Mol Therm Lecture Note 2016 PDF
Mol Therm Lecture Note 2016 PDF
Mechanics
Lecture 1
Reading: 3.1-3.5 Chandler, Chapters 1 and 2 McQuarrie
This course builds on the elementary concepts of statistical mechanics that
were introduced in Physical Chemistry and provides a more in-depth exploration
of topics related to simple and complex uid systems at equilibrium.
* Discuss course info sheet / oce hrs / web site
* Discuss syllabus
Lets now launch into our introductory lecture, reviewing the basic principles
of equilibrium statistical mechanics:
liquid state, rather than the gas state and the solid state (which is treated in
courses on solid state physics). Note the coarse-graining of our cartoons
as we move up in lengthscale.
Imagine that you made a large number M of such measurements over a long
time period and the system was at equilibrium. Then, you could compute the
time average of the number of particles N inside the permeable barrier by
M
1
Ni
M i=1
= lim
N
one measurement on a large number (ensemble) of identical equilibrium system at the same time, weighting the observation of a particular microstate by
its frequency of occurrence.
The ergodic hypothesis of statistical mechanics says that
= N
N
This is generally met in most systems that you will encounter in your scientic
or engineering careers. However, there are notable exceptionse.g., glasses.
Generally, systems are ergodic when there is largely unrestricted ow and access to the various microstates throughout their dynamical trajectories. This
ow is often closely associated with the trajectories becoming chaotic due to
nonlinearities in the dynamical equations. Thus, it turns out that there is a close
connection between statistical mechanics (a relatively old subject) and one of
the hottest areas of contemporary theoretical research: nonlinear dynamics.
Dierent ensembles are convenient for calculations in statistical mechanics,
depending on what constraints are imposed on a system. To mimick an isolated
system that cannot exchange energy or mass with its surroundings, we use the
C. Microcanonical Ensemble (N , V , E)
To begin we dene a quantity known as the microcanonical partition function:
(N, V, E) = number of states with N, V, and energy E xed
Next, we invoke a fundamental postulate of statistical mechanics that all
states of the same energy are equally likely at equilibrium. In other words, these
states are characterized by a uniform probability distribution, P . It follows
that
1
states of ensemble
P = (N,V,E)
0
/ states of ensemble
Using this distribution function, we can compute ensemble averages of some
property G:
G =
P G
We can now use standard thermodynamics relations to get all the other
equilibrium properties. Recall that the internal energy E in an isolated system
satises
dE = dq dw
1st Law
= T dS p dV
closed system
T =
E
S
=
S
p = T V
V,N
ln
E
E,N
or
1
1 S
= kB
kB T
E
V,N
V,N
gives temperature
ln
= 1 V
E,N
gives pressure
In this ensemble, dierent states can have dierent internal energies and the
probability of occurrence of a state of energy E is now not uniform, but
Boltzmann-distributed:
P
Q(N, V, T )
=e Q
= eE
= Q
N,V
= dE T dS SdT
= SdT pdV 1st Law
= A
T
V,N
Thus,
E = A + TS = A T
A
T
V,N
= kB T 2
ln Q
T
Notice that the last two expressions can be viewed as a 1st order, inhomogenous,
linear ordinary dierential equation for A as a function of T at xed V and N .
A particular solution is
A(N, V, T ) = kB T ln Q(N, V, T )
which represents the fundamental connection to thermodynamics in the canonical ensemble. In this important ensemble A is thermodynamic potential that is
logarithmically related to the canonical partition function Q. Notice that there
is also a homogeneous solution const. T to the dierential equation, but the
constant coecient must vanish or the third law of thermodynamics would be
violated (vanishing entropy of a perfect crystal at T = 0).
The pressure can be computed in the canonical ensemble according to
ln Q
A
= kB T
P =
V N,T
V
N,T
and similar expressions can be used to calculate all other thermodynamic quantities.
Finally, we return to the entropy expression
A
S=
T V,N
Lets show that this is equivalent to the Gibbs entropy formula
S = kB
P ln P
P ln P
P ln P
kB P ln P
There are many other ensembles that one can choose besides the microcanonical and canonical. For example, it is natural in many experiments to maintain
control over the variables P, T, N . The (P, T, N ) ensemble is referred to as the
isothermal-isobaric ensemble and you will have opportunity to work with it
in your homework exercises. We now turn to consider the most important other
one: the grand canonical ensemble.
Lecture 2
Reading: 3.63.7, 7.1 Chandler, Chapter 3 McQuarrie
Recap: ergodic hypothesis
ensembles, kB T
ensemble
micro. can.
(NVE)
canonical
(NVT)
grand can.
(V T )
part. fnc
(N, V, E) =
Q(N, V, T ) =
eE
QG (, V, T ) =
S = kB
thermo connection
S = kB T ln
e(E N )
P ln P
A = kB T ln Q
pV = kB T ln QG
Gibbs entropy formula
Thus, N and E both uctuate from state to state. We control the N uctuations by imposing a chemical potential . Recall from the thermodynamics
of open systems
dE = T dS pdV + dN
dA = SdT pdV + dN,
etc.
that is the conjugate thermodynamic variable to N . Indeed, (T, S), (p, V ),
and (, N ) are all conjugate pairs.
One might guess that the appropriate distribution is thus
P = e(E N ) /QG
where
QG (T, V, ) =
e(E N )
is the grand partition function, which is usually given the symbol (but I hate
to draw this on the board!). Indeed, one can prove that
this is the distribution
that maximizes the Gibbs entropy formula S = kB P ln P subject to the
following constraints:
1.
P = 1
2.
See Chandler 3.7
E P = E
3.
N P = N
You may also want to review the material in Chapter 1 of McQuarrie on
the method of Lagrange multipliers. With this choice of P , the Gibbs formula
gives
S= kB P [E + N ln QG ]
= + T1 E T N + kB ln QG
or
kB T ln QG = T S E + N
Now, recall the following denition of the Gibbs free energy:
G = A + pV = E T S + pV
However, G is an extensive function that can be expressed in terms of its partial
molar property .
G
G = N = N
N T,p
Thus,
kN T ln QG = T S E + G
or
pV = kB T ln QG (T, V, ),
7
Thus,
Q1
G
Similarly
N=
QG ,V = E N
N P =
1
ln QG T,V
F. Fluctuations
It is important to appreciate that the variables that are not held xed will
uctuate in the various ensembles. We can often relate these uctuations to
higher-order derivatives of the thermodynamic potentials. Such formulas are
particularly useful in computer simulations. As an example, consider energy
uctuations in the canonical ensemble:
E E E
(E)2 = (E E)2
= E 2 2EE + E2
=
E 2 E2
= E2 P ( E P )2
1 2
1
2
=Q
2 Q ( Q Q)
2
=
2 ln Q)N,V
=
E
const. V heat capacity
N,V
T
kB T 2
E
2
kB T Cv =
= (E)2
N,V
Thus, in the canonical ensemble, we can calculate the heat capacity via
Cv =
1
2
kB T 2 (E)
This is referred to as a uctuation formula for the heat capacity. From the
expression on the right hand side, we might navely think that (E)2 N 2 ,
as one increases the size of a system. However, Cv is extensive (recall that
Cv = 32 N kB for a structureless ideal gas), so
(E)2 N or (E)2 1/2 N 1/2
Alternatively,
(E)2 1/2
N 1/2
1
1/2 0
E
N
N
2. Classical Fluids
A. Coarse-graining and the classical limit
For concreteness, lets now focus on a uid phase of a simple monatomic substance, e.g. Argon.
eE
where the sum is over all quantum states that the electrons and nuclei of
the atoms can be in. This is a hard sum to do, since it requires solving a
many-body quantum mechanics problem to nd the states. Instead, we use
our physical intuition that the light electrons are moving much faster than the
heavy nuclei, so that for a given nuclear conguration, the quantum states of the
electrons are nearly at local equilibrium (Born-Oppenheimer approximation).
E
ER,i
Q= e
R
e
i
= e ER
R e ER
where R is a sum over nuclear
states (i.e. position and momentum treat
with classical mechanics) and i eER,i is a sum over electronic states (use
quantum mechanics).
R , the eective energy (actually free
The second line is a denition of E
energy, since it is dependent) of just the nuclear states with the electronic
states averaged out. This is an example of the notion of coarse-graining in
statistical mechanics
small objects
(electrons, nuclei)
fundamental
interactions
larger objects
(atoms)
eective
interactions
coarsegrain
We have now made a lot of progress in getting rid of the quantum mechanics
in our problem, assuming that we can evaluate the eective interactions between
R . Indeed, this is what ab initio quantum chemistry tries
atoms that enter E
to do. There are now a large number of user-friendly software packages for
computing the eective interactions between atoms and molecules that are very
useful in deducing classical descriptions of uids.
Q=
e ER C dr1 . . . drN dp1 . . . dpN e ER
R
R = E(r
N , pN ) H(rN , pN )
E
eH(r ,p )
drN dpN eH
which is normalized such that drN dpN f = 1. Ensemble averages in the
canonical ensemble are thus dened by
P (rN , pN ) drN dpN f (rN , pN )P (rN , pN )
f (rN , pN ) =
for any property P (rN , pN ) of interest. For example, the average energy is
E= H(rN ,pN ) = drN dpN f H
1 Q
= Q
as before!
V,N
The expression A(N, V, T ) = kB T ln Q(N, V, T ) remains the fundamental thermodynamic connection for this ensemble.
