You are on page 1of 100

- 1-

PCSURF01. DOC BUTT PC UNI SI EGEN 23. 10. 2000


The physics and chemistry of interfaces
1 INTRODUCTION
An interface is the area which separates two phases from each other. A
surface is a special case. It separates a liquid from a gas or a solid from a gas
or liquid.
Often interfaces and colloids are discussed together. Colloids are disperse
systems, in which the particles have dimensions varying from 1 nm to 1 m.
The word "colloid" comes from the Greek word that stands for glue and has
been used the first time in 1861 by Graham.
1nm-1m
Dispersion
medium
Disperse
phase

How are topics like "interfaces" and "colloids" related? Colloids have an
enormous specific surface. More precisely: their interface-to-volume relation is
so large, that their behavior is determined by their surface properties. Gravity is
negligible in the majority of cases.
Example: This becomes clear
observing the following image
taken by a scanning electronic
microscope (SEM). The image
shows aggregates of quartz
particles ( 0.9 m). These
particles were blown as dust into a
chamber filled with gas. While
sinking down they formed fractal
aggregates. On the bottom they
were collected. These aggregates
are stable for weeks and months.
Gravity, which rules the
macroscopic world, is not able to
bend down the particle chains.
Surface forces are much stronger.
Why is there an interest in
interfaces and colloids? First, for a better understanding of natural processes.
Examples:
Biology: Surface tension of water allows to form lipid membranes. This
leads to "clustering", and subsequently to any higher form of life.
Geological sciences: Swelling of clay upon the addion of water.
- 2-
PCSURF01. DOC BUTT PC UNI SI EGEN 23. 10. 2000
Meteorology: Formation of clouds and rain due to nucleation of water
around small dust particles.
Food, like butter and milk, are colloids.
Second, there are many technological applications. Examples:
Flotation, for the enrichment of ore or for the bleaching of scrap paper.
Washing processes.
Production of new materials, e.g. latex-films.
Coatings, to paint and protect a large number of surfaces.

Introductory books on this topic are refs. 1,2,3.




1
A.W. Adamson, A.P. Gast, Physical Chemistry of Surfaces, John Wiley & Sons, New York, 1997.
2
G. Brezesinski, H.-J. Mgel, Grenzflchen und Kolliode, Spektrum Akademischer Verlag, 1993.
3
R.J. Hunter, Foundations of Colloid Science I+II, Clarendon Press Oxford, 1995.
- 1-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
2 THE SURFACE OF LIQUIDS
2.1 Microscopic representation of the liquid surface
A surface is not an infinitesimal sharp boundary in the direction of its normal, but it has
a certain thickness.
For example, if we consider the density along the height z, one can observe that within the
space of a few molecules the density decreases to that of the gas.
The density is only one criterion to define the thickness of an interface. Another possible
parameter is the orientation of the molecules. For example, water molecules at an interface all
have a preferential orientation. This orientation fades when the distance from the surface
increases. After 1-2 nm the molecules are again statistically oriented.
Moreover, the thickness of the interface depends on the selected parameter. Which thickness
has to be assumed for a phase-interface depends on the examined problem. In case of doubt,
one should always choose a large value for the thickness.
The surface of a liquid is a very turbulent place.
Example: Water at 25C has a vapour pressure p of 3168 Pa. Following the kinetic
gas theory the number of water vapour molecules hitting a surface area A per second
is
T mk
pA
B
2

With a molecular mass m of kg mol kgmol
26 1 23 1
10 3 10 02 , 6 018 , 0

= , 10
7
water
molecules per second hit a surface of 10
2
. If there is an equilibrium, the same
number of molecules leave the liquid phase. 10
2
is approximately the area covered
by a water molecule in liquid water. That means that the average lifetime of a water
molecule on the surface is in the order of 0.1 s.
2.2 The surface tension
The following experiment helps us to define the surface tension: A liquid film is
spanned over a frame, which has a mobile slider. The film is relatively thick, so that
the distance between the two surfaces is large enough to avoid overlapping of the
front and back interfacial regions. Practically, the film should have a minimum
thickness of 0.1 to 1 m. In the absence of gravity, this can easily be achieved.

Liquid
phase
Interfacial region
Gaseous
phase
z
- 2-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
Liquid
film
dA b dx = 2
dx
b

If the slider is moved by an infinitesimal distance dx in order to increase the area of the liquid, a
certain work needs to be done. This work dW is proportional to the increase of the surface dA.
This is equal to two times bdx, because the film has a front and a back side. The proportionality
factor is the surface tension
dW dA =

The equation is an empirical law and a definition at the same time. The law states that the work
is proportional to the surface variation. The proportionality constant depends only on the
composition of the liquid (and in a smaller part of the gas composition), and is totally
independent from the area. The definition says that the proportionality constant is called
"surface tension".
The surface tension can as well be defined by the force F that is required to hold the
slider in place and to balance the surface tensional force:
F b = 2
Both forms of the law are equivalent, provided that the process is reversible. Then one
can write
b
dx
dW
F 2 = =
The force is directed to the left while x increases to the right. Therefore we have the negative
sign.
The term "surface tension" is tied to the conception that
the surface stays under a tension. In the same way as
for a rubber balloon, a force is required to increase the
surface area against this tension.
Molecular level interpretation:
For molecules it is energetically convenient to be surrounded
by other molecules. If this would not be the case, the
condensed phase didnt exist at all. There would be only the
gaseous phase. Molecules on the surface are only partially
surrounded by other molecules. To say it with other words: the
number of adjacent molecules is smaller than in the volume. It
is energetically unfavourable. In order to bring a molecule from
the bulk liquid to the surface, one thus must perform work.
With this view can be interpreted as the energy, which is
necessary to create a certain surface.

Substance
Argon Heptane Ethanol Aceton Acetic acid Phenol Water
Temperature 90 K 20C 20C 22C 22C 22C 20C
/ mJ m
-2
11.9 20.1 22.4 23.3 27.5 42.2 72.8
Surface tensions of some liquids.
Liquid
Gas
- 3-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
Example: If a water film is formed on a frame with a slider length of 1 cm, then the film
pulls on the slider with a force of
N Jm m
3 2
10 44 , 1 072 , 0 01 , 0 2

=
That corresponds to a weight of 0,15 g.
2.3 The equation of Young
a
and Laplace
b

If a liquid surface is curved, there is in equilibrium a pressure difference across it.
Intuitive reason: We consider a circular window of the surface having the radius d
(illustration next page). The surface tension tries to pull that surface as small as
possible, which results in a plane. In order to curve the surface, the pressure on one
side must be larger than on the other side. The Young-Laplace equation sets a
relation between the surface tension and this pressure difference p:
p
R R
= +

1 1
1 2

R
1
and R
2
are the two radii of curvature. They must be measured in planes which are
perpendicular to each other. Otherwise their orientation is arbitrary. A statement of
differential geometry says that the value 1 1
1 2
R R + for an arbitrary surface does not
depend on the orientation.
Example: How large is the pressure in a bubble with diameters of 2 mm and a bubble
of 20 nm diameter in pure water compared with the pressure outside?
R R R = =
2 1
, therefore R p 2 = and
R = 1 mm = =

m m
J
p
3 2
10
2
072 , 0 144 Pa
R = 10 nm = =

m m
J
p
8 2
10
2
072 , 0 1.4410
7
Pa = 144 bar
The pressure inside the bubbles is therefore 144 Pa and 1.4410
7
Pa, respectively,
higher then the outside pressure.
In many cases one has to deal with a rotational symmetric structure. Assuming that
the axis of symmetry is identical to the y axis of an orthogonal cartesian coordinate
system, then it is usual to put a bending radius in the plane of the xy coordinate. This
radius is given by
( )
3
2 1
' 1
' ' 1
y
y
R
+
= ,
where y and y are the first and second derivatives with respect to x. The plane for the
second bending radius is perpendicular to the xy plane. It is
2
2
' 1
' 1
y x
y
R
+
=

a
Thomas Young, 1773-1829.
b
Pierre-Simon Laplace, 1749-1827.
- 4-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
With the help of the Young-Laplace equation it is possible to calculate the form of the
surfaces of liquids in equilibrium. In this case the pressure difference on every point of
the surface is the same. Otherwise there would be a flow of liquid to regions of low
pressure. The equation says also that, in the absence of external fields (e.g. gravity),
the surface of liquids has the same curvature at all places.
Derivation of Young-Laplace equation
R
1
R
1
R
2
R
2
A B
C
D d
dl

X

First, we draw a line around point X which is characterised by the fact that all points on that line
have it the same distance d from X. On a plane this would be a circle. On this line we take two
cuts that are perpendicular to each other (AXB and CXD). Consider in B a small segment on the
line with length . The surface tension pulls with a force . Therefore:
Vertical force on segment: sin dl
For small surface areas (and small ) this becomes

1
R d dl
The sum of the four vertical components at points A, B, C, and D is

+ =

+
2 1 2 1
1 1
2
2 2
R R
d dl
R
d
R
d
dl
This expression is independent of the orientation of AB and CD. Integration over the borderline
(only 90 rotation of the four segments) gives the total vertical force, caused by the surface
tension.
Total vertical force

+
2 1
2
1 1
R R
d
This force that pulls downward must be compensated by an equal force in the
opposite direction. The upward force is produced by a pressure p in the bottom part.
A comparison of the forces gives

+ =

+ =
2 1 2 1
2 2
1 1 1 1
R R
p
R R
d d p
These considerations are valid for any small part of the liquid surface. Since the part is
arbitrary the Young-Laplace equation must be valid everywhere.
- 5-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
A difficulty is to determine the sign of the bending radii and the sign of the pressure
difference. The pressure on the concave side of the interface is larger. Two examples:
Liquid
Gas
A
R
1
R
2
B

Both inside a bubble and inside a drop the pressure is larger than outside. Thus we cannot say
that the pressure inside the liquid is always bigger. The pressure difference over the surface of
a drop that hangs between two cylinders can be zero if 1/R
1
=-1/R
2
, even if the surface is bent.
Example: When connecting two soap bubbles with a tube, the smaller shrinks and the
bigger expands.
The form of a bubble or of a drop is determined by the Laplace-equation. In big
bubbles one has to consider also the hydrostatic pressure:
gh
R R
p +

+ =
2 1
1 1

Dehydration of a polyeder foam. On the contact point of the lamellae a channel is
formed, and it is called plateau's border. Due to the small bending radius in this place,
in the channel there is a depression. This causes liquid to be aspired from the
lamellae, and contributes to the dehydration of the lamellae.
2.4 Methods to measure the surface tension
There are several techniques to
measure the surface tension of
liquids. Only few important ones
are described. Before doing so
we need to introduce the so
called contact angle . Now
we only need to know what it is.
Latter contact angle pheno-
mena are discussed in detail.
For a wetting surface we have
=0.
Liquid
Contact angle
Three phase contact
Solid
- 6-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
One possibility is to measure optically the contour of a sitting or pendant drop. The
measured contour is then compared with a contour calculated with the Young-Laplace
equation.
The drop must be large enough so that gravitation plays a significant role. Why?
The same method is applied with a pendant or sessile bubble. Using a bubble
ensures that the vapour pressure is 100%, a requirement for doing experiments in
thermodynamic equilibrium. Often problems caused by contamination are reduced.
Maximum-bubble-pressure method. From the value of the
pressure which is necessary to push a bubble out of a
capillary against the Laplace pressure, it is possible to find
the surface tension from 2 p r
K
= .
A capillary tube, with internal radius r
K
, is immersed into the liquid.
The surface of the capillary tube must be wetting and the liquid
spreads on the tube material. A gas is pressed through the tube,
so that a bubble is formed on its end. If the pressure in the bubble
increases, the bubble is pushed out of the capillary tube more and
more. In that way the bending radius of the bubble decreases
according to the Young-Laplace equation. The maximum pressure
is achieved when the bubble forms a half-sphere with a radius
r
B
=r
K
. If the volume of the bubble is further increased, also the
radius of the bubble would have to become larger. But a larger
radius corresponds to a smaller pressure. The bubble would thus become unstable and detach
from the capillary tube.
Drop-weight method. The liquid is allowed to flow out from
the bottom of a capillary tube. Drops are thus formed, and
they detach when they reach a critical dimension. A drop
falls off when the gravitational force, determined by the
mass M of the drop, is not anymore compensated by the
surface tension

K
r Mg 2 =
It is thus determined by the radius of the capillary (Tate's
law). Here, g is the gravitational constant. If the surface of
the capillary tube is wetting the external radius of the
capillary has to be taken. For completely nonwetting
surfaces (contact angle higher 90) the internal radius determines the drop weight.
In practice, the
equation is only
approximately valid,
and a weight less than
the ideal value is
measured. The reason
becomes evident when
the process of drop
formation is observed
closely.: A thin neck is
formed before the drop
is released. Correction
factors are therefore
used.
A similar technique is the ring-tensiometer, called also Du-Noy tensiometer.
1
The
force necessary to detach a ring from the surface of a liquid is measured. The force
necessary for the detachment is
( ) 2 + r r
i a

p
r
K

rB
not wetting
wetting
- 7-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
Condition is that the ring surface must be completely wetting. A platinum wire is often used, and
it can be annealed (for cleaning) before the measurement. In the
Wilhelmy-plate method a thin plate of glass, gold, platinum or filter paper (it is only
important that the whole surface is wetting) is vertically placed half way into the liquid.
One measures the force which pulls the plate into the liquid. This force is 2l ,where l
is the length of the plate.
The advantage of the Wilhelmy-plate method is that its is simple and no correction
factors are necessary. One has only to take care that the plate do not contaminate in
air.
Question: It is evident why the ring is pulled down in the ring tensiometer: the surface tension at
the meniscus produces a downward force. But what is responsible for this force in the Wilhelmi-
plate method? Close to the three phase contact line the liquid surface is oriented almost
vertically (provided the contact angle is 0). Thus the surface tension can exert a downward
force.
Finally there are dynamic methods to measure the surface tension. For example a
liquid jet is pushed out from a nozzle, which has an elliptic section. The relaxation to a
circular cross-section is observed. An advantage of this method is that it can measure
changes of the surface tension (for example caused by diffusion of matter to the
interface).
2.5 The Kelvin-equation
Every liquid has a certain vapour pressure at a given temperature and composition.
Values for the vapour pressure given in the literature refer to liquids with planar
surfaces. If the surface is curved the vapour pressure changes. The vapour pressure
of a drop is higher than that of a flat surface. In a bubble the vapour pressure is
reduced.
The cause for this change in vapour pressure is the Laplace pressure. The raised Laplace
pressure in a drop causes the molecules to evaporate more easily. In the liquid which surrounds
a bubble the pressure, with respect to the inner part of the bubble, is reduced. This makes it
more difficult for molecules to evaporate.
Quantitatively the change of vapour pressure for curved liquid surfaces is described by
the Kelvin
c
equation:
RT
p
p
V
R R
k
m
= +

ln
0
0 1 2
1 1


p
0
k
is the vapour pressure of the curved, p
0
that of the flat surface. The index "0"
should remind that everything is only valid in thermodynamic equilibrium. V
m
is the
molar volume of the liquid. For a sphere-like volume of radius r, the Kelvin equation
can be simplified:

c
William Thomson, later Lord Kelvin, 1824-1907, Physics professor at the University of Glasgow.
r
i

r
a

l
2l
Du-Noy ring tensiometer Wilhelmy plate
- 8-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
r
V
p
p
RT
m
k
2
ln
0
0
= or p p e
k
V
RTr
m
0 0
2
=


The constant 2V RT
m
is 1,06 nm for Ethanol (=0,0223 N/m, V
m
=58 cm
2
) and 1,08
nm for Water (=0,0728 N/m, V
m
=18 cm
2
) at NTP.
Derivation of the Kelvin equation
We consider the free molar enthalpy of the liquid, since in most practical cases pressure and
temperature are constant. The free molar enthalpy changes when the surface is being curved
because the pressure increases due to the Laplace-pressure.
The increase of the free enthalpy per mole of liquid upon curving (keeping the
temperature
d
constant) amounts to

+ = =

2 1
0
1 1
R R
V dp V G
m
p
m
,
where the molar volume V
m
is considered to remain constant. For a spherical drop in
its vapour we have simply r V G
m
2 = .
The molar free enthalpy of the vapour depends on the vapour pressure p
0
according to
0
0
ln p RT G G + =
For a liquid having a curved surface we have
k
p RT G G
0
0
ln ' + =
The variation of the free molar enthalpy inside the vapour due to curving the interface
is therefore
0
0
ln '
p
p
RT G G G
k
= =
Since liquid and vapour are supposed to be in equilibrium, the two expressions must
be equal.
By applying the Kelvin equation, it is instructive to distinguish two cases:
Drop in its vapour
The vapour pressure in a drop is higher than that of a liquid with a planar surface.
Consequence: Aerosol of drops in gas (fog) is unstable.
Let's assume we have a box filled with many drops in gaseous environment. Some drops are
larger than others. The small drops have a higher vapour pressure than the large drops. Hence,
more liquid evaporates from their surface. This tends to condense onto large drops. Within a
population of drops of different sizes, the bigger drops will grow at the expense of the smaller
ones.
For a given vapour pressure, there is a critical drop size. Every drop bigger than this
size will grow, every smaller drop evaporates. If a vapour is cooled to reach over-
saturation, it cannot condense (because every drop would instantly evaporate again),
except nucleation sites are present. In that way it is possible to explain the existence
of over-saturated vapours.
Bubbles in a liquid

d
Remember: SdT Vdp dG =
- 9-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
From the first relation of the derivation one can see that a negative sign has to be used for a
spherical bubble because of the negative curvature of the surface of the liquid:
r
V
p
p
RT
m
k
2
ln
0
0
=
r is the radius of the bubble. The vapour pressure inside a bubble is therefore
lowered. This explains, why it is possible to overheat liquids.
r / nm p
0
k
/p
0

