Sunteți pe pagina 1din 51

B2 Theories of Fluid Microstructures

G. Nagele
Institut fur
Festkorperforschung
Forschungszentrum Julich
GmbH

Contents
1 Introduction 2
1.1 Model systems and pair potentials . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Pair Distribution Function 6


2.1 Basic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Potential of mean force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Scattering experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4 Thermodynamic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3 Ornstein-Zernike Integral Equation Methods 18


3.1 Ornstein-Zernike equation and direct correlations . . . . . . . . . . . . . . . . 18
3.2 Theory of critical opalescence . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3 Various closure relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.4 Percus-Yevick solution for hard spheres . . . . . . . . . . . . . . . . . . . . . 29
3.5 Thermodynamic consistency and Rogers-Young closure . . . . . . . . . . . . . 33
3.6 Extension to mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4 Effective Colloid Interactions 38


4.1 Depletion-induced colloid attraction . . . . . . . . . . . . . . . . . . . . . . . 41
4.2 Effective electrostatic macroion potential . . . . . . . . . . . . . . . . . . . . 43

5 Summary and Outlook 47

Extended version of the article appeared in: Soft Matter: From Synthetic to Biological Materials, 39th
IFF Spring School 2008, Series: Key Technologies Vol. 31, Forschungszentrum Julich Publishing (2008)
B2.2 G. Nagele

1 Introduction
Atomic liquids and colloidal fluids are distinguished from dilute gases by the importance of
short-range correlations and particle collision processes, and from crystalline solids by the lack
of long-range order. The distinction between a liquid and a gas is only of quantitative nature,
since there is no change in spatial symmetry in going from the gas to the liquid phase.

The most simple liquid systems are monoatomic liquids consisting of spherically shaped atoms
(argon or neon) or quasi-spherical molecules such as methane. The interactions between the
atoms can be characterized by simple pair potentials which depend only on the distance between
an atomic pair. Colloidal fluids consist of mesoscopically large colloidal particles, typically a
few hundred nanometers in size, dispersed in a low-molecular solvent such as water. There are
strong local correlations in the positions of these particles (super atoms) so that they form a
colloidal liquid in the liquid (solvent). The omnipresence of colloidal dispersions in chemistry
and biology, and the unsurpassed variety and tunability of their particle interactions explains
their importance both in industrial applications and fundamental research. Well-studied exam-
ples of simple colloidal fluids are suspensions of silica and plexiglass spheres in organic sol-
vents, and aqueous suspensions of globular proteins or highly charged polystyrene latex spheres.

While the friction-dominated, diffusive dynamics of colloidal particles is quite different from
the ballistic one of atomic liquids, atomic and colloidal liquids are quite similar in terms of
their equilibrium microstructure, i.e., in terms of the average ordering of particles. There are
just orders of magnitude differences in their characteristic length scales.

In the present lecture, I will discuss a versatile class of liquid state theory methods which al-
low to determine theoretically the microstructural and thermodynamic properties both of atomic
and colloidal liquids, from the knowledge of the particle interactions. These so-called integral
equation schemes are based on the Ornstein-Zernike (OZ) equation. Some of the most relevant
OZ schemes will be introduced, and we will explore their predictive power in comparison with
computer simulations and experiment. The quantity of central importance calculated with in-
tegral equation schemes is the radial distribution function, g(r), of an isotropic liquid, and the
associated static structure factor S(q). The latter property is essentially the Fourier transform of
g(r), and it can be measured using scattering techniques. The radial distribution function (rdf)
quantifies the probability of finding a particle at some distance from any other one. It can be
indirectly measured by radiation scattering experiments and, for very large colloidal particles,
also directly using video microscopy.

Typical atomic and colloidal liquids of spherically shaped particles will be briefly characterized
in section 1 in terms of their interaction potentials. Section 2 provides an overview on general
properties of g(r) and S(q), and on their relation to scattering experiments and thermodynamic
properties. Section 3 introduces the fundamental Ornstein-Zernike equation and the associated
concept of direct correlations. As a first application of the OZ equation, a mean-field type
description of critical opalescence in near-critical liquids will be given. Various approximate
closure relations are discussed which lead to closed integral equations for g(r). Particular atten-
tion is paid to the hard-sphere fluid, since it serves as a reference system of many liquids, similar
to the harmonic solid in solid-state physics. In the final part of section 3, the Ornstein-Zernike
integral equation concept is generalized to fluid mixtures. In section 4, I will explain the con-
Theories of Fluid Microstructures B2.3

cept of effective colloid interactions in systems which include, aside from the large colloidal
particles and in addition to the solvent, components of small particles such as free polymer
chains or salt ions. In two important examples, I will describe how a coarse-grained, simpli-
fied description is achieved, in the form of a one-component suspension of decorated colloidal
particles, by integrating out the small-particles degrees of freedom. The examples discussed
are a binary mixtures of colloidal hard spheres and small polymer chains, and a suspension of
charged colloidal particles with added salt ions.

1.1 Model systems and pair potentials


In the following we exemplify pair potentials suitable for describing the pair forces acting in
simple and colloidal fluids of spherical particles. Simple fluids of non-polar atoms or molecules
will be considered first, followed by a discussion of pair interactions in suspensions of spherical
colloidal particles.

The Lennard-Jones 12-6 potential [1, 2]


[( ) ]
12 ( )6
u(r) = 4 for r > 0 (1)
r r

provides a fair description of the interaction between pairs of rare-gas atoms such as in argon,
krypton and xenon, and also of quasi-spherical molecules such as CH4 . A sketch of the potential
curve is provided in Fig. 1. Two parameters characterize the potential: the collision diameter
where u(r) = 0, and the depth, , of the potential minimum at r = 21/6 . The values of and
have been determined for a large number of atoms using, e.g., atomic scattering techniques.
For argon, Ar = 0.34 nm and Ar /kB = 119.8 K. The short-range repulsive part of the pair
potential proportional to r12 represents approximately the electronic repulsion of two atoms.
The longer-ranged van der Waals attraction between two atoms at a distance r is described by
the r6 part.

u(r)
21 / 6

r [nm]
-

Fig. 1: Lennard-Jones pair potential describing simple atomic liquids.

The most simple pair potential one can think of is the potential between hard spheres of diameter
, i.e.
{
for r <
u(r) = . (2)
0 for r >
B2.4 G. Nagele

While there exists no atomic fluid of hard atoms, colloidal suspensions of hard spheres are
realized within good approximation, by coated polymethyl-methacrylate (PMMA) spheres dis-
persed in a refractive index-matched non-polar solvent such as cyclohexane. The coating con-
sists of a thin layer, as compared to , of adsorbed polymer chains (cf. Fig. 2). The polymer
brush gives rise to a short-range repulsion between the colloidal spheres which counterbalances
the remnants of the van der Waals attraction. The sizes of the colloidal hard spheres are in the
range of several hundred to a few thousand nanometers.

Fig. 2: Model of colloidal hard spheres: PMMA spheres of radius a = /2 in a non-polar


solvent with surface-grafted, short polymer hairs.

PMMA particles in a non-organic solvent are a paradigm for sterically stabilized dispersions.

A well-studied example of charge-stabilized colloidal dispersions are polystyrene latex spheres


dispersed in a polar solvent such as water [3,4]. The latex particles acquire a high surface charge
through the dissociation of ionizable surface groups. Each colloidal particle is surrounded by
a diffuse layer of oppositely charged counterions, which are monovalent in the simplest case.
Overlap of the electric layers of two colloidal macroions leads to an electrostatic repulsion
which counteracts the van der Waals attraction and prevents the particles from irreversible ag-
gregation (cf. Fig. 3). In subsection 4.2, I will show that the screened electrostatic repulsion
between two charged colloidal spheres dispersed in a solvent of static dielectric constant is
approximately described by the effective pair potential
( a )2 r
2 e e
uel (r) = LB Z , r> (3)
1 + a r
which is the repulsive electrostatic part of the celebrated Derjaguin-Landau-Verwey-Overbeek
(DLVO) potential [4]. Here, Z is the an effective or renormalized surface charge number of a
spherical colloidal particle of radius a = /2, = 1/(kB T ), and LB = e2 /(kB T ) is the so-
called Bjerrum length. This length is the characteristic distance, for a pair of elementary charges
e, where their Coulomb interaction energy is equal to the thermal energy kB T . For water at
room temperature, LB = 0.71 nm. In case of a large and strongly charged colloidal particle
(macroion), the effective charge number in Eq. (3) can be substantially smaller than the bare
macroion charge as defined in a more refined many-component Primitive Model description of
spherical macroions and microions (cf. subsection 4.2).
The Debye-Huckel screening length, 1 , in a closed system is given by

2 4LB Z2
= (4)
1
Theories of Fluid Microstructures B2.5

u(r)
uel (r)
k BT

0
u vdW(r)
r

Fig. 3: Left: Electrostatic and van der Waals potential contributions to the total effective pair
potential u(r)(red curve). Right: colloidal macroions with counterions and coions.

where the sum is taken over all types of microions, i.e., surface-released counterions and salt
ions, of number densities and charge numbers Z . The factor 1/(1 ) corrects for the
free volume accessible to the microions, owing to the presence of colloidal spheres, where
= (4/3)a3 is the colloid volume fraction and the colloid number density. Notice here
that the range and strength of the potential can be controlled by adding or removing small ions,
and by changing the solvent or temperature.

The total effective pair potential, u(r), of charge-stabilized colloidal particles is the sum, u(r) =
uel (r) + uvdW , of uel (r) and the attractive van der Waals pair potential, uvdW (r). The van
der Waals attraction between two identical colloidal spheres can be described approximately
by [4, 5]
[ ( )]
Aef f 2 a2 2 a2 4 a2
uvdW (r) = + 2 + ln 1 2 r> (5)
6 r2 4 a2 r r
with uvdW (r) (Aef f 6 /36) /r6 for large r, and uvdW (r) (Aef f /24)/(r ) near
the contact distance r = (cf. Fig. 3). Clearly, the divergence of uvdW (r) at contact distance,
arising from the assumption of ideally smooth spheres surfaces with precisely two surface atoms
overlapping, is unrealistic.
Electrodynamic retardation and non-pairwise additivity effects on the dispersion forces are in-
corporated to some extent in the effective Hamaker constant Aef f . For non-metallic colloidal
spheres, Aef f is typically of the order of a few kB T [4, 5]. Van der Waals forces between iden-
tical particles are always attractive. Non-identical particles can repel each other however, when
the dielectric susceptibility of the solvent is in between that of the two particles.
For dispersions of highly charged colloidal particles at lower salt content, uvdW (r) becomes
completely masked by the electrostatic part uel (r). In this case, one frequently refers to the col-
loidal particles, with the associated diffuse layer of associated microions, as Yukawa spheres,
since their microstructural properties are determined by the Yukawa-like, exponentially screened
Coulomb potential uel (r).

It is often assumed that the potential energy, U (rN ), of a N -particle liquid system can be ap-
B2.6 G. Nagele

proximated by a sum of pair interactions


N
N
U (r )
N
u(|ri rj |) = u(rij ) , (6)
i<j i<j

for any configuration rN = {r1 , , rN } of position vectors {ri } pointing to the particles cen-
tres. The quality of this pairwise-additivity assumption depends on the choice of u(r), and
on how certain many-body aspects (e.g., non-additive dispersion forces, influence of solvent
molecules, electrostatic screening et cetera) are approximately included in the pair potential.
Typically, u(r) is density and temperature dependent. The dependence of the (effective) pair
potential on the thermodynamic state is a remnant of microscopic degrees of freedom which
have been averaged out on the level of coarse-graining where the potential applies. Colloidal
systems where U (rN ) is exactly pairwise additive are scarce. Examples of these exceptional
cases are ideal hard spheres, and suspensions of spheres with a specific short-ranged depletion
attraction that describes the integral influence of free polymers added to the system (cf. subsec-
tion 4.1).

Under the premise of Eq. (6), the microstructural properties of the fluid and, to some extent,
the thermodynamic properties are solely expressible in terms of u(r) and its associated radial
distribution function g(r). The latter is the most simple and most relevant example for a reduced
distribution function and will be discussed in the following.

2 Pair Distribution Function


In this section we discuss salient properties of g(r), and of its associated Fourier transform
pair S(q), referred to as the static structure factor. From knowing g(r), one can calculate
macroscopic thermodynamic properties and analyze the local microstructure. Furthermore, the
knowledge of static pair correlations is an essential ingredient in the theory of diffusion and
rheology of simple and complex fluids. A major task of the liquid state theory is therefore
to calculate g(r) from the information on the particle interactions, and to determine from it
scattering functions and thermodynamic properties.

2.1 Basic properties


The concept of reduced distribution functions has proven to be extremely useful in liquid state
theory. Consider a system of N identical spherical particles in a volume V at temperature T ,
i.e., a canonical N V T ensemble. The function

eU (r )
N
N
PN (r ) = , (7)
ZN (V, T )
with configurational integral

U (rN )
drN eU (r
N)
ZN (V, T ) = dr1 drN e = , (8)

is the probability density that the centers of the N particles are at the positions r1 , , rN . It
provides far more information than necessary for the calculation of scattering properties and
Theories of Fluid Microstructures B2.7

thermodynamic functions. What is really needed are the reduced distribution functions for a
small subset of n N particles irrespective of the positions of the remaining N n particles.

To this end we introduce the n-particle distribution function [1, 6]



(n) n
N (r ) = N (N 1) (N n + 1) drn+1 drN PN (rN ) (9)

of finding any set of n particles at a specified configuration rn = {r1 , , rn }, regardless


of how these n identical particles have been labelled. Of major importance are the reduced
distribution functions of order n = 1, 2. For a homogeneous system
(n) (n)
N (r1 , , rn ) = N (r1 + t, , rn + t) (10)
(1)
for an arbitrary displacement vector t. Then N = = N/V is equal to the average par-
(2) (2)
ticle number density, , and N (r1 , r2 ) = N (r1 r2 ) depends only on the vector distance
r12 = r1 r2 (to see this choose t = r2 ).

