T hci
Tc 0 1

Real liquid – gas transition


P T

Pc Tc P = Pc

T
Tc Density

Figure 14.7:

14.2 One–dimensional Ising Model


It is easy to solve the one–dimensional Ising model. Of course, we already know
that it cannot exhibit phase coexistance, from the theorm we proved earlier.
Hence we expect
hSi i = m = 0
for all T > 0, where m is the magnetization per spin. The result and the method
are quite useful, though.

Y +1
N X PN PN
J
Sj Sj+1 + k HT Sj
Z= e kB T j=1 B j=1

i=1 Si =−1
P
Note we have simplified the hiji for one dimension. We need, however, a
prescrition for the spin SN +1 . We will use periodic boundary conditions so that
SN +1 ≡ S1 . This is called the Ising ring. You can also solve it, with a little
Ising RingN 1 Periodoc Boundary Conditions

N −1 2

3
. . . .

Figure 14.8:
128 CHAPTER 14. ISING MODEL

more work, with other boundary conditions. Let the coupling constant

k ≡ J/kB T

and
h = H/kB T
be the external field. Then
Y +1
N X PN
Z(k, h, N ) = e j=1 (kSj Sj+1 +hSj )

i=1 Si =−1

or, more symmetrically,


N X
Y +1
1
P
Z= e j [kSj Sj+1 + 2 h(Sj +Sj+1 )]

i=1 Si =−1

Y N
+1 Y
N X
1
= ekSj Sj+1 + 2 h(Sj +Sj+1 )
i=1 Si =−1 j=1

Let matrix elements of P be defined by


1
hSj |P |Sj+1 i ≡ ekSj Sj+1 + 2 h(Sj +Sj+1 )

There are only four elements

h+1|P | + 1i = ek+h
h−1|P | − 1i = ek−h
h+1|P | − 1i = e−k
h−1|P | + 1i = e−k

Or, in matrix form, · ¸


ek+h e−k
P =
e−k ek−h
The partition function is
N X
Y +1
Z= hS1 |P |S2 ihS2 |P |S3 i...hSN |P |S1 i
i=1 Si =−1
+1
X +1
X +1
X
= hS1 |P ( |S2 ihS2 |) P ( |S3 ihS3 |)...P |S1 i
S1 =−1 S2 =−1 S3 =−1

But, clearly, the identity operator is


X
1≡ |SihS|
S
14.2. ONE–DIMENSIONAL ISING MODEL 129

so,
+1
X
Z= hS1 |P N |S1 i
S1 =−1

= T race(P N )
Say the eigenvalues of P are λ+ and λ− . Then,
Z = λN N
+ + λ−

Finally, say λ+ > λ− . Hence as N → ∞


Z = e−F/kB T = λN
+

and
F = −N kB T λ+
It remains only to find the eigenvalues of P .
We get the eigenvalues from
det(λ − P ) = 0
This gives
(λ − ek+h )(λ − ek−h ) − e−2k = 0
or,
ek+h + ek−h
λ2 − 2λ ( ) +e2k − e−2k = 0
| 2
{z }
ek cosh h
The solutions are
p
2ek cosh h ± 4e2k cosh2 h − 4e2k + 4e−2k
λ± =
p 2
= ek [cosh h ± cosh2 h − 1 + e−4k ]
So, p
λ+ = ek (cosh h + sinh2 h + e−4k )
is the larger eigenvalue. The free energy is
p
F = −N kB T [k + ln (cosh h + sinh2 h + e−4k )]
where k = J/kB T , and h = H/kB T .
The magnetization per spin is
1 ∂F 1 ∂F
m=− =−
N ∂H N kB T ∂h
Taking the derivative and simplifying, one obtains
sinh h
m(h, k) = p
sinh2 h + e−4k
As anticipated, m(h = 0, k) = 0. There is magnetization for h 6= 0, although
there is no broken symmetry of course. There is a phase transition at T = 0.
130 CHAPTER 14. ISING MODEL

+1 low T
high T
m

−1

Figure 14.9: Magnetization of 1-d Ising model

14.3 Mean–Field Theory


There is a very useful approximate treatment of the Ising model. We will now
work out the most simple such “mean–field” theory, due in this case to Bragg
and Williams.
The idea is that each spin is influenced by the local magnetization, or equiva-
lently local field, of its surrounding neighbors. In a magnetized, or demagnetized
state, that local field does not fluctuate very much. Hence, the idea is to replace
the local field, determined by a local magnetization, with an average field, de-
termined by the average magnetization. “Mean” field means average field, not
nasty field.
It is easy to solve the partition function if there is only a field, and no
interactions. This is a (very simple) paramagnet. Its energy is
X
EState = −H Si
i

and,
+1
Y X H
P
Sj
Z= e kB T j

i Si =−1
X H
S1
X H
SN
=( e kB T )...( e kB T )
S1 =±1 SN =±1
X H
S
=( e kB T )N
S=±1

so,

H N
Z = (2 cosh )
kB T
and
H
F = −N kB T ln 2 cosh
kB T
14.3. MEAN–FIELD THEORY 131

so since,
1
m=− ∂F/∂H
N
then
m = tanh H/kB T
For the interacting Ising model we need to find the effective local field Hi at
+1
1 H>0
m

H m

−1
T

Figure 14.10: Noninteracting paramagnet solution

site i. The Ising model is


X X
EState = −J Si Sj − H Si
hiji i

we want to replace this — in the partition function weighting — with


X
EState = − Hi S i
i

Clearly
∂E
− = Hi
∂Si
Thus, from the Ising model
∂E X
− =J (Sj δk,i + δj,i Sk ) + H
∂Si
hjki
X
=J ·2· Sj + H
j neighbors of i
(no double counting)

Now let’s get rid of the 2 and the double–counting restriction. Then
X
Hi = J Sj + H
j neighbors of i

is the effective local field. To do mean–field theory we let

Hi ≈ hHi i
132 CHAPTER 14. ISING MODEL

where
X
hHi i = J hSj i + H
j neighbors

but all spins are the same on average, so

hHi i = qJhSi + H

and the weighting of states is given by


X
EState = − (qJhSi + H) Si
| {z }
i
mean field, hHi

For example, the magnetization is


P hHi
S
S=±1 Se kB T
m = hSi = P hHi
S
S=±1 e kB T

where hSi = hS1 i, and contributions due to S2 , S3 ... cancel from top and bottem.
So
hHi
m = tanh
kB T
or

qJm + H
m = tanh
kB T
This gives a critical temperature

kB Tc = qJ

as we will now see.


Consider H = 0.
m
m = tanh
T /Tc
Close to Tc , m is close to zero, so we can expand the tanh function:

m 1 m3
m= − + ...
T /Tc 3 (T /Tc )3
One solution is m = 0. This is stable above Tc , but unstable for T < Tc .
Consider T < Tc , where m is small, but nonzero. Dividing by m gives

1 1 m2
1= − + ...
T /Tc 3 (T /Tc )3
14.3. MEAN–FIELD THEORY 133

Rearranging

T 2 T
m2 = 3( ) (1 − ) + ...
Tc Tc

Or,

T 1/2 T
m = (3 ) (1 − )1/2
Tc Tc
for small m, T < Tc . Note that this solution does not exist for T > Tc . For
T < Tc , but close to Tc , we have
T β
m ∝ (1 − )
Tc
where the critical exponent, β = 1/2. All this is really saying is that m is a
m
H=0
Close to Tc , mean
field gives m ∝ (1 − T /Tc )1/2

Tc T

Figure 14.11:

parabola near Tc , as drawn in fig. 14.12 (within mean field theory). This is

+1 Parabola near m = 0, T = Tc , H = 0
m

T “Classical” exponent β = 1/2


same as parabola
−1

Figure 14.12:

interesting, and useful, and easy to generalize, but only qualitatively correct.
Note 
2J
 ,d = 1
kB Tc = qJ = 4J ,d = 2


6J ,d = 3
The result for d = 1 is wildly wrong, since Tc = 0. In d = 2, kB Tc ≈ 2.27J, so
the estimate is poor. Also, if we take the exact result from Ousageris solution,
134 CHAPTER 14. ISING MODEL

m ∝ (1 − T /Tc )1/8 , so β = 1/8 in d = 2. In three dimensions kB Tc ∼ 4.5J


and estimates are β = 0.313.... Strangely enough, for d ≥ 4, β = 1/2 (although
there are logarithmic corrections in d = 4)!
For T ≡ Tc , but H 6= 0, we have
H
m = tanh (m + )
kB Tc
or

H = kB Tc (tanh−1 m − m)

Expanding giving

1
H = kB Tc (m + m3 + ... − m)
3
so
kB Tc 3
H= m
3
for m close to zero, at T = Tc . This defines another critical exponent

T = Tc Shape of critical isotherm


m
is a cubic H ∝ m3
in classical theory
H

Figure 14.13:

H ∝ mδ

for T = Tc , where δ = 3 in mean–field theory.