Drop
p
0
k
/p
0

Bubble
1000 1,001 0,999
100 1,011 0,989
10 1,114 0,898
1 2,950 0,339
Equilibrium vapour pressure over a curved water surface.
2.6 Capillary condensation
An important application of the Kelvin equation is the description of capillary
condensation. This is the condensation of vapour into capillaries or fine pores even
below the vapour pressure of a liquid (over a flat surface).
Capillary condensation is easily illustrated by the model of a conical pore with a totally wetting
surface. Liquid will immediately condense in the tip of the pore. Condensation continues until
the bending radius of the liquid has reached the value given by the Kelvin equation.
The situation is analogous to that of a bubble and we can write
k
m
k
r
V
p
p
RT
2
ln
0
0
=
The vapour pressure of the liquid inside the pore decreases to p
0
k
, with r
k
being the capillary
radius at the point where the meniscus is in equilibrium.
Many surfaces are
not totally wetted,
but they form a
certain contact
angle with the
liquid. The bending
radius increases in
that case. It is not
equal any more to
the capillary
radius, but to
= cos
k
r r .
Example: We have a porous material with pores of all dimensions. It is in water-vapour
at 20C. The humidity is 90%. What is the size of the pores which fill up with water?
nm
K JK
m Jm
RT
V
r
m
k
10
9 , 0 ln 293 31 , 8
10 18 072 , 0 2
9 , 0 ln
2
1
3 6 2
=


=


r
k

=0
r
k

>0
- 10-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
Attention has to be paid on what radius to put into the Laplace equation. Generally,
there is no rotational symmetric geometry. Then
k
r 2 has to be substituted by
2 1
1 1 R R + . In a fissure or crack one radius of curvature is infinitely large. Instead of
k
r 2 there should be r 1 in the equation, with r being the bending radius vertical to
the fissure direction
The capillary condensation has been analysed with various methods, and the validity
of the previous description has been confirmed for several liquids and radii of
curvature down to few nanometers.
2,3,4

An important consequence of the capillary condensation is that liquids are strongly
adsorbed at porous surfaces of solids. Another important consequence is the
existence of the capillary force. Capillary condensation often strengthens the
adhesion of fine particles and determines therefore in many cases the behaviour of
powders. If for instance two spherical particles with radius R
P
are getting into contact,
liquid (usually water) will condense into the gap of the contact zone.
5
The meniscus is
curved. As a consequence, the Laplace-pressure in the liquid is negative, and the
particles attract each other.
A fact that might be surprising at a first glance is that the capillary force is
approximately independent on the radius of curvature of the liquid surface over a wide
range. Why is this so?
For simplicity we assume that the liquid wets the surface of the particles. This is for instance the
case for clays and many other minerals and water. The total radius of curvature of the liquid
surface is
r x r R R
1 1 1 1 1
2 1
= +
In most practical cases we can safely assume that x>>r. The Laplace pressure is therefore
p=/r. It acts upon a cross-sectional area x
2
. Now we use Pythagoras theorem to express x
2

through r:
( ) ( )
2 2
2 2 2 2 2 2 2 2
2 2
2 2
x xr x rR
R r xr x r rR R R r x r R
P
P P P P P
+ =
+ + + = + + + + = +

For the last approximation we assumed x>>2r. From this follows x
2
=2rR
P
. Therefore the capillary
force is
P
R x p F 2
2
= =
R
P

x
r
- 11-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
The force only depends on the radius of the particle and the surface tension of the
liquid. It does not depend on the actual radius of curvature of the liquid surface nor on
the vapour pressure!
Explanation: With decreasing vapour pressure also the radius of curvature, and therefore x,
decreases. At the same time the Laplace pressure increases by the same amount.
Example: a quartz sphere hangs on a second similar sphere. Some water vapour is in
the room. Beyond which radius is the gravity strong enough to separate the two
spheres? Density: = 3000 kgm
-3
.
Weight
3
2 2
5 3
10 23 , 1
3
4
P P
R
s m
kg
g R =
Capillary force
P P
R
s
kg
R
m
J
=
2 2
45 , 0 072 , 0 2
Both are equal for mm m R
P
9 , 1
10 23 , 1
45 , 0
2
5
=

= .
In reality the capillary force is often much smaller than the calculated value.
Explanation: the surfaces are rough and touch only at some points. Capillary
condensation takes place only at this points.
The surface tension of the liquid itself gives a further contribution to the attraction due
to the Laplace pressure. It is x F 2 = . Since x<<R
P
, this contribution is usually small.
A calculation of the distance dependence of the capillary force and of adhesion is in
refs. 6
7
. The capillary force was measured experimentally with the surface force
apparatus.
8

2.7 Nucleation-theory
The Kelvin-equation has important consequences for the condensation of a vapour (or
the formation of bubbles during boiling). The formation of a new phase in the absence
of an external surface is called "homogeneous nucleation". In homogeneous
nucleation first small clusters of molecules are formed. These clusters grow due to
condensation of other molecules (plus due to aggregation with other clusters). Finally
macroscopic drops form. Usually this happens only if the vapour pressure is
significantly above the saturation pressure at a relatively well defined vapour pressure.
Condensation: classical theory of homogeneous Nucleation
The classical theory of homogeneous nucleation was developed around 1920-1930.
9,10

In order to describe nucleation, we consider the change of free enthalpy for the
transition of n moles vapour at a partial pressure p into a drop (n is much smaller than
one). The free enthalpy of the vapour is
p nRT nG G
v
ln
0
+ =
- 12-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
To obtain the free enthalpy we use the free enthalpy of a (hypothetical) vapour, which
would be in equilibrium with liquid drops. Since the drops have a curved surface of
radius r the vapour pressure p
0
k
is higher then that of the a flat liquid surface. The free
enthalpy of the liquid is
k m equilibriu
v l
p nRT nG G G
0
0
ln + = =
It is equal to the virtual free enthalpy of the vapour which is in (hypothetical) thermodynamic
equilibrium with the liquid in the form of drops with radii r. With the change of n moles of vapor at
actual vapour pressure p into liquid drops, the free enthalpy changes by
k
o
k
v l
p
p
nRT p nRT p nRT G G G ln ln ln
0
= = =
In addition, the drop has a certain surface tension which has to be considered:
2
0
4 ln r
p
p
nRT G
k
+ =
In a drop having radius r there are n r V
m
= 4 3
3
moles, where V
m
is the mole volume
of the liquid phase. Therefore
2
0
3
4 ln
3
4
r
p
p
V
RTr
G
k
m

+ =
This is change of free enthalpy upon condensation of a drop with radius r from a
vapour phase with partial pressure p.
Let us analyse the equation more in detail. For
k
p p
0
< the first term is positive and
therefore G always positive. No condensation can occur. For
k
p p
0
> , G increases
with increasing radius, has a maximum at the so-called critical radius r
c
and then
decreases again. At the maximum we have 0 = dr G d , which leads to
G
r
C
max
=
4
3
2


k
m
C
p
p
RT
V
r
0
ln
2

=


That means that the Kelvin equation ( )
C m
k
r V p p RT 2 ln
0
= applies at the
maximum, since 0 = dr G d and the system is formally in equilibrium.
As an example, the following illustration shows a plot of G versus the radius for water
at different vapour pressures. At 20C, the vapour pressure is equal to 2337 Pa. To
produce a double over-saturation, the vapour should initially be formed at 32C - here
the vapour pressure is almost 22337 Pa and then cooled down to 20C.
- 13-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
0 1 2 3 4 5
0
250
500
750
Water

=0,072 Jm
-2
V
m
=18 cm
3
T=20C
p/p
0
=2
p/p
0
=1,7
p/p
0
=1,4
Radius / nm

G/kT

Example: for water at T=0
0
C and p/p
0
=4, the critical radius is 8 . This corresponds
approximately to 70 molecules. There G
max
=1,910
-19
J.
Now we can picture the following situation: There are always a certain number of
clusters. Most of them are very small. Some are a little bit larger. When the actual
partial pressure p becomes higher than the equilibrium vapour pressure p
0
k
, large
clusters occur more frequently. If a cluster exceeds the critical size, thermal
fluctuations will enlarge it even more, until it becomes infinitely large and the liquid
condenses.
The aim of any theory of nucleation is to find a rate I with which clusters of critical size
are formed. This number is proportional to the Boltzmann factor exp( / )
max
G kT . A
complete description of the classical theory of nucleation
8,9
is not possible within this
lecture. The result is:
I Z K e
Zel
G
kT
=

max
, K
n
G
kT
Zel
C
=
1
3

max


K
Zel
is the so-called Zeldovich factor. Z is the rate at which clusters of two
molecules are formed. This corresponds roughly to the frequency of collision.
The classical nucleation theory is the bases to understand condensation.
Unfortunately, predictions do often not agree with experimental results.
11,12
The theory
predicts too low nucleation rates at low temperatures. At high temperatures the
calculated rates are too high. Ref. 13 reviews experimental methods. General
overviews are refs. 14,15,16.
Crystal-growth
Crystals often grow through Ostwald ripening (Ostwald Reifung). In the process large
crystals grow at the expense of small ones. This can also be explained with the Kelvin equation.
The immediately obvious cause is that small crystals have a bigger specific surface area than
large ones. Large crystals are therefore energetically more advantageous.
- 14-
PCSURF02. DOC BUTT PC UNI SI EGEN 30. 10. 2000
2.8 Exercises
Drop-weight method. A plastic box is filled with water to a height h=1 cm. A hole of
radius R=0.1 mm is drilled into the bottom. Does all water run out? The plastic is
supposed to be nonwetting.
Wilhelmy plate method. What is the force on a plate of 1 cm width having a contact
angle of 45in water?


1
L. du Noy, J. Gen. Physiol. 1919, 1, 521.
2
L.R. Fisher, J.N. Israelachvili, J. Colloid Interface Sci. 1981, 80, 528.
3
J.E. Curry, H.K. Christenson, Langmuir 1996, 12, 5729.
4
L.R. Fisher, R.A. Gamble, J. Middlehurst, Nature 1981, 290, 575.
5
N.L. Cross, R.G. Picknett, Trans. Faraday Soc. 1963, 59, 846.
6
A. Fogden, L.R. White, J. Colloid Interface Sci. 1990, 138, 414.
7
A. de Lazzer, M. Dreyer, H.J. Rath, Langmuir 1999, 15, 4551.
8
H.K. Christenson, J. Colloid Interface Sci. 1988, 121, 170.
9
M. Volmer, A. Weber, Zeitschr. f. phys. Chem 1926, 119, 277.
10
R. Becker, W. Dring, Annalen der Physik 1935, 24, 719.
11
V. Vohra, R.H. Heist, J. Chem. Phys. 1996, 104, 382.
12
A. Bertelsmann, R.H. Heist, Atmospheric Research 1998, 46, 195.
13
R.H. Heist, H. He, J. Phys. Chem. Ref. Data 1994, 23, 781.
14
A. Laaksonen, V. Talanquer, D.W. Oxtoby, Annu. Rev. Phys. Chem. 1995, 46, 489.
15
D.W. Oxtoby, J. Phys.: Condens. Matter 1992, 4, 7627.
16
D.W. Oxtoby, Accounts of Chemical Research 1998, 31, 91.
- 1-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
3 THERMODYNAMICS OF INTERFACES
3.1 The surface excess
The presence of an interface influences generally all thermodynamic parameters of a
system. To consider the thermodynamics of a system with an interface, we divide that
system into three parts: The two phases with volumes V

and V

, and an infinitesimal
thin interface . Since the interface is ideally thin, the total volume is
V V V = +


All other extensive quantities can be written as a sum of three components. Examples
are the internal energy, U, the number of particles of the ith material, n
i
, and the
entropy, S:
U U U U = + +


n n n n
i i i i
= + +


S S S S = + +


The contributions of the two phases and of the interface are derived as follows. Let u


and u

be the internal energies per volume of the two phases. The contribution of the
volume phase to the total energy of the system is u V u V

+ . The internal energy
of the interface is
U U u V u V

=
Other surface parameters like the surface entropy S

, the Helmholtz free energy F

,
and Gibbs free energy G

can be defined in the same way.


At an interface the molecular constitution changes. The concentration of the i
th

material is, in the two phases, respectively c
i

and c
i

. The additional quantity that is


present in the system due to the interface is
n n c V c V
i i i i

=
n
i
is the number of the total available moles of the material i in the system. With its
help it is possible to define something like a surface concentration, the so called
interfacial excess (Grenzflchenberschukonzentration):
A
n
i
i

=
A is the interfacial area. In the model of the ideal interface there is one problem. At a
real interface the density varies continuously. The position of the infinitesimal thin
interface is arbitrary. Depending on where the interface should be, we have different
interfacial excess concentrations. Also the other extensive quantities, such as the
volumes V

and V

, change.
Example: In order to get an impression of the importance of the position of the ideal
interface, we consider an equimolar mixture of ethanol and water (Page 25, ref. 1). If
the position of the ideal interface is so that
H2O
=0, one finds experimentally that

Ethanol
=9,510
-7
mol/m
2
. If the interface is placed 1 nm outward, then we obtain

Ethanol
=-13010
-7
mol/m
2
.
- 2-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
c
1
Liquid Interfacial region
Substance 1,
Solvent
Substance 2
Interface
c
2
Substance 2
Interface
c
2

The model of the ideal, infinitesimal thin interface was introduced by Gibbs
a
. There are
also other models. Guggenheim, for example, takes the extended interfacial region
including its volume explicitly into account.
2
In most applications it is easier to use the
Gibbs model.
3.2 Fundamental thermodynamic relations
3.2.1 Thermodynamic definition of the interfacial tension
Let us consider a system of two phases, and , which are divided by an interface.
Something with this system changes, e.g. work is done on that system. As a
consequence the state quantities change. If we want to describe the variation of the
extensive state quantities, we have to take the interface into account (except for the
volume).
Let us start the analysis with the internal energy. The variation of the internal energy
of a two-phase system is, according to the first and second principle of
thermodynamics,
dW dn pdV TdS dU
i i
+ + =
W is the work done on the system, that is no volume-work. This term contains the
surface work dA .The sum runs over of the chemical potentials multiplied with the
respective variations of the number of moles runs over all components, that means
over all the substances that are chemically different.
We first analyse the internal energy, and not the enthalpy, the free energy or the free
enthalpy, because the internal energy contains only extensive quantities (S, V, n
i
, A) as

a
Josiah Willard Gibbs, 1839-1903, American mathematician and physicist, Yale College.
- 3-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
variables. The free enthalpy would have for example p, T and n
i
as natural variables
( dW dn SdT Vdp dG
i i
+ + = ). This fact simplifies the following calculation. Now
we split the internal energy:
dA dn TdS dV p dn TdS
dV p dn TdS
dU dU dU dU
i i i i
i i




+ + + + +
+ =
+ + =

The TdS terms stands for the energy change, which is caused by an entropy change,
e.g. a heat flow. The
i i
dn terms consider the energy change caused by a change in
the composition. Both pdV terms correspond to the volume-work of the two phases.
The interface is infinitely thin, thus it can perform no work.
With

dV dV dV dV dV dV = + = and summing up the entropy terms the
equation simplifies:
( ) dA dn dn dn dV p p dV p TdS dU
i i i i i i


+ + + + =
Now we consider the free energy. The free energy of the system is F U TS = . It
follows that
( )
dF d U TS
SdT p dV p p dV dn dn dn dA
i i i i i i
=
= + + + +
( )



In case the temperature and volume are constant ( 0 = dV , 0 = dT ) the first two terms
are zero.
In equilibrium the equation can be simplified even further. In equilibrium the chemical
potentials in the three phases are equal. This I would like to demonstrate. Therefore
we assume that there is no exchange of material with the outside world (dn
i
= 0); we
have a closed system. Thus, the three parameters n
i

, n
i

and n
i

are not independent,


because n n n n
i i i i
= + +

is constant. Only two at a time, as an example n
i

and n
i


can be varied independently. n
i

is then determined by the other two because


dn dn dn
i i i

= . Therefore we can simplify:
( ) dF p p dV dA dn dn
i i i i i i
= + + +

( ) ( )
At equilibrium, with constant volume, temperature and constant amounts of materials,
the free energy is minimal. This means
dF
dn
i
i i