The correlation length (T ) is a characteristic distance over which two particles are correlated.
For fluids, is typically of the range of u(r) or larger to some extent, but under certain con-
ditions (i.e., near a liquid-gas critical point) it can become extremely large. For a n-particle
cluster with large mutual distances rij = |ri rj | , and N 1,

(n)

n
(1)
N (r1 , , rn ) N (ri ) = n (11)
i=1

since these particles are then uncorrelated. To describe pair correlations in a fluid relative to a
classical ideal gas of uncorrelated particles at the same density and temperature, we define the
pair distribution function, gN (r1 , r2 ), in the N V T -ensemble as
(2)
N (r1 , r2 )
gN (r1 , r2 ) := (1) (1)
, (12)
N (r1 ) N (r2 )
so that gN (r1 , r2 ) 1 for r12 . If the system is isotropic as well as homogeneous (i.e.,
(2)
no spatially varying external force field, no crystalline state), N and gN are functions of the
separation r = r12 only. Then,
(2)
N (r) N (N 1)
gN (r) = = dr3 drN PN (rN ) (13)
2 2
is denoted as the radial distribution function (rdf). It plays a central role in one-component
fluids, since it is indirectly measurable by radiation scattering experiments. Moreover, thermo-
dynamic quantities can be written as integrals over gN (r) and u(r), provided that the particles
interact by pairwise additive forces (cf. subsection 2.4).

As an important observation, we note that gN (r) is the average density of particles a distance
r apart from a given one. In fact, integration of g(r) over the system volume leads to

N (N 1)
dr gN (r) = dr12 dr3 drN PN (rN ) = N 1 (14)

B2.8 G. Nagele


since dr12 = V 1 dr1 dr2 . Likewise, Eq. (14) can be rewritten as

1+ dr [gN (r) 1] = 0 . (15)
V

Notice for the canonical rdf that gN (r ) = 1 N 1 , since there are N 1 particles left
aside from the one at r = 0. For large N , we can identify gN (r) as the conditional probability
of finding a particle a distance r from a given one. To see this we use that in a homogeneous
system, dr/V is the single-particle probability of finding the given particle in the volume ele-
ment dr. Then, dr/V gN (r) is the (unconditional) joint probability for two particles being
separated by the distance r, with one of them located inside dr. The sum (i.e., the integral)
of the joint probability over all accessible volume elements must give the value one. Precisely
this is stated in Eq. (14). The unconditioned single-particle and two-particle joint probabilities
can attain values between zero and one only. The conditional probability gN (r), on the other
hand, is the ratio of two unconditioned probalities and can thus attain values also larger than one.

Eqs. (14) and (15) are valid for a finite system of fixed N , without fluctuations in the number
of particles. In order to be independent of the specific statistical ensemble used in calculating
static properties of a macroscopically large system, it is understood that the thermodynamic
limit (T-limit, for short) of macroscopically large systems, i.e. N, V with = N/V kept
fixed, is taken at the end of each calculation. Then
g(r) := lim gN (r) lim gN (r) (16)
N,V

denotes the ensemble-independent radial distribution function, g(r), of a macroscopic system.

Let us summarize general properties of the g(r) for a fluid, which follow from the definition of
gN (r) in Eq. (13):
g(r) 0 , g(r ) = 1 (17)
g(r) 0 , for u(r) 1 (18)
g(r) = eu(r) + O() (19)
g(r) continuous for u(r) piece-wise continuous (20)
The typical behavior of g(r) for an atomic liquid with a soft pair potential of Lennard-Jones
type, and for a hard-sphere fluid is sketched in Fig. 4. Regions of r with g(r) > 1 (g(r) < 1)
have a larger (lower) probability of finding a second particle from a given one at r = 0, than for
an ideal gas at the same T and . The main features are a small-r region where g(r) = 0 owing
to strong repulsive forces exerted by the particle at the origin, and several peaks representing
increasingly diffuse shells, for increasing r, of next neighbors, second next neighbors and so
on. This shell layering manifests the granularity (non-continuum nature) of the fluid. The
oscillations in g(r) decrease in amplitude with increasing r. Eventually, g(r) approaches its
asymptotic value one for r > (T, ). The oscillations in g(r) become more pronounced with
increasing . Whereas the Lennard-Jones g(r) is continuous at any r, the hard-sphere g(r)
jumps from zero to values 1 at r = , due to the singular nature of the hard-sphere u(r).
For hard spheres, g(r) = (r ) + O() according to Eq. (19). Hard-sphere systems are
athermal (i.e., T -independent) since the probability for a given particle configuration is either
zero or one, independent of , depending only on whether two or more spheres do overlap or
not.
Theories of Fluid Microstructures B2.9

g(r)


r/
1 2

Fig. 4: The g(r) of a Lennard-Jones liquid (left), and a fluid of hard spheres (right).

2.2 Potential of mean force


There is a remarkable relation between g(r) and the so-called potential of mean force, w(r). In
the T-limit, w(r) is defined in terms of g(r) by [2]
g(r) =: ew(r) (21)
or
[ ( 2 )]
U (rN ) V
w(r12 ) = kB T ln g(r12 ) = kB T lim ln dr3 drN e + ln . (22)
ZN
To reveal the physical meaning of w(r), we take the derivative of w with respect to the position
vector r1 of a particle 1:
( )
dr3 drN r1 eU
U

1 w(r12 ) = lim = 1 U (rN
) . (23)
dr3 drN eU 1,2 FS

The quantity on the right-hand-side can be interpreted as the force on particle 1, if we hold
particle 2 fixed, and average over the positions of all the other particles (as denoted by 1,2 ).
Thus, w(r) is the potential for this force, and it is therefore called the potential of mean force.
Likewise, w(r) can be interpreted as the reversible work required in a process where two par-
ticles in the system are moved together, at constant N , V and T , from infinite separation to a
relative distance r. To see this explicitly, recall from statistical mechanics that

F (r12 ) = kB T ln dr3 rN eU (r ) + F0 (N 2, V, T )
N
(24)

is the Helmholtz free energy of a (N 2)-particle system in the presence of two fixed spheres
at a distance r12 . The spheres 1 and 2 influence the system through their excluded volume and
longer-ranged interactions. Here, F0 is an irrelevant configuration-independent part of the free
energy. Consequently
w(r12 ) F (r12 )
= (25)
r12 r12
B2.10 G. Nagele

and
w(r12 ) = F (r12 ) F () , (26)
since w(r) has been defined such that w(r ) = 0. The first and second laws of thermody-
namics tell us for a reversible small change of state that
dF = SdT pdV + dN + Wrev . (27)
This implies that dF = Wrev = 12 w(r12 ) dr12 is the reversible (maximal) work done
by the N 2 particle system to achieve an infinitesimal displacement, dr12 , of the two inner
boundary particles at fixed temperature, system volume and particle number.

The function w(r; , T ) is in general a considerably more complicated object than the mere pair
potential u(r), since it involves the effects of particles 1 and 2 on the configurations of the other
particles. It is only in the limit 0 that
w(r) u(r) i.e. g(r) eu(r) , (28)
as will be shown further down. A sketch of w(r) for hard spheres at finite density is shown in
Fig. 5. As seen, two hard spheres effectively attract each other at distances r 1.5. This
many-body effect arises from an unbalance of forces on the two spheres when the gap between
the two particles is depleted from the other ones. Depletion interactions of this kind occur in
any system with strong excluded volume interactions. This is the reason why depletion effects
have attracted considerable interest in the past few years (see also section 4).

g(r)

0
r
r < 2
w(r)

Fig. 5: Potential of mean force of hard spheres (left) and depletion attraction (right).

On assuming pairwise additive interactions so that



1 U (rN ) = 1 u(r1i ) , (29)
i>1

the force law for w(r) in Eq. (23) can be rewritten as


[ (3) ]
g (r1 , r2 , r3 )
1 w(r12 ) = 1 u(r12 ) dr3 1 1 u(r13 ) . (30)
g(r12 )
Here,
(3)
(3) (r1 , r2 , r3 )
g (r1 , r2 , r3 ) := lim N (31)
3 FS02_8

FS02_8
Theories of Fluid Microstructures B2.11

is the triplet distribution function of an isotropic and homogeneous liquid. Furthermore, g (3)
depends only on the two pair separations r12 , r13 , and the angle between r13 and r12 . The mean
force on particle 1, in presence of particle 2 at a distance r12 , is thus the sum of the direct in-
teraction between 1 and 2, and the interaction of 1 with a third particle at r3 , weighted by the
factor g (3) (r1 , r2 , r3 )/g(r12 ). The latter gives the probability of finding a particle at r3 , given
that there are particles at r1 and r2 . The equation given above is the lowest order one in the
so-called Yvon-Born-Green (YBG) hierarchy of equations relating reduced equilibrium proba-
bility density functions of consecutive order. We see now that Eqs. (19) and (28) are limiting
cases of Eq. (30), for 0.

To first order in , the triplet distribution function factorizes according to g (3) (r1 , r2 , r3 ) =
g(r12 )g(r13 )g(r23 ), since the probability of finding three particles simultaneously within the
correlation distance is of order 2 . The substitution of this factorization approximation into
Eq. (30) results into a closed, non-linear integro-differential equation for g(r) which can be
solved numerically for a given u(r). This so-called Born-Green integral equation, however,
gives poor predictions for the g(r) of dense liquids [2]. More reliable schemes to calculate g(r)
based on the Ornstein-Zernike equation will be discussed in section 3.

2.3 Scattering experiments


In the following, I briefly explain how pair correlations can be measured indirectly through
radiation scattering. A scattering experiment will have to probe distances of the order of the

particle sizes and next neighbor distances, which are Angstroms in case of atomic liquids, and
fractions of microns in case of colloidal dispersions. Therefore, X-rays and neutrons are used
for atomic liquids, whereas colloids can be probed by light scattering as well as small-angle
neutron and synchrotron radiation scattering.

A schematic view of a scattering experiment is shown in Fig. 6. Monochromatic radiation of


wavelength impinges on a fluid sample, and is scattered at an angle into a distant detector
which measures the average intensity, I(q), of scattered neutrons or photons.

ki
laser
q
r i (t) kf

detector

Fig. 6: Schematic light scattering setup.

For single and quasi-elastic scattering from spherical particles [2, 7],

I(q) N P (q) S(q) (32)


B2.12 G. Nagele

where q = (4/) sin(/2) is the modulus of the scattering wave vector q = kf ki , and N
is the average number of particles in the illuminated volume part of the sample. The so-called
form factor P (q), normalized as P (0) = 1, contains information on the scattering material
distribution inside a particle, i.e., information on the particle size and form.
The most relevant quantity in Eq. (32) including information on inter-particle correlations is
called the static structure factor S(q). Its statistical mechanical definition for q > 0 reads
2
N N
1 iq[rl rp ] 1 iqrl
S(q) = lim e = lim e 0 (33)
N N
l,p=1 l=1

where denotes an equilibrium ensemble average. The static structure factor is the T-limit
of the autocorrelation function,

1
S(q) = lim SN (q) = lim q q , (34)
N N

of the q-th Fourier component,


N
q = eiqrl N q,0 , (35)
l=1

of microscopic density fluctuations of N particles. The prefactor 1/N renders S(q) into an
intensive quantity. We have denoted the structure factor of the finite-volume NVT ensemble as
SN (q), and periodic boundary conditions have been used.
By expanding the double sum into self, l = p, and distinct, l = p, parts, it can be shown that

iqr sin(qr)
S(q) = 1 + dr e h(r) = 1 + 4 dr r2 h(r) , (36)
0 qr
where, h(r), with
h(r) = g(r) 1 (37)
is called the total correlation function. Notice that h(r ) = 0, and S(q ) = 1. As a
result, the static structure factor determines the Fourier transform, h(q), of h(r). Since Fourier
transforms are one-to-one mappings, S(q) can be inverted to determine h(r) and thus g(r):
[ ]
1 S(q) 1 1
g(r) = 1 + F (r) = 1 + dq q sin(qr)[S(q) 1] . (38)
2 2 r 0
In principle, this would require to measure S(q) for all wave numbers q where S(q) exhibits
significant oscillations. In light scattering experiments on colloids, this is usually not feasible
since the largest q value accessible is limited by qmax = 4/, which corresponds to backward
scattering.

We have used here the following basic result on Fourier transforms: suppose h(r) is an isotropic
function, with h(r) 0 sufficiently fast for r . The three-dimensional Fourier transform,
h(q), of h(r) is then defined as

sin(qr)
h(q) := F{h(r)} = dr e h(r) = 4
iqr
dr r2 h(r) (39)
0 qr
Theories of Fluid Microstructures B2.13

where the third equality follows from performing two angular integrals in spherical coordinates.
Fourier inversion leads then to

1 1 iqr 1 sin(qr)
h(r) = F {h(q)} = 3
dq e h(q) = 2 dq q 2 h(q) . (40)
(2) 2 0 qr
For notational simplicity, Fourier transformed functions are distinguished from their real-space
counterparts by their argument q only.

q=k

/ 2

/ 2 d = 2 / k

Fig. 7: First-order Bragg diffraction from a sinusoidal particle density wave of wavenumber
k = 2/d. Here, is the scattering angle, and the wavelength of light in the solvent.

For an intuitive understanding of scattering from a fluid-like particle system we show in the fol-
lowing that a measurement at a particular q corresponds to diffusive first-order Bragg diffraction
from a spatially sinusoidal particle density fluctuation around = N /V of the form

mk (r, t) = exp{k 2 Dc t} 0k exp{i k r} , (41)

provided that k = q, with a real-valued amplitude factor 0k . At each instant of time, a super-
postion of such density fluctuations is thermally induced with wave vectors k of varying length
and direction. Viewed from a coarse-grained level of resolution, the superposition of density
waves
m (r, t) = mk (r, t) , (42)
k

is the general solution of the macroscopic diffusion equation



(r, t) = Dc 2 (r, t) , (43)
t
describing large-scale mean density fluctuations (as indicated by the overline) characterized by
a collective diffusion coefficient Dc . For interacting particles, Dc can be substantially differ-
ent from the diffusion coefficient of an isolated particle. By linearity of the diffusion equa-
tion, also the real and imaginary parts of (r, t) are solutions. A special density fluctuation
of wavenumber k decays with the time constant k = 1/(k 2 Dc ). A fluctuation of larger wave-
length d = 2/k requires a longer time for its relaxation since the particles diffuse over a longer
B2.14 G. Nagele

distance d.

Consider now the scattering of a monochromatic radiation beam of wave vector q from such a
melange of density fluctuations. Note that Eq. (34) can be readily recast as
2
1
S(q) = lim dr exp{i q r} m (r) , (44)
N V

with the microscopic particle density fluctuation



m (r) = (r rj ) . (45)
j

On identifying m (r) with its coarse-grained average m (r, 0), one notices from Eq. (44) that
S(q) is non-zero only for a density fluctuation of special orientation and wavelength such that
k = q. The planes of constant phase of this sinusoidal density fluctuation are oriented at an
angle /2 with respect to the incident radiation, with denoting the scattering angle. Diffraction
from these planes is governed by the Bragg condition (see Fig. 7)

2d sin(/2) = m (46)

of constructive interference of radiation scattered from two equal-phase neighboring planes,


where is the radiation wavelength within the scattering medium. In this equation, we have to
set m = 1 since, different from an ideal crystal with sharply defined lattice planes, higher order
Fourier components are absent for an exactly sinusoidally varying density fluctuation of infinite
spatial extent. Indeed, using k = 2/d and k = q in Eq. (46) with m = 1, we recover the
equation
4
q= sin(/2) , (47)

which relates q for quasielastic scattering from a fluid-ordered colloidal system to the scattering
angle and radiation wavelength. Finally, we emphasize that the length scale, d = 2/q, of
probed density fluctuations is inversely proportional to q.