This is about as far as we can go without specifying the form of the Hamiltonian H. Evaluating any thermodynamic property or Q seems to involve doing
6N integrals, where N 1023 ! Indeed, equilibrium statistical mechanics is really
all about the evaluation of high dimensional integrals. Life gets a bit easier if we
think physically about the form of the eective Hamiltonian that is obtained
by removing the electronic degrees of freedom. We expect
K(pN ) =
N
p2i
, p2i = p2ix + p2iy + p2iz
2m
i=1
for the kinetic energy where m is the eective mass of the atom (m
mnucleus ). Then, noting that
H=
N
p2i
+ U (rN ),
2m
i=1
where
prob. of observing
system at conguration
space point rN
P (r ) =
eU (r )
drN eU (rN )
(pi ) =
epi /2m
,
2
dpi epi /2m
and
2
Maxwell-Boltzmann
momentum distribution
g(px )=
Thus, e.g.
p2x
=
dpp2x (p)
epx /2m
(2m/)1/2
This implies that the mean-squared velocity component of any atom in the uid
is given by vx2 = kB T /m and that
p2 = p2x + p2y + p2z = 3mkB T
We can now draw some conclusions about a classical uid at equilibrium:
At the same T , increasing m implies smaller average RMS velocities v 2 1/2
The kinetic energy is equally partitioned among the three translational
modes at equilibrium
2
K= N p
2m =
3
= 2 N kB T
N
2m
3 mkB T
x, y, z
0
4
v =
4
m(2m/)3/2
dp p3 ep
2m2 / 2
/2m
= [8kB T /(m)]1/2
(2mkB T )1/23N
or
Q=
1
N !3N
T
drN eU (r
Lecture 3
Recap: Classical Limit, canonical partition function
Q=
eE Q=
1
N !3N
T
Qc (N, V, T ), Qc =
T = h/ 2mkB T
P f (rN , pN )= P (rN )
N
3
i=1 =1
g(p1x )=
p2 /2m
1x
e
(2m/)1/2
drN eU (r
thermal wavelength
g(pi ), P (rN ) =
1 U (rN )
e
Qc
Maxwell-Boltzmann Distribution
where the width of the wave packet is denoted by x. This is the scale
of the uncertainty in position of the quantum particle at some instant in time.
The Fourier transform of such a wave function has a broad peak centered at
1
, which immediately gives the Heisenberg
k = 2/ with width k p/h x
uncertainty principle: xp h. This principle relates the characteristic scales
of position and momentum uncertainty of a quantum particle. In an equilibrium
1/2
system, we could thus estimate x by computing p p2 M B kB T m. It
follows that
N
N
1
1
eN H(r ,p )
3N
QG (, V, T ) N !h
eN
N =0
Q(N, V, T )
1
N !3N
T
Qc (N,V,T )
zN
Qc (N, V, T )
N!
N =0
drN
N =0
E=
ln QG
N =
ln QG
,V
,V
+
=
ln QG
ln QG
ln z
,V
,V
C. Intermolecular Potentials
The above results, while restricted to systems that obey classical mechanics, are
exact. However, any calculations based on these formulae require an explicit
form for the eective potential energy U (rN ) of interaction among atoms or
molecules.
It is often the case that the biggest limitation on theoretical calculations
for uids is obtaining an accurate representation of U (rN ). Ab initio quantum
chemical methods are advancing rapidly, but many systems (hydrogen-bonding
uids, molten metals and salts) remain challenging for the purpose of parameterizing U (rN ).
Most calculations on liquids and gases are based on the notion of pairadditive potentials, namely
U (rN )
=
1
2
N
N
i = j
i <
u(ri , rj )
u(ri , rj )
N (N 1)
pairs
2
pairs
In the case of nearly spherical atoms like argon, or molecules like methane, the
pair potential u depends only on the distance rij |ri rj | between a pair
of atoms. Thus, u(ri , rj ) = u(rij ).
This would seem restrictive, but non-spherical molecules can be treated in
an interaction-site model by superposing spherically symmetric potentials at
dierent atomic sites in a molecule. For example, in the case of the nitrogen
molecule we can express the potential energy as a sum of spherically symmetric
site-site interactions:
N
N
2
2
i <
u (|ri rj |),
Mathematically, however, we expect u(r) to be a continuous, smooth function. Neutral molecules at large separations have dipole-induced dipole attractions that vary as r6 . A robust form that ts experiments on argon, methane
and other simple quasi-spherical molecules is
12 6
u(r) = 4
r
r
Lennard-Jones
6-12 potential
Finally, we note that eective ion-ion pseudo potentials for liquid metals,
e.g. sodium or potassium, look like
The pair approximation for U (rN ) is also often seriously in question in such
systems!
10
3. Theory of Gases
A. The Ideal Gas
As we shall see, gases are much easier to deal with than liquids, because the low
density of gases makes collisions and pair interactions of molecules infrequent.
The strategy will be to derive an expansion in powers of particle density known
as the virial expansion. The leading term, accurate at innite dilution, is the
ideal gas and follows by setting U (rN ) = 0.
We shall begin by pursuing the thermodynamic properties of a monatomic
ideal gas in the canonical ensemble.
Qc (N, V, T ) = drN eU = V N
Q=
VN
N !3N
T
A = ln Q = ln N ! + N ln(3T /V )
We are interested in the thermodynamic limit of this expression, which corresponds to considering a system of macroscopic extent, i.e. with Avogadros
number of atoms. We take this limit by simultaneously taking N ,
V , while holding the average particle number density = N
V constant.
Stirlings asymptotic approximation for the logarithm of large factorials, ln N !
N ln N N + . . ., gives
A N ln N N ln V + N (ln 3T 1)
N ln(N/V ) + N (ln 3T 1)
The ideal gas equation of state follows from:
A
p= V
N,T
p= N
1
V
Thus, the ideal gas equation of state can be written in any of the alternate
forms: p = , pV = N kB T, pV = nRT where n is the number of moles of
atoms.
Next, we repeat this calculation in the grand canonical ensemble:
zN
(zV )N
Qc (N, V, T ) =
QG (, V, T )=
N!
N!
N =0
N =0
= ezV
The thermodynamic connection is pV = ln QG = zV or p = z = e 3
T .
We thus have the pressure expressed as p(z, T ), but a conventional equation of
state is of the form p(, T ). Thus, we need an expression relating the activity z
to the average density , i.e. we need the function z(). Recall that
N =
z
ln QG
=V
=Vz
,V
ln z
ln z ,V
So z = N
V is the functional relationship between z and in the ideal gas
limit. It follows immediately that the equation of state for an ideal gas in the
grand canonical ensemble is given as before by p = .
Notice that we needed Stirlings approximation in the canonical ensemble,
but not in the grand canonical ensemble. Thus, results from the two agree only
in the thermodynamic limit! This, however, is as expected, since uctuations in
the two ensembles are dierent for systems of nite size and these will contribute
small corrections to thermodynamic quantities that go to zero in the limit of an
innite system.
Lecture 4
Reading: McQuarrie Chp. 12, Andersen handout
Qc = drN exp
u(rij )
i<
p =
virial
expansion
Bi (T )i
i=1
zN
drN
N!
N =0
eu(rij )
1i<jN
eu(r) 1 + f (r)
where f (r) = eu(r) 1 is the so-called Mayer f -function. This function has
the property that f (r) 0 for r , so volume integrals over powers of f
will remain nite even for V . With this simple change:
ZN
[1 + f (rij )]
Qc =
drN
N!
N =0
1i<jN
1 + f12 +f13 + f23 + f12 f13 + f12 f23
+f13 f23 + f12 f13 f23
3
where we have introduced some obvious shorthands. In particular, d1 dr1
and fij f (rij ).
Our next step is to introduce a Mayer cluster diagram shorthand for the
various terms:
= S1 z V dr1
= zV
1 2
= S1 z 2 V dr1 V dr2
2 1 S=2
z2 2
2 V
1 2
Sz
V
1 3
Sz
dr1
dr1
dr2 f (r12 )
dr2
dr3 f (r12 )
,S = 2
Note that
1
S
1
2
1
3!
1 3
z
dr1 dr2 dr3 f (r12 )f (r23 )f (r13 )
S
All 3! permutations are equivalent, so S = 6.
Now, notice that the disconnected graphs
all grow like V 2 or V 3 for V , while the connected graphs grow only like V
for V . For example,
2
= z2 dr1
dr2 f (r12 )
indep
of
r
for
V
1
z2
2 V
dr2 f (r12 )
nite
The disconnected graphs look like they are going to be a problem in the thermodynamic limit, but when one takes the logarithm and re-expands, magically
all the disconnected diagrams disappear!
ln QG = +
bi z i
i=1
where bi can be interpreted as 1/V times the sum of all distinct, connected
diagrams with i 1-vertices and f-bonds. The bi are now O(1) for V .
= V1 dr = 1
b1 = V1 {}
b2 =
..
.
1
V
}=
1
V
1
2
1
2V
dr2 f (r12 )
drf (r)
We can turn this into a density expansion by guring out the relation between
z and .
<N >
V
1 ln QG
V ln z
= z p
z
,V
or
z ln QG
V
z
,V
,V
= z ibi z i1
i=1
ibi z i
i=1
p =
Bi (T )i
i=1
where
B1 (T )
B2 (T )
= b1 = 1
= b2 = V1 {
B3 (T )
..
.
= 2 V1 {
} = 12
} = 13
dr2
dr2 f12
Notice that the b3 diagram has dropped out of the B3 expression. This is
because it contains an articulation vertex, namely a vertex that when removed
disconnects the diagram. Magically, the only diagrams appearing in the Bi are
those that are both connected and free of articulation vertices!