= = 0,
dF
dn
i
i i

= = 0

i i i i
= =
Hence, in equilibrium the chemical potentials are the same everywhere in the system.
This also applies in the case of reversible processes, in which the system is always in
equilibrium. This means that even if one changes the amounts of material in a
reversible way, the chemical potentials in all three phases are the same. With this, the
equation can be further simplified:
( ) dF p p dV dA dn
i i
= + +


With this equation, we get a thermodynamic definition of the surface tension:

F
A
T V V n
i
, , ,

- 4-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
Question: Why is it not possible, using the same argumentation, e.g. dF/dA=0, to
conclude that at equilibrium must be zero? Explanation: A is not an independent
parameter. The surface A and the volume V

are connected. If the volume of a body


changes, in general also its surface changes. V

and A can thus not be varied


independently. The condition dF/dA=0 is perfectly valid in equilibrium. It leads to
( )
dF
dA
F
A
F
V
V
A
p p
V
A
= + = =

0
It is possible to come to the same equation requiring that the system cannot perform
any mechanical work when it is in equilibrium:
( ) ( )
A
V
p p dA dV p p dW
mech

= = + =


0
From that it is easy to derive the Laplace equation. In fact, a statement of differential
geometry says that ( )
1
2 1
1 1

+ = R R A V . If we apply this expression, we
immediately obtain the Laplace equation.
At this point we should note that fixing the bending radii, we define the location of the
interface. A possible choice for the ideal interface is the one that is defined by the
Laplace equation. If the choice for the interface is different, also the value for the
surface tension must be changed accordingly. Otherwise the Laplace equation would
not be valid any more.
All this can be illustrated with the example of a
spherical drop.
3
We can for instance consider
the evaporation or the condensation of liquid
from or to a drop of radius r. There we have
V r


=
4
3
3
, A r = 4
2

=

= V
A V
A
r

4
3 4 2
3 2

If the interface is set at a radius r', then the corresponding value is r' 2. The pressure
difference

p p can in principle be measured. This implies that p p r

= 2
and p p r

= 2 ' ' are both valid at the same time. This is only possible if,
dependent on the radius, one accepts a different interface tension. Therefore I used '
in the second equation. In case of a curved surface, the interfacial tension
depends on the location of the ideal interface! In case of flat surfaces this problem
does not occur. There, the pressure difference is zero and the surface tension is
independent on the location of the ideal interface.
A possible objection could be that the surface tension is measurable and thus the
Laplace equation assigns the location of the ideal interface. But this is not true. The
only thing that can be measured is mechanical work and equivalently - the forces
acting during the process. For curved surfaces it is not possible to divide volume and
surface work. Therefore it is not possible to measure only the surface tension.
For the free enthalpy at equilibrium we have
dG d U TS pV SdT V dp V dp dn dA
i i
= + = + + + + ( )


With help of this equation it is also possible to define the interfacial tension:
Vapour
Drop
- 5-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000



G
A
T p p n
i
, , ,


3.2.2 Surface tension, free surface energy, and free surface
enthalpy
First we want to derive a relation between the surface tension and the surface free
energy F

. Therefore we start from the generally valid expression of the internal


surface energy:
dA dn TdS dU
i i


+ + =


The term pdV

disappears, because the ideal interface has no volume. One integrates


the expression keeping the intrinsic parameters T,
i
and constant.
b
This integration
is allowed because it represents, in principle, a feasible process, e.g. through simply
increasing the size of the system. In other words: We assume that the system is
homogeneous. Result:
A n TS U
i i


+ + =


For the free energy of the surface we use TS U F = and get
+ = + =
i i i i
A
F
n A F



The same expression can be derived for the enthalpy:
+ =
i i
A
G

and G

are equal. Explanation: the difference is pV

. Since V

=0, there's no
distinction. Note: For one-component systems, the surface tension is equal to the free
surface energy, or respectively to the free surface enthalpy per area, since one can
choose =0. Otherwise, this identity is not valid.
3.3 The surface tension of pure liquids
In pure liquids the surface tension is equal to the free surface energy per unit area:

= f
Thereby A F f

= . Since the volume of the interfacial region is zero in the Gibbs
model we also find that the surface tension is equal to the free surface enthalpy:

= g
g

is the free surface enthalpy per unit area A G g



= .
Now we consider how the internal surface energy changes when the surface
increases. This increase can be imagined as represented in the following illustration:

b
In the Gibbs model all volume terms disappear. In the Guggenheim model one must also keep p
constant during the integration.
- 6-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
During the surface increase the volumes of the liquid and of the vapour phase are
supposed not to change: 0 = = = dV dV dV

. Therefore, in the expression for the
internal energy of the system
( ) dA dn dV p p dV p TdS dU

+ + =
the 2
nd
and the 3
rd
term are zero. Moreover, if we assume that the system is closed,
0 = dn , also the 4
th
term disappears.
These conditions, that seem to be rather severe at a first glance, are easily satisfied for pure
liquids. For the majority of closed systems, V

and V

are indeed constant. If the liquid is


uncompressible and if no molecules pass to the vapour phase (or from the vapour phase into
the liquid phase) then from a constant total volume follows a constant V

and therefore a
constant V

. The change in internal energy is therefore


dA TdS dU + =
The internal energy depends on how the entropy varies as the surface increases.
When this happens, only the entropy of the surface increases (because the surface is
the only changing thing). The entropies of the volume phases remain constant. If the
entropy per unit area is A S s

= the entropy of the system changes by dA s dS

= .
Remembering the generally valid equation S T G
A p
=
,
we obtain

S T G
A p
=
,
or per unit area
A p
T g s
,
=

. Finally, with

= g we get
an important expression for the surface entropy:
A p
T
s
,


The surface entropy per unit area is given by the change of the surface tension with
temperature.
Inserting this expression in the equation for the internal energy we get
A p A p
T
T
dA
dU
dA
T
T dA dU
, ,


or
A p
T
T u
,


It is thus possible to determine the internal surface energy and the surface entropy by
simply measuring the dependence of the surface tension from the temperature.
For the majority of liquids, the surface tension decreases with increasing temperature.
This behaviour was already observed by Etvs, Ramsay und Shields at the end of
the last century.
4,5
The entropy on the surface of a liquid is thus increased, that means
the molecules are less ordered than in the volume phase.
or
- 7-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
How does the heat flow during the considered process? In a reversible process TdS is
the heat Q that the system absorbes. The heat absorption is proportional to the
surface increase and we can write qdA Q = . Here, q is the heat per unit area that is
absorbed by the system.
With dA s dS

= and T s =

we get
dA
T
T dA Ts TdS Q qdA

= = = =



or
T
T q

=


This is the heat per unit area absorbed by the system during an isothermal increase of
the surface. Since T is mostly negative the system usually takes up heat.
At this point it is instructive to discuss a possible objection. The heat that has to be provided in
order to have a surface increase is in principle measurable. In the same way , and therefore
/T, are measurable. The surface entropy depends on the choice of the ideal interface. It is
also arbitrary, in a certain way.
The basic relation between the quantities to be measured is q=-T/T. It is independent from
the position of the chosen interface. If the interface is chosen so that =0, then s

=q/T=-/T
applies. If the interface is in another position, the last equation becomes invalid. The specific
choice of the interface is useful because in that case the surface entropy is related to and q in
a simple way.
The following table lists the surface tension, surface entropy, surface enthalpy and
internal surface energy of some liquids (values from ref. 6, page 233 and ref. 7):

Nm
-1

s

=-d/dT
Nm
-1
K
-1

Ts


Nm
-1

h

=u


Nm
-1

water 72,7510
-3
1,6010
-4
46,910
-3
119,710
-3

n-hexane 18,40 1,02 29,9 48,3
n-nonane 22,85 0,94 27,5 50,4
n-dekane 25,35 0,89 26,1 51,5
benzene 28,88 1,30 38,1 67,0
3.4 The Gibb's adsorption isotherm
3.4.1 Derivation
The Gibb's adsorption isotherm is a relation between the surface tension and the
excess interfacial concentrations. One can derive it as follows:
The surface internal energy is
dU TdS dn dA
i i

= + +
Again we can easily integrate since all integrants are constant intrinsic parameters
U TS n A
i i

= + +
- 8-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
Differentiation leads to
dU TdS S dT dn n d dA Ad
i i i i

= + + + + +
Equating with the first expression we get
0 = + + S dT n d Ad
i i


At constant temperature it can be simplified to:
d d
i i
=
This and the previous equation are called "Gibb's adsorption isotherms". In general
isotherms are state functions plotted versus pressure, concentration, etc. at constant
temperature.
Remark: The given equation is only valid for those surfaces whose deformation is
reversible and plastic, i.e. liquid surfaces. In solids changes of the surface are usually
accompanied by elastic processes. In order to consider elastic tensions, there is
another term to add to the equation:
0 = + + +

S dT n d Ad dA
i i plas

( )

plas
is the proportionality factor in front of dA, and it describes the interfacial work in
the case of pure plastic, reversible work. If the actual proportionality factor is
different then the last term becomes effective.
3.4.2 System of two components
The simplest application is a system of two components, e.g. a solvent 1 and a solute 2. In this
case we have
d d d =
1 1 2 2

The ideal interface is conveniently defined in order to have
1
=0. The interface is placed where
it would be also in the pure solvent. Then we get
d d =
2
1
2
( )
,
where the index (1) should remind of the special choice of the interface. The chemical
potential of the solute is described by the equation
0
0
2 2
ln
a
a
RT + =
Here, a is the activity and a
0
is a standard activity, e.g. 1 mol/l. Differentiating with respect to a/a
0

leads to
( )
d RT
d a a
a a
RT
da
a

2
0
0
= =
This can be substituted in the equation above. It follows
T
a RT
a

=
) 1 (
2

This is a very important equation. It implies:
0
) 1 (
2
> : When a substance is enriched at the interface, the surface tension decreases by
adding the substance into the solvent;
- 9-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
0
) 1 (
2
< :When a substance does not like to go into the interface, the surface tension increases
by adding the substance.
If by adding a substance to a solvent one observes
a decrease of the surface tension, the substance is enriched in the interface;
an increase of the surface tension, then the substance "avoids" to go in the interface.
Substances, whose addition to a liquid reduces the surface tension, are labeled as
surface active (grenzflchenaktiv). Moreover, if small amounts of the substance
achieve a drastic reduction of the surface tension, it is called a surfactant (Tensid).
Example: Adding 1 mM SDS (sodium dodecylsulfate, NaSO
2
(CH
2
)
11
CH
3
) in pure water
at 25C leads to a decrease of the surface tension from 73 mJ/m
2
to 68 mJ/m
2
. What
is the surface excess of SDS?
2 2
5
) 0 001 , 0 (
) 073 , 0 068 , 0 (
Mm
J
Mm
J
c a
=


It follows that
2
6
2
10 2 ) 5 (
298 31 , 8
001 , 0
m
mol
Mm
J
J
Mmol
a RT
a

=

= =


Every molecule has a surface of 1.2 nm
2
.
The choice of the ideal interface in the Gibb's adsorption isotherm for a two
component system is, in a certain view, arbitrary. The choice is, however, is
convenient. There are two reasons.
First: on the right side there are physical measurable quantities (a, , T), which are
related in a simple way to the so-defined interfacial excess. Any other choice of the
interface would lead to a more complicated expression.
Second, the choice of the interface is intuitive self-evident, at least for c
1
>>c
2
. One
should, however, keep in mind that different spatial distributions of the solute can lead
to the same
) 1 (
2
. The following figure shows two examples of the same interfacial
excess concentration
) 1 (
2
. In the first case the distribution of the molecules 2
stretches out of the interface, but the concentration is nowhere increased. In the
second case, the concentration of the molecules 2 is actually increased.
More information can be found in the literature under the keyword relative
adsorption, e.g. in ref.1.
Interface
c
x
Substance 1
Substance 2
Interface
c
x
Substance 1
Substance 2
- 10-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
3.4.3 Experimental aspects
Three typical adsorption isotherms are shown in the following figure.
Type I is observed for lyophobic substances, i.e. substances which do not like to stay in
solution. They get enriched in the interface and decrease the surface tension. If water is the
solvent, most organic substances show such a behaviour.
Type II adsorption isotherms are characteristic for lyophilic substances. The most ions in water
show such a behaviour.
Type III adsorption isotherms are typical for amphiphilic substances, i.e. with substances
which have a lyophilic and a lyophobic tail. They readily go into the interfacial region. In many
cases above a certain critical concentration (typically 1-10 mM) micells are formed. This
concentration is called the critical micellar concentration, cmc. Above the cmc the surface
tension does not change significantly any further. Explanation: Addition of the substance does
not contribute to more free substance, but only to the formation of more micells.
In order to describe the influence of a substance on the surface tension, one could
specify the gradient of the adsorption isotherm for c0. In the following table are
listed values for some substances in relation to the solvent (water) at room
temperature.
8

Substance
( ) d dc
mN/m
HCl -0.28
LiCl 1.81
NaCl 1.82
CsCl 1.54
CH
3
COOH -38
Example: putting 1 mM NaCl in water results in a slight surface tension increase of
m
N
m
N
6 3
10 82 , 1 001 , 0 10 82 , 1

= = . Upon addition of 1 mM CH
3
COOH the
surface tension decreases by 3,810
-5
N/m.
Tears of wine
9

Often a typical shape of wine on the rim of a wine glass is
observed. Reason: A wine drop is thicker at the bottom than at
the top because of its weight. Part of the liquid, mainly the
ethanol, evaporates. This rate of evaporation is roughly
proportional to the surface area. Since the drop is thinner at
Surface tension
Solute concentration
II





I
III

0

cmc
- 11-
PCSURF03. DOC BUTT PC UNI SI EGEN 19. 11. 2000
the top the concentration of ethanol decreases faster at the top. This causes the
surface tension at the top to increase. This effect is in general called Marangoni
effect, namely, the carrying of bulk material through motions energized by surface
tension gradients.
Determination of the interfacial excess concentration
Two procedures are generally used to measure the interfacial excess concentration:
The dissolved substance is radioactively labelled. The radioactivity close to the
surface is measured. -emitters (
3
H,
14
C,
35
S) are suitable, because of their short
range. I.e. any recorded radioactivity comes from molecules from the interface, or
close below.
10

One measures the surface tension of a liquid in relation to the concentration of the
added substance. With Gibb's adsorption isotherm the interfacial excess can be
calculated.

1
Defay, R.; Prigogine, I.; Bellemans, A.; Everett, D.H. Surface Tension and Adsorption, Longmans &
Green, London, 1966.
2
Aveyard, R.; Haydon, D.A. An Introduction to the Principles of Surface Chemistry, University Press,
Cambridge, 1973.
3
Kahlweit, M. Grundzge der Physikalischen Chemie VII: Grenzflchenerscheinungen, Steinkopff Verlag
Darmstadt, 1981.
4
Etvs, R.V. Wied. Ann. 1886, 27, 456.
5
Ramsay, W.; Shields, J. Phil. Trans. A 1883, 184, 647.
6
Hunter, R.J. Foundations of Colloid Science, Vol. 1, Oxford Science Publications, 1986.
7
Jasper, J.J.; Kring, E.V. 1955, 59, 1019.
8
Weissenborn, P.K.; Pugh, R.J. Langmuir 1995, 11, 1422.
9
Vuilleumier, R.; Ego, V.; Neltner, L.; Cazabat, A.M. Langmuir 1995, 11, 4117.
10
Nilsson, G. J. Phys. Chem. 1957, 61, 1135.
- 1-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000
4 THIN FILMS ON SURFACES OF LIQUIDS
The adsorption of a solved material at the surface of the liquid can be so strong
that a compact monomolecular layer is formed. Many amphiphilic materials
practically do not dissolve in the underlying liquid. This leads to unsoluable
monolayers. Examples are fatty acids or phospholipides in water. In this case
the surface excess corresponds to the added amount of material divided by
the surface area.
If there is still a significant proportion of the substance dissolved in the liquid
we talk about Gibb's monolayers. In this case is determined from the
reduction of the surface tension.
4.1 Monolayers - Introduction
Unsoluable monolayers are often examined with the help of a film balance,
also called Langmuir trough:
If the material is significantly soluble in the liquid, one uses a PLAWM
a
trough.
In the PLAWM trough a membrane, which is fixed to the barrier, separates the
two compartments. This membrane is easily movable, so that the barrier is
affected only by surface effects.
If the barrier can freely move, it drifts in the direction of the liquid with higher
surface tension. In this way the system can reduce its entire free energy and
free enthalpy. One can imagine that this movement is caused by a film
pressure (also called "lateral pressure"). The film pressure is defined as the
difference between the surface tensions of the pure solvent
0
and the surface
tension of the solution :