2.4 Thermodynamic properties


There exist various routes through which thermodynamic properties of the liquid can be related
to integrals involving g(r). In the following, we review the three most common ones.

The energy equation



3 1
E = Tkin + U (r ) = N KB T + N
N
dr 4r2 g(r) u(r) (48)
2 2 0

expresses the internal energy, E, of a one-component N -particle system with pair-wise addi-
tive U (rN ) in terms of u(r) and g(r). The internal energyis the sum of a kinetic ideal gas
part, (3/2)N kB T , and an interaction or excess part, U (rN ) . The latter can be understood on
physical grounds as follows: For each particle out of N , there are 4r2 g(r) dr neighbors in
a spherical shell of radius r and thickness dr, and the interaction energy between the central
particles and these neighbors is u(r). Integration from 0 to gives the interaction energy part
Theories of Fluid Microstructures B2.15

of E, with the factor 1/2 included to avoid double counting of particle pairs.

The pressure equation,



2 2
p = pid + pex = kB T dr r3 g(r) u (r) , (49)
3 0

relates the thermodynamic pressure, p, of a liquid with pairwise additive U to an integral over
g(r) and the derivative, u (r), of the pair potential. Here, pid = kB T is the kinetic pressure of a
classical ideal gas. The excess pressure contribution, pex , due to particle interactions can be de-
rived along the same lines as the energy equation using, e.g., the classical virial equation [1, 8].
For a repulsive pair potential where u (r) < 0, pint is a positive pressure contribution originating
from the enhanced thermal bombardment of container walls by the mutually repelling particles.

For a non-pairwise additive potential energy which can be decomposed as


N
N
N
U (r ) = u(rij ) + u3 (rij , rik ) + , (50)
i<j i<j<k

in terms of two-body, three-body and higher-order interaction terms, the energy and pressure
equations generalize to [10]
( )
Eex 2
lim = dr u(r) g(r) + dr dr g (3) (r, r ) u3 (r, r ) + (51)
N 2 6
and

pex 2 u3 (r, r ) (3)
= dr r g(r) u (r) dr dr r g (r, r ) + (52)
pid 6 18 r

The compressibility equation ,



T
= lim S(q) = 1 + dr [g(r) 1] , (53)
id
T
q0

relates the isothermal compressibility, T , of a one-component system, defined by


( ) ( )
1 V 1
T := = (54)
V p T p T
1
to an integral involving only g(r). Here, id
T = (kB T ) is the compressibility of an ideal gas.
The compressibility equation is more general than the energy and pressure equations, since it re-
mains valid even when the interparticle forces are not pairwise additive. According to Eqs. (36)
and (53), T can be determined experimentally from measuring S(q) in the long-wavelength
limit q 0, i.e., for q 1.

The compressibility equation can be derived in the grand canonical ensemble representing an
open system, at constant V and T , which allows for fluctuations in the particle number. This is
perfectly appropriate since S(q) is related to the intensity of quasi-elastically scattered radiation.
B2.16 G. Nagele

The radiation beam samples only a fraction of the system volume, and in this sub-volume the
number of particles, while macroscopically large, fluctuates.
To derive the compressibility equation, we need thus the definition of the radial distribution
function in the grand canonical ensemble. For a given N , we know from Eq. (13) that

N gN (r) = N (N 1) dr3 rN PN (rN )
2


N 1
= N (N 1) dr3 rN PN (r ) dr1 dr2 (r r12 )
V
N
N (N 1) 1
= (r r12 )N = (r rij ) (55)
V V i=j
N

where N denotes the fixed N , canonical ensemble average, and N = N/V . The canonical
gN (r) has been formulated by the most right equality in terms of an ensemble average invoking
Dirac functions. To obtain the grand-canonical rdf, denoted by gV (r), we merely have to
replace the canonical average by the grand canonical one, denoted by gc :


( )N = drN PN (rN )( ) ( )gc := P (N )( )N . (56)
N

We have introduced here the grand canonical probability, P (N ), of finding a system with ex-
actly N particles. The grand-canonical pair distribution function is thus given by
N
1
2
gV (r) := (r rij ) = P (N )2N gN (r) (57)
V i=j N =2
gc

with = N gc /V . Note here that gV (r) = gN (r) + O(1/N gc ). For a fixed N , we know from
Eq. (14) that
2 N (N 1)
N dr gN (r) = . (58)
V
For an open system of volume V , this leads to
[ ]
N (N 1) gc N 2
gc N 2
gc N 2
gc
2 dr [gV (r) 1] = = 1 (59)
V V N gc

The variance of the number fluctuations in an open, macroscopic system is related to the isother-
mal compressibility by the thermodynamic relation [6, 8]
( )2
N N gc
gc
kB T T = . (60)
N gc

This completes our derivation of the compressibility equation, on noting that g(r) = lim gV (r)
in the thermodynamic limit of N gc and V with fixed. The contradiction
between the compressibility equation and the canonical ensemble result in Eq. (15) is only
apparent. On first sight the latter might suggest that limq0 S(q) = 0 is exactly valid. However,
Theories of Fluid Microstructures B2.17

Eq. (15) applies only to a closed, finite-volume system with zero particle number fluctuations.
Notice further, for N and V finite and fixed (at given density ), that physically allowed wave
numbers are restricted to q > V 1/3 : It makes no sense to consider particle density fluctuations
of wave lengths ( q 1 ) larger than the system size. Therefore, the thermodynamic limit of a
macroscopic system should be performed first, followed by the limit q 0. In summary,
lim S(q) = lim lim SN (q) = lim SN (q = 0) = 0 , (61)
q0 q0

where SN (q) is the NVT static structure factor defined in Eq. (34). In the integral of the
compressibility equation which relates to a thermodynamic quantity, one must use the radial
distribution function in the thermodynamic limit (see also [9]).

The particles of a liquid near the triple point of gas-liquid-solid coexistence are densely packed
so that T is very small. Furthermore, the compressibility is nearly zero for a classical crystal
near T = 0, since there are hardly any vibrations of the atoms around their equilibrium posi-
tions. In contrast, T diverges at a critical point which is the terminal point of the gas-liquid
coexistence line. The divergence is accompanied by a long distance tail in h(r) which causes
the phenomenon of critical opalescence observed in light scattering studies near critical points.
We will study critical opalescence in section 3.2. Summarizing, we note

T 0, fluid near triple point
0, ideal crystal (62)
id
T
, fluid at critical point .
The reader should be warned that Eqs. (51) and (52) for E and p, and also the compressibility
equation as stated in Eq. (53), apply to state-independent, i.e., (, T )-independent pair inter-
actions only. Special considerations and care are required for state-dependent pair potentials
u(r; , T ) and, more generally, for a state-dependent U (rN ) [1117]. Effective potentials of
this kind occur in size-asymmetric colloidal mixtures, and in liquid metals, in the process of av-
eraging out the particle degrees of freedom associated with all but a single component of large
particles, which may be colloidal particles in the case of colloid-polymer mixtures, or metal
ions in case of liquid metals. The mixture is hereby mapped onto a one-component system of
pseudo-particles, or decorated particles, whose interactions are governed by an effective, state-
dependent interaction energy U ef f (rN ). This mapping on an effective one-component system
will be explained in section 4, and two prominent examples will be discussed.

In many practical situations, a mixture of large colloidal particles and small colloidal or poly-
meric particles, dispersed in a solvent is in osmotic equilibrium with a reservoir of the small
particles and solvent by means of a semi-permeable membrane impermeable to the large parti-
cles. Since this fixes the chemical potentials, { }, of the small particles and the solvent, it is
most convenient to treat them grand-canonically in order to derive U ef f (rN ; { }), so that, in
case of a phase separation, U ef f remains in the same form in both phases [18]. It is remarkable,
that the osmotic compressibility of the non-contracted-out large particles component is deter-
mined by their pair correlations alone, independent of the large-small and small-small particle
correlations, and independent of the so-called volume term (see section 4), through the osmotic
compressibility equation of Kirkwood and Buff [19]
( )
1 1
= . (63)
kB T T,N,{ } S(q 0)
B2.18 G. Nagele

Here, is the osmotic pressure of large particles relative to the reservoir of small particles and
solvent, and S(q) is the structure factor describing the large-large particle correlations. More-
over, the { } denote the externally controlled chemical potentials of the small particles and sol-
vent in the reservoir. To obtain the thermodynamics and phase behavior of the whole mixture,
however, information on the pair correlations of all particulate components is required, includ-
ing the large-small and small-small ones. Eq. (63) can be obtained from the multi-component
extension of the compressibility equation in Eq. (53) in combination with the thermodynamic
Gibbs-Duhem relation for the osmotic pressure of the large particles.

3 Ornstein-Zernike Integral Equation Methods


We proceed to discuss theoretical methods which allow to calculate the g(r) and S(q) of dense
liquids from a given pair potential. All these methods are based on the so-called Ornstein-
Zernike (OZ) equation, initially introduced by Ornstein and Zernike (1914) in their investiga-
tions of critical opalescence in near-critical liquids. The OZ equation introduces the direct cor-
relation function, c(r), as a very useful concept. Closed integral equations determining g(r) can
be derived from the OZ equation, when c(r) is additionally related in some physically appealing
approximation to g(r) and u(r). These additional relations are known as closure relations. We
will introduce various closure relations, and discuss their merits and shortcomings.

3.1 Ornstein-Zernike equation and direct correlations


The Ornstein-Zernike equation of a homogeneous and isotropic system is given by [1]

h(r12 ) = c(r12 ) + dr3 c(r13 ) h(r23 ) . (64)

It introduces the direct correlation function, c(r), as a new function and can be viewed as the
definition of c(r) in terms of the total correlation function, h(r) = g(r) 1, of two particles a
distance r = r12 apart. Eq. (64) can be recursively solved for h(r12 ) to give

h(r12 ) = c(r12 ) + dr3 c(r13 )c(r23 ) + 2
dr3 dr4 c(r13 ) c(r24 ) c(r34 ) + O(c4 ) . (65)

This leads to the following physical interpretation of the OZ equation: the total correlations be-
tween particles 1 and 2, described by h(r12 ), are due in part to the direct correlations, c(r12 ), of
these particles but also to an indirect correlation propagated by direct correlations via increas-
ingly large numbers of intermediate particles.

Once information is available about c(r) in form of a closure relation involving u(r), the OZ
equation can be viewed also as a closed integral equation for h(r). Some information on c(r)
derives from Eq. (64) in the low-density limit 0 where c(r) h(r). With Eq. (19) follows
then
c(r) f (r) for 0 , (66)
where
f (r) := eu(r) 1 (67)
Theories of Fluid Microstructures B2.19

is called a Mayer-f function. It follows then that

c(r) = u(r) (68)

for r . Without proof we note that the long-distance asymptotic result (68) holds true for
a wide class of pair potentials even at finite densities. The range of c(r) is thus comparable
with that of u(r), and the fact that h(r) is generally longer ranged than u(r) can be ascribed
to indirect correlation effects. One word of warning: we refer here and in what follows to
one-component liquids of electrically neutral particles. Ionic fluids must be distinguished from
neutral fluids in that the effect of screening (cf. subsection 4.2) in such systems is to cause
h(r) to decay exponentially at large r, whereas c(r) still has the range of the infinite Coulomb
potential and therefore decays as r1 . In principle, ionic liquids consist of at least two charged
components of opposite sign to enforce overall charge neutrality.

To relate c(r) to S(q), we slightly rewrite the OZ equation as



h(r) = c(r) + dr c(|r r |) h(r ) (69)

using r = r12 , r = r23 and |r r | = r13 . Fourier transformation of both sides of the OZ
equation leads to
h(q) = c(q) + c(q) h(q) (70)
or
c(q)
h(q) = , (71)
1 c(q)
where c(q) is the three-dimensional Fourier transform of c(r). From noting that S(q) = 1 +
h(q), we obtain S(q) in terms of c(q):
1
S(q) = 0, (72)
1 c(q)
from which we learn that c(q) 1.

In the derivation of Eq. (70), we have employed the convolution theorem of Fourier transfor-
mation theory. The convolution (german: Faltung), f1 f2 , of two integrable functions f1 (r)
and f2 (r) is defined as

(f1 f2 )(r) := dr f1 (r )f2 (r r ) = dr f2 (r )f1 (r r ) .

(73)

The convolution theorem states that



dr eiqr (f1 f2 )(r) = f1 (q)f2 (q) (74)

i.e., the Fourier transform of the convolution of two functions is equal to the product of their
Fourier transforms.
Using Eq. (72), we finally obtain the compressibility equation in terms of c(q):

1
= 1 c(q 0) = 1 4 dr r2 c(r) . (75)
kB T T 0
B2.20 G. Nagele

3.2 Theory of critical opalescence


In the following, we explore the behavior of g(r) near a critical point. Consider a one-component
system with an attractive part in the pair potential, such as a Lennard-Jones-type system (say,
Argon) or a suspension of sticky colloidal spheres. For a purely repulsive u(r) such as the
hard-sphere potential, there is only one fluid phase and thus there is no liquid-gas critical point
in the one-component case. However, a critical point of a liquid-gas-type demixing transition
may occur in two-component, size-asymmetric dispersions of colloidal hard spheres due to the
depletion attraction effect discussed earlier (cf. Fig. 5 and section 4).

Schematic p V and p T phase diagrams of a Lennard-Jones system are displayed in Fig.