This exercise has illustrated an interesting and elegant feature of cluster
expansions, namely that we can represent a complicated and unwieldy perturbation series by a series of diagrams and that manipulations of the series can
often be described succinctly in terms of topological statements about the remaining diagrams. Similar features are present in perturbative treatments of
eld theories in theoretical physics, where the relevant diagrams are referred to
as Feynmann diagrams.
It turns out to be a lot of work to evaluate the successively higher terms:
Nevertheless, for special cases, e.g. hard spheres, people have worked this
out to very high order. Lets now consider this special case of the hard sphere
potential.
Recall that the hard sphere model has u(r) = , r < d; u(r) = 0, r > d.
This implies
1, r < d
u(r)
1=
f (r) = e
0, r > d
Coe
B2
B3
B4
B5
B6
B7
# diags
1
1
3
10
56
468
# integrals (3-d)
1
2
3
4
5
6
So,
B2 (T )
dr2 f (r12 ) = 12 dr12 f (r12 )
d
2d3
= 12 4
dr r2 (1) =
3
0
= 12
which is 4 times the particle volume Vp = 43 (d/2)3 = d3 /6 and is T independent! Note that for more realistic potential functions, the virial coecients are
normally temperature dependent. The leading term in the virial expansion for
the hard sphere uid is thus
p = +
or
2d3 2
+ O(3 )
3
p
= 1 + 4 + O( 2 )
where
=
Vp N
= d3
V
6
To how high a density can we trust this truncated series? The close-packed
face-centered-cubic
(fcc) crystalline density of hard spheres is 0 = 2/d3 , or
0 = 2/6 0.7405, and the hard sphere uid crystallizes into a solid at
LS 0.670 0.50. Since the virial series is expanded about 0, it is a
representation of the uid phase equation of state, so it is of little interest for
> LS 0.5. Lets see how the series does in the physical range < 0.5:
It appears that the virial series converges surprisingly well over the physical
range of . The solid curve is a very accurate equation of state developed by
Carnahan and Starling. It was derived by noting that the virial series is very
close to
p
1+
(i2 + 3i) i
i=1
(B2 = 4, B3 = 10, B4 18, B5 28, B6 40, B7 54)
This can be summed to
p
1 + + 2 3
=
(1 )3
CarnahanStarling EOS
which is believed very accurate over entire uid range. Now, using
A
A
= ()()
p=
V
V
T,N
or
p
A
N
A ex
N
, 1 =
A id
N
where Aex = A Aid denotes the excess Helmholtz free energy. We can thus
integrate over the packing fraction, starting from the ideal gas (dilute) limit to
obtain:
A ex
=
N
0
d
1 p
(4 3)
(
)
1
=
(1 )2
C-S excess
Helmholtz F.E.
Lecture 5
Reading: Chandler 7.2, 7.3, 7.4, 7.5
Recap:
dilute to moderately dilute gases treated by virial expansions about =
0 (z = 0) state.
easily extended to molecular uids and dilute mixtures.
Also applicable to dilute solutions of solute in solvents, including polymers
and colloids, but not electrolyte solutions as we shall see.
P (rN ) =
eU (r )
1 U (rN )
=
e
Qc
drN eU (rN )
N!
(N n)!
n = 1:
(1) (r1 ) should clearly be independent of position r1 in a uid that is homogeneous (uniform) on average. This is true in the gas and liquid states, but
clearly not in a crystal phase. To be normalized properly, (1) must correspond
to the average density
N
(1) (r1 ) =
V
n = 2:
(2) (r1 , r2 ) should depend only on |r1 r2 | r12 in a homogeneous, isotropic
uid (liquid or gas state). Hence (2) (r1 , r2 ) = (2) (r12 ).
For r12 , the two points in the uid are very far away, so there should
be no correlation in density between the two points. Thus
lim (2) (r12 ) =
r12
N (N 1)
2 .
V
V
as a function that 0 for r , hence focusing attention on particle separations r for which correlations exist. A nice physical way to think about g(r)
is:
average density of particles at r,
g(r) = given that a tagged particle is
placed at the origin
For weak, elastic scattering, |kin | |kout | and the momentum transfer or
scattering wavevector is given by
2
2
k = 2kin
2kin
cos
k kout kin , k 2 = k
2
k = 2kin 1 cos = 2kin 2 sin (/2) = 2kin sin /2
or
k=
4
in
sin 2
|f (k)|2
S(k)
form factor
structure factor
The form factor is related to the Fourier transform of the particle shape, so
contains only single-particle information. The structure factor contains the
information on particle-particle correlations:
S(k) = N 1
N
N
eik(rj r )
j=1 =1
S(k)
= N 1 N eik0 + N 1 N (N 1)eikr12
= 1 + N 1
Finally, noting
S(k)
dr2 eikr12 N (N 1)
dr1
dr3 . . .
drN P (rN )
1
2 2
0
dkk 2 j0 (kr)[S(k) 1]
where j0 (x) = (sin x)/x is the familiar spherical Bessel function of order zero.
p2
i
+
=
u(rij )
2m
i
i <j
3
1
= N kB T + N (N 1)u(r12 )
2
2
3
1
dr1 dr2 (2) (r1 r2 ) u(r12 )
= N kB T +
2
2
2 g(r12 )
E
dr1
dr12
or
E =
3
1
N kB T + N
2
2
drg(r)u(r)
Thus
E /N = 32 kB T +
1
2
drg(r)u(r)
energy equation
and we have derived an exact formula connecting the average internal energy to
a pair-potential weighted integral over g(r).
Now, what about the equation of state of a uid? In this case we use
A
A = kB T ln Q , p =
V
T,N
A = kB T ln(1/N !3N
T ) kB T ln Qc
A (no V-dep.)
N
(A A ) = ln drN eU (r )
Next, we rescale all of the coordinate integrals to xi = V 1/3 ri , dxi =
V dri , which has the eect of removing V from the limits of integration.
N
i<j
(A A ) = ln V + ln dxN e
1
= +
T,N
dxN e
i<j
dxN e
()
i<j
du
1 2/3
xij
i<j d(V 1/3 xij ) 3 V
T,N
1 du(rij )
rij
3 V i <j drij
du(r12 )
11
N (N 1)
r12
3 V 2
dr12
1 1
du
=
r12
dr1 dr2 (2) (r12 )
6 V
dr12
V
dr12
1
du
= dr2 g(r)r
6
dr
=
A
V
T,N
A
= V
T,N
= 1 16
drg(r)r du
dr
Thus, knowledge of g(r) also determines the equation of state, and by thermodynamic integration, the various free energies.
There is one more important formula relating g(r) to , the isothermal compressibility:
1 V
V p N,T
If we take k 0+ in the formula for the structure factor in the grand canonical
ensemble, one can show that
S(0+) =
N 2 N 2
N
dr[g(r) 1]
Lecture 6
Recap:
liquid structure quantied by g(r)
7
compressibility equation
E g
p g
g
energy eqn
virial eqn
compressibility eqn
kB T r1 dr3 . . . drN e
=
dr3 . .. drN eU
= kB T
ln dr3 . . . drN eU
r1
Add to both sides kB T r 1 ln[N (N 1)/ drN eU ] = 0,
N (N 1)
U
r
r
r1 U 1 2
= kB T r1 ln
dr3 . . . drN e
Qc
fixed
p2 g(r12 )
r1
= kB T
ln g(r12 )
r 1 w(r12 )
8
A way to get some insight into why c(r) is shorter-ranged than h(r) is to
employ cluster expansions. Just as we expanded the partition functions and
free energies in our study of gases, one can develop cluster expansions in z or
9
of the distribution functions. The series for h(r) in powers of is (see HCA
handout):
h(r12 ) = sum of all connected diagrams
with two root vertices (1 and
2), and number of eld vertices and f -bonds, and no articulation vertices.
Notice that the root vertices are not integrated over in evaluting diagrams;
thus, e.g.,
Note also that here an articulation vertex is one that when removed disconnects the diagram with at least one of the pieces having only eld points:
Although the h series has no articulation vertices, the graphs do have connecting
vertices, i.e., vertices that when removed disconnect the diagrams, e.g.,
The series for the direct CF, c(r), turns out to be:
c(r) = subset of h(r) graphs with no connecting
vertices
Finally we point out that many of the c diagrams have a f -bond between
the roots, which makes them short-ranged in r12 . The remaining diagrams also
turn out to be short-ranged, although this is less obvious.
Next, we note that if we view c(r) as our object of approximation, any approximatio nis summed to all orders in . To show this, we solve the OZ equation
by Fourier transformation
h(k)
= dreikr h(r) etc.
h(k)
= c(k) +
c(k)h(k)
c
(k)
h(k)
= 1c(k)
c(k)
1
dkeikr 1
h(r) = (2)
3
c(k)
The RHS has an expansion in to all orders!
Various integral equation theories of liquids are obtained by approximating
c(r) in dierent ways (closure approx.). The most important one is the socalled Percus-Yevick (PY) equation.
The PY equation is obtained by arguing that c(r) can be obtained by subtracting the indirect correlations from the full h(r).
c(r)
= h(r) hindirect
(r)
ew(r) 1 e[w(r)u(r)] 1
The PY equation is clearly very accurate for the HS uid and some empirical
corrections exist to make it nearly perfect. As with any approximate theory, it
is important to note that if we substitute into the various exact equations
11
relating g(r) to thermo. props., we will get dierent results. For example,
integrating the compressibility equation gives
1 + + 2
pc
=
(1 )3
PY compressibility EOS
(1 )2
PY virial EOS
2 c 1 v
p + p
3
3
HNC
long-ranged
u(r)
MSA
analytical results
for HS + attraction,
e.g., square wall uid
a
, = d3
1 4
6
attractive pert.