0

The film pressure is often measured with the Wilhelmy plate method or, more rarely directly, by
detecting the force acting on the barrier.
Monolayers as two-dimensional gases
Often monolayers behave in such a way, as one would expect from a two-dimensional gas.
Reason: For many substances at small concentration the surface tension decreases linearly
with the concentration of the added material c:

a
The name is composed by the first letters of Agnes Pockels (1862-1935, amateur scientist from
Braunschweig, Lower Saxony), Irvine Langmuir (1881-1957, American scientist, spent most of his time
at the General Electric Company, NP Chemistry 1932), Adam, Wilson and McBain.
Solvent, mostly water
Barrier
Monolayer of
amphiphilic
molecules
- 2-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000
bc =
0

Here, b is a constant which depends on the solvent and the solute. One can take this equation
as the first term of a serial expansion. From this follows: =bc. Now we use the Gibb's equation
RT
RT RT
bc
dc
d
RT
c
= = = =


and obtain
RT A
m
= =

or = kT
A
m
is the surface per mole, is the surface per molecule.
Instead of the ideal two-dimensional gas equation often a van der Waals type
equation is used, which contains an excluding surface A
0
:

=
A
A A RT
m
m 2 0
( )

Here is a material dependent constant.
4.2 Experimental procedures for the investigation of mono-
layers
A good introduction to the classic experimental techniques and results is ref.1.
4.2.1 The surface potential
A monolayer changes the electrical potential between the liquid and the gas.
The change of the Volta potential between the liquid phase and the gaseous
phase upon spreading of the layer is called the surface potential of the film. If
V
0
is the potential between the inner part of the pure solvent and the gaseous
phase and if V is the potential in presence of the monolayer, then
0
V V V
Surf
=
is the surface potential. V
0
, V and V
Surf
are positive, if the electrode in the
gaseous phase is more positive than the electrode in the liquid.
The surface potential is connected with the dipole moment of the molecules at
the interface. If

is the dipole moment perpendicular to the surface, then


25
0

A
V
Surf

=
is the dielectric constant of the liquid, A is the surface area occupied by one
molecule. Practically one often measures the change of the surface potential
when changing the film pressure. From this we can get information on the
dipole moment and orientation of the molecules on the surface.
2

- 3-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000
In order to make the equation plausible, we imagine that the dipoles form a
homogeneously distributed layer on the liquid surface. The dipoles are
supposed to have a dipole moment D Q = . Directly on the liquid the
monolayer starts with a charge density Q = . Then follows a region without
free charges of thickness D. Finally we have a second layer with the charge
density Q = .
Now we try to find out which electric field strength we have within the three
regions. For simplicity we assume that there are no free charges in the liquid
and we can set the potential to zero throughout the liquid phase. In the area
between the two charge layers the field amounts to
0
= E . In the gaseous
phase the field is zero again. To obtain the surface potential we integrate the
electric field starting from somewhere in the liquid phase and ending
somewhere in the gas phase. With this integration we get the potential:
0 0
0
0

A
D
dx Edx V
D
Gas
flssig
Surf

= = = =


For molecules with no net free charge the measured dipole moment
corresponds directly to the internal dipole moment of the molecule.
3
With
charged molecules an additional dipole has to be considered which is formed
in the liquid by the charge of the molecule plus the counter ions in the electrical
double layer.
Measurements of the surface potential were first done in the 1932-37.
45
Two
methods are commonly applied to measure the surface potential:
The ionising electrode method. The potential between a reference electrode
in liquid (e.g. an AgCl electrode) and another electrode, which is placed few
micrometers over the liquid is measured. Practically this is done by measuring
the current and applying an external voltage V
ext
which compensates the
surface potential. The voltage compensates the surface potential, so that no
current flow is zero. In order to increase the
conductivity of the air gap, the upper
electrode is coated with an emitter, e.g.
Polonium Po
210
.
The vibrating electrode method. The
procedure goes back to Kelvin and therefore
is called also Kelvin probe. If there is a
V
ext

I
x
D
Gas






Liquid
0 E V
V
Surf

- 4-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000
potential difference between the surface of the liquid and an electrode attached
over it, a current is generated when the position of the electrode is changed. In
order to measure the surface potential, the upper electrode is periodically
moved up and down. The generated alternating current is measured. The
current is proportional to the surface potential according to
4

dt
dC
V
dt
dQ
I
Surf
= =
From the capacity of a plate capacitor D A C
0
= follows
dt
dD
D
A
dt
dD
dD
dC
dt
dC
= =
2
0


A is here the surface area of the plate capacitor. Substitution results in
dt
dD
D
A
V I
Surf
2
0

=
Example: The electrode in air is 1 cm
2


large and 0.3 mm away from the
surface of the liquid. It is periodically (in a sinusoidal way) moved up and down
with an amplitude of D
0
=2,5 m and a frequency of =330 Hz. The used
ampermeter can barely resolve a current I =1 pA. How precise can the
surface potential be measured?
( ) ( ) t D
dt
dD
t D D 2 cos 2 2 sin
0 0
= =
The difference in height amounts, from maximum to minimum, to 4D
0
. Thus
V
m m AsV m s
m A
A D
D I
V
Surf
01 , 0
10 10 85 , 8 10 5 , 2 330 4
) 10 3 ( 10
4
2 4 1 1 12 6 1
2 4 12
0 0
2
=


=

=



4.2.2 Surface elasticity and viscosity
The mechanical characteristics of thin films on liquids are described similarly to
the three-dimensional case. So the surface elasticity
6
is defined as
E A
d
dA
A
d
dA
=


A is the surface. 1/E is also called compressibility. It represents a measure for
the resistance of the film against compression.
When measuring the surface elasticity one should pay attention if there exist
different phases in the film which may show a different behaviour.
7

The surface viscosity
S
is defined as follows: Two parallel line elements are
moved relatively to each other. between them (if we consider Newtonian
behaviour) a constant speed gradient dv dx will exist. The force, which is
necessary to maintain the parallel movement is
- 5-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000
F l
dv
dx
S
=
In the canal type viscosity-meter one
measures the surface which flows, with a
certain pressure per time, through a gap (or
through a two-dimensional channel).
With a surface torsion pendulum the
visco-elastic characteristics can be measured as well. One determines how
quickly the rotational oscillation of a disk lying on the surface is damped.
Thereby damping in pure liquid is always compared with the damping in the
monolayer.
In the viscous traction viscosity-meter there are two cylinders, which go
perpendicularly through the surface. One cylinder rotates. It is measured which
torque is transferred to the other cylinder
4.2.2 Optical methods
In fluorescence microscopy a small amount fluorescent dye is added to the
amphiphilic molecules studied. To go into the monolayer the dye also has an
amphiphilic character. The film is illuminated. The lateral distribution of the
fluorescent molecules are observed with an optical microscope
8
. Depending on
the phase condition of the monolayer the fluorescent molecules distribute
unevenly or have a different quantum yield. Usually the dyes are expelled from
solid phases. With the fluorescence microscopy was demonstrated for the first
time the coexistence of different phases in monolayers on water.
9,10


The picture of Mathias Lsche, Leipzig, shows L--DPPC (Dipalmitoyl-
phosphatidylcholin) on water at 22C and a film pressure of 12 mN/m. As
fluorophor 0,5%mol NDB-DPPC was added. It is pushed out of the crystalline
regions and is enriched in the fluid region. The width corresponds
approximately to 250 m.
A criticism of the fluorescence microscopy was that the presence of the
fluorescent dyes possibly changes the structure of the monolayer. The problem
is avoided in the BAM (Brewster angle microscopy).
11,12,13,14
In Brewster
angle microscopy we utilise the principle that incident light, which is polarised
parallel to the plane of incidence, is not reflected under the Brewster angle. For
v+v
v
x
l
- 6-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000
water the Brewster angle amounts
to 53. The presence and structure
of the monolayer slightly change
the Brewster angle. If one observes
the water surface under 53 the
bare water surface appears dark.
Regions which are covered with a
monolayer have a slighlty different
Brewster angle an appear bright.
With the help of Brewster angle
microscopy the phase behaviour of monolayers can be observed with a lateral
resolution in the m range.
15
Comparative investigations with the fluorescence
microscopy and the Brewster angle microscopy showed nearly no
differences.
16

4.2.3 X-ray diffraction
25

X-rays have a large penetration depth. In order to receive a sufficiently intense
signal from the superficial layer the incident beam is applied under a small
angle. Typical vertical irradiating angles are 0.1, which leads to penetration
depths of approximately 5 nm.
From the reflected radiation we get different information:
17

Specular refection (=, =0). The intensity of the directly reflected beam is
measured versus . The experiments give information about the roughness of
the surface.
18,19
It was used, for example, for measuring capillary waves.
20

From the dependence of the in-plane reflexes (=) from the angle one
obtains information about the two-dimensional packing of the molecules on the
surface. A perfect flat two-dimensional crystal on the water surface would
generate strips in vertical direction.
If one analyses these strips as a function of the angle information about the
electron density profile vertical to the surface and about possible tilting angles
of the molecules is obtained. The measurements are usually performed with
intensive X-ray sources such as with synchrotron radiation.
21,22,23,24
4.3 Phases of monomolecular films
The phase of a monolayer is most conveniently de detected in pressure-area
isotherms. These may look quite different for different substances. The
behaviour of relatively simple amphiphilic molecules, like long-chain alcohols,
amines or acids, was examined. Overview articles over the behaviour of


Lateral view top view
Water
Brewster
angle
- 7-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000
monolayers are refs. 25,26,27. Fatty acids, phospholipids, etc. on water often
show the following phases:
28,29

Gaseous (G)
Equation of state is = kT . This film pressure is, however, usually so small
that is almost not detectable. The average area per molecule on the surface is
much larger than the molecule size. The total area can be expanded
indefinitely without having a phase transition. There is a first order phase
transition to the L-state.
Liquid films (L)
There is a significant lateral interaction between the molecules. At least two
types of liquid phases exist: Liquid expanded (LE): - graphs extrapolate to
zero point to a surface that is larger than the molecule size. For long-chain
hydrocarbons with a polar head (e.g. acids, alcohols) this surface amounts to
about 40-70
2
. The molecules are lying flat on the water surface. The
equation of state is often of van der Waals type. In addition we may have a
tilted phase. This is relatively stiff film with water between the molecule heads.
The molecules are not flat on the water surface anymore but they start to stand
up with increasing compression. - graphs are typically linear.
Solid films (S)
There is no more water between the head groups. Pressure-area graphs are
liner. Extrapolation to zero film pressure results in a surface that corresponds
to the molecular cross section (e.g. with long-chain fatty acids, 20.5
2
). Solid
films have a high compressibility.
It is instructive to relate the film pressure a three-dimensional pressure.
On the barrier of length l the film
applies a force l. In the three-
dimensional case we estimate
the force from the pressure p
which acts upon a surface ld (d:
thickness of the monolayer). This
force is pld. If the forces are set
equal, we obtain
p d =

A
m

gaseous
liquid expanded
tilted phase
solid-condensed
l
d
Barrier
p
- 8-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000
Typical values for the G-phase: d=1 nm, =1 dyn/cm=10
-3
N/m p = 10
6
N/m
2

= 10 atm.
Pressure-surface isotherms of polymers or proteins often show no defined
phases. Even with very small film pressures the behaviour is not ideal. The
behaviour depends strongly on the specific condition of the polymers. Using
polymers with an amphiphilic character, pressure-surface isotherms can be
measured relatively well;
30
with other polymers such a measurement is more
difficult because the isotherms are often irreversible.
With small concentrations one can often describe the behaviour with the
following model:
The molecule is a chain with n
segments (e.g. amino acid residues).
Each segments has a certain flexibility
z (z=2 for a rigid, up to 4 for a flexible
binding). The most compact packing is b. When the pressure increases the
molecules pack closer but the entropy decreases.
4.4 Thick films - spreading of a liquid on another
What happens, when we put a drop of liquid B on the surface of liquid A? The
two liquids are supposed not to mix and the density of B should be smaller
than the density of A. That is true, for example, for many oils on water. There
are two possibilities: Either liquid B spreads or it forms a lens with a defined
edge.
Liquid B spreads spontaneously on liquid A, if the spreading coefficient, also
called spreading pressure
S
A B A B AB /

is bigger than zero.
A
and
B
are the surface tensions of the pure liquids.
AB
is
the interfacial tension of the interface between the liquids. The condition results
from a simple energy balance: For S
A/B
>0 it is energetically more favourable if
liquid B spreads. More precisely: The free energy and the free enthalpy of the
system decrease if the liquid B spreads and forms a homogeneous film (Note:
We are not talking about monomolecular films but thick films).
A complication now arises.
A
and
B
are values for the pure liquids. In the given situation,
however, molecules A are dissolved in liquid B, and vice versa. Therefore one should actually
use
A(B)
and
B(A)
instead of
A
and
B
. Here,
A(B)
is the surface tension of the liquid A, in which
molecules B are dissolved up to saturation.
If the spreading
coefficient is smaller
than zero, then a well
defined drop B forms
on the surface of A.
The shapes of the
water surface, the oil
surface and the oil-
water interface are determined by the Laplace equation. The angles
1
,
2
and

3
are boundary conditions. They are given by the equation


Vapour
Oil
Water

3


2


1

- 9-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000
2 1 3
cos cos cos + =
OW O W

W
,
O
and
OW
are the surface tensions of water and oil, as well as the interfacial tension of the
oil-water interface. The equation results after the following consideration: At equilibrium the
three-phase contact line do not move. Hence, the sum of the forces acting in horizontal direction
must be zero. The water surface pulls with a force (per unit length) of
W
cos
3
to the left. The
oil surface pulls to the right with a force
O
cos
1
and so does the oil-water interface
OW
cos
2
.
There is an interesting special case when liquid A has a much higher density
than liquid B. This is for example the case if a water drop is put on mercury. In
that case
2
0 and
3
0, and the boundary condition is simplified
WHg W Hg
+ =
1
cos
The same equation applies to liquid drops on solid surfaces.
Measurement of the interfacial tension between two liquids

The interfacial tension between two liquids can be measured with the spinning
drop method. The liquid with the higher density is filled into a horizontal
capillary. Through a septum, which closes the capillary, a drop of the second
liquid is injected with a syringe. When the capillary is brought into fast rotation
liquid B will go to the center forming a drop astride the axis of revolution. With
increasing speed of revolution the drops elongates, since the centrifugal force
increasingly opposes the surface tensional drive toward minimum interfacial
area. At sufficiently high speed of revolution, the drop approximates to an
elongated cylinder. From the shape one can calculate the interfacial tension. A
summary of interfacial tensions between water and organic liquids is contained
in ref. 31.


1
Gaines, G.L., Insoluble Monolayers at Liquid-Gas Interfaces, Interscience Publ., New York, 1966.
2
Oliveira, O.N., Bonardi, C. Langmuir 1997, 13, 5920.
3
Demchak, R.J. ; Fort, T. J. Colloid Interface Sci. 1974, 46, 191.
4
Porter, E.F. J. Am. Chem. Soc. 1937, 59, 1883.
5
Yamins, H.G.; Zisman, W.A., J. Chem. Phys. 1933, 1, 656.
6
Wantke, K.D.; Lunkenheimer, K.; Hempt, C. J. Colloid Interface Sci. 1993, 159, 28.
7
Bercegol, H.; Meunier, J. Nature 1992, 356, 226.
8
Lsche, M.; Mhwald, H. Rev. Sci. Instrum. 1984, 55, 1968.
9
Lsche, M.; Sackmann, E.; Mhwald, H. Ber. Bunsenges. Phys. Chem. 1983, 87, 848.
10
Weis, R.M.; McConnell, H.M. Nature 1984, 310, 47.
11
Honig, D.; Mbius, D. J. Phys. Chem. 1991, 95, 4590.
12
Hnon, S.; Meunier, J. Rev. Sci. Instrum. 1991, 62, 936.
13
Lheveder, C.; Hnon, S.; Mercier, R.; Tissot, G.; Fournet, P.; Meunier, J. Rev. Sci. Instrum. 1998, 69,
1446.
14
Harke, M.; Teppner, R; Schulz, O.; Motschmann, H.; Orendi, H. Rev. Sci. Instrum. 1997, 68, 3130.
15
Vollhardt, D. Adv. Colloid Interface Sci. 1969, 64, 143.
16
Rivire, S.; Hnon, S.; Meunier, J.; Schwartz, D.K.; Tsao, M.W.; Knobler, C.M. J. Chem. Phys. 1994,
101, 10045.
17
Kjaer, K. Physica B 1994, 198, 100.
18
Braslau, A.; Deutsch, M.; Pershan, P.S.; Weiss, A.H.; B.M.; Als-Nielsen, J.; Bohr, J. Phys. Rev. Lett.
1985, 54, 114.
- 10-
PCSURF04. DOC BUTT PC UNI SI EGEN 19. 11. 2000