8. The gas-liquid coexistence line in the p T diagram terminates in a critical point at the

V T
Fig. 8: Schematic p V and p T phase diagrams of a Lennard-Jones system (Argon).

critical temperature Tc and pressure pc . It is the location of a continuous (i.e., 2nd order) phase
transition. On approaching the critical point along the coexistence line, the (density-) differ-
ence between liquid and gas phases ceases to exist. At the critical point, strong and long-living
density fluctuations occur (cf. Eq. (60)) such that T becomes arbitrarily large in the thermo-
dynamic limit (cf. Fig. 8):
( )
p
= 0 i.e. T = . (76)
V T,N

According to Eqs. (53) and (72),



1
lim S(q) = 1 + dr h(r) = for T Tc , (77)
q0 1 c(q 0)

which means that S(q) becomes very large for small q as the critical point is approached. Then,
regions of larger and smaller densities develop. As the size of these regions approaches the
wavelength of visible light, there is in fact so much scattering that the fluid appears cloudy or
opalescent. This phenomenon is therefore called critical opalescence. It is due to the occurrence
of long-range spatial correlations between particles in the vicinity of a critical point such that
the volume integral over h(r) diverges in the limit of an infinite volume. As T Tc ,

1
c(q 0) = 4 dr r2 c(r) (78)
0
Theories of Fluid Microstructures B2.21

which means that, contrary to h(r) and g(r), c(r) remains short-ranged with finite second mo-
ment.

Since c(q) is well behaved at Tc , we assume that it can be expanded in a truncated Taylor series
around q = 0 up to O(q 2 ). Using sin(x)/x 1 x2 /6 + O(x4 ), c(q) is thus approximated, for
small q, by
sin(qr)
c(q) = 4 dr r2 c(r) = c0 c2 q 2 + O(q 4 ) (79)
0 qr
with expansion coefficients

c0 = 4 dr r2 c(r) 1 for T Tc (80)
0
2
c2 = dr r4 c(r) . (81)
3 0
The coefficient c2 is positive valued provided the attractive tail of u(r) is sufficiently long-
ranged. We restrict here our attention to small q and thereby to large distances r. Note that
c0 = c(q = 0) 1. Substitution of this truncated expansion into Eq. (72) then gives
1 1 1
S(q) = (82)
1 c0 + c2 q 2 c2 2 + q 2
as an approximation of S(q) for small q, more precisely for qRU 1. This is the small-q
approximation for the near-critical structure factor originally proposed by Ornstein and Zernike
around 1914. The range of the interaction potential is denoted as RU , and we have further
introduced the correlation length
( )1/2 ( )1/2
c2 1/2 T
(T ) := = [c2 S(0)] = c2 id , (83)
1 c0 T
by assuming the fourth moment, c2 , of c(r) to exist and to be positive. With the compressibility
diverging for T Tc in a power-law fashion as T (T Tc ) , it follows that the corre-
lation length diverges as (T Tc )/2 . The numerical value of the critical exponent is
predicted as = 1 in a simple Landau-type mean-field approximation. However, the accepted
value of for the liquid-gas transition is 1.24, as derived from renormalization group calcula-
tions of critical phenomena and verified by high-precision experiments on critical fluids [20,21].

Fourier inversion of the OZ approximation for S(q) gives the asymptotic form of h(r) for large
pair separation r RU

1 iqr 1 er/
h(r) = dq e [S(q) 1] . (84)
(2)3 4 c2 r
At the critical temperature, becomes infinite with
1
h(r)|T =Tc for r RU (85)
r
which corresponds to
1
S(q)|T =Tc for q RU 1 . (86)
q2
B2.22 G. Nagele

The total correlation function decays thus algebraically slow, and not exponentially fast, at the
critical point.

If the OZ approximation described by Eqs. (82) - (86) is valid, then


1 1 [ ]
= c2 2 + q 2 (87)
I(q) S(q)
i.e., a plot of the reduced inverse scattered intensity against q 2 should yield a straight line of
practically constant slope, c2 , and an intercept c2 / 2 that approaches zero for T Tc . Fig. 9
shows such an experimental Ornstein-Zernike plot for argon. It suggests that the mean-field-
type OZ approximation is valid to good approximation at least as long as q and |T Tc | are not
very close to zero.

1

I(q)

q2
Fig. 9: Experimental Ornstein-Zernike plot of argon. From Ref. [22].

Experiments performed very close to Tc reveal in fact deviations at small q from the OZ approx-
imation prediction for S(q). Very close to the critical point the data for 1/S(q) versus q 2 bend
slightly downwards for q 2 0. Improved modern renormalization group theory calculations of
critical phenomena lead to a corrected asymptotic scaling behavior of h(r) at Tc , given in three
dimensions by [20, 21]
1
h(r)|T =Tc 1+ (88)
r
or, equivalently,
1
S(q)|T =Tc 2 , (89)
q
with a small but nonzero critical exponent . The accepted theoretical value for is 0.04. The
reason for the failure of the OZ approximation very close to Tc is that, by assuming a truncated
small-q 2 expansion of c(q) in the neighborhood of q = 0 to hold true, one does not account for
Theories of Fluid Microstructures B2.23

the full spectrum of correlation fluctuations existing at all length scales. In the thermodynamic
limit, fluctuations of very long wavelengths become possible so that at a critical point the free
energy and related two-point correlation functions such as c(r) become non-analytic in their
dependence on the thermodynamic variables and r respectively. It should be noticed here that
the higher-order coefficients cn , with n = 4, 6, , appearing in the regular small-q expansion
of c(q), cease to be finite when u(r) has a van der Waals tail. We finally remark that the liquid-
gas transition of simple liquids at Tc belongs to the same universality class of second-order
phase transitions as the ferromagnetic-paramagnetic transition in uniaxial ferromagnets (cf. the
three-dimensional Ising model). All members of the same universality class show the same
critical exponents.

3.3 Various closure relations


After having explored the long-distance behavior of g(r) in a near-critical liquid, we discuss
now various closure relations which express c(r) approximately in terms of h(r) and a given
pair potential u(r). These relations, and a deeper understanding of the meaning of c(r), can
be obtained by diagrammatic and density functional theory methods. We take here a pragmatic
point of view and establish the closure relations most simply using plausibility arguments. In
combination with the OZ equation, the closures lead to closed integral equations for g(r). These
integral equations have been found, in comparison with computer simulation results and scat-
tering data, to be very useful in calculating the full r-dependence of g(r) and thermodynamic
properties of dense liquids.

For a system with a hard-core excluded volume part in u(r), any closure relation should be
consistent with the exact condition

h(r < ) = 1 , i.e. g(r < ) = 0 , (90)

which states that two spheres of hard-sphere diameter can not interpenetrate, and the asymp-
totic result
c(r) = u(r) , for r (91)
valid for a wide class of pair potentials.

Mean-spherical approximation (MSA):

The exact asymptotic form of c(r) forms the basis of the so-called mean-spherical approxima-
tion, first introduced into liquid state theory by Lebowitz and Percus (1966). In MSA, c(r) is
assumed to be given approximately by the closure relation

c(r) u(r) (92)

for all non-overlap distances r > . Together with Eqs. (90) and (92), the OZ equation (69)
becomes a linear integral equation determining g(r) for r > , and c(r) for r < . The most
attractive feature of the MSA closure, as compared to other ones, is that analytic solutions exist,
even in the many-component case, for various pair potential models, namely for the hard and
sticky hard-sphere potentials, the square well potential, the Coulomb potential, attractive and
repulsive Yukawa-type potentials, and for the dipolar hard-sphere potential. These potentials
B2.24 G. Nagele

are of particular interest for molten salts, electrolyte solutions and in colloid science. No ana-
lytic MSA solution exists for the Lennard-Jones potential.

While the MSA is well suited for short-range attractive and repulsive potentials, it can predict
non-physical negative values for g(r) close to contact distance in case of dilute systems of
strongly repelling particles. At very low density (more precisely, small volume fractions ), the
MSA predicts that

g(r) = 1 + c(r) + O() 1 u(r) + O() , r> , (93)

with a negative g(r) for u(r) > 1. This unphysical prediction should be contrasted with
the exact zero-density form of g(r) given in Eq. (19). Recall here, that the volume fraction
= (/6) 3 is defined as the fraction of the system volume filled by the spherical particles.

Rescaled MSA:

For fluids of (colloidal) particles, where the physical hard core is masked by strong and long-
range repulsive forces, there exists an improved variant of the MSA which preserves the positive
semi-definiteness of g(r). This variant is called the rescaled MSA (Hansen and Hayter, 1982).
It is based on the fact that the g(r) of such systems is continuous at all distances r. Moreover
and most importantly, two particles in these systems are virtually never closer to each other than
a certain distance > , so that g(r) 0 for r < .

g(r)

T
r/
Fig. 10: Radial distribution function of a charge-stabilized Yukawa-type dispersion. Compari-
son between RMSA-g(r) and Monte Carlo computer simulation results. From Ref. [3].

In RMSA, the actual system is replaced by a fictitious system consisting of particles of enlarged
diameter > , at the same number density and, for r > , with the same pair potential
u(r) than the original one. The (density-dependent) effective diameter, , is determined from
the continuity of g(r) at r = , by demanding that

g(r = ; , ) = 0 , (94)
Theories of Fluid Microstructures B2.25

with g(r) calculated in MSA for a larger rescaled volume fraction = ( /)3 > . The
RMSA-g(r) is positive semi-definite since the volume fraction, , of the fictitious system is so
much larger than that Eq. (93) does not apply any more.

Fig. 10 includes the RMSA-g(r) for a dilute aqueous suspension of highly charged polystyrene
spheres interacting by the Yukawa-type DLVO potential of Eq. (3). The charge number,
ZRM SA = 257, employed in the RMSA calculation has been selected such that the height
of the principal peak of the RMSA g(r) is coincident with that of the exact g(r), gener-
ated by Monte Carlo (MC) computer simulations using a smaller charge number ZM C = 205.
Since ZRM SA > ZM C , the RMSA underestimates the structural ordering in systems of strongly
correlated particles. However, once Z has been adjusted to fit the actual peak height, the over-
all shape of g(r) is remarkably well predicted by the semi-analytical RMSA solution. The
RMSA has been extended to multi-component systems of mutually repelling Yukawa particles
(cf. Ref. [3]). The accuracy of the RMSA in its prediction for the effective charge number Z
can be further improved by correcting for the penetrating background of a uniformly assumed
microion density which maintains electro-neutrality [23].

Percus-Yevick (PY) closure relation:

Aside from the linear MSA, there exist a variety of non-linear integral equation schemes. The
Percus-Yevick approximation (Percus and Yevick, 1958) is among the most popular ones. To
introduce the PY closure relation, we reformulate the OZ equation as
[ ]

c(r) = g(r) 1 + dr c(r ) {g(|r r |) 1} =: g(r) gind (r) . (95)

The term in brackets, gind (r), describes the indirect part of the pair correlations. Since g(r) =
exp[w(r)], one can approximate gind (r) by
gind (r) = 1 + (c h) (r) e[w(r)u(r)] , (96)
or, equivalently, c(r) by
[ ]
c(r) g(r) 1 eu(r) = g(r) y(r) = f (r)y(r) . (97)
This is the PY closure relation for c(r). We have introduced here the so-called cavity function,
y(r), defined as
y(r) := eu(r) g(r) . (98)
Contrary to g(r), which for hard spheres has a jump discontinuity at r = due to the factor
exp[u(r)], y(r) is continuous for all r. It agrees with g(r) for all r where u(r) = 0. In PY
approximation, c(r) is thus assumed to be zero at distances where the pair potential vanishes.
The continuity of y(r) is easy to see, using Eq. (13), from noting that
[ ]
V2 N

y(r12 ) = lim dr3 drN exp u(rij ) (99)
ZN
i<j

for pairwise additive forces, with the pair {i, j} = {1, 2} omitted from the sum. Hence y(r) is
a smooth continuation of g(r) into the overlap region (cavity) r < .
B2.26 G. Nagele

Substitution of Eq. (97) into the OZ equation gives the non-linear PY integral equation
[ ][ ]
u(|rr |) u(r )
y(r) = 1 + dr e y(|r r |) 1 e 1 y(r ) (100)

for y(r) or, likewise, g(r). This equation can be solved analytically in three dimensions for
the important case of hard spheres (cf. section 3.4), and by numerical methods for arbitrary
pair potentials. In contrast to the HNC approximation, the PY approximation can not exclude
unphysical negative values in g(r). Although the PY-g(r) of mondisperse hard-sphere fluids is
non-negative at all distances r, negative values of certain PY partial radial distribution functions
of strongly size-asymmetric hard-sphere mixtures can be found at some intermediate distances.
However, the PY predicted contact values in these asymmetric systems remain positive valued.

The PY approximation is exact to first order in the particle density. For a proof expand the YBG
equation in Eq. (30) to linear order in , using

g (3) (1, 2, 3) = [f (12) + 1][f (13) + 1][f (23) + 1] + O() (101)

with Mayers f function, and the definitions i := ri and ij := rij . This gives
[ ]
1 [w(12) u(12)] = lim 1 d3 f (13) f (23) + O(2 ) (102)

so that
[w(r) u(r)] = (f f ) (r) + O(2 ) . (103)
The T-limit of an infinite system should be taken after differentiation with respect to r1 , with
the consequence that the constant terms coming from the factorization of (f 1)(f 1) drop
out. Expansion of the PY-closure in Eq. (96) to linear order in gives the same result, showing
that g(r) in PY approximation is exact to first order in density. Since
[ ]
g(r) eu(r) = eu(r) e(w(r)u(r)) 1 eu(r) [w(r) u(r)] + (2 ) , (104)

we obtain the leading order virial expansion result


[ ]
g(r) = eu(r) 1 + (f f ) (r) + (2 ) . (105)

Inserting this into the OZ equation gives


[ ]
c(r) = f (r) 1 + (f f ) (r) + (2 ) (106)

for the low-density form of the direct correlation function. The common factor f (r) in the virial
expansion of c(r) indicates that its range for non-ionic systems is essentialy that of u(r).

Hypernetted-chain (HNC) approximation:

Another frequently used approximate integral equation scheme is the hypernetted-chain ap-
proximation (van Leeuwen et al., 1959). The name stems from its diagrammatic derivation.
The HNC closure relation in terms of c(r) is

c(r) u(r) + h(r) ln [1 + h(r)] = h(r) ln y(r) . (107)


Theories of Fluid Microstructures B2.27

Exponentiation gives
g(r) eu(r)+h(r)c(r) , (108)
showing that in HNC approximation the positive definiteness of the exact g(r) is preserved at
any density. Since from Eq. (84)

c(r) u(r) , r (109)

the HNC approximation leads further to the correct asymptotic behavior of c(r) for arbitrary .
On the other hand the PY-c(r) gives the correct long-distance behavior in general only for small
densities.

The HNC closure combined with the OZ equation leads to



[ u(r) ]
ln e g(r) h(r) c(r) = dr c(|r r |) h(r ) . (110)

Introducing the cavity function, Eq. (110) can be re-expressed as



ln [y(r)] = dr h(r ) [u(|r r |) + h(|r r |) ln g(|r r |)] . (111)

This is the non-linear HNC integral equation for g(r). It can be solved only numerically even
for hard spheres. Like in the PY approximation, the HNC approximation predicts g(r) correctly
to first order in the density.