The susceptibility is
∂m
XT ≡ ( )T
∂H
This is like the compressability in liquid–gas systems
1 ∂V
κT = − ( )T
V ∂P
Note the correspondence between density and magnetization per spin, and be-
tween pressure and field. The susceptibility “asks the question”: how easy is it
to magnetize something with some small field. A large XT means the material
is very “susceptible” to being magnetized.
14.3. MEAN–FIELD THEORY 135

Also, as an aside, the “sum rule” for liquids


Z
d~rh∆n(r)∆n(0)i = n2 kB T κT

has an analog for, magnets


Z
d~rh∆m(r)∆m(0)i ∝ XT

We do not need the proportionality constant.


From above, we have
kB Tc m + H
m = tanh
kB T
or,
H = kB T · tanh−1 m − kB Tc m
and
∂H 1
( )T = kB T − kB Tc
∂m 1 − m2
so
1
XT =
T − Tc
close to Tc . Or
XT ∝ |T − Tc |−γ
where γ = 1. This is true for T → Tc+ , or, T → Tc− . To get the energy, and so

χT
χT , susceptibility
diverges near Tc

Tc

Figure 14.14:

the heat capacity, consider


X X
hEi = h−J Si Sj − H Si i
hiji i

All the spins are independent, in mean–field theory, so

hSi Sj i = hSi ihSj i

Hence
Jq
hEi = − N m2 − HN m
2
136 CHAPTER 14. ISING MODEL

The specific heat is


1 ∂E
c= ( )N
N ∂T
Near Tc , and at H = 0,
(
2 (Tc − T ) T < Tc
m ∝
0 T > Tc

So, (
−(Const.)(Tc − T ), T < Tc
hEi =
0, T > Tc
and (
Const. T < Tc
c=
0 T > Tc
If we include temperature dependence away from Tc , the specific heat looks
more like a saw tooth, as drawn in fig. 14.15. This discontinuity in c, a second
C

Tc

Figure 14.15: Discontinuity in C in mean field theory.

derivative of the free energy is one reason continuous phase transitions were
called “second order”. Now it is known that this is an artifact of mean field
theory.
It is written as
c ∝ |T − Tc |−α
where the critical exponent
α = o(disc.)
To summarize, near the critical point,

m ∼ (Tc − T )β , T < Tc , H = 0
δ
H∼m , T = Tc
−γ
XT ∼ |T − Tc | , H=0

and

c ∼ |T − Tc |−α , H=0
14.3. MEAN–FIELD THEORY 137

The exponents are:

mean field d = 2 (exact) d = 3 (num) d≥4


α 0 (disc.) 0 (log) 0.125... 0 (disc.)
β 1/2 1/8 0.313... 1/2
γ 1 7/4 1.25... 1
δ 3 15 5.0... 3
Something is happening at d = 4. We will see later that mean–field theory
works for d ≥ 4. For now we will leave the Ising model behind.

0.5 non trivial


0.4 no exponents
transition
β 0.3 mean field
theory works
0.2
0.1
1 2 3 4 5
d

Figure 14.16: Variation of critical exponent β with spatial dimension


138 CHAPTER 14. ISING MODEL
Chapter 15

Landan Theory and


Ornstien–Zernicke Theory

Nowadays, it is well established that, near the critical point, one can either solve
the Ising model
X X
E = −J Si Sj − H Si
hiji i

with
Y X
Z= e−E/kB T
i Si =±1

or, the ψ 4 field theory


Z Z
k r 2 u 4
F = d x{ (∇ψ) + ψ + ψ } − H dd x ψ(~x)
d 2
2 2 4

where k, u, and H are constants, and

r ∝ T − Tc

with the partition function


Z −Z
Z= D ψ(·) e−F

where the kB T factor has been set to kB Tc , and incorporated into k, r, u and
H (this is ok near Tc ).
Look back in the notes for the discussion of the discrete Gaussian model
versus the continuous drumhead model of interfaces. Although it may not look
like it, the field theory model is a very useful simplification of the Ising model,
wherein no important features have been lost.

139
140CHAPTER 15. LANDAN THEORY AND ORNSTIEN–ZERNICKE THEORY

To motivate their similarities, consider the following. First, the mapping


between peices involving external field only require getting used to notation:
Ising Model ψ 4 Field Theory
N
X Z
−H Si ⇐⇒ −H dd xψ(~x)
i=1

One is simply the continuum limit of the other. The square gradient term is
likewise straightforward, as we now see.
Ising Model ψ 4 Field Theory
N Z
JX X
(Si − Sj )2 ⇐⇒ K dd x(∇ψ)2
2 i=1
j neighbors of i

yusing Si2 = 1
X
const. − J Si Sj
hiji
R
The terms harder to see are the dd ~x ( 2r ψ 2 + u4 ψ 4 ) terms. Let us define
r 2 u 4
f (ψ) = ψ + ψ
2 4
This is the free energy for a homogeneous system in the ψ 4 field theory (that is,
~ ≡ 0), in the absence of an external field, per unit volume. Note
where ∇ψ
r > 0, T > Tc
r < 0,T < Tc
So we have the double well form. This may look familiar, it was originally due

f f
T > Tc T < Tc
r>0 r<0

ψ ψ

~
Figure 15.1: “free energy”/volume. No ∆ψ, no H

to Van der Waals. (In fact, the whole ψ 4 theory is originally due to Van der
Waals.)
Note that, since ψ is determined by the minimum of F , this single or double
well restricts ψ to be within a small range. Otherwise, ψ could have any value,
since Z −Z YZ ∞
Dψ(·) = dψ(~x)
− ~
x −∞
141

These unphysical values of ψ are suppressed in the free energy. In the Ising
model, the restriction is Si = ±1. We can lift this restriction with a kronecker
delta
+1
Y X ∞
Y X
= δSi2 −1
i Si =−1 i Si =−∞
Y X ∞
− ln (1/δS 2 −1)
= e i

i Si =−∞

So, although this is awkward and artificial, it is sort of the same thing. Expand-
ing it out, as Si2 → 0,

− ln (1/δSi2 −1 ) ≈ Si2 − O(Si4 )

the double–well form. This can be done in a very precise way near Tc . Here I
am just trying to give an idea of what is going on.
So the ψ 4 theory is similar to the Ising model, where ψ(~x) is the local
magnetization. We will be concerned with the critical exponents α, β, γ, δ, η,
and ν defined as follows

F (T ) ∼ |T − Tc |2−α ,H = 0

∂2F
So that C = ∂T 2 ∼ |T − Tc |−α

ψ ∼ (Tc − T )β , H = 0 , T < Tc
∂ψ
( )T = XT ∼ |T − Tc |−γ ,H = 0
∂H
ψ ∼ H 1/δ , T = Tc
1
hψ(~x)ψ(0)i ∼ d−2+η , T = Tc , H = 0
x
(
e−x/ξ , T > Tc
hψ(~x)ψ(0)i ∼ 2 −x/ξ
hψi + O(e ) , T < Tc

and

ξ ∼ |T − Tc |−ν ,H = 0

This is the first time we have mentioned η and ν in this context.


Landan motivates the ψ 4 theory in a more physical way. He asserts that a
continuous phase transition involves going from a disordered phase, with lots
of symmetry, to an ordered phase, with a lower symmetry. To describe the
transition we need a variable, called the order parameter which is nonzero in
the ordered phase only. For example, in a magnet, the order parameter is the
magnetization. In general, identifying the order parameter, and the symmetery
142CHAPTER 15. LANDAN THEORY AND ORNSTIEN–ZERNICKE THEORY

broken at Tc can be quite subtle. This will take us too far afield, so we will
only consider a scalar order parameter, where the symmetry broken is Z2 , that
is, where ψ ↔ −ψ symmetry is spontaneously broken. (Note that a magnetic
spin confined to dimension d = 2, wherein all directions are equivalent, has an
order parameter ψ ~ = ψx x̂ + ψy ŷ. The group in u(1). A spin in d = 3 has
~ = ψx x̂ + ψy ŷ + ψz ẑ, if all directions are equivalent; the group in O(3) I think.)
ψ
So Landan argues that the free energy to describe the transition must be an
analytic function of ψ. Hence
Z
F = dd x f (ψ)

where
r 2 u 4
f (ψ) = ψ + ψ + ...
2 4
No odd terms appear, so that the

ψ ↔ −ψ

Symmetry is preserved. In the event an external field breaks the symmetry,


another term is added to the free energy, namely
Z
−H dd x ψ

At high temperatures, in the disordered phase, we want ψ = 0. Hence

r>0

for T > Tc . However, at low temperatures, in the ordered phase, we must have
ψ > 0. Hence
r<0
(To make sure everything is still well defined, we need u > 0. There is no need
for u to have any special dependence in T near Tc , so we take it to be a constant.
Or, if you prefer

: unimportant
u(T ) ≈ u(Tc ) + (Const.)(T − Tc ) + ...

near Tc . So we simply set u to be constant.)


To find r’s evident T dependence, we Taylor expand

r = r0 (T − Tc ) + ...

keeping only the lowest–order term, where r0 > 0.