19
Schwartz, D.K.; Schlossmann, M.L.; Kawamoto, E.H.; Kellog, G.J.; Pershan, P.S; Ocko, B.M. Phys.
Rev. A 1990, 41, 5687.
20
Braslau, A.; Deutsch, M.; Pershan, P.S.; Weiss, A.H.; B.M.; Als-Nielsen, J.; Bohr, J. Phys. Rev. Lett.
1985, 54, 114.
21
Helm, C.A.; Mhwald, H.; Kjaer, K.; Als-Nielsen, J. Biophys. J. 1987, 52, 381.
22
Li, Z.et al., Langmuir 1995, 11, 4785.
23
Barton, S.W.; Thomas, B.N.; Flom, E.B.; Rice, S.A.; Lin, B.; Peng, J.B.; Ketterson, J.B.; Dutta, P. J.
Chem. Phys. 1988, 89, 2257.
24
Kenn, R.M.; Bhm, C.; Bibo, A.M.; Peterson, I.R.; Mhwald, H.; Als-Nielsen, J.; Kjaer, K. J Phys.
Chem. 1991, 95, 2092.
25
Mhwald, H. Annu. Rev. Phys. Chem. 1990, 41, 441.
26
McConnell, H.M. Annu. Rev. Phys. Chem. 1991, 42, 171.
27
Knobler, C.M.; Desai, R.C. Annu. Rev. Phys. Chem. 1992, 43, 207.
28
Harkins, W.D.; Young, T.F.; Boyd, E. J. Chem. Phys. 1940, 8, 954.
29
Harkins, W.D.; Copeland, L.E. J. Chem. Phys. 1942, 10, 272.
30
Ahrens, H.; Frster, S.; Helm, C.A. Macromolecules 1997, 30, 8447.
31
Freitas, A.A.; Quina, F.H.; Carrol, F.A. J. Phys. Chem. B 1997, 101, 7488.
- 1-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
5 THE ELECTRICAL DOUBLE LAYER
5.1 Introduction
Charging of interfaces
This chapter is about electrically charged solid surfaces in liquids. The most important liquid
is water. In water, surface charges are generated when ions are adsorbed at surfaces or
when surface groups dissociate.
1
Amino groups for instance bind protons (NH
2
+H
+
NH
3
+
).
Many oxides, like SiO
2
, TiO
2
, Al
2
O
3
, have hydroxyl groups at their surface. The hydroxyl
groups can dissociate and release a proton (OHO
-
+H
+
). As a result the surface becomes
negatively charged. Another cause for surface charges is an external applied electrical
potential at a metal as is typically done in electrochemical experiments.
Electrical double-layer, Helmholtz, Gouy
a
and Chapman
b

Surface charges cause an electrical field. This electrical field attracts counter ions. The layer
of surface charges and counter ions is called electrical double layer. The first theory for the
description of electric double layers comes from Helmholtz. Helmholtz started with the fact
that a layer of counter ions binds to the surface charges. The counter ions are directly
adsorbed to the surface. The charge of the counter ions exactly compensates the surface
charge. The electrical field generated by the surface charges is accordingly limited to the
thickness of a molecular layer. Helmholtz could interpret measurements of the capacity of
double layers; electrokinetic experiments, however, contradicted his theory.
Solid
x Helmholtz
Gouy-Chapman

Gouy and Chapman went a step further. They considered a possible thermal motion of the
counter ions. This thermal motion leads to the formation of a diffuse layer, which is more
extended than a molecular layer. For the one-dimensional case of a planar, negatively
charged plane this is shown in the illustration. Gouy and Chapman applied their theory on the
electric double layer of planar surfaces.
2
Later, Debye and Hckel calculated the behaviour
around spherical solids.
3

5.2 Gouy-Chapman theory of the electrical double layer
5.2.1 The Poisson-Boltzmann equation
The aim is to calculate the electric potential near charged interfaces. Therefore we
considers a plane with a homogeneous distributed electrical charge density , which is in

a
Louis George Gouy, 1854-1926, french physicist, professor in Lyon.
b
David Leonhard Chapman, 1869-1858, english chemist, professor in Manchester and Oxford.
- 2-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
contact with a liquid. This charge generates a surface potential,
0
. Generally charge density
and potential are related by the Poisson
c
equation:
0
2
2
2
2
2
2
2

e
z y x
=

e
is the electrical charge density at a certain place in C/m
2
. With the Poisson equation the
potential distribution can be calculated once the positions of all charges are known. The
complication in our case is that the ions in solution are free to move. Since their distributions,
and thus the charge distribution in the liquid, is unknown, the potential cannot be found only
by applying the Poisson equation. Additional information is required.
This additional formula is the Boltzmann
d
equation. If we have to bring an ion in solution from
far away closer to the surface, electric work W
i
has to be done. The local ion density would
be
T k W
i i
B i
e n n

=
0

n
i
0
is the density of the i
th
ion sort in the volume phase, given in particles/m. The local ion
concentration depends on the electrical potential at the respective place. For example, if the
potential at a certain place in the solution is positive, then at this place there will be more
anions, while the cation concentration is reduced.
Now we assume that only electrical work has to be done. We neglect for instance
that the ion must displace other molecules. In addition, we assume that only a 1:1
salt is dissolved in the liquid. The electrical work required to bring a charged cation
to a place with potential is =
+
e W . For an anion it is =

e W .
The local anion and cation concentrations n
-
and n
+
are related with the local potential
through the Boltzmann factor:
T k e
B
e n n

=
0
,
T k e
B
e n n
+
=
0

n
0
is the volume concentration of the salt. The local charge density is
|
|
|
.
|

\
|
= =

+ T k
e
T k
e
e
B B
e e e n n n e
0
) (
Substituting the charge density into the Poisson equation gives the Poisson-
Boltzmann equation:
|
|
|
.
|

\
|
=

T k
z y x e
T k
z y x e
B B
e e
e n
) , , ( ) , , (
0
0 2


This is a partial differential equation of second order. In most cases, it cannot be
solved analytically. Nevertheless, some simple cases can be treated analytically.
5.2.2 One dimensional geometry
A simple case is the one-dimensional situation of a planar, infinitely extended plane. In this
case the Poisson-Boltzmann equation only contains the coordinate vertical to the plane:
|
|
|
.
|

\
|
=

T k
x e
T k
x e
B B
e e
e n
dx
d
) ( ) (
0
0
2
2



c
Denis Poisson, 1781-1840, French mathematician and physicist, professor in Paris.
d
Ludwig Boltzmann, 1844-1906, Austrian physicist, professor in Wien.
- 3-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
Before we solve this equation for the general case, it is illustrative to treat a special
case:
Low potentials
How does the potential change with distance for small surface potentials? "Small"
means, strictly speaking T k e
B
<<
0
. At room temperature that would be 25 mV.
Often the result is valid even for higher potentials, up to approximately 50-80 mV.
With small potentials we can expand the exponential functions into a series and neglect all
but the first (i.e. the linear) term:
=
|
|
.
|

\
|

T k
e n
T k
e
T k
e e n
dx
d
B B B 0
2
0
0
0
2
2
2
1 1

K
This is sometimes called the linearized Poisson-Boltzmann equation. The general solution of
the linearised Poisson-Boltzmann equation is
x x
e C e C x

+ =

2 1
) (
with
T k
e n
B 0
2
0
2

=

C
1
and C
2
are constants which are defined by boundary conditions. For a simple
double layer the boundary conditions are ( ) x = 0 and ( ) x = = 0
0
. The
first boundary condition guarantees that with very large distances the potential
disappears and does not grow infinitely. From this follows 0
2
= C . From the second
boundary condition follows
0 1
= C . Hence, the potential is given by
x
e

=
0

The potential decreases exponentially. The typical decay length is given by
1
=
D
. It is called the Debye
e
length.
The Debye length decreases with increasing salt concentration. That is intuitively
clear: The more ions are in the solution, the more effective is the shielding of the
surface charge. If one quantifies all the factors for water at room temperature, then
for a monovalent salt with concentration c the Debye length is
D
c = 3 , with c
in mol/l.
f
Example: at a concentration of NaCl of 0.1 M the Debye length is 0.95 nm.
In water
D
cannot be longer than 900 nm. Through the dissociation of water, in
accordance with 2H
2
OH
3
O
+
+OH
-
,

the ion concentration cannot decrease below
10
-7
M.
For the general case the inverse Debye length is given by

=
i
i i
B
z n
T k
e
2 0
0
2


Here, z
i
is the valency of the i
th
ion sort. Example: For a 1 M CaCl
2
solution
3 23 0
10 6

=
+
m n
Ca
,
3 23 0
10 12

=

m n
Cl
, 2 = +
Ca
z , 1 =
Cl
z .
Arbitrary potential

e
Peter Debye, 1884-1966, American physicist of Dutch origin. Professor in Zrich, Utrecht, Gttingen,
Leipzig, Berlin and Ithaca. NP for chemistry, 1936.
f
n is the number of particles per m
3
.
- 4-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
Now comes the general solution of the one-dimensional Poisson-Boltzmann
equation. It is convenient to treat the equation with the dimensionless potential
kT e y . The Poisson-Boltzmann equation becomes thereby
( ) ( ) y e e
T k
e n
e e
T k
e n
dx
y d
y y
B
y y
B
sinh
2
1 2
2
0
2
0
0
2
0
2
2
= = =



To obtain this we used:
2
2
2
2
dx
d
T k
e
dx
y d
B

= and ( )
y y
e e y

= 2 1 sinh
To solve the differential equation we multiply both sides with 2dy/dx:
y
dx
dy
dx
y d
dx
dy
sinh 2 2
2
2
2
=
One can also write the left side as
2
2 |
.
|

\
|
dx
dy
dx
d
. Now we integrate:

=
|
.
|

\
|
' sinh
'
2 '
' '
2
2
dx y
dx
dy
dx
dx
dy
dx
d

C y dy y
dx
dy
+ = =
|
.
|

\
|

cosh 2 ' ' sinh 2


2 2
2

C is an integration constant. It is determined by the boundary conditions: At large
distances x the potential y and dy/dx are zero. Since 1 cosh = y for y=0, C=-2
2
.

From this follows:
( ) 2 cosh 2 1 cosh 2
2
2
= =
|
.
|

\
|
y
dx
dy
y
dx
dy

In front of the square root there is a minus sign because y decreases with positive
potential and increasing distance, i.e. 0 0 < > dx dy y .
Now we remember the mathematical relation ( ) 1 cosh
2
1
2
sinh = y
y
. Thus
2
sinh 2
y
dx
dy
=
Separation of the variables and integration leads to:

= = ' 2
2
'
sinh
'
2
2
sinh
dx
y
dy
dx
y
dy

C x
y
2 2
4
tanh ln 2 + = |
.
|

\
|

C is again an integration constant. Written out:
C x
e e
e e
y y
y y
+ =
|
|
.
|

\
|
+

4 4
4 4
ln
- 5-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
Multiplying denominator and numerator in brackets with
4 y
e leads to
C x
e
e
y
y
+ =
|
|
.
|

\
|
+

1
1
ln
2
2

The potential must correspond to the surface potential for x=0, that means
0
) 0 ( y x y = = . With the boundary condition one gets the integration constant
C
e
e
y
y
=
|
|
.
|

\
|
+

1
1
ln
2
2
0
0

Substituting results in
x
e e
e e
e
e
e
e
y y
y y
y
y
y
y
=
|
|
.
|

\
|
+ +
+ +
=
|
|
.
|

\
|
+

|
|
.
|

\
|
+

1 1
1 1
ln
1
1
ln
1
1
ln
2 2
2 2
2
2
2
2
0
0
0
0

1 1
1 1
2 2
2 2
0
0
+ +
+ +
=

y y
y y
x
e e
e e
e


Solving the equation for
2 y
e leads to the alternative expression
x y y
x y y
y
e e e
e e e
e

+
+ +
=
) 1 ( 1
) 1 ( 1
2 2
2 2
2
0 0
0 0

The potential and the ion concentrations are shown as an example in the illustration.
A surface potential of 50 mV and a salt concentration (monovalent) of 0.1 M were
assumed.
0 1 2 3 4
0
1
20
30
40
50
Distance / nm
0,0
0,2
0,4
0,6
Potential/mV Concentration/M
potential
conc. counter-ions
conc. co-ions

We see that
the potential decreases approx. exponentially with increasing
distance
the salt concentration of the counter ions decreases more rapidly
than the potential
the total ion concentration close to the surface is increased
In the following illustration the potential, which is calculated with the linearised form
of the Poisson-Boltzmann equation (dashed), is compared with the potential which
- 6-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
results from the complete expression. We see that the decrease of the potential
becomes steeper with increasing salt concentration. This reflects that the Debye-
length decreases with increasing salt concentration.
0 2 4 6 8 1
0
1
20
30
40
50
60
70
80
Distance / nm
Potential / mV
1 mM
10 mM
0.1 M
full
linearized

5.2.3 The Grahame equation
How are surface charge and surface potential
0
related? This question is
important, since the relationship can be examined with the help of a capacity
measurement. Theoretically this relation is described by the Grahame equation.
One can deduce the equation easily from the so-called electroneutrality condition.
This condition demands that the total charge, i.e. the surface charge plus the
charge of the ions in the whole double layer, must be zero. The total charge in the
double layer is

0
dx
e
and we get
4

0
0
0
2
2
0
0
=


=

= =

x
e
dx
d
dx
dx
d
dx
In the final step we had used 0 =
= x
dx d . With
2
sinh 2
y
dx
dy
= and
( )
dx
d
T k
e
dx
T k e d
dx
dy
B
B

=

= follows
|
|
.
|

\
|
=
T k
e
kT n
B
2
sinh 8
0
0 0

For small potentials one can expand sinh into a series (sinh ... x x x = + +
3
3! ) and
break off after the first term. That leads to the simple relation:
D

0 0

=
The following illustration shows the calculated relation between surface tension and
surface charge for different concentrations of a monovalent salt.
- 7-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
0,00 0,05 0,1 0,1 0,20
0
20
40
60
80
100
120
140
Surface charge density / e/nm
2
Surface potential / mV
0.1 M
10 mM
1 mM

We see that
for small potentials is proportional to
0

for high potentials rises more steeply then
0

the capacity d d
0
grows with increasing surface potential
depending on the salt concentration, the linear approximation (dashed) is
valid till
0
40...80 mV.
The capacity of the double layer
A plate capacitor has the capacity
d A dU dQ C
0
= =
Q: Charge, U: applied voltage, A: Area, d: Distance. The capacity of an electrical double layer
per area is thus
|
|
.
|

\
|
=
|
|
.
|

\
|
=

=
T k
e
T k
e
T k
n e
d
d
C
B D B B
2
cosh
2
cosh
2
0 0 0 0 0
2
0



We can expand cosh into a series (cosh ! ! ... x x x = + + + 1 2 4
2 4
) and for small
potentials we can break off after the second term. Then results
C
D
=

0

Result: The electrical double layer behaves like a plate capacitor, in which the plate
distance is given by the Debye length!
The capacity of a double layer - that is the ability to store charge - thus rises with
the salt concentration. Two examples with water as medium:
c = 0.001 M
D
= 9.5 nm C = 0.075 F/m
2

c = 0.1 M
D
=0.95 nm C = 0.75 F/m
2
(1 F = 1 C/1 V).
5.2.4 How realistic is the Gouy-Chapman Theory?
The following was assumed in the calculation:
5

Water and ions were treated like a continuous medium with continuous charge distribution.
The fact that discrete, extended molecules were treated, whose amount of charge can be
only a multiple of the unity charge, was ignored.
Likewise an ion-ion interaction was neglected, apart from pure Coulomb interaction.
- 8-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
A change of the dielectric constant in proximity of the surface or due to the presence of ions
was not considered.
The surface charges were assumed as continuously distributed.
Despite these strong hypotheses, the Gouy-Chapman theory usually describes very well the
behaviour of electrical double-layers. The reason is that two errors with opposite effect just
compensate themselves:
The ion-ion interaction and the assumption of a constant dielectric constant lead to the fact
that the actual surface potential is lower than the calculated one.
The hydration of the surface and of the ions, and the associated more inefficient shielding of
the surface charges, cause an underestimation of the surface potential.
For aqueous environment one can say that the Gouy-Chapman theory provides
relatively good predictions for monovalent salts at concentrations below 0.2 M and
for potentials below 50-80 mV. The fact that the surface charge in reality is not
continuously but discretely distributed, leads, according to experience, to deviations
only with bivalent and trivalent charges. Often, however, the surface charges lie not
precisely in one plane. This is true for example for biological membranes. In this
case larger deviations might result.
5.3 The Debye-Hckel Theory
The Debye-Hckel theory describes the electrical double-layer around a sphere.
With radial symmetry it is convenient to use polar coordinates. In polar coordinates
the Poisson-Boltzmann equation becomes
|
|
|
.
|

\
|
= |
.
|

\
|

T k
r e
T k
r e
B B
e e
e n
dr
d
r
dr
d
r
) ( ) (
0
0 2
2
1


This equation cannot be solved analytically. For small potentials one can expand
the exponential functions in a series and omit all but the linear term. Then the
equation is simplified to
= |
.
|

\
|
2 2
2
1

dr
d
r
dr
d
r

In order to solve this differential equation u r must be first substituted. Using
the boundary conditions 0 and ( ) r R = =
0
the differential equation can be
solved. The result is
) (
0
R r
e
r
R