The PY is quite successful for hard spheres or, more generally, for particles with short-range
hard-sphere-like interactions. In contrast to the MSA, however, it does not work so well for
systems with attractive tails. The HNC is complementary to the PY in the sense that it is un-
satisfactory for hard spheres but appears to account satisfactorily for the effects of soft cores
and, in particular, for long-range repulsive potential tails as given in ionic fluids and dispersions
of Yukawa particles. All three integral equation schemes have severe deficiencies close to a
gas-liquid critical point.

HNC results for a Yukawa system (cf. Eq. (3)) of moderately charged colloidal particles (Z =
107) of diameter = 160 nm and fixed screening parameter in an organic solvent ( = 10)
are shown in Fig. 11. The charge Z was determined from a fit of the HNC peak height of S(q)
to the experimentally given one. There is then rather good agreement between the theoretical
and experimental S(q). The deviations at small q and around the minimum can be attributed
to polydispersity effects, that is to a spread in the experimental particle sizes. With increasing
volume fraction , there is increasing ordering visible through more pronounced undulations
in g(r), and the system becomes less compressible (decreasing S(0)). The particles avoid each
other as much as possible because of the strong and longer-ranged electrostatic repulsion so that
g(r < 1.5) = 0. In monodisperse systems with long-range repulsion, the position, qm , of the
principal peak of S(q) increases with volume fraction approximately as qm 1/3 . Away from
a critical point, the location, rm , of the main peak of g(r) is approximately related to qm by
2
rm . (112)
qm
B2.28 G. Nagele

= 0.101

S(q)
g(r)

r/ q [107 m-1 ]

Fig. 11: HNC radial distribution function g(r) at various volume fractions (left), and static
structure factor S(q) (right) of charge-stabilized dispersions of silica spheres. Open circles:
light scattering results of S(q). From Ref. [24].

Random phase approximation (RPA):

Suppose we can separate the pair potential of a liquid system into a short-range reference part,
u0 (r), and a soft and long-range perturbational part ul (r) (cf. Fig. 12),

u(r) = u0 (r) + ul (r) , (113)

where ul (r) is Fourier-integrable. Let us further assume that the direct correlation function,
c0 (r), of the reference system (where u = u0 ) is known exactly or to a good approximation. For
a reference system of hard spheres, e.g., we could use the analytic PY solution for c0 (r) derived
in section 3.4. The true direct correlation function, c(r), of the system can then be approximated

u 0 (r )

1 r

u1 (r)
Fig. 12: Pair potential consisting of hard-sphere reference part, u0 (r), and a longer-ranged,
attractive perturbational part ul (r).

by
c(r) c0 (r) ul (r) , r > 0 (114)
Theories of Fluid Microstructures B2.29

which is asymptotically correct at long pair separations. For historical reasons, this closure
relation is referred to as the random phase approximation. In this approximation one obtains
1 1
S(q) = . (115)
1 c(q) 1 c0 (q) + ul (q)

The static structure factor is thus expressed in terms of the structure factor, S0 (q), of the refer-
ence system and the Fourier transform,

ul (q) = dr eiqr ul (r) , (116)
0

of the perturbational part of u(r). Eq. (115) can be rewritten in the mnemonic form

1 1
+ ul (q) . (117)
S(q) S0 (q)

The RPA is the most simple perturbation theory for fluid microstructures, usually suited only for
small wave numbers. It has been successfully used, also in its multi-component extension, for
the calculation of (partial) structure factors in dense systems with (ultra-) soft potentials where
u0 (r) = 0, such as polymer blends and star polymers [25, 26]. For an ideal gas as reference
system, one has c0 (r) = 0 and S0 (q) = 1. The RPA reduces then to the MSA for point-like
particles, referred to in the literature as a version of the Debye-Huckel approximation. One
should notice that the perturbation must be sufficiently weak, or the density sufficiently low, to
ensure that ul (q)S0 (q) > 1 with S(q) > 0. Furthermore, the RPA does not ensure that
g(r < ) = 0 in case of a hard-sphere reference system. This non-physical feature of the RPA
is related to an ambiguity in the choice of the perturbation potential ul (r) in Eq. (116) for r < .
The true g(r) should not depend on this choice.

In the optimized random phase approximation (ORPA), ul (r) is extended into the hard-core
regime r < such that g(r < ) = 0. While the ORPA is a considerable improvement of the
RPA, there is a price to pay in form of a much larger numerical effort to calculate S(q).

3.4 Percus-Yevick solution for hard spheres


Hard spheres serve as a reference system in the theory of uncharged liquids, as an ideal gas does
in the theory of dilute gases, and a harmonic solid in solid-state physics. The PY approxima-
tion leads to an integral equation for the hard-sphere cavity function y(r) which can be solved
analytically.

The solution proceeds as follows. For hard spheres,


[ ]
c(r) = g(r) 1 eu(r) = 0, r> (118)

in PY approximation, i.e. the hard-sphere direct correlation is approximately set equal to zero
for non-overlap distances. As a matter of fact, the true c(r) has a small but non-vanishing tail
for r > . As can be noticed here, the PY closure is identical to the MSA closure (cf. Eq. (92))
in the case of a hard-sphere fluid.
B2.30 G. Nagele

The hard-sphere cavity function reads



g(r), r > exact
y(r) = eu(r) g(r) = (119)

c(r), r < PY aproximation
where the lower equality follows from the PY closure in Eq. (97). It follows that c(r) and g(r)
share a jump discontinuity of equal magnitude at r = , with g(r = + ) = c(r = ), since
y(r) is continuous everywhere.

Upon inserting the hard-sphere potential into Eq. (100), one obtains a quadratic integral equa-
tion for y(r) of the form


y(r) = 1 + dr y(r ) dr y(r ) y(|r r |) . (120)
r < r <, |rr |>

It is required to solve this integral equation within r < for c(r) = y(r), since c(r > ) = 0
is known already. Following Wertheim [27], we use a third-order polynomial
c(r) = a0 + a1 r + a2 r2 + a3 r3 (121)
as a trial solution of c(r < ), with yet unknown density-dependent expansion coefficients {ai }.
This ansatz is suggested from the low density form of c(r), which is a third order polynomial
in case of hard spheres. For a proof of this statement use Eq. (65) to show that to first order in
density, y(r) is given by

y(r) = 1 + dr f (r ) f (|r r |) + O(2 ) . (122)

For hard spheres, the Mayer-f function is f (r) = 1 for r < and zero otherwise. The
convolution integral in Eq. (122) is then equal to the volume of overlap of two spheres of equal
radii with centres separated by r. As a consequence
[ ]
3 1 3
y(r) = 1 + 8 1 x + x (2 x) + O(2 ) , (123)
4 16
with x = r/. The overlap volume is zero for x > 2 as expressed by the unit step function
(2 x). Recall that the PY approximation describes g(r) exactly to first order in . In using
the polynomial ansatz in Eq. (121), it is assumed that the functional form of c(r) is valid for all
volume fractions.

The four expansion coefficients, {ai }, are determined by employing the continuity of y(r) and
its first two derivatives at r = . Their continuity follows from Eq. (120) and its first two
derivatives. A fourth condition follows from the PY integral equation (120) evaluated at r = 0:

y(0) = 1 + dr y(r ) . (124)
r <

After inserting Eq. (121) in Eq. (120) and making use of the four boundary conditions to
determine the {ai }, a lengthy calculation gives the following PY result for the hard-sphere c(r):
[ ( ) ]
1
c(r < ) = 1 1 + x + 2 x 3
(125)
2
Theories of Fluid Microstructures B2.31

with
(1 + 2)2 6 (1 + 0.5)2
1 = , 2 = . (126)
(1 )4 (1 )4
The PY result for c(r) reduces, for small , to the correct first order density form of y(r)
given in Eq. (123).

g(r)

c(r)

0 1 2 3 4 5
r/ r/
Fig. 13: Percus-Yevick direct correlation function (left) and radial distribution function (right)
of hard spheres.

Fourier transformation of c(r) leads with Eq. (72) to an analytic expression for S(q). This
expression reads explicitly [3]
1
S(y) = 2 (127)
X (y) + Y 2 (y)
with
X(y) = 1 12 [ A f1 (y) + B f2 (y) ] (128)
Y (y) = 12 [ A f3 (y) + B f4 (y) ] , (129)
where
1 + 2 1 + 0.5
A= , B= , (130)
(1 )2 (1 )2
and
y sin(y) cos(y) 1
f1 (y) = , f2 (y) = (131)
y3 y2
f2 (y) 1
f3 (y) = + , f4 (y) = yf1 (y) . (132)
y 2y
We have introduced here the reduced wave number y = q.
The reduced isothermal compressibility follows as
(1 )4
lim S(q) = , (133)
q0 (1 + 2)2
which is a monotonically decreasing function in . For given analytical S(q), the hard-sphere
g(r) can be determined in principle by numerical Fourier-inversion. However, to avoid prob-
lems caused by the jump discontinuity in g(r), it is safer to calculate first the function (r) :=
B2.32 G. Nagele

S(q)

r/

q
Fig. 14: Percus-Yevick static structure factor of hard spheres.
6
= 0.49
5
MD
4 PY-VW
g(r) PY
3 RY

0
2 3 4 5
r/a
Fig. 15: Hard-sphere g(r) for a volume fraction = 0.49 close to the freezing transition.
Comparison between PY, RY, Verlet-Weis corrected PY and MD computer simulations (filled
circles). From Ref. [28].

h(r) c(r) by Fourier-inverting (q) = c(q)h(q) = [S(q) 1]2 /(S(q)). The hard-sphere
g(r) follows then in PY approximation from g(r) = y(r) = 1 + (r) for r > . Contrary to
g(r), (r) and its first two derivatives are continuous also at r = as one can deduce from
the OZ equation. Notice here that the identity y(r) = 1 + (r) holds true only within PY ap-
proximation. The relation between y(r) and (r) is approximated in the HNC approximation
by y(r) = exp{(r)}, which agrees with the corresponding PY relation for small (r) only.
While the full PY-g(r) of hard spheres can not be represented analytically, one can derive closed
expressions for the contact values of g(r) and its first derivative from the continuity of y(r) and
its derivative at r = :
1 + 0.5
g(r = + ) = , (134)
(1 )2
and
dg 4.5 (1 + )
(r = + ) = . (135)
dr (1 )3
PY results for the hard-sphere c(r) and g(r), and for the static structure factor, are shown in
Figs. 13 and 14, respectively, for various volume fractions. For = 0.1, we further show the
Theories of Fluid Microstructures B2.33

cavity function with y(r) = c(r) for r < . The PY approximation provides a quite good
representation of the true hard-sphere S(q) and g(r) for volume fractions 0.35. At larger
values of it underestimates the contact value of g(r) and its oscillations are slightly out of
phase, as can be seen from Fig. 15 in comparison with Molecular Dynamics (MD) computer
simulation results. The PY approximation further fails to predict the liquid-solid freezing tran-
sition which occurs for hard spheres at f = 0.494. This failure is not restricted to the PY
approximation: The Ornstein-Zernike integral equations discussed in this lecture have been de-
signed to describe the homogeneous fluid state, and are thus not suited to describe the density
inhomogeneities and the symmetry change in a first-order liquid-crystal phase transition.

On the basis of the analytic PY solution, Verlet and Weis [29] provide a simple prescription to
obtain results for the hard-sphere g(r) and S(q), which are in good agreement with computer
simulation results even up to the freezing volume fraction (see also [30]).
For a given and , the Verlet-Weis corrected g(r) is given by
(1)x
e
gV W (x; ) = gP Y (x ; ) + A( ) cos [ (x 1)] , (136)
x

with x = r/, and a rescaled volume fraction and diameter = ( /)1/3 . The three
parameters A, , and are determined on demanding that gV W ( + ) = (pCS /pid 1) /4 =
(1 0.5 ) / (1 )3 and SV W (q = 0) = kB T (/pCS )T in agreement with the Carnahan-
Starling equation of state for the hard-sphere pressure given in Eq. (140), and from minimizing
the integral over |gM D (r/; ) gP Y (r/ ; )| for an interval ranging from r/ = 1.6 to 3.0.
Here, gM D (r) is the exact rdf obtained from Molecular Dynamics (MD) simulations. This
leads to = (1 /16), and to A and expressed as simple functions of [29]. Fig. 15
compares the rdf predictions of various integral equation schemes, including the VW-corrected
PY and RY schemes, for a hard-sphere system at volume concentration = 0.49 close to
freezing. The principal peak height of S(q) at wave number qm , predicted by the VW-corrected
PY scheme, is well described by the expression [7, 31]

(1 0.5)
S(qm ) = 1 + 0.644 , (137)
(1 )3

and conforms with the empirical Hansen-Verlet freezing criterion [32]. According to this cri-
terion, a three-dimensional fluid freezes into a solid when the principal peak height, S(qm ), of
S(q) is located in between 2.8 to 3.1. The precise height depends to some extent on the range
of the pair potential, and amounts to S(qm ; f = 0.494) = 2.85 in the case of hard spheres.
For dispersions of strongly repulsive colloidal particles, this criterion becomes equivalent to a
dynamic freezing criterion [33], as has been shown on the basis of a dynamic mode coupling
theory [34]. For an alternative analytic parametrization of the rdf of hard-sphere fluids see [35].

3.5 Thermodynamic consistency and Rogers-Young closure


One important reason for the failure of the PY approximation at higher densities is its ther-
modynamic inconsistency : due to the approximate nature of the PY-g(r), the thermodynamic
routes in Eqs. (48-53) give pressure curves, p(), which become increasingly different from
each other with increasing density. The results for the thermodynamic properties obtained via
the three routes are in general different for all OZ integral equations discussed so far. This lack
B2.34 G. Nagele

of thermodynamic consistency is a common feature of approximate theories.