HS ref.
We should be able to do even better if we use the best HS ref. the CS
EOS:
modied VDW EOS
1 + + 2 3
p
=
a
(Longuet-Higgins & Wido??)
(1 )3
12
Lecture 7
Notes on HW:
pgm gives gd (r) from analytical solution of PY equation. Parameter in
pgm is redly(?) d.
Vikas will hold a copy of my notes, plus supplemental reading for checkout.
There are two successful approaches in the ilterature for doing this more systematically: the Barker-Henderson method and the Weeks-Chandler-Andersen
(WCA) method. We adopt the latter, which is superior, and illustrate the ideas
in the context of the LJ 612 model:
WCA method for LJ 612:
1. Recall u(r) = 4e (/r)12 (/r)6 has a minimum at r = r0 = 2116 ,
u(r0 ) = . Separate repulsive and attractive parts: u(r) = u0 (r) + u1 (r)
u0 (r) =
u(r) + ,r < r0
0
,r > r0
u1 (r) =
u(r)
repulsive
,r < r0
,r > r0
attractive
13
f0 fd
blip function
1 2
+
dru1 (r)g0 (r)
N
N
2
like VDW a term!
AD
CS EOS
from
with d(T, )
N
a = a(T, )
T >
3 and <
0.6, when the attractive forces start to play a more
signicant role.
14
System
atomic or molecular
liquid
Technique
MC (statics)
MD (dynamics)
Comments
MC faster
for statics
polymer solution
or melt
MC (statics)
BD/MD (dynamics)
coarse-grained
models often used
colloids
MC (statics)
BD (dynamics)
solvent is
continuum.
When it comes to computing forces, one also uses the minimum image
convention:
1. translate a cell of the same size (L) and shape to be centered around the
molecule of interest.
2. Sum the forces with the N 1 molecules in the translated cell. These are
the closest periodic images of these molecules.
Frequently, one also truncates the potential at some cuto distance rc (e.g.,
2.5). It is crucial that
rc L/2
for consistency with the minimum image convention as shown in the attached
gure:
B. The MD Method
The basic idea here is to solve Newtons equations of motion:
mri = Fi
i = 1, 2, . . . , N
forward in time, starting from some set of initial positions {ri } and velocities
{ri }. Many nite dierence algorithms are available; the simplest and most
popular is the Verlet algorithm. Start with:
ri (t t) = ri (t) r i (t) +
t2
ri (t) . . .
2!
Fij =
Now, consider
x u(|u|),
u(rij ) =
u(|rij |)
ri
rij
x = (x1 , x2 , x3 )
x
x
u=
u(x) =
u (x)
x1
xi x
xi
x2 =
x2i , 2x
xi
u(x) = u (x)
xi
x
Thus,
Fij =
1
rij
du
drij
x
= 2xi
xi
u (x)]
[or u(|x|) = x
rij =
w(tij )
rij
2
rij
where w(r) = r du
dr is the so-called pair virial function. For the LJ 612:
Fij =
24
6
2 [2(/rij ) ]rij
rij
Lecture 8
MD summary
ri =
1
m Fi
1
m
Fij (rij )
j=i
du
Fij = r1ij dr
rij
ij
24
= r2 [2(/rij )12 [/rij )6 ]rij
for LJ 612
ij
Verlet
ri (t + t)= ri (t t) + 2ri (t) +
r i (t)=
1
2t [ri (t
t2
m
Fij + (t4 )
j=i
+ t) ri (t t)] + (t2 )
Now, dene
=
For argon,
m 2
1/2
L-J 612 time constant
.34nm,
m = 6.7 1023 g
/kB = 120K
1.
ri (t t )
ri (t + t )= 2ri (t )
8
) ]rij
+24(t )2 [2(rij
j=i
2.
1
[r (t + t ) ri (t t )]
2t i
For accuracy, we need t 1, typically t <
0.01.
Recall that Newtons equations conserve energy, so as described, this is a
microcanonical algorithm. We can check accuracy by the extent of energy conservation.
Now, how to get thermo properties?
vi (t ) =
E =K +U
and
K =
1 2
3
mvi = N kB T
2 i
2
1
3
K = sumi (vi )2 = N T
2
2
2
1
u (rij
)
U =
i<
E = E = K + U
12
6
E = 32 N T +
4[(rij
)
(rij
) ]
i<
j
1
w(rij )
3N
i< j
P
T =
w (r )
1 1
3 NT
i<
w (rij
)
= 24[2(r )12 (r )6 ]
for EOS
Thus, we have seen how a basic NVE MD algorithm and simulation works.
We leave the subject with a few comments:
1. The MD equations are sti at (close approach of particles) and numerically unstable. Particle trajectories are not computed accurately (e.g.,
dierent t trajectories exponentially dierent in time) and chaos ensues. However, this is what we want in order to sample constant E phase
space! All that matters is that we conserve E accurately along the overall
trajectory.
2. Potential truncation and neighbor lists can reduce the force calculation
work from 12 N (N 1) N 2 to (N ) for short-ranged potentials.
3. Long-ranged potentials (e.g., Coulomb 1/r) are dicult. Ewald sums
and other specialized techniques have been developed.
4. Other ensembles can also be simulated, e.g., methods for NVT and NPT
MD have been devised. You can read about these in the Allen & Tildesley
reference.
Lecture 9
MC Method
Sample conguration states i rN (i) from Boltzmann distribution P (i) =
U
1
U (i)
/ i eU (i) importance sampling aves f (rN ) = M
e
i=1 f (i).
How to do sampling? Generate Markov chain of states from
P (i, t) =
Wij P (j, t)
Wji P (i, t)
r
i=1
i=1
microscopic reversibility
dxf (x).
f (x)
f (x)
=
(x)
(x)
b
a
dx(x) =
1 f (xi )
+ (M 112 )
M i=1 (xi )
(b a)
f (xi )(M 112 )
M i=1
which gives a simple scheme for evaluating integrals. This Monte Carlo shcme
is poor for low-dimensional integrals. The errors are
error (MC) (M 112 )
while a deterministic quadrature method, e.g., trapezoidal rule, has errors
that are much smaller:
error (TR) (M 2 )
Consider now evaluating the canonical conguration partition function
N
Qc = drN eU (r ) .
This is a 3N-dimensional integral. If we used a deterministic method with say
M 10 points in each dimension, this would require 103N function evaluations,
which is enormous! Instead, we might try the MC method, using a uniform
distribution of points in rN space:
M
Qc
V N U (rN (i))
e
M i=1
7
drN eU (r
N)
=
f (rN )
NQc N
N
= dr P (r )f
(r N) N
N
)
= dr (rN ) P (r(r)fN(r
)
N
N
)
= P (r(r)fN(r
)
1
M
M
P (i)f (i)
i=1
(i)
as before. Now, suppose that our random congurations rN (i) are chosen not
from a uniform distribution, but from one with the canonical pdf:
(rN ) = P (rN ).
Then,
f (rN )
M
1
f (rN (i)).
M i=1
This choice gives a simple formula for calculating averages and is the original
choice by Metropolis and coworkers. The drawback of this importance sampling is that we need a scheme for sampling the M congurations rN (i) from
= P.
The method for sampling from the pdf P (rN ) is to invent a ctitious
stochastic dynamics that has as its steady state, P (rN ). To simplify the notation, let i stand for a particular 3N-conguration space point rN (i). Then,
eU (i)
P (i) = U (i)
ie
is the PDF we would like to sample. Consider the following stochastic dynamics,
known as a master equation:
P (i, t) =
Wij P (j, t)
Wji P (i, t)
t
j(=i)
j(=i)
gain
loss
8
and the RHS is consistent with this. However, we have considerable freedom
in selecting the Wij and, hence, dening a dynamical model. In particular, we
would like this dynamics to have the Boltzmann P (i) as its steady-state solution.
For consistency, this requires for t :
P (i, t) = 0 =
Wij P (j)
Wji P (i).
t
j=i
j(=i)
One way, but not the only way, for the RHS to be zero is to demand microscopic
reversibility:
Wij P (j) = Wji P (i).
This gives a constraint on the ratios of the corresponding o-diagonal elements
of W :
U (i)
Wij
(i)
= PP (j)
= eeU (j) = e[U (i)U (j)]
Wji
eij
,
ij U (i) U (j)
A suitable model is thus:
Wij =
1
eij
, ij < 0
, ij > 0
Lecture 11
Recap:
MF theories can be unied via Landau expansions.
10
11
Here the bead positions rN are used to approximate the string shape and
1
2
ds2 = du2 + dx2 = dx2 1 + du
dx
2
2
ds= dx 1 + [u (x)] dx 1 + 12 (u (x))
The work involved in stretching that segment is
dU = (ds dx)
12 [u (x)]2 dx
So,
2
L
dU
1
dx
2 0
dx
Here we have to use functional notation. U is said to be a functional of the
function u(x). For each shape u(x) is associated a value of U U clearly depends
on the values of u(x) for all D x L.
U [u]
B. Calculus of Functionals
Before studying the statistical mechanics of our string, lets discuss the notion
of a functional derivative. Suppose we perturb the string shape slightly: u(x)
u(x) + u(x), where u(x) is small, arbitrary perturbation.