=


We can calculate the surface potential and the particle charge Q just as we did in
Gouy Chapman theory by integrating the total charge in the double layer:
( ) ( )
( )
( ) 1 4
1
4
1
4
4 4
0 0
2
0
2
0
0
2
0
2 2
0
2
+ =
(

=
(
(

+ =
= =


R R e
R
R
dr e e
r
R
dr r dr r Q
R
R r
R
R r
R
R r
R R
e



- 9-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
The Debye-Hckel theory is used in case of particles, whose radius is smaller or in
the same order of magnitude with the Debye length. For large particles, for R
D
>> ,
the Gouy-Chapman theory is more practical.
5.4 Additional description of the electrical double layer
5.4.1 The model of Stern
g

The Gouy-Chapman model is unrealistic for high surface potnetials
0
and small distances
because then the ion concentration becomes too high. Stern tried to handle this problem by
dividing the region close to the surface into two regions:
6

The Stern-layer, consisting of a layer of ions, which, as assumed by
Helmholtz, are directly adsorbed to the surface.
A diffuse Gouy-Chapman layer.
In reality all models can only describe certain aspects of the electric double-layer.
The real situation at a metal surface might for instance look like in the following
illustration.
The surface potential
0
is determined by the external potential and by adsorbed
ions. These ions are not distinguishable from the metal itself (e.g. Ag
+
or I
-
on AgI).
Next comes a layer of relatively tightly bound, hydrated ions (e.g. H
+
or OH
-
on
oxides or proteins). If this layer exists it contributes to
0
. The so-called inner
Helmholtz plane marks the centre of these ions.
The Stern layer consists of adsorbed, hydrated ions. The outer Helmholtz plane
goes through the centre of these ions.
Finally there is the diffuse layer. The potential at the distance where it originates
corresponds roughly to the Zeta-potential (see latter).

g
Otto Stern, 1888-1969, German physicist, Prof. in Hamburg.
x


Stern-layer
GC-layer
-
-
- 10-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
hydrated
cations
specific adsorbed
anions
metal
=
secondary bond water
32
primary bond water
6
"volume
water
= 78
diffused layer

-potential
OHP
IHP

0
,

0
Stern-layer

metal

M

5.5 The free enthalpy of the electrical double layer
The free enthalpy, which is connected with the formation of a double layer, must be
negative, since the double layer forms spontaneously. The energy originates from
the chemical energy, which is obtained if surface groups dissociate or if ions bind to
the surface. In order to calculate the free enthalpy, one proceeds in three steps:
7,8

1. First we calculate the chemical energy which is gained when the potential-
determining ions either bind to the surface or dissociate from surface groups.
This chemical energy corresponds to the electric energy q
0
, which is necessary
to bring an ion of charge q to a potential
0
. The associated Gibbs free energy
per unit area is:
G =
0

2. Then, step by step, we bring the counter ions to the interface. The number of the
counter ions is equal to the number of ions of the diffused double layer. In order
- 11-
PCSURF05. DOC BUTT PC UNI SI EGEN 03. 12. 2000
to bring counter ions to the interface a work d dG '
0
= is required.
0
is the
surface potential, at a certain time of the charging - or better discharging -
process.
3. In the third step the counter ions are released from the surface. Stimulated from
the thermal movement, they partially diffuse away from the surface and form the
diffused double layer. Entropy but at the same time energy increases. One can
show that both terms compensate, so that in the third step no contribution to the
Gibbs free energy results.
The total Gribbs free energy of the diffuse double-layer is

+ =


0
0 0
' ' d G
Now one uses the purely mathematical relationship
( ) ( )

+ = + = ' ' ' ' ' ' ' ' ' ' ' '
0 0 0 0 0 0
d d d d d d
in order to get:


= + =
0 0 0
0
0
0
0
0
0 0
' ' ' ' ) ' ' ( d d d G


The integral can be solved with the help of the Grahame equation (thus in the
framework of the Gouy Chapman theory):
(

|
|
.
|

\
|
=
(

|
|
.
|

\
|
=

|
|
.
|

\
|
= =



1
2
cosh
8
2
'
cosh
2
8
'
2
'
sinh 8 '
0 0
0
0
0 0
0
0
0
0 0
0
0
0
0 0
T k
e T k n
T k
e
e
kT
T k n
d
T k
e
T k n d G
B
B
B
B
B
B



For very small potentials the relation
0 0
= applies and we get
0
2
0
0
0
0 0 0
2
1
2
' '
0
= = =

d G


1
Lyklema, J. Adsorption of Small Ions, in: Adsorption from Solution at the Solid/Liquid Interface, Ed.
Parfitt, C.H. Rochester, Academc Press, London, 1983.
2
Gouy, G., Journ. de Phys. 1910, 9, 457; Chapman, D.L., Phil. Mag. 1913, 25, 475; Gouy, G., Ann.
Phys. 1917, 7, 129.
3
Debye, P.; Hckel, E., Physikalische Zeitschrift 1923, 24, 185.
4
Grahame, D.C., Chem. Revs. 1947, 41, 441.
5
Cevc, G., Biochim. Biophys. Acta 1990, 1031-3, 311.
6
Stern, O., Zeitschrift fr Elektrochemie 1924, 30, 508.
7
Verwey, E.J.W.; Overbeek, J.T.G., Theory of the Stability of Lyophobic Colloids, Elsevier, New York,
1948, S. 51.
8
Chan, D.Y.C.; Mitchell, D.J., J. Colloid Interface Sci. 1983, 95, 193.
- 1-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
6 EFFECTS AT CHARGED INTERFACES
6.1 Electrocapillarity
Electrocapillarity: The change of the interfacial tension at a metal-electrolyte interface
upon variation of the applied potential. It was known for a long time that the shape of
a mercury drop which is in contact with an electrolyte depends on the electric
potential. Lippmann
a
examined this electrocapillary effect in 1875 for the first time. He
succeeded in calculating the interfacial tension as a function of an applied potential
and he measured it with mercury.
6.1.1 Theory
The change of the interfacial tension can be
calculated with the help of the Gibbs-Duhem
equation even when a potential is applied. In
order to use the equation, we first need to find
out which molecular species are present.
Evidently, only those which are free to move
are of interest. In the electrolyte we have the
dissolved ions. In the metal the electrons can
move and have to be considered.
Since additionally to the chemical potentials

i
also electrical potentials affect the charged
particles, electrochemical potentials
i

must
be used. The Gibbs-Duhem equation for
changes of state functions at constant
temperature is
d d d
i i
i
n
e e
=


1

where
d d z F d
d d F d
i i i A
e e A

= +
=


The first term refers to the electrolyte. Accordingly the sum goes over all particle
types which are present in the electrolyte. The second term contains the contribution
of the electrons in the metal.
i
und
e
are the interfacial excess concentrations of the
particles in solution and of the electrons in the metal, respectively. The
electrochemical potential of the ith sort of particles.
i
is the chemical potential of the
particle type i, z
i
is its charge number. F
A
=96485 C mol
-1
designates the Faraday
constant.
e

is the electrochemical potential of the electrons. Substitution leads to


d d F z d d F d
i i
i
n
A i i
i
n
e e A e


= +
= =

1 1

und

are the electrical (galvanic) potentials in the two phases.



a
Gabriel Lippmann, 1845-1921, french physicist, NP in physics1908.

Metal
Electrolyte


- 2-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
Which potentials have to be used? In the metal the potential is everywhere the same
and therefore one can describe it with one value. This is not the case in the
electrolyte. The potential changes in the electrical double-layer. Nevertheless the
electrochemical potential in the liquid phase is everywhere the same assuming that it
is in equilibrium. At large distance away from the surface the potential is constant.
Thus we use the electrochemical potential of the volume phase, that means far away
from the interface.
The concentrations of ions and electrons are not independent. With the electro-
neutrality condition one degree of freedom is lost. It is practical to regard the
electrons in the metal as the dependent component. With the help of the
elctroneutrality condition
i i e
z = the last term can be rewritten as F z d
A i i

.
Then it is possible to summarise the second and fourth term and one gets:
d d F z d d F z d
d d d
i i
i
n
A i i
i
n
e e A i i
i
n
i i
i
n
e e




= +
=
= = =
=



1 1 1
1
( )

This is the fundamental equation for the description of classical electrocapillarity
curves. Thereby = F z
A i i
is identified with the surfacial charge density, which is
produced from the ions in the solution. This identification is generally doubtful,
because the values
i
depend on the position of the interface. If, however, the
electrode is totally polarisable, that means that no electrons are exchanged between
the metal and the electrolyte, then the positioning of the interface is trivial and
represents the surfacial charge density, which is produced by the ions.
The difference of potential

is not necessarily equal to the externally applied
voltage U. Usually both differ by a constant, which is determined by the reference
electrode. A possible voltage drop in the electrolyte, in particular with small salt
concentrations, should be considered too. Since

and U differ only by a
constant, d dU ( )

= is still valid.
From the last equation the Lippman equation

U
=
and the relationship

2
2
U U
C = =
are directly obtained. C is the capacity (differential capacity) of the electrical double-
layer. The index means that the composition of the electrolyte and that of the metal
is kept constant.
In the derivation we supposed that no electrochemical reactions take place, i.e. the
electrode should be totally polarisable. A derivation of the Lippmann equation, which
explicitely considers electrochemical reactions, is found in electrochemistry text
books.
Lippmanns classical derivation was much simpler. Lippmann took, as a model for the
surface with an electrical double-layer, a plate capacitor with the area A. One plate is
the metal, the other the layer of counterions in the electrolyte. The potential difference
- 3-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
between the two plates is U. The reversible work dG upon a change of the surface A
or, respectively, of the charge Q is
UdQ dA dG + =
The second term indicates the work which is necessary to bring the charge dQ from a
plate to the other, against the potential U. Integration with constant and U results in
G A UQ = + .
A comparison of the differential (dG dA Ad UdQ QdU = + + + ) with the original
equation shows that
0 = + Ad QdU
A
Q
dU
d
=

.
The simplest electrocapillarity curves are obtained under the assumption that C is
constant (see the following illustration). Then we have
=
0
2
1
2
CU
because

= = = ' ' ' '
'
dU CU dU
A
Q
dU
dU
d

Remember: C is capacitance per unit area.
0
is the interfacial tension at the point, at
which the interface is uncharged, and the potential accordingly disappears (point of
zero charge).
U

0
pzc

The interfacial tension decreases with increasing surface potential. The reason
is the increased interfacial excess of the counter ions in the electric double-layer. In
accordance with the Gibb's adsorption isotherms the interfacial tension must
decrease with increasing interfacial excess. At charged interfaces ions have an effect
similarly to surfactants at liquid surfaces.
6.1.2 Measurement of the electrocapillarity
Fundamental knowledge about the behavior of charged surfaces comes from
experiments with mercury. How can an electrocapillarity curve of mercury be
measured? A usual arrangement, the so-called dropping mercury electrode, is
represented in the following illustration.
1
A capillary filled with mercury and a counter
electrode are placed into an electrolyte solution. A voltage is applied between both.
The surface tension of mercury is measured with the pressure bubble method.
Mercury is thereby pressed into the electrolyte solution under constant pressure p.
The number of drops per time unit is measured as a function of the applied voltage.
- 4-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
The number of drops, which detach per time unit, depends on the surface tension of
mercury.
N
2
gas with
pressure p
Tunable
voltage U
Not polarisable
electrode
Electrolyte
solution
Mercury
manometer

There are different methods to measure the electrocapillarity of solid metals.
However, all these methods are by far not as precise and easy to carry out as with
mercury. The fundamental reason is: The molecules in the solid are not mobile and
the formation of new surface is different than in a liquid. Thus also the theoretical
treatment is substantially more difficult.
2,3

The deflection of a spring bar, which is coated on one side with the metal, is
measured in a further method.
4,5
The surface tension of the metallic side changes
when the applied voltage varies. Assumed that the surface tension of the metallised
side decreases, then the spring bends to the other side, and vice-versa.
Metal
U

In an alternative method for the measure of the electrocapillarity of solid metals, one
detects the expansion, or respectively the compression, of a strained wire under the
effect of the surface tension.
6,7
Finally, also a change of the contact angle with
variation of the potential can be measured. It contains information about the surface
tension of the solid.
8,9,10

6.2 Examples of charged surfaces
6.2.1 Mercury
Mercury has one of the best examined surfaces. The substantial explanation is that it
is a liquid metal. On metals a voltage can be easily applied. Differently than for a
solid, the surface tension in a liquid can be measured simply and precisely, and then
- 5-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
the surface charge can be calculated with the help of Lippmann equation.
Additionally, a fresh and not contaminated surface can be continuously produced.
If one measures electrocapillarity curves of mercury in aqueous medium, which
contains KF, NaF, or CsF, then it is observed that the typical parabolas become
narrower with increasing concentration. Explanation: with increasing salt
concentration the Debye-length becomes shorter, the capacity of the double-layer
increases. The maximum of the electrocapillarity curve, and thus the point of zero
charge (pzc), remains constant. I.e. neither the cations nor fluoride adsorbs strongly
to mercury.
Another behaviour is observed in solutions with KBr, KI, or KCl. An increase of the
concentration leads to a shift of the maximum to negative potentials. Explanation: the
anions bind specifically to mercury and shift the pzc. Since there is no charge on the
surface, a negative potential must be applied, in order to get the anions away from
the surface.
Anions bind also to other metals, e.g. gold, platinum or silver.
11,7
Why can anions
specifically bind to metals, while cations normally do not? The explanation is the
strong hydration of cations. A cation would have to lose its hydration shell for an
adsorption. This is energetically disadvantageous. Anions are barely hydrated and
can therefore bind more easily to metals.
12
Another possible explanatuion is the
stronger van der Waals force between anions and metals. The binding of ions to
metallic surfaces is not yet understood and even the idea that cations are not directly
bound to the metal, was brought into question by molecular-dynamic calculations.
13

6.2.2 Silveriodide
If AgI or AgCl is submerged in water, a certain number of molecules dissolve. An
equilibrium between the ions in solution and the crystal is established:
+
+ I Ag AgI
The concentrations of
+
Ag and

I in the solution is very small, since the solubility


product
2 16
10 M a a K
I Ag

= + is small.
The surface of AgI can be imagined like a regular lattice of
+
Ag and

I ions. With the


same number of
+
Ag as

I , the surface would be uncharged. It is however clear that


this does not agree with the case where also in solution equal ion concentrations are
present. Iodide has a somewhat higher affinity for the AgI surface than
+
Ag . In the
case [ ] [ ] M I Ag
8
10
+
= = , AgI therefore carries a negative surface charge. The
No adsorption Specific adsorption

+ U -
10
-4
M
10
-3
M
10
-2
M

10
-4
M
10
-3
M
10
-2
M
+ U -
- 6-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
surface can be neutralised by increasing the
+
Ag concentration in solution, e.g. by
addition of AgNO
3
. Already the small quantity of 10
-5,5
M AgNO
3
is sufficient to reach
the pzc of AgI. One has then M
5 , 10
10

I in solution, i.e. 100.000 times less than
+
Ag .
The surface charge of AgI can thus be changed by addition by
+
Ag or

I . The two
ions are called the "potential determining ions" of AgI.
How is the surface potential quantitatively related to the concentration of the potential
determining ions? In equilibrium the electrochemical potential of
+
Ag ions at the
crystal surface is equal to that in solution:
K A K K L A L L
F Ag a RT Ag F Ag a RT Ag + + = + +
+ + + +
) ( ln ) ( ) ( ln ) (
0 0

) (
0 +
Ag
K
and ) (
0 +
Ag
L
are the chemical standard potentials at the crystal surface
and in solution, respectively.
K
and
L
are the Galvani potentials in the crystal and in
the solution. In particular, the equation is valid at the pzc:
pzc
A
pzc
K K
pzc
L L
F Ag a RT Ag Ag a RT Ag + + = +
+ + + +
) ( ln ) ( ) ( ln ) (
0 0

pzc
is the difference of the Galvani potentials, which is caused solely by dipoles in
the interface, not by free charges. If we subtract the two equations from each other
we get
( )
pzc
L K A
pzc
L
L
F
Ag a
Ag a
RT =
+
+
) (
) (
ln
It is assumed that ) ( ) (
+ +
= Ag a Ag a
pzc
K K
. The expression in brackets is called
"surface potential"
0
. Thus one obtains the Nernst equation:
) (
) (
ln
0
+
+
=
Ag a
Ag a
F
RT
pzc
L
L
A