To illustrate the thermodynamic inconsistency of the PY approximation for the case of hard
spheres we integrate the compressibility in (133) with respect to . This yields the compress-
ibility equation of state
pc 1 + + 2
= . (138)
pid (1 )3
Using instead the pressure (or virial) equation of state, Eq. (49), one obtains

pv 1 + 2 + 32
= 1 + 4 g(r = + ) = , (139)
pid (1 )2

which agrees with the pressure, pc , derived from the compressibility equation only up to third
order in the volume fraction. The pressure Eq. (49) is referred to also as virial equation of
state, since it can be derived from the virial theorem of classical mechanics. The first equality
in Eq. (139) relating the pressure to the contact value of g(r) is an exact relation stating that p =
kB T [1 + 4 g( + ; )]. There is thus only a trivial temperature dependence of the pressure in
the equation of state of hard spheres which arises solely from the ideal gas contribution.
The energy equation, Eq. (48), leads for hard spheres to the exact result E = Eid = N kB T ,
with zero excess internal energy coming from particle interactions. Thus, the excess pressure
can not be derived using the energy equation, since the internal energy of hard spheres is of
purely kinetic origin. As a consequence, the phase diagram of hard spheres is temperature in-
dependent and depends only on (athermal system). Note that the pressure and the elastic bulk
and shear moduli are measures of the energy density and thus scale for hard spheres as kB T
times a function of alone.

The phase transitions in a hard-sphere fluid are purely entropy driven, since Aex = T Sex , so
that a stable phase at a given has maximal entropy. A system of monodisperse hard spheres
is in a supercritical fluid state for < f 0.494, where f is the freezing volume frac-
tion, followed by a fluid-crystal coexistence region for f < < m 0.54 where m is
the melting volume fraction. In thermodynamic equilibrium, a fully crystalline face-centered
cubic (fcc)phase, characterized by 12 nearest neighbors, is found for m < < cp . Here,
cp = / 18 0.7404 is the maximal packing density allowed for monodisperse spheres,
as conjectured by Johannes Kepler already in the year 1611. However, the mathematical proof
that fcc is the densest packing of monodisperse, non-deformable spheres was given no earlier
than in 1998 by Thomas Hales, about three hundred years past its conjecture. For comparison,
the closest-packed volume fractions
for a simple cubic (sc) and a body-centered cubic (bcc)
crystal are /6 0.52 and 3/8 0.68, respectively. A hexagonal cubic crystal has the
same closest packed volume fraction as a fcc crystal. However, a fcc hard-sphere crystal is
thermodynamically more stable than a hc crystal [36]. While the pressure in the fluid branch
of a hard-sphere fluid is well described by the Carnahan-Starling equation of state given in Eq.
(140), a decently good parametrizations of the fcc branch, valid for > m , has been provided
by Wood [37] and Hall [38]. The orientation-averaged pair distribution function, gf cc (r), of the
fcc hard-sphere phase has been parameterized by Kincaid and Weis [39].

Real suspensions of colloidal hard spheres are always polydisperse to a certain degree, and nu-
cleation often becomes exceedingly slow so that for 0.58 < < rcp 0.634 the system is
Theories of Fluid Microstructures B2.35

trapped in a glass-like, non-equilibrium phase without long-range ordering. Here, rcp is the
volume fraction of random close packing [40] where a sphere has on the average six contacts
with neighboring ones. Both at random close packing and crystal close packing the particles get
immobilized so that the pressure at concentrations close to these two concentrations diverges
like p 3kB T /( rcp ) and p 3kB T /( cp ), respectively.

Contrary to the three-dimensional case where rcp < cp , and consistent with the occurrence
of a glassy phase, the local rule of starting from a single sphere and placing the next one such
that the (areal) density is maximised leads precisely to the areal volume fraction, 2D , of a
two-dimensional hexagonal lattice. The latter is the one of the maximal packing configuration
with 2D
cp = / 12 0.9069.

p
1
pid


Fig. 16: Hard-sphere compressibility and pressure (virial) equations of states in PY and HNC
approximations. Dashed line: exact pressure curve. From Ref. [1].

The PY and HNC compressibility and pressure (virial) equation of states for hard spheres are
plotted in Fig. 16, in comparison with the exact pressure curve obtained from computer
simulations. The exact hard-sphere equation of state is very well described in the fluid regime
( 0.494) by the Carnahan-Starling (CS) formula [1, 2]
[ ]
pCS 1 1 2 1 + + 2 3
= pv + pc = . (140)
pid pid 3 3 (1 )3
As can be noticed from Fig. 16, the exact pressure is bracketed by pc and pv , with the difference
between pc and pv increasing for increasing . The PY solution is obviously a better approxi-
mation for hard spheres than the HNC. The fluid-solid coexistence pressure obtained from the
CS formula is pCS /pid (f ) = 12.48 corresponding to g( + ; f ) = 5.81.
FS02_22
B2.36 G. Nagele

Rogers-Young (RY) approximation:

Several hybrid integral equation schemes have been proposed in the past which partially restore
thermodynamic consistency. Out of these schemes, we discuss here only the one proposed
by Rogers and Young [41], which interpolates between the PY and HNC approximations and
removes part of their thermodynamic inconsistencies. It is based on the observation (cf. Fig.
17) that computer simulation data for the S(q) of particles with purely repulsive pair potentials
are bracketed, near qm , by the PY and HNC structure factors. The RY closure relation is given
by { }
u(r) 1 [ f (r)[h(r)c(r)] ]
g(r) e 1+ e 1 (141)
f (r)
with a mixing function,
f (r) = 1 er , (142)
including a mixing parameter {0, }. The closure relation is constructed in such a way
that for

r or 0 : f (r) 0 RY PY (143)
r or : f (r) 1 RY HNC .

Hence the RY y(r) reduces to its PY value for r 0 and to its HNC value for r . This

3
PY
RMSA
HNC
2 RY
MC
g(r)

0
1 2 3
r/
Fig. 17: Radial distribution function of a charge-stabilized, Yukawa-like suspension with =
0.2 and Z = 100. Comparison of RY, HNC, RMSA and PY predictions with MC simulations.
From Ref. [28].

is consistent with the observation that, while the HNC closure is correct at large separations,
the PY approximation is expected to be more reliable at small r, at least for strongly repulsive
potentials. The parameter determines the proportion in which HNC and PY are mixed at
intermediate r. Its numerical value follows from requiring partial thermodynamic consistency
by demanding the equality, pT = cT , of the compressibilities obtained from the pressure and
compressibility equations od state. Since T is directly related to the long-wavelength limit of
the static structure factor, one may expect that the RY approximation will provide quite reliable
results for S(q) at finite q.
Theories of Fluid Microstructures B2.37

The RY mixing scheme has been found to perform very well for three-dimensional liquids with
purely repulsive pair potentials, unless the system is very close to the freezing transition line.
Its predictions of the pair structure are less precise for two-dimensional systems. Moreover, the
RY scheme is less accurate for ultra-soft potentials such as the one describing star polymers,
since its closure is PY-like at small particle separations [26].
Fig. 17 includes a comparison of RY, HNC and PY results for g(r) with Monte Carlo (MC) sim-
ulations of a three-dimensional colloidal dispersion of Yukawa spheres. The true Monte-Carlo
g(r) is strongly overestimated by the PY approximation, whereas the fluid g(r) is somewhat
underestimated by the HNC approximation, and to an even larger extent by the RMSA. The RY
approximation, on the other hand, reproduces the MC data quite well.

3.6 Extension to mixtures


The concept of pair and direct correlations, and the Ornstein-Zernike integral equation schemes
described in earlier sections, can be generalized without difficulty to multi-component liquids
and polydisperse (colloidal) systems. We will discuss these generalizations in the following.

Partial correlation functions

Consider an atomic or colloidal liquid consisting of m components of spherical particles of


diameters and partial number densities = N /V . We employ greek symbols, with =
1, . . . , m, to label the m components which build up the liquid. The particles within each
component are identical. We assume that the particle interactions can be described by pairwise
additive forces. The following replacements

u(r) u (r)
g(r) g (r)
c(r) c (r)
h(r) h (r) = g (r) 1 (144)

are needed for a generalization to a m-component mixture. There are m(m + 1)/2 partial pair
potentials u (r), radial distribution functions g (r), total correlation functions h (r), and
direct correlation functions c (r) (with , {1, . . . , m}) necessary to characterize the fluid
pair structure. Here, u (r) is the pair potential of two particles belonging to component
and , respectively. Furthermore, g (r) gives the relative conditional probability of finding
an -type particle a distance r apart from a given -type particle. The partial pair potentials
are obviously symmetric in the two component indices, i.e. u (r) = u (r). The remaining
functions in Eq. (144) are thus also symmetric in their component indices.

The one-component OZ equation is replaced in mixtures by a set of m(m + 1)/2 coupled OZ


equations, one for each h (r). In case of a homogeneous and isotropic liquid mixture, these
OZ equations are given by


m
h (r) = c (r) + dr c (|r r |) h (r ) . (145)
=1
B2.38 G. Nagele

Fourier-transformation gives the OZ equations in q-space:


m
h (q) = c (q) + c (q) h (q) . (146)
=1

The total correlation function, h , between two particles of components and is thus written
as the sum of a direct correlation part, c , and an indirect correlation part mediated through all
other particles of components = 1, , m with relative density weight .

Since there are m(m + 1)/2 unknown functions, c (r), in the OZ equations one needs the
same number of closure relations to obtain a complete set of integral equations determining the
partial radial distribution functions. We quote here only the multi-component generalizations
of the MSA relations. All the other integral equation schemes discussed previously can be
generalized accordingly to mixtures. The multi-component MSA closure relations are [3]

c (r) u (r)/kB T , r > ( + ) /2. (147)

Combined with the exact non-overlap conditions

h (r) = 1 , r < ( + ) /2 , (148)

and the OZ equations, one obtains a closed set of integral equations for the h (r). In place of
a single static structure factor, there are now m (m + 1) / partial static structure factors related
to the partial rdfs by

1/2
S (q) = + ( ) dr eiqr h (r) , (149)

and defined such that S (q ) = .

4 Effective Colloid Interactions


In the framework of many-component liquid state theory, I will outline the derivation of two
widely used effective pair potentials, namely the potential due to Asakura-Oosawa [42, 43] and
Vrij [44, 45], that describes the attraction of colloidal hard spheres induced by free polymer
chains, and the electrostatic potential in Eq. (3) describing the repulsion of charged colloidal
spheres in the presence of screening microions. The effective colloid potentials will be obtained
by contracting the small particles (i.e., the polymer coils and microions, respectively) out of the
description.

Consider a binary suspension of N large colloidal spheres (component L) at positions rN and


a second component (S) of small particles in a macroscopic volume V. Suppose that the small
particles and the solvent, the latter treated here as an incompressible, structureless continuum,
are in osmotic equilibrium with a reservoir of small particles and solvent, through a semi-
permeable membrane impenetrable to the large colloids. This fixes the chemical potential, S ,
of the small particles rather than their number density. The system under consideration is most
conveniently described in terms of a semi-grand canonical ensemble, where the number N
Theories of Fluid Microstructures B2.39

Fig. 18: Left: Asakura-Oosawa-Vrij (AOV) model of colloidal hard spheres in osmotic equilib-
rium with a solution of ideal, freely overlapping polymer coils. Right: lens-shaped depletion
zones overlap volume V2 (r) (shadowed) of two colloidal spheres.

of large particles in the suspension volume V , and the chemical potential of small particles are
fixed.1 The semi-grand total free energy, F (N, V, T, S ), of the mixture is then given by [13,46]
{ }
dr N (zS )NS
eF = eULL (r ) dxNS e [ ULS (r ,x )+USS (x ) ]
N N NS NS

N ! 3N NS =0
N S !

drN ef f N
= 3N
eULL (r ) , (150)
N!
where U = ULL + ULS + USS is the total potential energy of the mixture, zS = exp{S }/3S
is the activity coefficient of a small particle at fixed reservoir chemical potential S , and
and S are the thermal wavelengths of large and small particles, respectively. We have used
the grand-canonical average over the set of position vectors, xNS , of small-particle systems of
particle number NS ranging from zero to infinity.
In the second equation, I have introduced the effective interaction energy,
ef f N
ULL (r ; V, T, S ) = ULL (rN ) + (rN ; V, T, S ) , (151)
of the N large particles in the mixture. It is the sum of the bare potential contribution, ULL , and
a grand-free energy contribution, , with
{ }
(zS )NS
NS [ ULS (rN ,xNS )+USS (xNS ) ]
(r ; V, T, S ) = kB T ln
N
dx e , (152)
N =0
NS !
S

of an inhomogeneous dispersion of small particles in the external field of N large particles fixed
at positions rN . Similar to the discussion of Eq. (30), we can convince ourselves that
( )
(zS )NS
NS 0 NS ! dx NS
U
r1
eU
N
F1 (r ) = (zS )NS
= 1 ULL
ef f N
(r ) (153)
dx N S e U
NS 0 NS !

1
Different ensembles become equivalent in the T-limit, and one can change then from one description to the
other by a Legendre transformation. However, it is conceptually advantageous to select that ensemble compatible
with the experimentally controlled variables and the physical constraints.
B2.40 G. Nagele

is the mean force on a large sphere 1, for the positions of the N large particles fixed, and grand-
canonically averaged over the positions of the small particles. Here again, the total potential en-
ef f
ergy in the mixture is denoted by U (rN ; xNS ). Hence, ULL plays the role of a state-dependent
ef f
N -particle potential of mean force. Once ULL is known, any configurational average over a
function, f (rN ), of large-particle positions, such as the colloidal pair distribution function, can
ef f ef f
be formally expressed as a one-component average involving ULL only. However, ULL is state-
dependent, since it includes a free-energy contribution, , of the small particles. Therefore, and
ef f
even for a pairwise-additive ULL , ULL contains in general the whole sequence of many-body
contributions.

ef f
In the framework of the semi-grand canonical treatment, the ULL of a homogeneous system
can be uniquely decomposed in a sum of independent many-body contributions [13],


N
N
ef f N
ULL e0 (N, V, T, S ) +
(r ; V, T, S ) = U e2 (rij ; S ) +
u e3 (rij , rik ; S ) + (154)
u
i<j i<j<k

Here, ue2 (rij ; S ) is the grand-free energy-change required to bring two colloidal spheres in the
system from infinity to positions r1 and r2 . Likewise, u e3 (rij , rik ; S ) is the free energy change
to bring three infinitely distant particles to (r1 , r2 , r3 ), minus the sum of the three pairwise work
contributions, and so on. The many-body terms are constructed such that u en 0, whenever
the distance of a pair of particles in a n-cluster is very large. It is crucial to realize that, pro-
vided the LS and SS interactions embodied in ULS and USS are short-ranged as compared to
ef f
the LL interactions, one can expect that the three-body and higher-order contributions to ULL
are quite small. This allows then to map the mixture on an one-component system of large par-
ticles interacting pairwise via the state-dependent pair potential u e2 . Incidentally, if the degrees
of freedom of the large particles would be integrated out instead of the small ones, then the
ef f
resulting USS (x) would include non-negligible many-body contributions of any order, which
renders such a contraction rather useless.