U [u + u]=
1
2
[u + (u) ]2
dx
= U [u] +
dx u (x) u +D(u2 )
dx
L
dxu (x)u(x)
L
+u (x)u(x)
0
} is often denoted
U [u]
u(x)
since then
U =
dx
U
= u (x)
u(x)
The minimum P.E. is achieved when u(x) is replaced by u (x) which gives the
smallest U . At the minimum
U [u]
u(x)
=0
uk
u (0) = 0
u (L) = 0
u (x) = 0
eU [u]
D[u]eU [u]
un sin(nx/L)
n=1
where un =
2
L
D[u] =
n=1
dun
In this representation,
L
dx[u (x)]2
U [u]= 12
0
L
nx
mx
n m
1
cos
cos
= 2
dx
un um
L L
L
L
0
n m
2
nm L
m,n
un um
= 12
L2 2
n m
2 2
n
= 12
n u2n
,
n
2L
n=1
1
2
Thus, P [u]
e 2 n un and each Fourier component has its own independent
n
Gaussian distribution!
Now we can compute quantities of interest.
2
1
dun un e 2 n un
=0
un =
1 n u2
dun e
u2n = 1n = 2 2L
n2
un um = n,m 1n
, etc.
Clearly, u(x) =
sin(nx/L)un = 0
n
u(x)u(x )=
n
1
=
sin(nx/L) sin(nx /L)
n
n=1
1
= 12
cos[n(x x )/L]
n
n
1
12
cos[n(x + x )/L]
n
m
These are sums that can be evaluated by conversion to contour integrals in the
complex plane. The result is
L
1
1 1
1 1
1
2
2
(x + x ) +
u(x)u(x ) =
(x
x
)
(x
+
x
)
|x x | +
2L
2L
4 L2
4 L2
Consider the special use of x = x :
u2 (x) =
x
1
x 1
L
5
7. Critical Phenomena
The subject of critical phenomena, i.e., the study of the equilibrium and nonequilibrium properties of systems near a critical point (e.g., a gas-liquid critical
point), has a long history, but many notable recent advances (culminating in a
Nobel prize to Robert Wilson in 1982).
pT plane:
T plane:
Lets now locate the critical point in the VDW model. Clearly
p
2p
=0 ,
=0 ,
VDW equation
2
T
b
a
ab
ab
Above Tc , this has 1 real, 2 imaginary roots; these merge to 3 equal real roots
at Tc . Thus, at (Tc , pc ):
pc
kB Tc
1 2
pc
3
+
=0
+
b
a
ab
ab
or
= (p pc )3 = p3 32 c + 32c 3c
c =
1
3b
pc =
a
27b2
kB Tc =
8a
27b
Zc =
c pc
c
3
8
pr = p/pc
Tr = T /Tc
T Tc
Tc
c
c
= Tr 1
= r 1
ppc
pc
= pr 1
temp dierence
order parameter
(density dierence)
pressure dierence
2
1+
1
1 2 3
4t 1 + 2
= 4t
+ 32 3 111
2
+ . . . + 32 3 + . . .
for |t| 1, || 1.
Next, we do a thermodynamic integration to get the Helmholtz f.e.: write
A V AV
A
p= V
, AV = AV (, T )
N,T
= AV V
AV
V
= AV V
pc ( + 1)= AV + (1 + )
AV
AV
T
N,T
V
= AV + A
T
F (, t) =
an (t) n
n=0
where
a0 (t)
a1 (t)
a2 (t)
a3 (t)
a4 (t)
..
.
= F0 (t)
= 1 + 4t + F0 (t)
= 3t
= 0 + (t)
= 38 + (t)
The a0 and a1 terms constitute constant shifts in the f.e. and chemical potential,
and thus do not contribute to the phase behavior. To see their eort on , note
AV
A
V
N
= V A
= V N
N
V,T
V,T
V
T
pc F
= c
t
so
c
= a1 (t) + 2a2 (t) 2 + 3a3 (t) 3 + . . .
pc
Landau F.E.
,
a4 = 3/8
6t
2
+ 32 3 = 0
= 4t ,
= 2(t)1/2 .
Lecture 10
Critical phenomena
t
c
c
= T T
,
= pp
,
=
Tc
pc
3
3 3
4t(1 + 2 + . . .) + 2 + . . .
AV /pc = (3t) 2 + 38 4 + . . .
t < 0 (t)
c
c
Landau
sion
= 1/2
order param
expan-
= 1/2
is known as a critical scaling relationship and VDW theory gives us the value
= 1/2.
There are many scaling laws that can be derived, which correspond to
looking at dierent thermo quantities or approaching the critical point from
dierent directions. For example, the isothermal compressibility is calculated
as:
= V1 V
V1 V
p
p
N
T
N,T
1
= p
=
T
1
6t+ 92 2 +...
and the compressibility diverges at Tc ! This has implications for light scattering
experiments. Recall
kB T =
N 2 N 2
= S(0+)
N
h(k)
=
c(k)
.
1
c(k)
Recall that
S(k) = 1 + h(k)
=
1
1
c(k)
One nice feature of c(r) you will recall is that it is shorter-ranged than g(r) or
h(r). This remains true near Tc , so we might guess that c(r) is well-behaved in
the critical region. Thus, to study small-angle scattering (long-ranged correlations) in the critical uid, we expand
c(k) c(0) c2 k 2 + (k 4 ).
Now
S(k)
But
S(0+)
S(k)
1
.
[1
c(0)] + c2 k 2
1
= kB T
t
S(0)
2
1+(k)
correlation length
tv
(VDW), so
c2
1
, 2 1
c(0) t
1
S(0) = 1c(0) t1
, v = 12 (VDW)
1
dkeikr [S(k) 1]
(2)3
S(0)
1
dkeikr 1+(k)
2 (r)
(2)3
S(0) 1 r/
(r)
4 2 r e
Thus,
Thus, we see that not only does the scattering intensity diverge, but the
correlation range diverges like tv . In particular, at Tc (t = 0):
h(r) r(d2+)
= 0 (VDW)
Let i = +1 (up), 1 (down) describe the state of the ith spin. A simple
model for the energy of a spin conguration = (1 , 2 , . . . N ) is:
E() = H
i=1
Jij i j .
2
j
i=
The rst term describes the energy of interaction between the spins and an
external magn. eld H and favors the spin up arrangement. The second term
with jJij 0 describes the interactions between spins that tend to cause them to
align in the same direction. Note N V in this model. The partition function
is:
H
i +
Jij i j
2
i
j
i
=
e
Q(, N, H) =
where
1 =1 2 =1
...
N =1
E
=H+
Jkj j Hke ()
k
j(=k)
where H is the direct eld and Jkj is the eld created by neighbors. This is
really a conguration-dependent eective eld that acts on spin k . We make
a MF theory by simply averaging this uctuating eective eld:
Jkj j .
Hke () Hke () = H +
j(=k)
1D 2D 3D
The energy contributed by spin k is thus Ek = Hke k and the dierent
spins are decoupled. We can determine the order parameter (magnetization)
self-consistently by
k eEk
= k =
k =1
eEk
k =1
T > Tc , = 0 is the only root and we have the paramagnetic (or disordered)
c zJ
c
= (zJ)
:
phase. Switching to t = T T
Tc
zJ
Thus, we have a coexistence curve much like that predicted from the VDW
model!
The Helmholtz free energy can be expressed in terms of by (see Handout)
A = kB T N ln
2
1 2
1
+ N zJ 2
2
2 cos h(H+zJ)
The rst term is kB T ln Q, Q = e k Ek ; the second term correct for
overcounting the pair interactions.
Perhaps not surprisingly, if we develop a Landau expansion of F = A/N
in powers of and switch to t = (T Tc )/Tc :
F (, t) a2 (t) 2 + a4 (t) 4
then a2 (t) t, a4 (t) 1 as before! Thus, all the same classical or MF
exponents are recovered, e.g.,
(t) , = 1/2
This same result could also have been obtained directly from the relation =
tan h(zJ) by expanding the RHS to ( 3 ).
We summarize with a few observations:
1. MF theories have been developed for many dierent systems including
uids, magnets, superconductors, superuids, etc.
2. These theories can be put in a common language by identifying an order
parameter and developing a Landau expansion of the FE. in the critical
region.
3. MF exponents for similar and seemingly dissimilar systems are in agreement and are independent of the space dimension d (see attached table).
4. MF theory suggests a universality of critical phenomena i.e., that
the molecular details of a uid, the lattice structure of a magnet or superconductor, etc. are irrelevant as far as the scaling behavior near a critical
point is concerned.
9
C. Failure of MF Theory
While the MF predictions are attractive and simple, they prove to be wrong in
general. How do we know this?
1. Experiments Numerous experiments have been carried out on a variety
of uid and magnetic, and other systems. The precision is such that the
MF exponents can be ruled out in many cases. See the attached table.
2. Exact calculations A few nontrivial models with critical points have
been solved exactly in one and two dimensions. For exaple, the shortranged ID Ising model can be exactly solved, as discussed in 5.1 of Chandler. At zero eld,
Q(, N, 0) = [2 cos h(J)]N
which gives ferromagnetic order only at Tc = 0, J ! Thus, the MF
theory is qualitatively wrong for the ID Ising model in that it predicts
ordering at the nite temperature:
(J)c =
MF
(z = 2d)
=2
1
J
1
= =
kB Tc
z
2
The 2D Ising model for H = 0 was also solved exactly by Lars Onsager in
the 1940s by a heroic eort. He found
Exact
2D Ising
(J)c = 0.441 . . .
1
8
which clearly doesnt agree very well with the MF predictions:
m (t) , =
(J)c =
1
1
=
z
4
MF
(z = 2d = 4)
1
2
No one has solved the 3D Ising model exactly (a sure-red way to win a
Nobel prize!), but the best numerical estimates give
m (t) , =
3D Ising
(numerics)
(J)c 0.25
m (t) , 0.313
which is to be contrasted with (J)c =
We can make some observations:
1
6
10
Jkj j
Hke () = H +
j(=k)
will uctuate less if the j sum contains more neighbors. In particular, the error
of writing Hke () Hke () is (z 1/2 ) for z . Thus, we expect MF
theory to be exact in the limit z = 2d .