Finally, we know that spatial fluctuations should be suppressed. Expanding
in powers of ∇~ and ψ, the lowest order nontrivial term is
Z
1
K dd r |∇ψ|2
2
143

R
where K > 0. Similar terms like ψ∇2 ψ turn into this after parts integration.
~ 6= 0) add free energy.
Note that this term means inhomogeneities (∇ψ
Now we’re done. The Landan free energy is
Z Z
k
F = dd x {f (ψ) + |∇ψ|2 } − H dd x ψ
2

with
r 2 u 4
f (ψ) = ψ + ψ
2 4
In high–energy physics, this is called a ψ 4 field theory. The ∇2 term gives
kinetic energy; r ∝ mass of particle; and ψ 4 gives interactions. Van der Waals
wrote down the same free energy (the case with k ≡ 0 is what is given in the
text books). He was thinking of liquid–gas transitions, so ψ = n − nc , where n
is density and nc is the density at the critical point. Finally, from the theory of
thermodynamic fluctuations, we should expect that (in the limit in which the
square gradient and quartic term can be neglected)

r ∝ 1/κT

or equivalently
r ∝ 1/XT
It is worth emphasizing that the whole machinery of quadratic, quartic, and
square gradient terms is necessary. For example, note that the critical point is
defined thermodynamically as

∂P
( )T = 0
∂V c

and

∂2P
( )T = 0
∂V 2 c
in a liquid–gas system. In a magnet:

∂H ∂2H
( )Tc = 0, and( )T = 0
∂m ∂m2 c
Trying to find the equation of state for T ≡ Tc in a liquid we expand:

∂P 1 ∂2P 1 ∂3P
P (n, Tc ) = Pc + ( )Tc (n − nc ) + ( 2 )Tc (n − nc )2 + ( 3 )Tc (n − nc )3 + ...
∂n 2 ∂n 6 ∂n
For a magnet

0 ∂H 0 1 ∂2H 1 ∂3H
H(m, Tc ) = H>
c+( )Tc (m − m*
c) + ( )T m 2
+ ( )T m3 + ...
∂m 2 ∂m2 c 2 ∂m3 c
144CHAPTER 15. LANDAN THEORY AND ORNSTIEN–ZERNICKE THEORY

So, at the critical point for both systems — from thermodynamics and analyticity
only — we have
(P − Pc ) ∝ (n − nc )δ
and
H ∝ mδ
where
δ=3
Experimentally, δ ≈ 5.0... in three dimensions, while δ ≡ 15 in d = 2. So we
have a big subtle problem to solve!

15.1 Landan Theory


~ = 0,
We now solve for the homogeneous states of the system. In that case ∇ψ
so we need only consider

F/V = f (ψ) − Hψ
r0 u
= (T − Tc )ψ 2 + ψ 4 − Hψ
2 4
For H = 0, we find the solution by minimizing F/V (or f ):

∂f
= 0 = r0 (T − Tc )ψ + uψ 3
∂ψ

There are 3 solutions

ψ = 0, (stable f or T > Tc )
r
r0 (Tc − T )
ψ=± , (stable f or T < Tc )
u

(To check stability, consider ∂ 2 f /∂ψ 2 .) for H 6= 0, only one solution is stable
for T < Tc . You can check this yourself.

∂ ∂f
(f (ψ) − Hψ) = 0 = −H
∂ψ ∂ψ
or
H = r0 (T − Tc )ψ + uψ 3
Note that this gives
∂H
1/XT = ( )T = r0 (T − Tc )
∂ψ
close to the critical point, or

XT ∝ |T − Tc |−1
15.2. ORNSTEIN — ZERNICKE THEORY OF CORRELATIONS 145

Also, at T = Tc , this gives


H ∝ ψ3
Finally, note that (at H = 0)
1 2 1 4
f= rψ + uψ
2 4
but (
2 0 , T > Tc
ψ =
− ur , T < Tc
so, (
0 , T > Tc
f= 2
− 41 ru , T < Tc
or (
0 , T > Tc
f∝
(Tc − T )2 , T < Tc
but c ∝ ∂ 2 f /∂T 2 , so the heat capacity
(
0 , T > Tc
c∝
Const. , T < Tc

In summary, these are the same results as we obtained from the mean field
theory of the Ising model. The relationship between analyticity of equations of
state, and mean–field theory, is worth thinking about.

α 0 (disc.)
β 1/2
γ 1
δ 3

15.2 Ornstein — Zernicke Theory of Correla-


tions
Ornstein and Zernicke give a physically plausible approach to calculating corre-
lation functions. It works well (except near the critical point).
In a bulk phase, say T >> Tc , and we shall only consider H = 0, ψ has an
average value of zero:
hψi = 0
So we expect
ψ 4 << ψ 2
and, above Tc , Z
k r
F ≈ dd x[ (∇ψ)2 + ψ 2 ]
2 2
146CHAPTER 15. LANDAN THEORY AND ORNSTIEN–ZERNICKE THEORY

Far below Tc , we expect ψ to be closse to its average value


r
r
hψi = ± −
u
where r = r0 (T − Tc ) = negative below Tc . Let

∆ψ ≡ ψ − hψi

Then
~ = ∇(∆ψ
∇ψ ~ ~
− hψi) = ∇(∆ψ)
Also,
ψ 2 = (∆ψ − hψi)2 = (∆ψ)2 − 2hψi∆ψ + hψi2
and

ψ 4 = (∆ψ − hψi)4
= (∆ψ)4 − 4hψi(∆ψ)3 + 6hψi2 (∆ψ)2 − 4hψi3 (∆ψ) + hψi4

Hence, grouping terms by powers of ∆ψ,


r 2 u 4 r 3u
ψ + ψ = (−rhψi − uhψi3 )∆ψ + ( + hψi2 )(∆ψ)2 + O(∆ψ)3
2 4 2 2
p
But hψi = ± −r/u, so
r 2 u 4
ψ + ψ = −r(∆ψ)2 + O((∆ψ)3 )
2 4
But (∆ψ)3 << (∆ψ)2 , so, below Tc
Z
k
F ≈ dd x [ (∇∆ψ)2 + |r|(∆ψ)2 ]
2
where we have used the fact that r is negative definite below Tc to write

−r = |r|

Both approximate free energies are pretty much the same as the free energy for
the interface model: Z
σ ∂h
dd−1 x( )2
2 ∂x
They can be solved exactly the same way. The definitions of Fourier transforms
are the same (except for dd−1 x → dd x, and Ld−1 → Ld , for example), and the
arguments concerning translational invariance of space, and isotropy of space
still hold. Let
C(x) ≡ hψ(x) ψ(0)i
−F
Then we have (putting the kB T back in e kB T )
kB T
Ĉ(k) = , T > Tc
Kk 2 + r
15.2. ORNSTEIN — ZERNICKE THEORY OF CORRELATIONS 147

and since
|r|
h∆ψ(x) ∆ψ(0)i = hψ(x) ψ(0)i −
u
using hψi2 = |r|/u, we have
kB T |r|
Ĉ(k) = − (2π)d δ(k) , T < Tc
Kk 2 + 2|r| u
These can be Fourier transformad back to real space.
As an asside, remember the Green function for the Helmholtz equation sat-
isfies
(∇2 − ξ −2 )G(x) = −δ(x)
Hence
Ĝ(k) = 1/(k 2 + ξ −2 )
The Green function is given in many books

−x/ξ
ξ/2 e
 , ,d=1
1
G(x) = 2π K0 (x/ξ) ,d=2

 1 −x/ξ
4πx e ,d=3

for example. Recall the modified Bessel function of zeroth order satisfies K0 (y →
0) = − ln y, and K0 (y → ∞) = (π/2y)1/2 e−y . This is more than we need. In
discussions of critical phenomena, most books state that if
1
Ĝ(k) ∝
k 2 + ξ −2
then
1
G(x) ∝ e−x/ξ
xd−2
The reason for the seemingly sloppiness is that proportionality constants are
unimportant, and it is only long–length–scale phenomena we are interested in.
Indeed, we should only trust out results for large x, or equivalently small k
~
(since we did a gradient expansion in ∇ψ).
In fact, so far as the Fourier transforms go, here is what we can rely on:
If
1
Ĝ(k) ∝
k 2 + ξ −2

then

G(x) ∝ e−x/ξ

unless ξ = ∞, in which case

1
G(x) ∝
xd−2
148CHAPTER 15. LANDAN THEORY AND ORNSTIEN–ZERNICKE THEORY

for large x.
Using these results, we finally obtain the correlation function up to a constant
of proportionality,


−x/ξ
e
 , , T > Tc
1
C(x) = xd−2
, T = Tc

 2
hψi + O(e−x/ξ ) , T < Tc

where the correlation length:

q
 k , T > Tc
ξ = qr
k

2|r| , T < Tc

Note that the result for T < Tc

C(x) = hψ(x) ψ(0)i = hψi2 + O(e−x/ξ )

is a rewriting of

h∆ψ(x) ∆ψ(0)i = e−x/ξ

p
In cartoon form: figure 15.2. Note the correlation length satisfies ξ ∼ 1/ |r|

ξ Correlation length ξ
ξ ∼ √1
|r|
or
ξ ∼ |T − Tc |−1/2

Figure 15.2:

above and below Tc , so

ξ ∼ |T − Tc |−1/2

in Ornstein–Zernicke theory. The correlation function becomes power–law–like


at Tc .
15.2. ORNSTEIN — ZERNICKE THEORY OF CORRELATIONS 149

C(x)
T < Tc

T > Tc T = Tc

Figure 15.3: Correlation function above at and below Tc

1
C(x) = , T = Tc , (H = 0)
xd−2
Since
ξ ∼ |T − Tc |ν
and
1
C(x) ∼
xd−2+η
we have ν = 1/2 and η = 0.
This completes the list of critical exponents.
In summary:

Mean Field Ising Landan 0 − Z d = 2 Ising d = 3 Ising d ≥ 4 Ising


α 0 (disc.) 0 (disc.) 0 (log) 0.125... 0
β 1/2 1/2 1/8 0.313... 1/2
γ 1 1 7/4 1.25... 1
δ 3 3 15 5.0... 3
η 0 0 1/4 0.041... 0
ν 1/2 1/2 1 0.638... 1/2
So something new is needed.
150CHAPTER 15. LANDAN THEORY AND ORNSTIEN–ZERNICKE THEORY
Chapter 16

Scaling Theory

Historically, the field of phase transitions was in a state of termoil in the 1960’s.
Sensible down–to–earth theories like the mean–field approach, Landan’s ap-
proach, and Ornstein–Zernicke theory were not working. A lot of work was put
into refining these approaches, with limited success.

For example, the Bragg–Williams mean field theory is quite crude. Refine-
ments can be made by incorporating more local structure. Pathria’s book on
Statistical Mechanics reviews some of this work. Consider the specific heat of
the two–dimensional Ising model. Since Ousager’s solution, this was known
exactly. It obeys C ∼ ln |T − Tc |, near Tc . The refinements make for much
better predictions of C away from Tc . At Tc , all mean field theories give C
to be discontinuous. Comparing to the exact result shows this as well. (This

Crude mean field Better mean field better . . .

C C C

T T T
Tc Tc Tc

All mean theories give C ∼ discontinuous

Figure 16.1:

151
152 CHAPTER 16. SCALING THEORY

Figure 16.2:

is taken from Stanley’s book, “Introduction to Phase Transitions and Critical


Phenomena”.)
Experiments consistently did not (and do not of course) agree with Landan’s
prediction for the liquid–gas or binary liquid case. Landan gives
n − nc ∼ (Tc − T )β
with β = 1/2. Three–dimensional experiments gave β ≈ 0.3 (usually quoted as
1/3 historically), for many liquids. The most famous paper is by Guggenheim
in J. Chem. Phys. 13, 253 (1945).

Figure 16.3:

This is taken from Stanley’s book again. The phenomena at Tc , where


153

ξ → ∞ is striking.

Figure 16.4:

Here I show some (quite old) simulations of the Ising model from Stanley’s
book. Note the “fractally” structure near Tc . You can easily program this
yourself if you are interested. Nowadays, a PC is much much faster than one
needs to study the two–dimensional Ising model. Strange things are also seen
experimentally.
For a fluid, imagine we heat it through Tc , fixing the pressure so that it
154 CHAPTER 16. SCALING THEORY

corresponds to phase coexistence (i.e., along the liquid–vapor line in the P–T
diagram as drawn). The system looks quite different depending on temperature.

liquid heat this way

vapor

Figure 16.5:

It is transparent, except for the meniscus separating liquid and vapour for T <

Transparent Opaque Transparent

P ≡ Pliquid vapor coexistance


Vapor Milky
Looking Gas
Liquid Mess

T < Tc T = Tc T > Tc

Figure 16.6:

Tc , for T < Tc and T > Tc . But at Tc , it is milky looking. This is called “critical
opalescence”. The correlation length ξ, corresponding to fluctuations, is so big
it scatters light.
Rigorous inequalities were established with the six critical exponents. (I
should mention that I am assuming any divergence is the same for T → Tc+ or
T → Tc− , which is nontrivial, but turns out to be true.) For example,
Rushbrooke proved
α + 2β + γ ≥ 2

Griffiths proved
α + β(1 + δ) ≥ 2

These are rigorous. Other inequalities were proved with some plausible assump-
tions. These all work for the exponents quoted previously — as equalities! In
fact, all the thermodynamic inequalities turned out to be obeyed as equalities.
These inequalities were very useful for ruling out bad theories and bad experi-
ments.
16.1. SCALING WITH ξ 155

Here is a list of all the independent relations between exponents

α + 2β + γ = 2
γ = ν(2 − η)
γ = β(δ − 1)

and
νd = 2 − α
So it turns out there are only two independent exponents. The last equation is
called the “hyperscaling” relation. It is the only one involving spatial dimension
d. The others are called “scaling” relations. This is because the derivation of
them requires scaling arguments.

16.1 Scaling With ξ


Note that near Tc , ξ diverges. Also, correlation functions become power–law
like at Tc according to the approximate Ornstein–Zernicke theory as well as the
exact result for the Ising model in two dimensions.
So let us make the bold assertion that, since ξ → ∞ at Tc , the only observable
information on a large length scale involves quantities proportional to ξ. That
is, for example, all lengths scale with ξ.
Let us now derive the hyperscaling relation. The free energy is extensive,
and has units of energy, so
V
F = kB T d
ξ
One can think of V /ξ d as the number of independent parts of the system. But

ξ ∼ |T − Tc |−ν

so
F = |T − Tc |νd
But the singular behavior in F is from the heat capacity exponent α via

F = |T − Tc |2−α

Hence
νd = 2 − α
This back–of–the–envelope argument is the fully rigorous (or as rigorous as it
comes) argument! The hyperscaling relation works for all systems, except the
mean field solutions must have d = 4.
To derive the three scaling relations, we assert that correlations are scale
invariant:
1 ∼
C(x, T, H) = d−2+η C(x/ξ, Hξ y )
x
156 CHAPTER 16. SCALING THEORY

Note that, since blocks of spin are correlated on scale ξ, this means only a little
H is needed to magnetize the system. So the effect of H is magnified by ξ. The
factor is ξ y , where y must be determined. In a finite–size system we would have

1 ∼
C(x, T, H L) = C(x/ξ, Hξ y , L/ξ)
xd−2+η
In any case, in the relation above

ξ ∼ |T − Tc |−ν

For ease of writing, let


t ≡ |T − Tc |
So that ξ ∼ t−ν . Then we have

1 ∼
C(x, T, H) = C(xtν , H/t∆ )
xd−2+η
where
∆ = yν
is called the “gap exponent”. But the magnetic susceptibility satisfies the sum
rule
Z
XT ∝ dd xC(x, T, H)
Z
1 ∼
= dd x d−2+η C(xtν , H/t∆ )
x
Z ∼
ν ∆
ν −2+η d ν C(xt , H/t )
= (t ) d (xt ) ν d−2+η
(xt )
so,

XT (T, H) = t(−2+η)ν X(H/t∆ )
where Z ∞
1 ∼ ∼ H
dd (xtν ) C(xtν , H/t∆ ) ≡ X( )
−∞ (xtν )d−2+η t∆
But
X(T, H = 0) = t−γ
Hence
γ = (2 − η)ν
called Fisher’s law.
Now note
∂ψ
(T, H) = XT (T, H)
∂H
16.1. SCALING WITH ξ 157

so integrating, we have the indefinite integral


Z
ψ(T, H) = dH X(T, H)
Z ∼ H
= dH t−γ X( ∆ )
t
Z
H ∼ H
= t∆−γ d( ∆ ) X( ∆ )
t t

= t∆−γ ψ(H/t∆ )

But
ψ(T, H = 0) = tβ
So
β =∆−γ
and the gap exponent
∆=β+γ
or equivalently, y = (β + γ)/ν. So we have
∼ H
ψ(T, H) = tβ ψ( )
tβ+γ
Here is a neat trick for changing the dependence “out front” from t to H. We

H
don’t know what the function ψ is. All we know is that it is a function of tβ+γ
.

Any factor of H/tβ+γ that we multiply or divide into ψ leaves another unknown
function, which is still a function of only H/tβ+γ . In particular, a crafty choice
can cancel the tβ factor:
H β H −β ∼ H
ψ(T, H) = tβ ( ) β+γ ( ) β+γ ψ( )
tβ+γ tβ+γ tβ+γ


We call the last term on the right–hand side ψ, and obtain

β/β+γ
∼ H
ψ(T, H) = H ψ( )
tβ+γ

∼ ∼
Note that H −β/β+γ ψ(H) = ψ(H). Now

ψ(Tc , H) = H 1/δ

is the shape of the critical isotherm, so δ = 1 + γ/β, or

γ = β(δ − 1)

called the Widom scaling law.