The decomposition of ULL ef f


in Eq. (154) includes a configuration-independent contribution Ue0 .
This state-dependent, so-called volume term has no bearing on the microstructure and thus on
gLL (r), but it contributes to thermodynamic properties like the system pressure. In certain cases,
the volume term may influence the phase behavior. A case in point are low-salt suspensions of
charged colloidal particles where, due to the electro-neutrality constraint and the infinite range
of the bare Coulomb forces, U e0 reveals a nonanalytic dependence on the colloid concentration
(see, e.g., [4750]).

For a binary mixture of neutral particles, the thermodynamically extensive volume term is given
by [13]
e0 = V S (S ) + N u
U e0 (S ), (155)

where S (S ) is the osmotic pressure of small particles in the system void of colloids, which
is equal to the reservoir polymer pressure, and u e0 (S ) is the grand-free energy change arising
from inserting a single colloidal particle to the polymer solution. Since U e0 /V is here linearly
dependent on L = N/V , it only adds a constant to the pressure and chemical potential of the
effective one-component system, with zero influence on its phase behavior.
Theories of Fluid Microstructures B2.41

4.1 Depletion-induced colloid attraction


The effective pair potential, ue2 (r), and the volume term can be straightforwardly calculated for
the Asakura-Oosawa-Vrij (AOV) model of colloidal hard spheres (L = c) of diameter = 2a
and density c = N/V , in osmotic equilibrium with a solution of small and non-adsorbing
free polymer chains (S = p) (see Fig. 18). In this idealizing model, originally discussed by
Vrij [44, 45], the polymer coils are simply described as hard spheres of diameter p = 2ap ,
as far as their interactions with the colloidal spheres is concerned, but with respect to their
mutual interactions they are treated as ideal point particles, i.e., as freely overlapping spheres.
Explicitly, the partial pair potentials in the mixture are

ucc (r) = , r<


ucc (r) = 0, r>
ucp (r) = , r < ( + p ) /2
ucp (r) = 0, r > ( + p ) /2
upp (r) = 0, r0. (156)

The grand free energy of the ideal gas of polymers in the system volume V in presence of N
hard spheres fixed at positions rN , and in contact with a reservoir of ideal polymers, is given
by [51]

id (rN ; V, p ) = id N
p (p ) Vf ree (r ) . (157)

Here, the T -dependence has been hidden for conciseness. According to Fig. 18,

4 N N
Vf ree (r ) = V N (a + ap ) +
N 3
V2 (rij ) + V3 (rij , rik ) + (158)
3 i<j i<j<k

is the free volume in the mixture accessible to the polymer sphere centers. Moreover, id p =
kB T zp is the osmotic pressure in the reservoir of the ideal-gas polymers at the given reservoir
activity zp , which for ideal polymers that do not
( r interact
) among themselves is equal to the reser-
r 3
voir polymer density, p , with p = kB T ln p p . Note here that the polymer concentration
in the reservoir is in general different from the one in the mixture (see Fig. 18 and [51,52]). The
reason why the reservoir osmotic pressure of polymers appears in Eq. (157), and not the os-
motic polymer pressure in the mixture, is that work has to be done against the polymer pressure
of the reservoir, when the free volume in the system decreases due to a change in the positions
of the colloidal spheres.
Each colloidal sphere excludes a volume (4/3) (a + ap )3 from the polymer centers. When
two or more exclusion (depletion) volumes overlap, more free volume becomes available for
the polymers with a corresponding loss in the grand-free energy. The overlap volumes, V2 ,
V3 , and so on of two, three and more excluded volumes correct for this gain in free volume.
In summary, the suspension can lower its free energy by the clustering of colloidal particles,
which implies that the spheres feel an effective attractive interaction commonly referred to as
depletion attraction. For the present model, where only temperature-independent excluded vol-
ume effects are operative, this attraction is purely entropic, i.e., the loss in free energy is due to
the gain in entropy caused by overlapping depletion zones.
B2.42 G. Nagele


For size ratios q = p / < 2/ 3 1 0.155, three-body and higher-order excluded volume
overlaps are exactly zero.2 When the ratio q is increased beyond this threshold, three-body and
higher-order interaction terms come successively into play. If three excluded volume zones can
overlap, the total gain in three volume is smaller than would be estimated on basis of the pair
interactions alone. Thus, three-body interactions lead to a repulsive correction to the depletion
interaction. The four-body interactions, in turn, are attractive again.
Quite notably, for q < 0.155, the AOV model of a colloid-polymer mixture is mapped on an
effective one-component colloid system with exactly pairwise additive effective interactions.
ef f
On noting that Ucc = Ucc + id , and using Eq. (158), we obtain the result

[ 3]

N
ef f N
Ucc (r ; V, p ) = kB T zp V 1 (1 + q) + [ uHC (rij ) + uAOV (rij ) ] , (159)
i<j

for the effective colloid interaction energy, with a purely attractive AOV depletion pair potential
given by [44, 45]

uAOV (r; p ) = id
p (p ) V2 (r) (160)
3 [ ]
(1 + q) 3 3r r3
= id
p (p ) 1 + , (161)
6 2 (1 + q) 2 (1 + q)3 3

for < r < (1 + q). The AOV pair potential vanishes for non-overlap distances, r >
(1 + q), of two excluded-volume spheres. Fig. 19 shows uAOV (r) for various (reduced)
osmotic polymer pressures, red = id 3
p , at a fixed q = 0.1 (left), and for various polymer-
colloid size ratios at fixed red = 50 (right).
The hard-core potential associated with is denoted by uHC (r). It prevents the colloid spheres
from overlapping. The volume term is seen to be linear in the colloid concentration. Therefore,
it has no effect on the phase behavior of the colloids in the mixture. Since p is externally con-
trolled, the AOV potential can be considered as state-independent. The range of the attractive
AOV potential in this simplifying model is equal to p , independent of the polymer concentra-
tion, but its depth can be controlled by changing the density and thus the osmotic pressure of
the polymer solution in the reservoir.

For interacting (self-avoiding) polymers, which deform at non-zero concentrations, the range
and depth of the depletion potential is smaller than for the ideal polymers considered here. In
addition, there is a slight repulsive barrier in front of the attractive well, caused by the corre-
lations between the interacting polymers. At high enough polymer concentration, depletion-
induced attraction can cause the suspension to phase-separate into colloid-poor and colloid-rich
phases. Depending on the range of attraction, the colloidal particles in the colloid-rich phase
can be either in liquid-like arrangements for longer-ranged attractions, or in a crystalline state
(for very short-ranged attractions) [52]. For very short-range attractions of range typically less
than 5 % of the colloid diameter, a first-order iso-structural solid-solid transition is predicted in
simulations [54, 55].

To make contact with subsection 2.2, note that the potential of mean force associated with the
rdf of large particles, wLL (r) = kB T ln gLL (r), contains in general many-body contributions.

2
The value q = 2/ 3 1 characterizes a small polymer sphere that fits exactly into the space in between three
colloid spheres touching each other in an equilateral configuration (see [53]).
Theories of Fluid Microstructures B2.43

10 0.1

30 q = 0.1 red = 50
0.2

red = 50 q = 0.3

Fig. 19: AOV depletion potential in units of kB T for values of reduced osmotic polymer pres-
sure, red = id 3
p , as indicated (left), and polymer-colloid size ratios, q = p /, as indicated
(right).

Different from ue2 (r), its definition involves also an average over the positions of all large parti-
cles with the exception of two kept fixed at the distance r. Therefore, it reduces to u e2 (r) only
in the zero- concentration limit of the large particles.

4.2 Effective electrostatic macroion potential


We wish to calculate the effective electrostatic interaction between two charge-stabilized col-
loidal spheres of radius a = /2 and charge Ze, immersed in a supporting electrolyte solution
(cf. Fig. 20). For this purpose, we model the colloidal particles, the counterions of charge zc e
dissociated from the colloidal particle surfaces, and the added salt ions as uniformly charged
hard spheres dispersed in a solvent described as a structure-less, uniform continuum of dielec-
tric constant . This is the so-called Primitive Model which is frequently used as a model for
electrolytes and charge-stabilized colloidal dispersions. It is primitive in the sense that the struc-
ture of the solvent and particle surfaces, and polarization effects are completely disregarded.

Fig. 20: Primitive Model of charged colloidal hard spheres (macroions) of radius a and charge
Ze and point-like counterions and coions (microions).

The surface-released counterions and added salt ions (i.e., the microions) are much smaller than
the colloidal particles (macroions). Therefore we can assume all microions to be point-like, i.e.,
of zero diameter. For simplicity, and although we want to study a system with added salt, let
us restrict for the time being to a two-component system of macroions and point-like, surface-
B2.44 G. Nagele

released counterions. Overall charge neutrality requires then that

1 Z + c zc = 0 , (162)

where, typically, |zc | = 1 and |Z| |zc |. According to Eq. (146), there are thus three coupled
Ornstein-Zernike equations

h (q) = c (q) + 1 c1 (q) h1 (q) + 2 c2 (q) h2 (q) (163)

for components 1 and 2, i.e., for the colloidal macroions of density 1 , and for the point-like
counterions of density 2 = c . The subscript c is used here to label the counterions, and not the
colloidal particles. The three OZ equations include information on macroion-macroion (11),
macroion-counterion (12) and counterion-counterion (22) pair correlations. However, we are
only interested here in the macroion-macroion correlations (11). To eliminate explicit reference
to the counterions, we define an effective one-component Ornstein-Zernike equation for the
macroions alone as
h11 (q) = cef f (q) + 1 cef f (q) h11 (q) . (164)
By demanding h11 to be the same as in the original two-component OZ equations, the effective
direct correlation function, cef f (q), is determined by

c c12 (q)2
cef f (q) = c11 (q) + . (165)
1 c c22 (q)
It is thus related to all three partial direct correlation functions. Eq. (164) is formally identical
with the OZ equation of a genuinely one-component system. It describes the microstructure
of particles 1 (macroions) immersed in a bath of particles of component 2 (counterions). The
counterions do not appear explicitly in Eq. (164). Their effects are hidden in cef f (q).

Suppose cef f (q) and its Fourier-inverse, cef f (r), would be already known from Eq. (165). We
can associate an effective pair potential, uef f (r), describing the interaction of two counterion-
dressed macroions from noting that

cef f (r) = uef f (r) for r , (166)

valid at large pair separation. In fact, according to a general theorem due to Henderson [56], at
a given state point (1 , T ), there is a one-to-one correspondence,

g11 (r; 1 ) uef f (r; 1 ) , (167)

between a given pair distribution function g11 (r) and an effective one-component system with
exactly pairwise interactions described by a pair potential uef f (r; 1 ), that exactly reproduces
g11 (r) irrespective of the underlying many-body interactions. In general, however, the knowl-
edge of uef f (r) alone is insufficient to give higher-order distribution functions such as g111 (r, r ),
(3)

and to obtain the volume term. The potential appearing in the asymptotic relation in Eq. (166)
is identical to the uniquely determined effective pair potential associated with g11 (r). The form
of uef f (r) may be different at different state points (densities) and must be re-established in
each case. For a given pair distribution function and state point, the associated effective pair po-
tential can be determined by an inversion procedure using, e.g., an appropriate one-component
OZ closure relation where u(r) is now the searched-for quantity, or using computer simulations
Theories of Fluid Microstructures B2.45

or density functional theory schemes [57, 58].


ef f
At non-zero 1 > 0, uef f (r; 1 ) includes in general higher-order contributions to U11 in an
averaged way, and agrees therefore neither with u e2 (r), nor with the potential of mean force
associated with g11 (r). In general,

lim uef f (r; 1 ) = lim w11 (r; 1 ) . (168)


1 0 1 0

However, salt-free macroion suspensions are an exemption to the zero-density rule in Eq. (168)
since, in three dimensions and due to Eq. (162), the counterions cease to screen the Coulomb
interaction between a pair of macroions when 1 0. A salt-free macroion suspension is thus
not weakly coupled even in the infinite dilution limit (see [47] for a lucid discussion of this
subtle point).
We can easily generalize Eqs. (163) and (165) to a more than a two-component ionic system by
including the effect of point-like added salt ions of number density s and charge number zs .
For an overall electro-neutral system,

s zs = 0 , (169)
s

where the sum extends over all components of salt ions.

Our task is to calculate first all partial direct correlation functions c of the mixture, from solv-
ing the coupled Ornstein-Zernike equations with appropriate closure relations. In a next step,
cef f (r) can be determined using the many-component extension of Eq. (165), with uef f (r)
deduced from its long-distance behavior. To determine cef f (r), we will use here the many-
component linear MSA closure because of its analytical simplicity.

The direct correlation functions of point-like salt and counterions (i.e., components > 1) are
approximated in the MSA by

c (r) = u (r)/kB T , r>0 (170)

where
z z
u (r) = LB , r>0, (171)
r
and z {zc , zs }. Fourier transformation leads to
z z
c (q) = 4LB . (172)
q2

Notice in this context that in the limit of point-like ions, or for ionic systems at infinite dilution,
the MSA reduces to what is known in electrolyte theory as the Debye-Huckel (DH) approxi-
mation. When dealing with direct correlation functions of charged particles, which have short-
range (e.g., excluded volume) interaction contributions aside from the long-range Coulomb
interactions, it is helpful to partition c (r) into a short-range part, cs (r), and a long-range
Coulomb part according to
z z
c (r) = cs (r) LB , (173)
r
B2.46 G. Nagele

since it holds quite generally that


z z
c (r) u (r) = LB r. (174)
r
This partitioning into long-range and short-range parts in combination with the MSA closure
for the point-like microions leads to
[ ]2
1
cef f (q) = cs11 (q) + [cs1 (q)]2 4LB Z + z cs1 (q) (175)
=c,s =c,s
2 + q2

where we have introduced the Debye-Huckel screening length, 1 , with


[ ]

2 = 4LB c zc2 + s zs2 . (176)
s

The screening parameter is determined by the concentrations and charges of all microion com-
ponents in the system.
1
The large-r asymptotic behavior of cef f (r) is determined by the factor [2 + q 2 ] in Eq. (175),
since cs1 (q) and cs11 (q) are Fourier transforms of short-range functions. Fourier inversion leads
thus to the intermediate result

1 iqr er
cef f (r) = dre c ef f (q) g r (177)
(2)3 r

with a yet undetermined interaction strength g. The magnitude of g is determined by the pre-
1
factor of [2 + q 2 ] which includes the so far unspecified colloid-microion direct correlations
cs1 (q). Thus, the effective macroion pair potential is asymptotically of the Yukawa-type, with
a screening parameter determined by the Debye-Huckel relation in Eq. (176). The microions
distribute themselves around the macroions to screen the colloid-colloid Coulomb repulsion.
The only approximation used to obtain the asymptotic result in Eq. (177) has been the DH
approximation for the microionic direct correlations.