The concentration of
+
Ag (and thus that of

I ) determines the surface potential. An


increase of the concentration by a factor of 10 causes an increase of the surface
potential of 59 10 ln
0
= =
A
F RT mV at 25 C.
During the derivation we assumed that ) ( ) (
+ +
= Ag a Ag a
pzc
K K
. That means that
during the charging of the AgI surface the activity of the
+
Ag ions on the surface
does not change. This assumption is justified to a large extent, because the number
of
+
Ag ions on the surface changes only slightly. The relative number of ions, i.e. the
number of the additionally adsorbed ions upon a variation of the potential, is very
small.
In addition, an estimation: in an uncharged, perfectly crystalline AgI surface, the
distance between
+
Ag ions amounts to about 0.4 nm. This corresponds to a surface
area for each
+
Ag ion of 0.16 nm
2
. In every nm
2
are thus 6.25
+
Ag ions. How much
does the density of the
+
Ag ions increase, if the potential increases by 100 mV?
Conditions: 25C, 1 mM KNO
3
background electrolyte. In order to estimate the surface
charge, we use the Grahame equation ( ) kT e kT n 2 sinh 8
0 0 0
= . With
- 7-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
95 , 1
298 10 38 , 1 2
1 , 0 10 60 , 1
2
1 23
19
0
=


=

K JK
V As
kT
e

and
3 23 1 20
0
10 02 , 6 10 02 , 6

= = m l n we get
2 2
3
23
2
12 3 23
0122 , 0 95 , 1 sinh 10 55 , 3
95 , 1 sinh 298 10 38 , 1 10 85 , 8 72 10 02 , 6 8
m
C
m
C
K
K
J
Jm
C
m
= =
=


If we divide this value by the unity charge we get an ion density of 0.076
+
Ag ions
per nm
2
. That means that the number of the ions at the surface changes only by
1.2%.
In the previous section the mercury electrode has been presented. If no redox pairs
(e.g. Fe
2+
, Fe
3+

) are in solution and if one excludes gas reactions, then the mercury
electrode is completely polarisable. If a potnetial is applied a current flows only for a
short time. If the electric double-layer is formed the current stops. No electrons are
transferred from mercury to molecules in the solution and vice versa.
The AgI electrode is an example for the other extreme of a completely reversible
electrode. Each attempt to change the potential of an AgI electrode leads to a current
because the equilibrium potential is fixed by the concentrations of
+
Ag or

I
according to the Nernst equation.
As example we can consider the electrochemical cell
Ag | H
2
(Gas) | HNO
3
, AgNO
3
| AgI | Ag
With [ ] [ ]
3 3
AgNO HNO = . Assumed a negative potential is applied to the right
electrode. Then the reactions
AgI + e
-
Ag +

I and
+
Ag + e
-
Ag
take place. Iode ions go into solution, silver precipitates. The associated Faraday
current flows until the concentration of
+
Ag increases so much and that of

I
decreases so much that the Nernst equation is satisfied.
6.2.3 Oxides
A third mechanism of surface charging dominates among oxides (e.g. SiO2, TiO2,
Al2O3)
14
, proteins and many polymers. At the surface of these substances there are
groups that can dissociate. They take up or release a proton depending on pH.
Examples are hydroxyd-, carboxyd-, sulfate- and amino-groups. The potential
determining ions are OH
-
and H3O
+
.
To calculate the surface potential we consider the simplest example of a surface with
one dissociable group. Dissociation leade to a negatively charged group according to
+
+ H A AH
[ ][ ]
[ ] AH
H A
K
local
a
+
=
The surface concentrations of the negatively charged dissociated groups and the
neutral nondissociated groups are given in mol/surface and not in mol/volume.
- 8-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
[H
+
] is the local proton concentration in the solution directly at the surface. This
concentration can differ from the concentration in the volume phase. Explanation: if
the surface is charged it either attracts protons (by negative surface charge) or it
repels protons (by positive surface charge). A relation between the local
concentration and the volume concentration can be found with the help of the
Boltzmann factor:
[ ] [ ]
kT
e
local
e H H
0

+ +
=

0
is the surface potential. Substituting and taking the logarithm
[ ] [ ]
[ ]
[ ]
[ ]
[ ]
kT
e e
H
AH
A
K
AH
e H A
K
a
kT
e
a
log
log log log
0
0

+ =

=
+

+

From this follows
[ ]
[ ] kT
e
pH
AH
A
pK
a
0
434 , 0 log

=


or
( )
[ ]
[ ] AH
A
F
RT
pH pK
F
RT
A
a
A

+ = log 30 , 2 30 , 2
0

At 25C we have
( )
[ ]
[ ]
(
(

+ =

AH
A
pH pK mV
a
log 59
0

Instead of the potential-determining ions in the Nernst equation for AgI, here the pH
formerly replaces the pK
a
of the dissociable group.
The model was extended later, and also the binding of other ions at certain binding
sites on the solid, or at a certain distance from the surface of the solid, was allowed.
15

The surface charges of mica
16
or silicon nitride
17
are treated for example in such a
way.
6.3 Electrokinetic phenomena: The Zeta potential
If a liquid moves tangential to a charged surface, then so-called electrokinetic
phenomena arise.
18
Electrokinetic phenomena can be divided into four categories:
Electrophoresis, electro-osmosis, streaming potential and sedimentation potential.
19

In all these phenomena the (Zeta) potential plays a crucial role. If a liquid moves
along the surface of a solid the shear plane is often not directly on the surface. The
surface binds one, two or several layers of molecules of the liquid and possibly ions
more or less strongly. Only at a distance from the surface the molecules start to
move. The potential at this distance is the -potential. The classic theory of
electrokinetic effects comes from Smoluchowski
b
.
20

Electrokinetic phenomena can be understood with the help of two equations: The
known Poisson equation and the Navier
c
-Stokes
d
equation.

b
Marian von Smoluchowski, 1872-1917, polish physicist, professor in Lemberg and Krakowia.
c
Claude Louis Marie Henri Navier, 1785-1836, french engineer, professor in Paris.
- 9-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
6.3.1 The Navier-Stokes equation
The Navier-Stokes equation describes the movement of a Newtonian liquid, i.e. a
liquid whose viscosity does not depend on shearing. In order to make the equation
plausible we consider an infinitesimal quantity of the liquid having a volume
dV dx dy dz = and a mass dm. Different forces affect this quantity:
A viscous force, caused by a change of fluid velocity gradients, i.e. by shearing:
v v grad div
r r
2
) ( =
A possible pressure gradient causes the force: ( ) p dV
An electrical force, caused by the action of an electric field on the ions in solution:
dV E
e
r

p: pressure, :viscosity, v: velocity of the liquid at a certain place, E: electric field
strength,
e
: charge density.
The sum of these forces is equal to the mass dm times its acceleration
( )
dt
v d
dm dV E p v
e
r
r
r
= +
2

In many cases a steady state flow is considered. Then we have dv dt
r
= 0 and the
equation simplifies
0
2
= + E p v
e
r
r

It is a vectorl equation. Transformed into cartesian coordinates it reads
0
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
=
|
|
|
.
|

\
|
+
|
|
|
|
|
|
.
|

\
|

|
|
|
|
|
|
|
|
|
.
|

\
|

z
y
x
e
z z z
y y y
x x x
E
E
E
z
p
y
p
x
p
z
v
y
v
x
v
z
v
y
v
x
v
z
v
y
v
x
v

Additionally the equation of continuity applies to incompressible liquids
0 =

=
z
v
y
v
x
v
v
z
y
x
r

6.3.2 Electro-osmosis and streaming potential
We consider a fluid on a planar, charged surface. Parallel to the surface we apply an
electric field. The liquid begins to move under the influence of the field. This
phenomenon is called electro-osmosis. Explanation: The charged surface causes
an increase of the concentration of counter ions in the liquid close to the surface. This
surplus of counter ions is moved by the electric field towards the corresponding
electrode. Thereby they drag the surrounding liquid with them and a fluid flow arises.

d
Sir George Gabriel Stokes, 1819-1903, english mathematician und physicist, professor in Cambridge.
- 10-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
applied electr. field
Fluid flow x
z

As an example the simple, linear case, as it is shown in the figure, is treated. In the
linear case the y component of the Navier-Stokes equation disappears. Therefore all
y derivatives are zero, because no quantiity can change with y due to the symmetry.
We further assume that the liquid flows only parallel to the x coordinate. Then we
apply 0 0 = = z v v
z z
. From the symmetry we know that 0 = y v
y
. From the
equation of continuity follows that 0 = x v
x
. The flow rate does not change with the
position along the x direction. Since this is valid at all x, it follows 0
2 2
= x v
x
. From
the three-component Navier-Stokes equation the two remaining equations become:
0
2
2
= +

x e
x
E
x
p
z
v
and 0 = +

z e
E
z
p

The electric field in x direction is applied externally. In z direction the field results from
the surface charges. We also assume no pressure in x direction is externally applied.
Then x p disappears and after rearranging the first equation we get
2
2
z
v
E
x
x e

=
Now we use the expression from the Poisson equation (
0
2 2

e
dz d = ) for
the charge density:
E
d
dz
d v
dz
x
x

0
2
2
2
2

=
This equation can be integrated twice in z. The integration starts at a point far away
from the surface, where =0 and v
x
has a stationary value v
0
, up to the shear plane at
a distance from the surface where 0 =
x
v and = . Note that far from the plane
d dz = 0 and dv dz
x
= 0.
1
st
integration
dz
dv
dz
dz
v d
dz
d
E dz
dz
d
E
x
z
x
x
z
x
=


'
'
'
'
2
2
0
2
2
0

2
nd
integration
0 0 0
'
'
'
'
v dz
dz
dv
E dz
dz
d
E
x
x x


=

= =



It follows

x
E
v
0 0
=
The flow velocity is proportional to the -potential and to the applied field.
One can observe electro-osmosis directly with an optical microscope using liquids,
which contain small, but yet visible, particles as markers. Most measurements are
made at capillaries. An electric field is tangentially applied and the quantity of liquid
transported per time unit is measured. Capillaries have typical diameters from 10 m
- 11-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
up to 1 mm. The diameter is thus much larger than the Debye-length. Then the flow
rate will change only at the boundary. Some Debye-lengths away of the boundary the
flow rate is constant:
Approximately the liquid volume V transported per time amounts to


E
r v r
dt
dV
K K 0
2
0
2
= =
Example: We apply a potnetial of 1 V across a
capillary of 1 cm length and of 100 m diameter. The
field should decrease linearly, i.e. 100 = E V/m. The
capillary is filled with water. The -potential amounts to
0,05 V. Then
( )
s
m
sm
kg
V
m
V
J
C
m
3
14
2
12
2
6
10 5 , 2
001 . 0
05 . 0 100
10 85 . 8 72 10 50 14 , 3

=



flow through the capillary every second. As a reminder: J VC 1 1 = . With a volume of
the capillary of ( )
3 11 6
10 85 , 7 01 , 0 10 50 14 , 3 m m m

= the liquid is exchanged in
the capillary in 52 min.
The inverse effect to the electro-osmosis is the rise of a streaming potential. In this
case the liquid is pressed through a capillary (more generally along a charged wall).
Generation of the potential
Fluid flow

The liquid drags the charges of the electrical double-layer along with it. Thus a
difference of potential - the streaming potential - arises between the two ends of
the capillary. For capillaries, whose radius is substantially larger than the Debye-
length, it is possible to calculate a streaming potential between the beginning and the
end of the capillaries:
21

p
e
=


0

p is the applied pressure
e
is the electrical conductivity.
6.3.3 Electrophoresis and sedimentation potential
In electrophoresis we consider the motion of charged particles in electric fields (on the
contrary, the movement of ions falls under the keyword "ionic conductivity"). As an
example the motion of a spherical particle is treated. is the potential at a distance
from the particle surface. The distance R+ is the hydrodynamic radius of the
particle. It can be larger than the particle radius due to the binding of liquid molecules
v
0

r
K

- 12-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
or ions. Up to this distance the surface charge density
amounts to

. The entire charge of the particle is




2 2
4 ) ( 4 R R Q + = . An electric E causes a
force QE. With constant drift velocity v, the force is
compensated by the friction force. Equating results in
R
QE
v
6
=
In the treatment of the Debye-Hckel theory a relation between the surface potential
and the surface charge was established: ( ) R R Q + = 1 4
0 0
. With the help of this
relation we can replace the charge with the potential:
( ) R
E
v


+ = 1
3
2
0

During this consideration an important factor had not been taken into account: The
electric field affects not only the surface charges of the particle, but also the ions in
the electrical double-layer. The counter ions in the double-layer move in an opposite
direction with respect to the motion direction of the particle. The liquid transported by
them inhibits the particle motion. This effect is called electrophoretic retardation.
Therefore the equation is only valid for R
D
<< . Generally the term in brackets must
be replaced by a complicated function ) ( R f and one gets:
22,23

( ) R f
E
v


=
3
2
0

Sedimentation potential
If a charged particle moves in the gravity field or in a centrifuge, an electric potential
arises: The sedimentation potential. The effect is similar to the
electrophoretic retardation. While the particle moves the
electrical double-layer is somewhat pulled back by the flowing
liquid. Thus a dipole develops which causes the sedimentation
potential.
The isoelectric point and the point of zero surface charge
Both names are used in the literature. At the isoelectric point
(iep) we have 0 = . It is determined by electrokinetic
measurements. At the point of zero charge (pzc) the surface
charge is zero. Both can be distinguished, since the -potential
refers to the hydrodynamic interface while the surface charge is
defined over the solid-liquid interface. The surface charge can be measured e.g. in
electrocapillarity experiments.
6.4 Definition of the potentials
To each thermodynamically stable phase can be assigned an internal or Galvani-
potential . The internal potential is defined over the work, which is necessary for
the transport of a test charge from an infinite distance and from vacuum into the
interior of the phase far from the phase boundary; is the work divided by the test
charge.


R

- 13-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
This transport can be imagined to occurr in steps:
First the charge is brought close (approx. 1 m) to
the interface. With this step is connected the Volta-
Potential (external potential).
In the second step the charge arrives to the internal
phase passing through the interface. The associated
potential is called surface potential (surface
potential jump). It is determined by dipoles aligned at
the interface or by surface charges. It is not identical
to the surface potential that has so far been used for
the description of the electrical double-layer! For the
treatment of the electrical doube-layer dipoles did not
play a role. In particular in water, however, the aligned water molecules contribute
substantially to the surface potential . It applies
= +
The Volta-potentials of individual phases and Volta-potential differences between
two phases can be measured. In contrast to it, the Galvani-potential and accordingly
the surface potential of a single phase cannot be measured. Explanation: If we bring
a test charge (e.g. an ion or an electron) into a medium there is always a chemical
work in addition to the electrical work, e.g. by van der Waals forces or by image
charge effects. This we can not distinguish.
For the same reason differences of the Galvani-potential between two materials
and differences of the surface potential cannot be measured. What is measurable
are changes in the differences and , e.g. by measuring the electrocapillarity.
The electro-chemical potential

i
is measurable. It is defined as the work
necessary to carry a particle of the sort i from the vacuum into the phase . It can be
divided into a chemical part, the chemical potential

i
, and an electrostatical part
z e
i

.
Here it is instructive to mention the thermoionic work function , whereby e
indicates the work that is necessary to carry an electron from the Fermi-level of a
solid into vacuum far away from the surface. The thermoionic work function can be
measured.
As an example let us consider the potential in an electrochemical cell. For the sake of
simplicity an electrode A' is made of the same metal as the connection to the second
electrode.

U
A
B
C
A

The measured potential corresponds to the difference of the Galvani-potentials:
U
A A A A
=
' '

Vacuum
Medium
1m




- 14-
PCSURF06. DOC BUTT UNI SI EGEN 03. 12. 2000
The last equality is only to introduce the notation used from now on. One could object
that now differences of the Galvani-potential can be measured, although it has been
stated that this is impossible. The original statement is correct. In reality, the Galvani-
potential difference between material A and material A is not measured (it is naturally
zero), but the change of the potential difference due to the presence of the other
materials.
The sum of the voltages over all transitions must be zero, in accordance with the
Kirchhoff law. From this follows:
U
A A A B B C C A
= = + +
' '

The voltage can also be expressed by the electrochemical potentials of the electrons:
e e
U
B
e
A
e
A
e
A
e

=

=
' '

The last equation is valid since metal A and metal B are in contact and the electrons
can flow freely over the border. Therefore
e
A
e
B
= .