There is one other limit where MF theory is obviously exact. Consider a
model with long-ranged interactions, e.g., an Ising model where
Jij = Jf (|rij |)
and f (r) decays on some large lengthscale :
There is no substitute for an actual picture, so the attached page from Chandler (page 167) shows typical congurations of an Ising model (2D) that has been
sampled by the MC method. In summary:
11
12
If we approach along = 0:
|t|v , = 0
where v is the scaling exponent we saw in our discussion of the VDW uid.
However, we expect a dierent divergence if we approach along t = 0:
||y , t = 0
where y is some new scaling exponent. Since v = y in general, constant
contours in the t plane will not be circles.
Now the basic idea behing the scaling hypothesis is that if only matters,
then the singular part of the free energy density (either Helmholtz or Gibbs)
should scale as
Av Gv [(, t)]d kB T
From this simple statement, all kinds of wonderful things ow. For example,
suppose we want
the shape of the coexistence curve, |t| . Recall from
thermo that
G
p
GV
GV
d 1 .
However, if the system only cres about and not , t separately, then we can
write
d 1 d ( 1/y )1 1/yd
or
|t|v(1/yd)
(1) Therefore, = vd v/y
gives in terms
of v, y
Next, suppose
we want in |t| , expressed in terms of v, y. Recall,
. Then,
t
G
Gv
d 2
v
t 2
d 1/y 2
2/yd
(
)
|t|v(2/yd)
gives in terms
of v, y
Here is one more example. Recall the exponent dened by the RDF:
(2) Therefore, = 2v/y vd
h(r) r(d2+)
13
It follows that
h() (d2+)
1
( + vd)
2
vd = 2 +
The attached sheet shows that the scaling theory indeed works very well and is
believed correct.
Thus, while the scaling theory does not allow us to compute exponents from
scratch, it does permit one to derive non-trivial relations among the exponents
indicating that only two are truly independent.
Lets nish our discussion by showing how Landau theory fails
Av d (|t|v )d |t|vd
(1/ )vd vd/
Since, vd/ is not an integer, Av () is not an analytic function at = 0! The
Landau theory (all MF theories), incorrectly lead to
Av t 2 + 4 + . . . 4
Lecture 12
Recap: RG Theory A framework for analyzing critical systems
RG Transformation:
1. coarse graining
2. rescaling
14
R = RG operator
15
D[]eH[]
lim
N
Q3 0
N a3 = V
N
i=1
di
16
1
V
eikr k ,
The latter is more convenient and will be used here. Note that arbitrarily
long-wavelength (small k) density uctuations are allowed, which are important
near the CP!
The rst step in an RG calculation is to do the coarse-graining. In real space
we form block spins at small scales; in mmomentum space we remove (integrate
out) k with the largest |k|!
H[]
e
dk eH[] , b >
1
b <|k|<
1
2
1
d
(r0 + k )k k
dd k(r0 + k 2 )k k
= 2V
2(2) |k|<
k
for
V
so the integrals over k factor into independent Gaussian integrals for each k.
The algebraic details of carrying out the integrals over the momentum shell
are complex, so we simply give the results to (u20 ). Note also that we cite
results for an arbitrary space dimension d: To (u20 ), one nds an expression for
We shall also want to rescale k so that the coecient of ||2 does not change:
q = z 1 k
where we wil gure out what the scaling factor z needs to be in a moment. Now,
in k-space
1
dd k(r + k 2 )k k .
H2 [] =
2(2)d |k|< b
After the rescalings, we have a new renormalized Hamiltonian:
1
d
d2 2
dd q(z 2 b
r + z 2 b
H2 [ ] =
q )q q .
2(2)d |q|<
r
are the RG ow equations, which describe how the eective Hamiltonian transforms under a dierential coarse graining and change of scale.
Lecture 13
RG momentum space recap:
Field theory model
1
1
1
2
4
2
r0 + u0 + ||
H[] = dr
2
4
2
v
SM of order parameter uctuations from: Q =
D eH[]
dk
|k|<
$
%
= u u9 2 u
= 2r + 83 2 u(2 r)
We can now see how the eective Hamiltonian transforms by solving these
ODEs parametrically in s. The relevant IC is
r(0) = r0 , u(0) = u0
initial
model
parameters
(r , u )
Gaussian FP
2
2
(2) r = 16
, u =
2 u = 6
8 2
9
19
both ()!
Wilson-Fisher FP
Now, lets look in the phase plane to see the RG ows near the FPS:
d > 4 ( < 0):
The physical CP is the Gaussian xed point, because there is one eigendirection where the point is stable. To get critical exponents, we linearize about
the GFP:
r(s) = r(s) r = r(s)
u
(s) = u(s) u = u(s)
Let
r(s)
u
(s)
d
2
P = MP , M =
0
ds
P(s)
Solution:
3
2
8 2
P(s) = c1 v1 e,s + c2 v2 e2 s
= v
, 2 = , v1 =
32
1
1 162
, v2 =
0
1
The eigenvalues have mixed sign for < 0, so this is indeed a saddle point.
The v1 direction is unstable (r-direction); the v2 direction is stable and is the
critical manifold.
Next, we show how the eigenvalues can be related to critical exponents.
Consider the correlation length relation tr . In our model, = (r, u) or
= (e1 , e2 ), where
e1 = v1 P = r
32
e2 = v2 P = 16
2r + u
Note that the coarse-graining step of the RG does not eect , since we are
integrating out ner scales. Thus,
(e1 (s + s), e2 (s + s)) =
b1
es 1s
20
=
e1 ds
e2 ds
ds
Thus, (e1 (s), e2 (s)) = es (e, (0), e2 (0)) or
(e1 (0), e2 (0)) = es (e, (s), e2 (s)).
Now,
dei
ds
= i ei ei (s) = ei s ei (0)
(e1 (0), e2 (0)) = es (e1 s e1 (0), e2 s e2 (0))
(r0 , e2 (0)) = r0
2
1, r0 1 e2 (0) .
The roles are now reversed, so the physical CP is the WF FP! Linearizing
the ow equations about that point:
1
2
= 2 13 + (2 )
= + (2 )
Thus,
v=
1
1
1
(1 + + . . .)
=
1
2(1 /b)
2
6
21
The two eigenvalues now determine all other critical exponents to (), e.g.
= 1 + 6
+ . . .
= 12 6
+ . . .
Thus, we seem to have a method for computing nontrivial corrections to exponents! Lets see how good they are when we blindly set d = 3( = 1!):
22
n+2
= 12 + 4(n+8)
+ (2 )
3
+ ()2
= 12 2(n+8)
n+2
= 1 + 2(n+8)
+ (2 )
These formulae reduce to those given previously for n = 1. However, generally, we see that we must sub-divide our notion of universality into so-called
universality classes, e.g.:
n=1
n=2
23
Lecture 14
8. Polymer Statistical Mechanics
A. Coarse-Graining
Polymer science is a particularly rich area of application of statistical mechanics. Before we can talk about the statistical thermo of polymers in solution
or in the melt state, we need to discuss some basic models for summing the
conformational states of a single polymer chain.
Common polymers are:
where N , the degree of polymerization can range from 103 105 . How are
we to describe such molecules within a CM framework?
1. An atomistic approach. Here we construct pair potentials, usually within
an interaction site framework for all the atoms composing the macromolecule. Rotations around backbone C-C bonds are permitted, which
leads to large numbers of conformational states eN . While in principle,
we can apply our MD and MC methods to such models, in practice they
are very computationally demanding! For example, a polymer melt has a
relaxation time that scales as
0 N 3.4
Next, we write ri = ni b, where the ni are unit vectors, |ni | = 1. In the FJC
model, all bond orientations are equally likely. Thus, we compute conformational averages as:
N
1
f ({r})
unit dni f ({nb})
4
i=1
sphere
Lets try this out. Notice the end-to-end vector R of a chain is i ri = b i ni :
R
=
sumN
i bni
=
bni = 0
i
So, on average, the chain is pointing in no particular direction. The meansquared end-to-end vector is:
R2 R R = i bni j bnj
= b2 i j ni nj
2
=b
N n1 n1 +N (N 1) n1 n2
So,
R2 1/2 = bN v , v = 1/2
More sophisticated models than the FJC model have been built and studied.
The most realistic are the so-called rotational isomeric state (RIS) models, which
attempt to capture real bond angles and trans-gauche conformers for specic
polymers. Perhaps surprisingly, the result for R2 1/2 is the same for N
R2 1/2 bN v , v = 1/2
except that now b has a more complex connection to the local geometry of
the chain. In fact, all such models with only short-ranged constraints (along
the chain) on conformations give this universal result. Asymptotically for
N , one can also show that all such models also lead to a Gaussian PDF
for the end-to-end vector R.
3
Because the local details appear not to be crucial, people working in polymer
SM like to work with a simple model that is Gaussian, both locally and globally:
N
3
(Ri Ri1 )2 .
2b2 i=1
eU [R]
D[R]eU [R]
where D[R] means a functional integral over all polymer conformations
R(s). This is our second example of the SM of elds as opposed to SM of
particles (rst example was critical
phenomena / RG theory).
What is the meaning of D[R]? One approach is simply to undo our
discretization:
N
dRi .