158 CHAPTER 16. SCALING THEORY

Note that ψ = ∂F/∂H, so


Z
F = dH ψ
Z ∼ H
= dH tβ ψ( β+γ )
t
Z
2β+γ H ∼ H
=t d( β+γ )ψ( β+γ )
t t
or,
∼ H
F (T, H) = t2β+γ F ( )
tβ+γ
But
F (T, 0) = t2−α
so 2 − α = 2β + γ, or
α + 2β + γ = 2
which is Rushbrooke’s law.
The form for the free energy is
∼ H
F (T, H) = t2−α F ( )
tβ+γ
Note that
∂nF
lim = t2−α t−n(β+γ)
H→0 ∂H n
or
F (n) = t2−α−n(β+γ)
So all derivatives of F are separated by a constant “gap” exponent (β + γ).

yn = 2 − α − n(β + γ)

yn free energy

2
magnetization
1
0 1 2 3
Susceptability
−1
−2

Figure 16.7: Gap Exponent

Finally, we will give another argument for the hyperscaling relation. It is


not, however, as compelling as the one above, but it is interesting.
16.1. SCALING WITH ξ 159

Near Tc , we assert that fluctuations in the order parameter are of the same
order as ψ itself. Hence
h(∆ψ)2 i
= O(1)
hψi2
This is the weak part of the argument, but it is entirely reasonable. Clearly

hψi2 ∼ t2β

To get h(∆ψ)2 i, recall the sum rule


Z
dd xh∆ψ(x)∆ψ(0)i = XT

The integral is only nonzero for x ≤ ξ. In that range,

h∆ψ(x)∆ψ(0)i ∼ Oh(∆ψ)2 i

Hence Z
dd xh∆ψ(x)∆ψ(0)i = ξ d h(∆ψ)2 i = XT
or,
h(∆ψ)2 i = XT ξ −d
But

XT ∼ t−γ

and

ξ ∼ t−ν

so,

h(∆ψ)2 i = tνd−γ

The relative fluctuation therefore obeys

h(∆ψ)2 i
= tνd−γ−2β
hψi2

But if, as we argued


h(∆ψ)2 i
= O(1)
hψi2
we have
νd = γ + 2β
which is the hyperscaling relation. (To put it into the same form as before, use
Rushbrooke scaling law of α + 2β + γ = 2. This gives νd = 2 − α.)
160 CHAPTER 16. SCALING THEORY

As a general feature
h(∆ψ)2 i
hψi2
measures how important fluctuations are. We discussed this at the beginning of
the course. Mean field theory works best when fluctuations are small. So we say
that mean–field theory is correct (or at least self–consistent) if its predictions
give
h(∆ψ)2 i
<< 1
hψi2
This is called the Ginzburg criterion for the correctness of mean–field theory.
But from above
h(∆ψ)2 i
= tνd−γ−2β
hψi2
Putting in the results from mean–field theory (ν = 1/2, γ = 1, β = 1/2) gives

h(∆ψ)2 i 1
= t 2 (d−4)
hψi2

Hence, mean–field theory is self consistent for d > 4 as T → Tc (t → 0). Mean–


field theory doesn’t work for d < 4.
This is called the upper critical dimension; the dimension at which fluctua-
tions can be neglected.
Chapter 17

Renormalization Group

The renormalization group is a theory that builds on scaling ideas. It gives a


way to calculate critical exponents. The idea is that phase transition properties
depend only on a few long wavelength properties. These are the symmetry of
the order parameter and the dimension of space, for systems without long–range
forces or quenched impurities. We can ignore the short length scale properties,
but we have to do that “ignoring” in a controlled carefil way. That is what the
renormalization group does.
The basic ideas were given by Kadanoff in 1966, in his “block spin” treatment
of the Ising model. A general method, free of inconsistencies, was figured out
by Wilson (who won the Nobel prize for his work). Wilson and Fisher figured
out a very useful implementation of the renormalization group in the ǫ = 4 − d
expansion, where the small parameter ǫ was eventually extrapolated to unity,
and so d → 3.
At a critical point, Kandanoff reasoned that spins in the Ising model act
together. So, he argued one could solve the partition function recursively, by
progressively coarse–graining on scales smaller than ξ. From the recursion rela-
tion describing scale changes, one could solve the Ising model. This would only
be a solution on large length scales, but that is what is of interest for critical
phenomena.
The partition function is
X P P
Z(N, K, h) = eK hiji Si Sj +h i Si

{States}

where K = J/kB T and h = H/kB T .


Now, let us (imagine we) reorganize all of these sums as follows.

1. Divide the lattice into blocks of size bb (we’ll let b = 2 in the cartoon
below).

2. Each block will have an internal sum done on it leading to a block spin

161
162 CHAPTER 17. RENORMALIZATION GROUP

S = ±1. For example, the majority–rule transformation is

 P
+1, d Si > 0

 Piǫb
Si = −1, Piǫb
d Si < 0

 ±1
 , iǫbd Si = 0
randomly

Clearly, if N is the number of spins originally, N/bd is the number of spins after
blocking.
It is easy to see that this is a good idea. Any feature correlated on a finite
length scale l before blocking becomes l/b after blocking once. Blocking m times
means this feature is now on the small scale l/bm . In particular, consider the
correlation length ξ. The renormalization group (RG) equation for ξ is

ξ (m+1) = ξ (m) /b

If the RG transformation has a fixed point (denoted by *), which we always


assume is true here,

ξ ∗ = ξ ∗ /b

The only solutions are

(
∗ 0, T rivial : T = 0, F ixed P oint : T = ∞
ξ =
∞, N ontrivial f ixed point : T = Tc

The critical fixed point is unstable, while the T = 0 and T = ∞ fixed points

Arrows show flow


under RG

Tc

Figure 17.1:

are stable.
163

Note that the coarse–graining eliminates information on small scales in a


controlled way. Consider figure 17.2. Let us consider a blocking transformation

Less Fluctuations
(lower T )

T > Tc blocking Renormalized System


System with
a few fluctuations

Figure 17.2:

(m+1)
(fig. 17.3). We know the states in the new renormalized system, Si = ±1,

Original System blocking with b = 2 Renormalized Configuration


RG level m RG level (m + 1)

Figure 17.3:

for the renormalized spins on sites

i = 1, 2, ..N (m+1)

Note,
N (m+1) N (m) /bd
In the cartoon, N (m+1) = 16, bd = 22 , and N (m) = 64.
This is all pretty clear: we can easily label all the states in the sum for Z.
But how do the spins interact after renormalization? It is clear that, if the
original system has short–ranged interactions, the renormalized system will as
well.
(m)
At this point, historically, Kadanoff simply said that if EState was the Ising
(m+1)
model, so was EState . The coupling constants K (m) and h(m) would depend
on the level of transformation, but that’s all. This is not correct as we will see
shortly, but the idea and approach is dead on target.
164 CHAPTER 17. RENORMALIZATION GROUP

In any case, assuming this is true


X (m) P (m) (m)
+h(m)
P (m)
Z (m) (N (m) , K (m) , h(m) ) = eK hiji Si Sj i Si

{States}

But
(m)
f (m) (h(m) ,t(m) )
Z (m) = e−N
where t(m) is the reduced temperature obtained from K (m) , and f (m) is the free
energy per unit volume. But we must have Z (m) = Z (m+1) , since we obtain
Z (m+1) by doing some of the sums in Z (m) . Hence

N (m) f (m) (h(m) , t(m) ) = N (m+1) f (m+1) (h(m+1) , t(m+1) )

or,
f (m) (h(m) , t(m) ) = b−d f (m+1) (h(m+1) , t(m+1) )
The only reason h and t vary under RG is due to B. We assume

t(m+1) = bA t(m)

and

h(m+1) = bB t(m)

So (and dropping the annoying superscripts from t and h) we have

f (m) (h, t) = b−d f (m+1) (hbB , tbA )

Now we use Kadanoff’s bold, and unfortunately incorrect, assertion that f (m)
and f (m+1) are both simply the Ising model. Hence

f (h, t) = b−d f (hbB , tbA )

This is one form of the scaling assumption, which is due to Widom. Widom
argued that the free energy density was a homogeneous function. This allows
for the possibility of nontrivial exponents (albeit built into the description), and
one can derive the scaling relations.
For example, the factor b is not fixed by our arguments. Let

tbA ≡ 1

Then
f (h, t) = td/A f (ht−B/A , 1)
or,

f (h, t) = td/A f (ht−B/A )
which is of the same form as the scaling relation previously obtained for F .
165

vs.

E (m) (m) (m) E (m+1) (m+1) (m+1)


= −K (m) = −K (m+1)
P P
kB T hiji Si Sj vs. kB T hiji Si Sj
+(langer range interactions)

Figure 17.4:

Nearest neighbor interaction shown by solid line


Next – nearest neighbor interaction (along diagonal)
shown by wavy line

Figure 17.5:

Kadanoff also presented results for the correlation function, but we’ll skip
that.
Let us consider the interactions between spins again–before and after appli-
cation of RG. For simplicity, let H = 0. The original system has interactions
to its nearest neighbors as drawn in fig. 17.4. The blocked spin treats 4 spins
as interacting together, with identical strength with blocked spins. The nearest
neighbor interaction is as a cartoon, in fig. 17.5. Note that the direct part of
the nearest–neighbor interaction to the right in the renormalized system has 4
spins interacting equivalently with 4 other spins. There are 16 cases, which are
replaced by one global effective interaction. Those 16 cases have: All K (m+1)
contributions treated equally in block spin near neighbor interaction.