To determine g analytically, we apply the MSA closure additionally to the colloid-colloid and
colloid-microion direct correlation functions:
Z2
c11 (r) = u11 (r) = LB , r > = 2a (178)
r
Zz
c1 (r) = u1 (r) = LB , r>a. (179)
r
In MSA, the short-range direct correlation parts are thus zero outside the overlap region, i.e.
cs11 (r) = 0 for r > = 2a and cs1 (r) = 0, for r > a. As a consequence, the MSA predicts the
microion-colloid short-range direct correlation part to be a linear function of the colloid charge
number. This linearity in Z is exactly valid for the true cs1 only in the limit of weak particle
charges. For larger macroion charges, the MSA underestimates the accumulation of counterions
close to the surface of a colloidal macroion.
Theories of Fluid Microstructures B2.47

Using these properties of the MSA closure leads, in the limit 1 0, to the final result
( )2
ea er
uef f (r) = lim cef f (r) = LB Z 2
, r> , (180)
1 0 1 + a r

for the effective macroion pair potential at infinite dilution, as quoted already in Eq. (3). Note
here that is non-zero for a finite salt ion concentration. Within the linearization approximation
of weak particle charges,

g11 (r) = ew11 (r) 1 w11 (r) , (181)

which is consistent with the DH-MSA approximation of direct correlations, uef f (r) can be
identified, for non-zero salt content, with the macroion - macroion potential of mean force
w11 (r). From the mean-spherical approximations made in its derivation, one might expect that
this form of the effective pair potential applies only at long distances r, and for weakly charged
colloidal particles. However, it has been empirically found that the range of applicability of
Eq. (180) can be substantially extended when Z is treated not as the bare macroion charge, but
as an effective or dressed colloid charge smaller than the bare one. In this way one accounts
approximately for the stronger screening by counterions close to strongly charged macroion
surfaces. Methods to determine the effective macroion charge by means of so-called cell-model
and jellium approximations are discussed in [59,60]. For an interesting alternative derivation of
the effective macroion pair potential, and volume energy, of charged colloidal spheres based on
linear response theory, see [61, 62]. This work gives also some justification for the free-volume
correction factor 1/(1 ) in the screening constant appearing in Eq. (4).

5 Summary and Outlook


The aim of my lecture has been to give an introduction to integral equation theories of the liq-
uid state. These theories are based on the Ornstein-Zernike equation with its concept of direct
correlations. The OZ based integral equation schemes provide a versatile framework for pre-
dicting microstructural and thermodynamic properties of simple and colloidal liquids, from the
knowledge of the pair potential. Moreover, I have explained how effective interaction potentials
of colloidal particles can be obtained from averaging out the degrees of freedom of the small
particle components, e.g., polymers or microions.

Due to the approximate nature and, usually, non-perturbative character of integral equation
schemes, one can not make decisive a priori statements on their accuracy. The accuracy of a
specific scheme depends in general on the range and the attractive or repulsive nature of u(r),
the system dimensionality, and on the degree of thermodynamic consistency. The extent of ther-
modynamic inconsistency is particularly influential when critical phenomena are studied. The
existence and location of the spinodal instability line, e.g., depends on fine details of the un-
derlying integral equation scheme. For a more recent discussion of liquid-gas criticality in the
context of the OZ integral equation approximations, see [63]. Partially self-consistent integral
equation schemes such as the RY approximation, in which different routes to the same thermo-
dynamic properties are enforced, are superior to standard integral equation schemes, however,
for the price of a somewhat larger numerical effort. Yet, even this enlarged numerical effort
is usually quite small as compared to computer simulation studies, which makes the integral
B2.48 G. Nagele

equation schemes well suited to study liquid state properties over a broad range of system pa-
rameters.

The importance of OZ integral equation schemes goes well beyond the calculation of g(r) and
S(q) from a given pair potential. In combination with powerful density functional theory meth-
ods [6466] and perturbation schemes, one can study first-order liquid-solid phase transitions,
and calculate structural properties of inhomogeneous fluids in external electric or magnetic
fields, and near confining walls or liquid-gas interfaces (e.g., layering and wetting phenomena).
Furthermore, integral-equation-calculated distribution functions play an important role as static
input to theories dealing with the dynamics of fluids.

Integral equation approaches form also the basis of inversion schemes to deduce information on
the (effective) pair potential from an experimentally determined g(r) or S(q). From knowing
g(r), c(r) can be determined using the OZ equation. The pair potential of the effective one-
component system (cf., Eq. 167) follows then directly, but approximately, from using a closure
relation such as the HNC closure. I should also add that the OZ equation for two-particle cor-
relation functions, that has been discussed in this lecture, can be generalized to higher-order
correlation functions using functional calculus methods [67]. The OZ equations for higher-
order correlation functions, in turn, are the basis for extended integral equation schemes that
allow to calculate, e.g., the triplet distribution function g (3) (r, r ) [68].

My lecture has been restricted to fluids of spherical particles with spherically symmetric pair
interactions. The integral equation schemes and their closure relations have been broadened
in the past to deal also with fluids of non-spherical rigid and flexible particles, such as more
complex molecules, rod-shaped viruses and polymers. For an example, the so-called refer-
ence interaction-site model (RISM) method has been used to calculate the site-site distribution
function of rigid molecules, whose interactions are modelled by an interaction-site potential.
The molecule is hereby represented by a discrete set of interaction sites located at the places
of the atomic nuclei [1]. The site-site distribution of non-rigid molecules like polymer chains
and polyelectrolytes has been successfully determined on the basis of the polymer reference
interaction-site model (PRISM) [69]. Furthermore, the microstructure of liquids adsorbed in a
porous medium, such as in a gel or in an arrested particle matrix, can be approximately predicted
using so-called replica Ornstein-Zernike equations [70, 71].
Theories of Fluid Microstructures B2.49

References
[1] J.-P. Hansen and I.R. McDonald, Theory of Simple Liquids (2nd Edition, Academic Press,
London, 1986)

[2] P.A. Egelstaff, An Introduction to the Liquid State (2nd Edition, Clarendon Press, Oxford,
1992)

[3] G. Nagele, Phys. Rep. 272, 215-372 (1996)

[4] W.B. Russel, D.A. Saville and W.R. Schowalter, Colloidal Dispersions (Cambridge Uni-
versity Press, Cambridge, 1989)

[5] J. Mahanty and B.W. Ninham, Dispersion Forces (Academic Press, New York, 1976)

[6] V.I. Kalikmanov, Statistical Theory of Fluids (Springer, Berlin, 2001)

[7] G. Nagele, The Physics of Colloidal Soft Matter, Lecture Notes 14, Institute of Funda-
mental Technological Research (Polish Academy of Sciences publication, Warsaw, 2004)

[8] J.-L. Barrat and J.-P. Hansen, Basic Concepts for Simple and Complex Fluids (Elsevier,
Amsterdam, 2003)

[9] T.L. Hill, Statistical Mechanics (McGraw-Hill, New York, 1956)

[10] P. Attard, Phys. Rev. A 45, 3659 (1992)

[11] A.A. Louis, Phil. Trans. R. Soc. Lond. A 359, 939 (2001)

[12] A.A. Louis, J. Phys.: Condens. Matter 14, 9187 (2002)

[13] M. Dijkstra, R. van Roij and R. Evans, J. Chem. Phys. 113, 4799 (2000)

[14] J. Dobnikar, R. Castaneda-Priego, H.H. von Grunberg and E. Trizac, New J. of Phys. 8,
277 (2006)

[15] E. Trizac, L. Belloni, J. Dobnikar, H.H. von Grunberg and R. Castaneda-Priego, Phys.
Rev. E 75, 011401 (2007)

[16] C.F. Tejero and M. Baus, J. Chem. Phys. 118, 892 (2003)

[17] F.H. Stillinger, H. Sakai and S. Torquato, J. Chem. Phys. 117, 288 (2002)

[18] H.N.W. Lekkerkerker, W.C.K. Poon, P.N. Pusey, A. Stroobants and P.W. Warren, Euro-
phys. Lett. 20, 559 (1992)

[19] J.G. Kirkwood and F.P. Buff, J. Chem. Phys. 19, 774 (1951)

[20] P.M. Chaikin and T.C. Lubensky, Principles of Condensed Matter Physics (Cambrigde
University Press, Cambrigde, 1995)

[21] J.J. Binney, N.J. Dowrick, A.J. Fisher and M.E.J. Newman, The Theory of Critical Phe-
nomena (Oxford University Press, New York, 1993)
B2.50 G. Nagele

[22] H.E. Stanley, Introduction to Phase Transitions and Critical Phenomena (Oxford Univer-
sity Press, New York, 1971)

[23] I.K. Snook and J.B. Hayter, Langmuir 8, 2280 (1992)

[24] G. Nagele, O. Kellerbauer, R. Krause and R. Klein, Phys. Rev. E 47, 2562 (1993)

[25] P.-G. de Gennes, Scaling Concepts in Polymer Physics (Cornell University Press, Ithaca,
1979)

[26] C.N. Likos, Phys. Rep. 348, 267 (2001)

[27] M.S. Wertheim, J. Math. Phys. 5, 643 (1964)

[28] A.J. Banchio and G. Nagele, unpublished

[29] L. Verlet and J.J. Weis, Phys. Rev. A 5, 939 (1972)

[30] W.R. Smith, D.J. Henderson, P.J. Leonard, J.A. Barker and E.W. Grundke, Mol. Phys.
106, 3 (2008)

[31] A.J. Banchio, G. Nagele and J. Bergenholtz, J. Chem. Phys. 113, 3381 (2000)

[32] J.P. Hansen and L. Verlet, Phys. Rev. 184, 151 (1969)

[33] H. Lowen, T. Palberg and R. G. Simon, Phys. Rev. Lett. 70, 1557 (1993)

[34] R. Pesche, M. Kollmann and G. Nagele, J. Chem. Phys. 114, 8701 (2001)

[35] A. Trokhymchuk, I. Nezbeda, J. Jirsak and D. Henderson, J. Chem. Phys. 123, 024501
(2005); Erratum, J. Chem. Phys. 124, 149902 (2006)

[36] A.D. Bruce, Phys. Rev. E 61, 906 (2000).

[37] W.W. Wood, J. Chem. Phys. 20, 1334 (1952).

[38] K.R. Hall, J. Chem. Phys. 57, 2252 (1971).

[39] J.M. Kincaid and J.J. Weis, Mol. Phys. 34, 931 (1977).

[40] C. Song, P. Wang and H. A. Makse, Nature 453, 629 (2008).

[41] F.J. Rogers and D.A. Young, Phys. Rev. A 30, 999 (1984)

[42] S. Asakura and F. Oosawa, J. Chem. Phys. 22, 1255 (1954)

[43] S. Asakura and F. Oosawa, J. Polym. Sci. 33, 183 (1958)

[44] A. Vrij, Pure Appl. Chem. 48, 471 (1976)

[45] H. de Hek and A. Vrij, J. Colloid Interface Sci. 84, 409 (1981)

[46] M. Dijkstra, J.M. Brader and R. Evans, J. Phys.: Condens. Matter 11, 10079 (1999)

[47] L. Belloni, J. Phys.: Condens. Matter 12, R549 (2000)


Theories of Fluid Microstructures B2.51

[48] B. Zoetekouw and R. van Roij, Phys. Rev. E 73, 021403 (2006)

[49] D.Y.C. Chan, Phys. Rev. E 63, 061806 (2001)

[50] G. Ruiz, J.A. Anta and C.F. Tejero, J. Phys.: Condens. Matter 15, S3537 (2003)

[51] H.N.W. Lekkerkerker and G.J. Vroege, in Fundamental Problems in Statistical Mechan-
ics, Vol. VIII, edited by H. van Beijeren and M.H. Ernst (Elsevier Science B.V., Amster-
dam, 1994)

[52] A. Vrij and R. Tuinier, in Fundamentals of Interface and Colloid Science, Vol. IV, edited
by J. Lyklema (Elsevier, Amsterdam, 2005)

[53] A.P. Gast, C.K. Hall and W.B. Russel, J. Colloid Interface Sci. 96, 251 (1983)

[54] P.G. Bolhuis and D. Frenkel, Phys. Rev. Lett. 72, 221 (1994)

[55] D. Frenkel in Soft and Fragile Matter: Nonequilibrium Dynamics, Metastability and Flow,
Proceedings of the 53d Scottish Universities Summer School in Physics, St. Andrews,
1999, edited by M.E. Cates and M.R. Evans (SUSSP Publications, Edinburgh, 2000)

[56] R.L. Henderson, Phys. Lett. A 49, 197 (1974)

[57] N.G. Almarza and E. Lomba, Phys. Rev. E 68, 011202 (2003)

[58] C.P. Royall, A.A. Louis and H. Tanaka, J. Chem. Phys. 127, 044507 (2007)

[59] E. Trizac, L. Bocquet, M. Aubouy and H.H. von Grunberg, Langmuir 19, 4027 (2003)

[60] S. Pianegonda, E. Trizac and Y. Levin, J. Chem. Phys. 126, 014702 (2007)

[61] A.R. Denton, Phys. Rev. E 62, 3855 (2000)

[62] B. Lu and A.R. Denton, Phys. Rev. E 75, 061403 (2007)

[63] See articles in New Approaches to Problems in Liquid State Theory, Eds.: C. Caccamo,
J.P. Hansen and G. Stell, (Kluwer, Dordrecht, 1998)

[64] H. Lowen, Coarse-graining Colloids and Polymers, 33. IFF-Ferienkurs 2002, Schriften
des Forschungszentrums Julich, Vol. 10 (2002)

[65] J. Wu, AIChe Journal 52, 1169 (2006)

[66] J. Wu and Z. Li, Annu. Rev. Phys. Chem. 58, 85 (2007)

[67] L.L. Lee, J. Chem. Phys. 60, 1197 (1974)

[68] J.L. Barrat, J.P. Hansen and G. Pastore, Mol. Phys. 63, 747 (1988)

[69] T. Hofmann, R.G. Winkler and P. Reineker, J. Chem. Phys. 114, 10181 (2001)

[70] J.A. Given and G. Stell, J. Chem. Phys. 97, 4573 (1992)

[71] B. Hribar, V. Vlachy and O. Pizio, Mol. Phys. 100, 3093 (2002)

S-ar putea să vă placă și