D[R]
i=1
R(s) = X0 + 2 Xp cos(ps/N )
p=1
Rouse
modes
1
N
ds cos(ps/N )R(s)
p = 0, 1, . . .
3
2b2
ds
dR
ds
2
b 2 2
= 12
kp X2p , kp
p
N b2
p=1
and
D[R]
dXp ,
p=0
so
P [R] P [X] =
p
12
kp X2p
12
dXp e
kp X2p
R = R(N ) R(0) = 2 Xp [cos p 1]
p=1
odd
= 4
Xp
p
odd odd
R = 16
Xp Xq
p
q
3
Xp Xq
=1
2
1
kp
p,q
odd 1
= 16 3
kp
p
odd 1
= 82 N b2
= N b2 !
2
p
p
2 /8
Thus, we have set up the Gaussian model with a b denition that agrees with
the FJC. If this were all we could compute, there would be no advantage of
the GCM. However, we can calculate structure factors of polymers, S(k), and
many other quantities of interest. One other interesting quantity is the radiusof-gyration, i.e., the second moment of the mass distribution of the polymer:
Rg2
1
N
1
N
=
=
12
N
12
N
0
N
2
ds[R(s) X0 ]2
ds4
cos(ps/N ) cos(qs/N ) Xp Xq
p
q
p
p
3p,q k1p
kp1 = N b2
1 1
]
2 p=1 p2
2
6
Rg2 =
N b2
6
R2
6
Lecture 15
Recap: Coarse-grained models
-freely jointed
-Gaussian (discrete and continuous)
R = R2 1/2 = bN v , b = 1/2 ideal chain
long-ranged interferences doesnt account for excluded volume
short-range interferences does account for local stiness, bond constraints,
etc.
Rg2 = R2 /6 , Rg bN v , v = 1/2 ideal chain
Dist. vector of end-to-end
2
2
3R2
P (R) e3R /2N b , N ; F kp T ln P , F 2N
b2 intrinsic
elasticity
would show ideal chain scaling behavior. This is wrong! Experiments have
shown that
R Rg N v , v 0.59
regardless of the good solvent used or the chemical details of the polymer!
The reason for the departure from ideal behavior is that the CG chain models
have neglected the excluded volume eect, namely that far away monomers
along the chain cannot occupy the same region in space:
R2
2
Nb
R3
B2 c2
repulsive coil
f. energy volume
density
entropic
stretching
(connectivity)
2
N
3
R2
N b2 + b
R3
R3
b3 N 2
R3
Minimize WRT R:
R bN vF , vF 3/5 = 0.6
# chains
c
=
vol.
N
where c = # monomers/vol. Also, two coils in a good solvent will repel as hard
spheres
Thus,
AV
c
N
ln c+ Nc (ln(3T /N ) 1)
+vN 3v2 c2 + . . .
8
T,np
np N
V
c
V
= AV V V
A
c
AV
= AV + c c
3v2 2
= Nc +
vN
c + ...
A2 (T,N )
Thus, the osmotic pressure of dilute polymer solutions is the sum of an ideal
gas term and an excluded volume term that scales as
A2 N 3vF 2 N 1/5
These predictions have all been conrmed! Note that as the solvent becomes
poorer, w(r) becomes more attractive and at the theta temperature T = ,
A2 (, N ) = 0. At this special point R bN 1/2 , chains are ideal; for T < it
they begin to collapse and aggregate:
Overall, we can classify solvent quality by this scheme:
T > , A2 (T, N ) > 0
T = , A2 (, N ) = 0
T < , A2 (T, N ) < 0
good solvent
theta solvent
poor solvent
For T < 0, because A2 < 0, we need to retain the next term in the virial expansion, A2 < 0. This augmented series can be used to map out a coexistence curve
for T < 0.
At the touching point, the so-called overlap concentration c , the concentration of monomers in the solution is the same as in an isolated coil:
c
N
1
3 N 13vF b3 N 4/5
R3
b
For high molecular weight, this can be a very small concentration! Furthermore,
for N large, there can be a big range of concentrations where
c
c
b3 the semi-dilute regime
where b3 is when = cb3 1 - pure polymer.
We have talked about the osmotic pressure of dilute solutions. Lets now try
to gure out how scales with c in the semi-dilute regime. Imagine that we
can represent versus c over both regimes in the sollowing form:
=
c
f (c/c )
N
where f (x) is a dimensionless functio nof x c/c . To recover our dilute results:
f (x) 1 + a1 x + (x2 ) for x
1
In the semi-dilute regime, one might expect to depend on the monomer concentration c, but not on N since chains are strongly overlapping:
1
3vF 1
5
4
cp+1 c9/4 , c
c
b3 . As you see from the attached experimental
plot (this would predict a slope of 5/4 = 1.25), the data collapse and scaling is
excellent!
Another way to recover this result is informative. Lets try to gure out how
the mesh size or correlation length scales with c. Imagine that
Rg(c/c ) bN v g(c/c )
10
kB T
Indeed
kB T
(c3/4 )3 c9/4 !
3
Lets now think a bit more physically about the meaning of . Suppose we
draw spheres of diameter around the strands of the mesh:
Call these spheres blobs. Within a blob, the piece of chain is nearly unaffected by neighboring bits of chain, so it should obey excluded volume statistics.
Thus, if there are g monomers per blob:
g vF g 3/5
or g 5/3 (c3/4 )5/3 c5/4 is the concentration dependence of g.
Now, each chain can be visualized as a pearl-necklace of blobs:
These pearl-necklace chains pack to form a dense melt of blobs; i.e., the blob
volume fraction, b = 1, but the polymer volume fraction = cb3
1. In
polymer melts (no solvent), we know from experiment and theory that chains
are ideal:
R bN 1/2 Flory theorem for melts
The physical reason for this is that the excluded volume eect is screened out
by the interactions with many other chains.
Now, back to semi-dilute solutions. For the same reason, we expect that our
pearl-necklaces, which are densely packed, should be ideal on large scales:
1/2
R Nb
, Nb
N
# blobs/chain
g
11
12
Lecture 16
Recap: semi-dilute, good solvent
kB T
3
c9/4
The density prole turns out not to be sharp, but has a width of order the
bulk correlation length .
We will prove that this is indeed the proper interfacial width scale shortly,
but now simply focus on the problem of determining the equilibrium (z) and
the interfacial energetics.
A. Thermo of Interfaces
Lets now review surface thermodynamics. Consider the gas-liquid interface as
above in an open system:
dE = T dS pdV + dN +
d
1st Law
work to create
more interface
area
so
E = T S pV + N +
Notice that the origin of the z-axis is arbitrary. Lets insert a so-called dividing
surface at zd , such that we call everything for z < zd gas and everything with
z > zd liquid. Then imagine a hypothetical uid with a step function (z):
N =
dz(z)
L
N
NS
L
zd
dzd (z)
L
zd
NG
dzd (z)
NL
dz[(z) d (z)]
where (z) =
1, z > 0
0, z < 0
unit
step fnc.
Lecture 17
Recap:
= AS / where AS = surface excess Helmh. F.E. = A AL AG
=
Equil:
A
(r)
0 (r)
V
dr(r) = N
A
Now, (r)
constant that can be lumped in with the chemical potential in []. The
2 A
can be shown to be equal to kB T [1
two-point function (r)(r
)
c (r
c
c(0)
both of these are nite for short-ranged c(r). Notice that [1c c(0)] t vanishes
at Tc and, indeed, the rst two terms are the beginning of our Landau expansion
of AV for a homogeneous system:
1
A[] = dr
A0V ((r))
+ kB T c dr||2 gradient
2
local
a2 (t) 2 (r)
+a4 (t) 4 (r) + . . .
Now, for a 1-D interface along z:
S
1
[] A [] =
dz{A0V ((z)) + kB T c [ (z)]2
2
(z)AV,L (z)AV,G }
Our DFT principle says that the equilibrium tension and prole (z) is given
by:
[]
=0
(z)
and we choose zd = 0 such that N S = 0, or
dz[(z)] ()(z)
(z)]
= 0 odd prole (z) = (z)
Lets now carry out the minimization, using the Landau expanded form of
F ()
A0V () for the VDW uid. Recall A0V () = pc
3t 2 + 38 4 +...
Noting that AV,L = AV,G = A0V ():
+ 4 [ (z)]2 }
[] = pc
dz{F () F ()
3
1
3
8
=
= ( 2 (z) 2 )(z) (z)
pc (z)
pc
2
3
We are interested in the solution of the ODE subject to the BCs: (z) z
and , z
0. To be consistent, clearly = 0 and recall that this is
consistent with the bulk chemical potential of our near-critical uid:
or
4
3
3
( )2 4 + 2 2
3
8
4
1
2
2 ( )
=0
:
4
3
3
3
( )2 4 + 2 2 4
3
8
4
8
32
( )2 = ( 2 2 )2
9
3
= ( 2 2 ) , sign is physical root.
4 2
6
z
d
3
=
dz =
2 ( 2 )2
4 2 0
0
+
1
= 3 ln + =
ln
2 4 2
(z)
3
z
4 2
3z
z/
2 2
ez/ 1
+ = ( )ez/ , = z/
, or
e
+1
(z) = tanh(z/2) , =
2 s
(t)1/2
with a width that is of the same scale as the bulk correlation length for
a homogeneous system! We also note that since the shaded areas are equal,
N S = 0, so the Gibbs dividing surface indeed is at z = 0.