2 interactions separated by 1 lattice constant

6 interactions separated by 2 lattice constants
8
 interactions separated by 3 lattice constants


2 interactions separated by 4 lattice constants

Let’s compare this to the next–nearest neighbor interaction, along the diagonals.
There are again 16 cases which have: All next nearest neighbor contributions
166 CHAPTER 17. RENORMALIZATION GROUP

Figure 17.6: Nearest neighbor interaction vs. Next–nearest neighbor

treated equally in block spin representation.



1 interactions separated by 2 lattice constants



4 interactions separated by 3 lattice constants


6 interactions separated by 4 lattice constants


4 interactions separated by 5
 lattice constants


1 interactions separated by 6 lattice constants

(m+1)
Let us call the block–spin next–nearest neighbor coupling Knn , which lumps
all of this into one effective interaction. It is clear from looking at the interac-
(m+1)
tions underlying the block spins that K (m+1) (nearest neighbor) and Knn
(m+1)
(next nearest neighbor) are quite similar, although it looks like Knn <
(m+1) (m+1)
K . However, Kadanoff’s step of Knn ≡ 0 looks very drastic, in fact
with the benifit of hindsight, we know it is too drastic.
One can actually numerically track what happens to the Ising model under
the block–spin renormalization group. It turns out that longer–range interac-
tions between two spins are generatedP (as we have described in detail above)
as well as three spin interactions (like Si Sj Sk ), and other such interactions.
Although the interactions remain short ranged, they become quite complicated.
It is useful to consider the flow of the Energy of a state under application of
RG. For H = 0, for the Ising model, the phase diagram is as in fig. 17.7. For
low T
0 Kc K Recall K = J/kB T
infinite T

Figure 17.7: Usual Ising model H = 0

the Ising model with only next nearest neighbor interactions (J ≡ 0) one gets
the same phase diagram at H = 0 (fig. 17.8). The phase diagram for an Ising

low T
0 (Knn )c Knn Knn = Jnn /kB T
infinite T

Figure 17.8: Ising model, no near neighbor interaction


167

model with both K and Knn namely,


E X X
= −K Si Sj − Knn Si Sj
kB T
hiji hhijii

for H = 0, is drawn in fig. 17.9. The flow under RG is hence as drawn in

Knn Low T
ordered
(Knn )c 2nd order transitions

disordered
0 Kc K
infinite T

Figure 17.9:

fig. 17.10. Of course there are many other axes on this graph corresponding to
Knn

Critical
fixed point
of RG transformation
K

Figure 17.10:

different interactions. One should also note that the interactions developed by
the RG, depend on the RG. For example, b = 3 would give a different sequence
of interactions, and a different fixed result, than b = 2. But, it turns out,
the exponents would all be the same. This is an aspect of universality: All
models with the same symmetry and spatial dimension share the same critical
exponents.
The formal way to understand these ideas was figured out by Wilson. The
RG transform takes us from one set of coupling constants K (m) to another
K (m+1) via a length scale change of b entering a recursion relation: K (m+1) =
R(k (m) ).
At the fixed point
K ∗ = R(k ∗ )
Let us now imagine we are close enough to the fixed point that we can linearize
around it. In principle this is clear. In practice, because of all the coupling
constants generated on our way to the fixed point, it is a nightmare. Usually it
is done by numerical methods, or by the ǫ = 4 − d expansion described later.
168 CHAPTER 17. RENORMALIZATION GROUP

Using subscripts, which correspond to different coupling constants

Kα(m+1) − Kα∗ = R(K (m) ) − R(K ∗ )

or, X (m)
δKα(m+1) = Mαβ δKβ
β

where
Mαβ = (∂Rα /∂Kβ )∗
is the linearized RG transform through which “b” appears, and

δK = K − K ∗

All that is left to do is linear algebra to diagonalize Mαβ .


Before doing this, let us simply pretend that Mαβ is already diagonal. Then
we can see everything clearly. If
 (1) 
λ 0

 λ(2) 

 ... 
0 ...

the transform becomes


δKα(m+1) = λ(α) δKα(m)
The form of λ(α) is easy to get. It is clear that acting with the RG on scale b,
followed by acting again by b′ , is the same as acting once by b · b′ . Hence:

M (b)M (b′ ) = M (bb′ )

somewhat schematically, or

λ(b)λ(b′ ) = λ(bb′ )

The solution of this is


λ = by
The remormalization group transformation becomes

δKα(m+1) = byα δKα(m)

If yα > 0, δKα is relevant, getting bigger with RG.


If yα = 0, δKα is marginal, staying unchanged with RG.
If yα < 0, δKα is irrelevant, getting smaller with RG.
The singular part of the free energy near the fixed point satisfies

f ∗ (δK1 , δK2 , ..) = b−d f ∗ (by1 δK1 , by2 δK2 , ...)


169

This is the homogeneous scaling form, where we now know how to calculate
the yα ’s (in principle), and we know why (in principle) only two exponents are
necessary to describe the Ising model (presumably the remainder are irrelevant).
In practice, the problem is to find the fixed point. A neat trick is the ǫ = 4−d
expansion. From our discussion above about the Ginzburg criteria, mean–field
theory works for d > 4 (it is marginal with logarithmic correlations in d = 4
itself). In fact we then expect there to be a very easy–to–find, easy–to–fiddle–
with, fixed point in d ≥ 4. If we are lucky, this fixed point can be analytically
continued to d < 4 in a controllable ǫ = 4 − d expansion. We will do this below.
I say “lucky” because the fixed point associated with mean–field theory is not
always related to the fixed point we are after (usually it is then a fixed point
related to a new phase transition at even higher dimensions, i.e., it corresponds
to a lower critical dimension).
Finally let us now return to the linear algebra:
X
δKα(m+1) = Mαβ δKα(m)
β

since Mαβ is rarely diagonal. We will assume it is symmetric to make things


easier. (If not, we have to define right and left eigenvalues. This is done in
Goldenfeld’s book.) We define an orthonormal complete set of eigenvectors by
X (γ)
Mαβ eβ = λ(γ) e(γ)
α
β

where λ(γ) are eigenvalues and e are the eigenvectors. We expand δK via
X
δKα = δ K̂γ e(γ)
α
γ

or, X
δ K̂γ = δKα e(γ)
α
α

where δ K̂ are the “scaling fields” — a linear combination of the original inter-
P (γ)
actions. Now we act on the linearized RG transform with α eα (...). This
gives X X X
e(γ)
α (δKα
(m+1)
)= e(γ)
α ( Mαβ δKα(m) )
α α β

It is a little cofusing, but the superscripts on the e’s have nothing to do with
the superscripts on the δK’s. We get
X (γ) (m)
δ K̂γ(m+1) = λ(α) eβ δKβ
β

so,
δ K̂γ(m+1) = λ(α) δ K̂γ(m)
This is what we had (for δK’s rather than δ K̂’s) a few pages back. Everything
follows the same from there.
170 CHAPTER 17. RENORMALIZATION GROUP

17.1 ǫ–Expansion RG
Reconsider Z
K r u
F [ψ] = dd x{ (∇ψ)2 + ψ 2 + ψ 4 }
2 2 4
near Tc , where K and u are constants and

r ∝ T − Tc

Near Tc we can conveniently consider


Z
F K r u
= dd x{ (∇ψ)2 + ψ 2 + ψ 4 }
kB T 2 2 4

by describing the kB T , which is nonsingular, within K, r, and u. We can remove


r F
the constant K by absorbing it into the scale lenght or letting r → K ,F → K
etc., so that Z
F 1 r u
= dd x{ (∇ψ)2 + ψ 2 + ψ 4 }
kB T 2 2 4
Now we can easily construct dimensional arguments for all coefficients. Clearly

F
[ ]=1
kB T

where [...] means the dimension of the quantity. Hence,


Z
[ dd x(∇ψ)2 ] = 1

or

[Ld−2 ψ 2 ] = 1, since [x] = L

or,

1
[ψ] = d
L 2 −1

Similarly,
Z
r
[ dd x ψ 2 ] = 1
2

or

1
[Ld r ]=1
Ld−2
17.1. ǫ–EXPANSION RG 171

or
1
[r] =
L2
Finally,
Z
u
[ dd x ψ 4 ] = 1
4
1
so [Ld u 2d−4 ] = 1
L
1
or [u] = 4−d
L
Eventually we will be changing scale by a length rescaling factor, so we need to
find out how these quantities are affected by a simple change in scale. These
are called engineering dimensions.
Note the following,
1
[ψ 2 ] = d−2
L
But at Tc
1
hψ 2 i = d−2+η
x
where η is nonzero! Also, since r ∝ T − Tc , near Tc
1
[T − Tc ] =
L2
but [ξ] = L, so
ξ = (T − Tc )−1/2
near Tc . But
ξ = (T − Tc )−ν
So our results of scaling theory give
1 1 1
[ψ]af ter avg. = = d
L(d−2+η)/2 L 2 −1 Lη/2

and
1 1 1
[r]af ter avg. = = 2 1 −2
L1/ν L Lν
The extra length which appears on averaging is the lattice constant a ≈ 5Ȧ.
Hence
1
hψ 2 i = d−2+η aη
x
and η/2 is called ψ’s anomalous dimension, while 1/η − 2 is the anomalous
dimension of r. By doing the averaging, carefully taking account of changes in
scale by L, we can calculate ν and η as well as other quantities.
172 CHAPTER 17. RENORMALIZATION GROUP