Next, we turn to the tension. Noting that 43 ( )2 = 38 ( 2 2 )2 for the
extremum prole,
8
8 pc 2
dz[ (z)]2 =
dx [tanh (x)]2
= pc
3
3 2
sec h4 (x)
4/3
or
=
2
16 pc
9
2
t , = 3/2
kB T
2
t2v t1.26
7
C. Capillary Waves
Fluctuations other than critical uctuations turn out to be important at all
nonzero temperatures for L-G and L-L interfaces. These are long-wavelength
undulations of the interface position, e.g., in a L-G system
The extra surface energy associated with additional surface area of the distorted interface is:
1 + (x )2 + (y )2 1
E = dx dy
2 dx dy()2
Also opposing the height uctuations is the gravitational potential energy:
1
Eg = dx dy m(L G )g[(x, y)]2
2
Lecture 18
Recap:
Capillary wave uctuations (x, y)
H[] =
1
2
dy{||2 + (mg) 2 }
gravity
surface
tension
dx
Q=
D[] H[]
where, e.g.
2nx
, nx = 0, 1, 2, . . .
L
So we take our region of interface to be a square patch with area L2 and impose
PBCs. This leads to
2
H[] =
{k + lambda2
c }k k
2L2
kx =
1/2
where c = mg
is the capillary length (typically 1 mm) and D
dk . In this representation, the interface broadening is given by:
k
w2 2 (r) =
1 i(k+q)r
e
k q
L4
q
k
However, due to the fact that H[] is diagonal in the Fourier representation:
k q
= k,q
Thus,
L2 kB T
[k 2 + 2
c ]
1
+ 2
c
k
2
kB T
1
1
d
k k2 +
2
L (2)2
c
k
kB T
1
2
W = (2)2 2
dk 2
, cuto
k + 2
c )
0
W 2=
kB T 1
L2
1 kB T
4
W =
k2
ln(1 +
1
2 2
2 ln(1+ c )
2 2
c )
As c 1 mm, 1/10
A 1/; c 1 W 2
1 kB T
2
ln(c ) , ln c 10
10
1-1 electrolyte
solution:
polyelectrolyte
solution:
We model these systems a bit dierently, depending on the lengthscales
involved. In many cases, it is convenient and accurate to replace the solvent
by a continuum, so that we do not have the added computational expense
of describing the solvent structuring around the solute. Lets start with the
simplest case of a solution of neutral particles.
Imagine that we have a solution of non-charged, neutral molecules or colloids,
dissolved in a low molecular weight solvent:
If such systems are dilute in solute, then we might think about using a virial
expansion to address the thermodynamic properties. The eective interactions
between solute molecules would correspond to the potential of mean force w(r)
between two solute molecules separated by some distance r in the pure solvent.
Indeed there is a theory known as the McMillan-Mayer theory that makes a
rigorous connection between dilute gases and dilute nonelectrolyte solutions:
Gases
p
=+
Bn (T )n EOS
kB T
n=2
u(r)
=+
Bn (T, )n EOS
kB T
n=2
= 1 + 2 +
Bij
(T, )i j + . . .
kB T
i=1 j=1
(T, )
Bij
1
=
2
dr[eWij (rj ) 1]
where Bij
is a 2 2 matrix of second virial coecients. The problem with this
is that these virial coecients diverge for the coulomb potential:
qi qj
2
Bij
int
e r 1 ] =
dr r [
r1 for r
r for r
Clearly virial expansions fail for electrolyte solutions! A new approach is needed.
We shall see that the problem is that the osmotic pressure is not analytic in the
i for charged solutions. Instead,
3/2
1 + 2 + (i )
kB T
3/2
The Debye-H
uckel theory was developed to evaluate this (i ) electrostatic
term.
B. Debye-H
uckel Theory
There are two ways to deal with the divergences caused by electrostatic interactions:
3
Lecture 19
Debye-H
uckel Theory for electrolyte solution:
Aex =
hard sphere d =
qi qj r
r e
(r )
0 r<
=
1 r>
convergence factor 0+
4
2
2
1
dr1 dr2 fij (r12 )i j
2 i=1 j=1
qi qj r
e
Wijd (r) +
(r )
r
1, r >
hard sphere coulomb with a
=
0, r <
d=
convergence factor,
0+
(ij )n /n!,
n=1
d
eWij 1 f d ( )
fij
ij
qi qj r
(r
r e
bond
)(
bond
It can be argued that this substitution in the diagrams leads to (see Andersen):
sum of top. distinct, irreducible
graphs that have 2 or more -eld
Aex = points, and at most one f d -bond
and any number of -bonds between
each pair of points
= Aex
d +
d
where Aex
d contains all the diagrams with only dashed f bonds. Notice that if
bonds connect any two points, there is now a factor of 1 in the value of a
diagram!
Lets now look at some individual diagrams:
Aex
+
d
V 3 2
(V 3 )
dominant term for
3 1
dilute
1
2
2
2
i
dr2 ij (r12 )i j
dr1
= 21 V ( qi i )2
r>
But,
2
dr
1 r
e
r
i=1
= 0. Now,
1
2
22! V
= 12 V
qi2 i
2
r>
2
qi i
dr
e2r
for 0
2 r2
drfd (r)
1 r
e
(r )
r
0
0
2
1
1
= 23!
3V
qi3 i
dr 3 3 e3r
r
r>
i
ln() , 0
0 for
11 electrolyte
3
2
1 3
dr12 dr23 13 r12 r113 r23
= 6 V
qi i
4 2
1 r
2
e
qi i ; (r)
4r
i
Then
1
2 2
Aex
dr12 (r12 22
(r12 )
ring = V ( )
1 2
+(1)3 32
dr3 (r13 )(r32 )
2
)
+(1)4 (
dr3 dr4 (r13 )(r34 )(r42 ) + . . .
42
or
2 2
Aex
ring = V ( )
=V
(k)=
dk(k)
C(k)
( ) 2
1
ikr 1 r
= 2 +k2
dre
4r e
n
= 12 (2)
(2 )
2
n
n=2
C(k)=
So
Aex
ring =
2 )
2 +ln(1
V
2(2)3
2 + . . .
dk{2 (k)
+ ln(1 2 (k))}
Let q = k/, 0+
aex
ring
3
= 2(2)
3
V
3
12
dq{q 2 + ln(1 q 2 }
3/2
A few comments:
1. The sume of the ring diagrams is nite for 0 (coulomb potential), even
though the individual terms are divergent. Mayer (see Andersen) showed
that the entire Aex series can be renormalized, eectively replacing
bonds by C bonds and eliminating all long-ranged divergences.
2. All remaining diagrams in the renormalization series, as well as Aex
d /V ,
are (2 ) which are smaller than the ring sum (3/2 ). Thus the ring
sum gives the rst correction to IG behavior for 0. Note that the
nite size of the ions, , does not enter until (2 ).
7
1 2
8
I , I
q i ionic strength
2 i i
+ (2 )
=
i
24
i
as previously announced! The chemical potentials are:
A
q 2
i
= ln(i 3T,i ) i
Ni V,T
2
Recalling the denition of the activity coecient i :
i = k T ln(3T,i i i )
we nd a purely electrostatic contribution:
q 2
ln i = 2ki T
As shown in the attached gures, this limiting law works nicely at low concentrations, but typically breaks down around 0.01 moles/.
Our next task is to explain the signicance of the Debye length D = 1 .
We shall do this by studying the structure of an electrolyte solution, i.e. gij (r).
First, however, note that the electrostatic contribution to both AV and can
be written:
kB T
e
3
Ae
V kB T 3
D
This is reminiscent of our blob picture of critical phenomena and semi-dilute polymer solutions!
Our starting point for examining the structure of a dilute electrolyte solution
is the OZ equation:
2
hij (r12 ) = cij (r12 ) + k=1 k dr3 cik (r13 )hkj (r32 )
In the dilute limit of i 0,
0
k dr3
Wkj (r32 )
r12
r13
k=1
qi qj
W (r)
W (r12 ) =
1
2
r12
4
dr3
1
W (r32 )
r13
4
2 1
k2 k2 W (k)
4
k2 +2
1
(k)
dkeikr W
(2)3
er
r
q q
i j r/D
Wij (r) = r
e
Thus, we see that while the bare coulomb potential is long-ranged:
1
r
the potential of mean force between two ions is screened by the surrounding
ions:
1
W (r) = er/D
r
On lengthscales greater than D = 1 I 1/2 , two ions no longer feel each
other! The physical picture is:
W 0 (r) =
Notice that D 30
A at = 0.01 mole/, which explains why 3
A is
not relevant.
Also note that the thermodynamics is evidently also controlled by these
3
clouds, which play the role of blobs, e Ae
V kB T /D !
Lecture 20
Recap: Debye-H
uckel Theory
For dilute electrolyte solutions:
3
2
Aex
)
V = 12 + (
8
1
2
qi2 i ionic strength
= I , I2
i
If these conditions are met, and the total charge on the colloid is +Q and the
counter ions are monovalent, then we can view the solution as a Q1 electrolyte:
Q2 r/D
e
+
W S (r)
r
short-ranged potential long-ranged
[hard-sphere (dia. )
electrostatic
and attractive VDW]
repulsion
We could then use W (r) as an eective pair potential and carry out MD or MC
simulatiosn to determine structure and thermo props of the solution. In such
a simulation, only the colloid positions are retainedthe water molecules and
counter ions enter only as a continuum through W (r)! Notice that the total
ionic strength enters D :
4
2
2
2
qi i
D = =
i
Q2 + e2 Qe
(Q2 + eQ)
Q 1 electrolyte
No added salt
where is the colloid number density. It is common to add an extra free salt,
e.g. NaCl to increase the ionic strength and thereby decrease D the range of
W (r). Then:
2
D
I (Q2 + eQ) +
qS2 S
S
11