In this spirit, let us measure quantities in units of their engineering dimen-


sions. Let
ψ ψ
ψ′ = = d
[ψ] 1/L 2 −1

and
x x
x′ = =
[x] L

so that (dropping the kB T under F ),


Z
1 r u
F = dd x′ { (∇′ ψ ′ )2 + ( L2 )ψ ′2 + ( L4−d )ψ ′4 }
2 2 4
we can of course let
r r
r′ = =
[r] 1/L2

and
u u
u′ = = = uL4−d
[u] 1/L4−d

so we have Z
1 r′ u′
F = dd x′ { (∇′ ψ ′ )2 + ψ ′2 + ψ ′4 }
2 2 4
Equivalently, since we are interested in the behaviour at Tc , let rL2 ≡ 1 (choos-
ing the scale by the distance from Tc ), then
Z
1 1 u′′ ′4
F = dd x′ { (∇′ ψ ′ )2 + ψ ′2 + ψ }
2 2 4
where
1 4−d
u′′ = u( ) 2
r
Note that u is negligable as L → ∞ for d > 4, and more revealingly, u′′ is

negligible as T → Tc for d > 4.


This is the Ginzburg criterion again, and note the equivalence between L →
∞ and T → Tc (i.e., r → 0).
It is worthwhile to consider how the coefficient of the quartic term (with its
engineering dimension removed) varies as scale L is varied. Consider the ”beta
function”:
du′
−L = −(4 − d)L4−d u
dL
= −(4 − d)u′
17.1. ǫ–EXPANSION RG 173


−L du
dL d>4 initial value of u′ is
desceased as L increases

d=4 u′

d<4 initial value of u′ is


increased as L increases

Figure 17.11:

The direction of the arrows is from


du′
= −(d − 4)u′ , d > 4
d ln L
du′
= (4 − d)u′ , d < 4
d ln L
Note that u′ = 0 is a stable fixed point for d > 4 and an unstable fixed point for
d < 4. This sort of picture implicitly assumes u′ is small. The renormalization
group will give us a more elaborate flow similar to
du′
−L = −(4 − d)u′ + O(u′2 )
dL
so that the flow near the fixed pt. determines the dimension of a quantity.


−L du
dL u′ = 0 fixed pt. d>4

new Wilson fixed pt.


u′ > 0
u′
d<4

Figure 17.12:

Clearly
du′
−L = (Dim. of u) u′
dL
and similarly
dr′
= (Dim. of r) r′
−L
dL
So the anomalous dimension can be determined by the “flow” near the fixed
point.
174 CHAPTER 17. RENORMALIZATION GROUP

The analysis above is OK for d > 4, and only good for u′ close to zero for
d < 4. To get some idea of behavior below 4 dimensions, we stay close to d = 4
and expand in ǫ = 4 − d. (See note below)
So consider Z
K r u
F = dd x{ (∇ψ)2 + ψ 2 + ψ 4 }
2 2 4
we will find
ǫ = 1 (d = 3) Exact (d = 3)
α = ǫ/6 0.17 0.125
β = 12 − 6ǫ 0.33 0.313
γ = 1 + ǫ/6 4 5.0
δ =3+ǫ 1.17 1.25
ν = 12 + ǫ/12 0.58 0.638
η = 0 + O(ǫ2 ) 0 0.041

(Note: Irrelevant Couplings


Consider Z
1 r u v w
dd x[ (∇ψ)2 + ψ 2 + ψ 4 + ψ 6 + ψ 8 ]
2 2 4 6 8
as before,
d
[ψ] = 1/L 2 −1
[ν] = 1/L2
[u] = 1/L4−d

But similarly,

[v] = 1/L2(3−d)
2
[w] = 1/L3(2 3 −d)

so,
Z
1 r u vL2(3−d) 6 2
F = dd x′ [ (∇′ ψ ′ )2 + ( L2 )ψ ′2 + ( L4−d )ψ ′4 + ψ + wL3(2 3 −d) ψ 8 ]
2 2 4 6

Hence, v, w look quite unimportant (at least for the Gaussian fixed point) near
d = 4. This is called power counting. For the remainder of these notes we will
simply assume v, w are irrelevant always.)
We will do a scale transformation in Fourier space,
Z
dd k i~k·~x
ψ(x) = d
e ψ̂k
|k|<Λ (2π)

where (before we start) Λ = π/a, since we shall need the lattice constant to
get anomalous dimensions. We will continually apply the RG to integrate out
17.1. ǫ–EXPANSION RG 175

ky ky
Λ ky
Λ/b
kx kx . . . . 0 kx
−Λ Λ
RG Λ/b RG
−Λ

domain of nonzero hatpsi k Scale transformation to k → 0


by factor of b

Figure 17.13:

small length scale behavior, which, in Fourier space, means integrating out large
wavenumber behaviour. Schematically shown in fig. 17.13. Let

ψ(x) = ψ̄(x) + δψ(x)


all degrees of freedom degrees of freedom degrees of freedom
after several for “slow” modes on for “fast” modes on
transformations |k| < Λ/b Λ/b < |k| < Λ
|k| < Λ to be integrated out

Also, for convenience,


Λ
= Λ − δΛ
b
So that the scale factor
Λ
b= ≥1
Λ − δΛ
Then we explicitly have
Z
dd k i~k·~x
ψ̄(x) = e ψ̂k
|k|<Λ−δΛ (2π)d

and Z
dd k i~k·~x
δψ(x) = e ψ̂k
Λ−δΛ<|k|<Λ (2π)d
Schematically, refer to fig. 17.14. This is called the “momentum–shell” RG. We
do this in momentum space rather than real space because it is convenient for
factors of Z Z
dd k 2
dd x(∇ψ)2 = k |ψ̂k |2
| {z } (2π)d
coupling different ψ(x)’s | {z }
fourier modes uncoupled
176 CHAPTER 17. RENORMALIZATION GROUP

ky ky ky
Λ Λ Λ
Λ − δΛ
Λ − δΛ
kx kx kx

ψ(x) ψ̄(x) δψ(x)

Figure 17.14:

which are trivial in momentum space.


The size of the momentum shell is

Z
dd k
= aΛd−1 δΛ
Λ−δΛ<|k|<Λ (2π)d

where a is a number that we will later set to unity for convenience, although
a(d = 3) = 1/2π 2 , a(d = 2) = 1/2π. Let us now integrate out the modes δψ

X
−F̄ [ψ̄]
|e {z } ≡ e|−F [{z
ψ̄+δψ]
}
scales |k| < Λ − δΛ States δψ scales |k| < Λ
| {z }
shell scales

Essentially we are — very slowly — solving the partition function by bit–by–bit


integrating out degrees of freedom for large k.
We will only do this once, and thus obtain recursion relations for how to
change scales.

Z
1 r u
F [ψ̄ + δψ] = dd x{ (∇(ψ̄ + δψ))2 + [ψ̄ + δψ]2 + [ψ̄ + δψ]4 }
2 2 4

We will expand this out, ignoring odd terms in ψ̄ (which must give zero contri-
bution since they change the symmetry of F ), constant terms (which only shift
the zero of the free energy), and terms of order higher than quartic like ψ 6 and
ψ 8 (which are irrelevant near d = 4 by power counting as we did above).
17.1. ǫ–EXPANSION RG 177

In that case
1 1 1
[∇(ψ̄ + δψ)]2 = (∇ψ̄)2 + (∇δψ)2 + odd term
2 2 2
r r r
[ψ̄ + δψ]2 = ψ̄ 2 + (δψ)2 + odd term
2 2 2
u u δψ 4
[ψ̄ + δψ]4 = ψ¯4 [1 + ]
4 4 ψ̄
u δψ (δψ)2 (δψ)3 (δψ)4
= ψ¯4 [1 + 4 +6 2 +4 3 + ]
4 ψ̄ ψ̄ ψ̄ ψ̄ 4
u 3u 2
= ψ̄ 4 + ψ̄ (δψ)2 + O(δψ)4 + odd term
4 2
We will neglect (δψ)4 , since these modes have a small contribution to the k = 0
behavior. This can be self–consistently checked. Hence
Z
1 1
F [ψ̄ + δψ] = F {ψ̄} + dd x[ (∇δψ)2 + (r + 3uψ̄ 2 )(δψ)2 ]
| {z } 2 2
|k|<Λ−δΛ

R dd k i~
but δψ = Λ−δΛ<|~
k|<Λ (2π)d
e k·~x ψ̂k , so

(∇δψ)2 = Λ2 (δψ)2

for a thin shell, and


Z
1
F [ψ̄ + δψ] = F [ψ̄] + dd x (Λ2 + r + 3uψ̄ 2 )(δψ)2
2
and we have,
X 2 +r+3uψ̄ 2
dd x Λ (δψ)2
R
e−F̄ [ψ̄] = e−F [ψ̄] e− 2

States δψ

where Z
X Y
= dδψ(x)
States δψ all δψ states

The integrals are trivially Gaussian in δψ, so


s
−F̄ [ψ̄] −F [ψ̄]
Y Y 2π
e =e
x total no. of δψ states
Λ2 + r + 3uψ̄ 2

we know the total number of δψ states because we know the total number of
modes in Fourier space in the shell are
Z
dd k
d
= aΛd−1 δΛ
Λ−δΛ<|k|<Λ (2π)
178 CHAPTER 17. RENORMALIZATION GROUP

so Y Y
dx)(aΛd−1 δΛ) ln (...)
R
(...) → e(
x total no.
and
dd x aΛd−1 δΛ 12 ln (Λ2 +r+3uψ̄ 2 )+const.
R
e−F̄ [ψ] = e−F [ψ̄] e−
Now consider
1 1 r 3u
ln [Λ2 + r + 3uψ̄ 2 ] = ln (1 + 2 + 2 ψ̄ 2 ) + const.
2 2 Λ Λ
ǫ2 ǫ3
we expand using ln (1 + ǫ) = ǫ − 2 + 3

1 r 3u 1 r 3u
= { + 2 ψ̄ 2 − ( 2 + 2 ψ̄ 2 )2 + ...}
2 Λ2 Λ 2 Λ Λ
const. const.
1 r7 3u 2 1 r * 3ur 9u2 4
= { 2 + 2 ψ̄ − ( 2 )2 − 2 ψ̄ 2 − ψ̄ + ...}
2 Λ Λ 2 Λ Λ 2Λ4
1 3u 3ur 1 −9u2
= ( 2 − 4 )ψ̄ 2 + ( 4 )ψ̄ 4
2 Λ Λ 4 Λ
So that
Z
1 3u 3ur 1 −9u2
F̄ [ψ̄] = F [ψ̄] + dd x{ ( 2 − 4 )(aΛd−1 δΛ)ψ̄ 2 + ( 4 )(aΛd−1 δΛ)ψ̄ 4 }
2 Λ Λ 4 Λ
so, Z
1 r̄ ū
F̄ [ψ̄] = dd x{ (∇ψ̄)2 + ψ̄ 2 + ψ̄ 4 }
2 2 4
where,
3u 3ur
r̄ = r + ( − 4 )(aΛd−1 δΛ)
Λ2 Λ
−9u2
ū = u + ( 4 )(aΛd−1 δΛ)
Λ

But
r̄ = r(Λ − δΛ) etc,
so
r(Λ − δΛ) − r(Λ) 3u 3ur
= ( 2 − 4 )aΛd−1
δΛ Λ Λ
and
u(Λ − δΛ) − u(Λ) −9u2
= ( 4 )aΛd−1
δΛ Λ
or,
1 ∂r 3u 3ur
− = ( 2 − 4 )Λd−1
a ∂Λ Λ Λ
1 ∂u −9u2 d−1
− = ( 4 )Λ
a ∂Λ Λ
17.1. ǫ–EXPANSION RG 179

Set a = 1 (or let u → u/a), and set Λ = 1/L, to obtain


∂r
−L = −3uL2−d + 3urL4−d
∂L
∂u
−L = 9u2 L4−d
∂L
We need to include the trivial change in length due to the overall rescaling
of space, with engineering dimensions [r] = 1/L2 , [u] = 1/L4−d , by letting
r′ = rL2 , u′ = uL4−d . Note
dr d(r′ /L2 ) dr′ r′ 1
= =( −2 ) 2
dL dL dL L L
so
dr dr′ 1
−L = (−L + 2r′ ) 2
dL dL L
and
du du′ 1
−L = (−L + (4 − d)u′ ) 4−d
dL dL L
But
L2−d 1
−3uL2−d = −3u′ = −3u′ 2
L4−d L
and

L4−d 1
−3urL4−d = −3u′ r′ 4−d 2
= −3u′ r′ 2
L L L
and

L4−d 1
9u2 L4−d = 9u′2 = 9u′2 4−d
(L4−d )2 L
Hence we obtain
dr′
−L = −2r′ − 3u′ + 3u′ r′
dL
and
du′
−L = −(4 − d)u′ + 9u′2
dL
Which are the recursion relations, to first order. Evidently there is still a trivial
Gaussian fixed point r∗ = u+ = 0, but there is a nontrivial “Wilson–Fisher”
fixed point which is stable for d < 4 ! From the u equation (using ǫ = 4 − d)
ǫ
u∗ =
9
and in the r equation (using u∗ ∼ r∗ ∼ ǫ with ǫ << 1), 2r∗ = −3u∗ , so
ǫ
r∗ = −
6
180 CHAPTER 17. RENORMALIZATION GROUP


−L du
dL d>4
d=4
Wilson–Fisher fixed pt.
Stable for d > 4
u∗
u′
Gaussian d<4
fixed pt.

Figure 17.15:

Linearlizing around this fixed point gives the dimensions of r and u as that fixed
point is approached.
Let δr = r′ − r∗ , δu = u′ − u∗ . In the u equation,

dδu 2δu
−L = −ǫu∗ − ǫδu + 9(u∗ )2 (1 + ∗ )
dL u
= [18u∗ − ǫ]δu
dδu
−L = ǫδu (Wilson–Fisher fixed pt.)
dL
So the dimension of u is ǫ.
In the r equation,

dδ r
−L = −2r∗ − 3u∗ − 2δr + 3u∗ δr
dL
but δr >> δu as L → ∞, so,

dδr
−L = (−2 + 3u∗ )δr
dL
or,

dδr ǫ
−L = −2(1 − )δr Wilson–Fisher fixed pt. (stable d < 4)
dL 6
So the dimension of r is −2(1 − ǫ/6).
Similarly, near the Gaussian fixed pt. where u∗ = r∗ = 0, we have (as we
did before)

dδu
−L = −ǫδu
dL
dδr
−L = −2δr
dL
Giving the dimensions of r and u there.
17.1. ǫ–EXPANSION RG 181

To get the physical exponents, recall

[r] = 1/L1/ν

Hence
1
ν= , Gaussian fixed pt.
2
1 ǫ
ν = (1 + ), Wilson–Fisher fixed pt.
2 6
To get another exponent we use
1
[ψ] = d
L 2 −1+η/2

or more precisely
1
[ψ] = d f (rL1/ν , uL−yu )
L 2 −1+η/2

where (
−ǫ, Gaussian
yu =
ǫ, Wilson–Fisher
is the dimension of u. Choosing the scale

rL1/ν ≡ 1

gives
ν
[ψ] = r 2 (d−2+η) f (1, urνyu )
for the Wilson–Fisher fixed point this is straight forward. Since νyu > 0, the
term urνyu is irrelavant as r → 0. Hence
ν : const.
[ψ] = r 2 (d−2+η) f (1, 0)
≡ rβ

Now it turns out η = 0 + O(ǫ2 ) (which can be calculated later via scaling
relations), so,
11 ǫ
β= (1 + )(2 − ǫ)
22 6
1 ǫ ǫ
= (1 + )(1 − )
2 6 2
or
1 ǫ
β= (1 − ) Wilson–Fisher fixed pt.
2 3
The Gaussian fixed point is more tricky (although academic since it is for d > 4).
We again obtain
ν
[ψ] = r 2 (d−2+η) f (1, urνyu )
182 CHAPTER 17. RENORMALIZATION GROUP

but now ν = 12 , yu = −ǫ = d − 4, η = 0, so
d−2 d−4
[ψ] = r 4 f (1, ur 2 )

This looks like β = d−2 4 ! Not the well–known β = 1/2. The problem is that
u is irrelevant, but dangerous, for T < Tc (where ψ is nonzero). Without this
quartic term, the free energy is ill defined. For the equation
p above, f (1, u → 0)
is singular. Its form is easy to see. By dimensions ψ ∼ r/u below Tc , so

1
f (1, u → 0) =
u1/2
and
d−2 1
[ψ]u→0 = r 4
d−4
(ur 2 )1/2
d−2 d−4
r 4 − 4
=
u1/2
r 1/2
=( )
u
So, β = 1/2 for the Gaussian fixed point. Hence d ≥ 4, the Gaussian fixed point
is stable and α = 0, β = 1/2, δ = 3, γ = 1, η = 0, ν = 12 (The remainder are
obtained carefully with the scaling relations).
For d < 4, the Wilson–Fisher fixed point is stable and since α + 2β + γ = 2,
γ = ν(2 − η), γ = β(δ − 1), νd = 2 − α we have:
ǫ=1 Ising d = 3
α = 6ǫ 0.17 0.125...
β = 12 (1 − 3ǫ ) 0.33 0.313...
γ = 1 + 6ǫ 1.17 1.25...
δ = 3(1 + 3ǫ ) 4 5.0...
ν = 21 (1 + 12
ǫ
) 0.58 0.638...
η = 0 + O(ǫ2 ) 0 0.041...

You might also like