1
D.M. Mohilner, T. Kakiuchi, J. Electrochem. Soc. 1981, 128, 350.
2
D.M. Mohilner, T.R. Beck, J. Phys. Chem. 1979, 83, 1160.
3
G. Lang, K.E. Heusler, J. Electroanal. Chem. 1994, 377, 1.
4
R.A. Fredlein, J. O`M. Bockris, Surface Science 1974, 46, 641.
5
R. Raiteri, H.-J. Butt, J. Phys. Chem. 1995, 99, 15728.
6
T.R. Beck, J. Phys. Chem. 1969, 73, 466.
7
K.F. Lin, T.R. Beck, J. Electrochem. Soc. 1976, 123, 1145.
8
I. Morcos, H. Fischer, J. Electroanal. Chem. 1968, 17, 7.
9
O.J. Murphy, J.S. Wainright, Langmuir 1989, 5, 519.
10
J.A.M. Sondag-Huethorts, L.G.J. Fokkink, Langmuir 1992, 8, 2560.
11
W. Paik, M.A. Genshaw, J.OM. Bockris, J. Phys. Chem. 1970, 74, 4266.
12
J. Wang, B.M. Ocko, A.J. Davenport, H.S. Isaacs, Phys. Rev. B 1992, 46, 10321.
13
L. Perera, M.L. Berkowitz, J. Phys. Chem. 1993, 97, 13803.
14
J.S. Noh, J.A. Schwarz, J. Colloid Interface Sci. 1989, 130, 157.
15
R. Charmas, W. Piasecki, W. Rudzinski, Langmuir 1995, 11, 3199.
16
P.M. Claesson, P. Herder, P. Stenius, J.C. Eriksson, R.M. Pashley, J. Colloid Interface Sci. 1986, 109,
31; P.J. Scales, T.W. Healy, D.F. Evans, J. Colloid Interface Sci. 1988, 124, 391; S. Nishimura, H.
Tateyama, K. Tsunematsu, K. Jinnai, J. Colloid Interface Sci. 1992, 152, 359.
17
D.L. Harame, L.J. Bousse, J.D. Shott, J.D. Meindl, IEEE Trans. Electron Devices 1987, 34, 1700; M.
Grattarola, G. Massobrio, S. Martinoia, IEEE Trans. Electron Devices 1992, 39, 813.
18
R.J. Hunter, Zeta Potential in Colloid Science, Principles and Applications, Academic Press, London,
1981.
19
H.-J. Jakobasch; F. Simon, C. Werner, C. Bellmann, Technisches Messen 1996, 63, 439.
20
M. Smoluchowski, in: Handbuch der Electrizitt und des Magnetismus Vol. II, S. 366, Barth, Leipzig,
1921.
21
R.J. Hunter, Foundations of Colloid Science, Vol. 1, Oxford University Press, S. 557.
22
D.C. Henry, Proc. Roy. Soc. London A 1931, 133, 106.
23
A.S. Dukhin, T.G.M. van den Ven, J. Colloid Interface Sci. 1994, 165, 9.
- 1-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
7 SURFACE FORCES
Surface forces are the forces which act between the surfaces of two solids. The
gravitational force is not a surface force since its origin, the mass, is distributed over
the whole volume.
In the strict sense also the van der Waals
a
force should not be counted to the group. It can,
however, be formally treated as a surface force. Moreover, it is important in many applications
of the research on colloids.
Surface forces are of great importance. They not only determine the behaviour of
dispersions, emulsions, foams and aerosols, but also that of phenomena like wetting,
adhesion and adsorption. Useful reviews are refs. 1,2.
7.1 Forces between atoms and molecules
The cause for all surface forces is the Coulomb force. The potential energy between
two electrical charges is
d
Q Q
V

=
0
2 1
4

Q
1
and Q
2
are the charges in C, d is the distance between them.
With Coulombs force law we can deduce the potential energy between a dipole and a
single charge
2
0
4
cos
d
Q
V



d

Q

The energy depends on the orientation. is the dipole moment. Often the old unit
Debye is used. 1 Debye = 1 unity charge 0.21 apart = 3.33610
-30
Cm.
In practice the molecule with the dipole moment is often mobile. For a freely rotating
dipole, which interacts with a single charge, the energy is
4 2
0
2 2
) 4 ( 6 d kT
Q
V


Also two free rotating dipoles interact with each other. This is often refered to as the
"Keesom energy":
b

6 2
0
2
2
2
1
) 4 ( 3 d kT
V

1

2
d

When a charge approaches a molecule without a static dipole moment, then all
energies considered so far would be zero. Nevertheless, there is an attractive force.
Explanation: The monopole induces a charge shift in the non-polar molecule. Thus an
induced dipole moment arises, which interacts with the monopole. The energy is

a
Johannes Diderik van der Waals, 1837-1923, Dutch physicist, professor in Amsterdam.
b
Wilhelmus Hendrik Keesom, 1876-1956, Dutch physicist, professor in Utrecht and Leiden.
- 2-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
4 2
0
2
) 4 ( 2 d
Q
V


is the polarisability in C
2
m
2
J
-1
. The polarisability is defined by E
ind
= .
Also a molecule with a static dipole moment interacts with a polarisable molecule. If
the dipole can freely rotate the energy is
6 2
0
2
) 4 ( d
V



This is called the "Debye energy".
All energies considered so far can be calculated using classical physics. Unfortunately they do
not convincingly explain the interaction between molecules in a gas. They for instance do not
explain the attraction between two nitrogen molecules. This interaction could be understood
only in the context of quantum mechanics. The static dipole moment, which occurs in the
equation of the Debye energy, is replaced by a fluctuating dipole moment in quantum
mechanics. The fluctuating dipole polarises neighbouring molecules, which leads to an
attraction.
The energy between two molecules with the ionisation energies h
1
and h
2
amounts
to
( )
2 1
2 1
6 2
0
2 1
) 4 (
2
3

=
h
d
V
It is called the "London
c
dispersion energy". The equation is an approximation. It is
crucial that the energy grows proportional with the polarisability of the two atoms, and
that the optical characteristics enter in the form of the excitation frequencies.
The names for the three last interactions are adopted in honour of the three
gentlemen, who helped understanding the intermolecular interactions.
3,4,5

7.2 The vdW force between two molecules
7.2.1 Contributions to the VdW force
The Van der Waals force is the Keesom plus the Debye plus the London dispersion
interaction, i.e. all the terms which consider dipole interactions. All three terms contain
the same distance dependency: The potential energy decreases with d
-6
. Usually the
London dispersion term is dominating. In the following table the contributions of the
individual terms for some gases are listed.
Polarisa-
bility

/4
0

(10
-30
m
3
)
Perm.
dipole
moment

(D)
Keesom
V
Kees
r
6


(10
-79
Jm
6
)
Debye
V
Debye
r
6


(10
-79
Jm
6
)
London
V
Lon
r
6


(10
-79
Jm
6
)
Total
vdW

Theory
(10
-79
Jm
6
)
Energy

From gas
state eq.
(10
-79
Jm
6
)
Ne 0.39 0 0 0 4 4 4
CH
4
2.60 0 0 0 102 102 101
HCl 2.63 1.08 11 6 106 123 157
HBr 3.61 0.78 3 4 182 189 207

c
Fritz London, 1900-1954, American physicist of german origin, Professor in Durham.
- 3-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
HI 5.44 0.38 0.2 2 370 372 350
CH
3
Cl 4.56 1.87 101 32 282 415 509
NH
3
2.26 1.47 38 10 63 111 162
H
2
O 1.48 1.85 96 10 33 139 174

7.2.2 Retarded VdW force
So far assumed that the molecules stay so close to each other that the propagation of the
electric field is instantaneous. In a molecule a dipole moment arises, which generates an
electric field. The electric field expands with the speed of light. It polarises a second molecule,
whose dipole moment in turn causes an electric field that reaches the first molecule with light
speed. In this way the two molecules interact.
The process takes place as calculated only if the electric field has enough time to cover the
distance d between the molecules. It takes a time t=d/c. c is the speed of light. If the alignment
of the first dipole is quicker than t, then the interaction becomes weaker. The time during
which the dipole moment changes is about 1/. Hence, only if

1
<
c
d

the interaction takes place as considered. From c= follows < d . Consequence:
For distances greater than 10 - 100 nm the VdW force drops more rapidly (i.e. for
molecules with d
-7
) as for smaller distances. The effect is known as retardation.
7.3 The vdW force between macroscopic solids
6

There are two approaches to calculate the vdW force between extended solids: The
microscopic and the macroscopic.
7.3.1 Microscopic calculation
The potential energy of the interaction between molecule A and molecule B is
6
) (
d
C
d V
AB
AB
=
The minus sign arises because of the attraction.
In order to determine the interaction between
macroscopic solids, in the first step we calculate
the vdW energy between a molecule A and an
infinitely extended surface. This is also important
to understand the adsorption of gas molecules on
surfaces. For this we sum the vdW energy
between the considered molecule and all
molecules in the solid B. Practically this is done
via an integration of the molecular density
B
over
the entire volume of the solid:
( )


+ +
= =
0 0
3
2 2
6
/
) (
2
'
r x d
rdrdx
C dV
d
C V
B AB
B
AB plane Mol


With ) ( 2
2
r d rdr = we get
d
d
x
r
- 4-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
( )
( )

+
=
+
=
(
(

+ +
=
+ +
=


0
3
0
4
0
0
2
2 2
0 0
3
2 2
2
/
) ( 3
1
2 ) (
1
2
) ( 2
1
) (
) (
x d
C dx
x d
C
dx
r x d
C
dx
r x d
r d
C V
B AB B AB
B AB
B AB plane Mol




Therefore
3
/
6d
C
V
B AB
plane Mol

=
The energy decreases less steeply than the energy between molecules! This can be easily
understood. The interaction is not only determined by the molecules in the solid B, which are
closest to A (a), but also by the parts of the solid, which are further away. With this we talk
about surface molecules (b) and also molecules, which are deep under the surface of the solid
(c). If the distance of the molecule A changes from d to d', then the contribution of the vdW
interaction changes relatively strong with (a), i.e. by (d/d)
6
. The interaction with (b) changes
only little, since the relative distance does not change significantly. Also the contribution to (c)
changes only by ((d+x)/(d+x))
6
.
Next we calculate the vdW energy between two infinitly extended solids which are separated
by a parallel gap of the thickness d. Therefore we use the above equation and integrate over all
molecules in the solid A:



+
=
+
=
0
3 3
) (
6
) (
6
x d
dzdydx C
dV
x d
C
V
A B AB A B AB


y and z are the coordinates parallel to the gap. The integral is infinite because the solids are
infinitely large. In order to get realistic values we have to refer to the interaction per unit area.
The vdW energy per unit area is
2
0
2
0
3
12 ) ( 2
1
6
) (
6
d
C
x d
C
x d
dx C
A
V
V
B A AB B A AB B A AB A

=
(
(

+
=
+
=


With the definition of the so-called Hamaker constants
B A AB AB
C A
2
=
we get
(a) (b)


(c)
d d
x
- 5-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
2
12 d
A
V
AB A


In the same way it is possible to calculate the vdW energy between solids having
different geometries. One important case is the interaction between two spheres with
radii R
1
and R
2
. The vdW energy is
7

(
(

|
|
.
|

\
|

+
+

+
+
=
2
2 1
2
2
2 1
2
2
2 1
2
2 1
2
2 1
2
2 1
) (
) (
ln
) (
2
) (
2
6
R R D
R R D
R R D
R R
R R D
R R A
V
AB

D is the distance between the centres of the spheres. The distance between the
surfaces is
2 1
R R D d = . Ref. 8 also consideres a short range repulsion.
If the radii of the spheres are substantially larger than the distance, i.e.
2 1
, R R d << ,
this can be simplified (ref. 2, vol. I, page 182):
2 1
2 1
6 R R
R R
d
A
V
AB
+
=
How large is the VdW force as a function of the distance? The force is equal to the
negative gradient of the potential:
2 1
2 1
2
6
R R
R R
d
A
dd
dV
F
AB
+
= =
Example: a quartz sphere hangs on a second similar quartz sphere caused by the
VdW attraction of
2
12d R A F
AB
= . The attraction thus increases linear with the
radius. The gravitatinal force increases cubically with the radius. As a conseuence,
the behavior of small spheres are dominated the VdW force. For large spheres the
gravity force is more important. Ideally one hangs a sphere onto a second similar
sphere. At which radius is the gravity force so strong that the sphere detaches? The
Hamaker constant is about 510
-20
J, density =3000 kgm
-3
. When the spheres are in
contact one can consider a distance of 2 , which corresponds to a typical interatomic
distance.
Gravitational force
3
2 2
5 3
10 23 , 1
3
4
R
s m
kg
g R =
VdW attraction R
m
J
R
m
J
=

2 2 10
20
104 , 0
) 10 2 ( 12
10 5

The two are equal for m m R 919
10 23 , 1
104 , 0
2
5
=

=
7.3.2 Macroscopic calculation - Lifshitz theory
In the microscopic consideration additivity was assumed. This is not correct: The vdW
force between two molecules is changed by the presence of a third molecule. For
example the polarisability can change. In the macroscopic theory, developed by
Lifshitz, additivity is not assumed. One neglects the atomic structure and the solids
are treated as continuous with certain optical properties. Fortunately, the same
expressions are obtained, especially the distance dependencies turn out to be
correct. Only the Hamaker constant is calculated in a different way.
- 6-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
The Hamaker constant is the sum over many frequencies. The sum can be converted
into an integral. Finally, for a solid 1 which interacts with a solid 2 over a medium 3,
the non-retarded Hamaker constant is obtained:
A kT
h i i
i i
i i
i i
d
132
1 3
1 3
2 3
2 3
1 3
1 3
2 3
2 3
3
4
0 0
0 0
0 0
0 0
3
4
1


+
|
\

|
.
|

+
|
\

|
.
|
+

+
|
\

|
.
|

+
|
\

|
.
|

( ) ( )
( ) ( )
( ) ( )
( ) ( )
( ) ( )
( ) ( )
( ) ( )
( ) ( )

A complete equation of the vdW force, which also contains retarded parts, is
described in Dzyaloshinskii et al.
9

The first term in the equation for the Hamaker constant is the entropy term. It
corresponds to the Keesom plus the Debye energy. It plays an important role for
forces in water since water molecules have a strong dipole moment. Usually the other
terms dominate, i.e. the integral. Since
13
1
10 4 2 = h kT Hz is relatively large,
only the electrical states contribute significantly.
For the dielectricity constants there are fortunately relatively simple expressions.
Often we can apply
( ) ( )


( ) i
C n
IR
IR e
= +
+
+

+
1
1
1
1
2
2
2

As an example the integrand for lipid in interacting with lipid across a water-filled gap
is shown in the following illustration. Spontaneously one expects that the small
frequencies contribute substantially, since in that range the dielectricity constants
differ strongly. That, however, is not the case. The integral goes linear over the
frequency. Therefore the behaviour at high frequencies dominates.
Dielectricity constants of lipid and water and dashed
( )
1 3 1 3
2
( ) ( ) ( ) ( ) i i i i + versus frequency in Hz.
0.1
1
10
100
1.00E+09 1.00E+11 1.00E+13 1.00E+15 1.00E+17
0.1
1
10
0.00E+00 4.00E+15 8.00E+15

If two different media interact over a third medium, then the vdW force is not always
attractive. Repulsive forces were for instance measured for the interaction of silicon
nitride with silicon oxide in diiodomethane.
10
An important example of a repulsive vdW
force is the force between a solid particle interacting in water with an air bubble. In the
- 7-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
case of thin liquid films on solid surfaces there is often a repulsive VdW force
between the solid-liquid and the liquid-gas interface.
11

In the following table non-retarded Hamaker constants are listed for different material
combinations. Hamaker constants, calculated from spectroscopic data, are found in
many publications.
12,13,14,15,16
A recent overview of Hamaker constants of inorganic
materials is in ref.17.
Medium 1 Medium 3 Medium 2 Calculated
Hamaker
Constant
10
-20
J
Measured
Hamaker
constant
10
-20
J
Au/Ag/Cu Vacuum Au/Ag/Cu 20-50
12

Au/Ag/Cu Water Au/Ag/Cu 10-13
12

Mica Vacuum Mica 7,0
14
10-13.5
14

Al
2
O
3
Vacuum Al
2
O
3
14,5
14
-15,2
17

SiO
2
Vacuum SiO
2
6,4
17
-6,6
14
5-6
14

Si
3
N
4
Vacuum Si
3
N
4
16,2
17
-17,4
14

TiO
2
Vacuum TiO
2
14,3
17
-17,3
14

Perfluorocarb. Vacuum Perfluorocarb. 3,4-6,0
15

Carbonhydr. Vacuum Carbonhydr. 2,6-3,0
15

Polymer Air Polymer 4-7
16

Mica Water Mica 0,29
14
2.2
14

Al
2
O
3
Water Al
2
O
3
2,8
14
-4,7
17
6.7
14

SiO
2
Water SiO
2
0,16
14
-1,51
17

Si
3
N
4
Water Si
3
N
4
4,6
14
-5,9
17
2-8
14

TiO
2
Water TiO
2
5,4
17
-6,0
14
4-8
14

Perfluorocarb. Water Perfluorocarb. 0,36-0,74
15

Carbonhydr. Water Carbonhydr. 0,39-0,44
15
0.3-0.6
Polymer Water Polymer 0,2-0,8
16

Polystyrene Water Polystyrene 0,9-1,3
SiO
2
Water Air -1,0
BSA (Albumin) Water SiO
2
0,7
18


An important question concerns the effect of electrolyte in aqueous solutions. Ions
hinder the water molecules in their hydration shell to orient in an external electric field.
However, they only effect the first term in the equation of the Hamaker constant. The
first term can lead substantially to an interaction in water. In addition, the salt
concentration is often much higher on surfaces than in the volume phase.
Consequence: The dielectric constant can be smaller than in the volume phase.
Example:
19
For the interaction of lipid bilayers over a layer of pure water a Hamaker
constant of 7,510
-21
J is calculated. Measured was only a value of 310
-21
J. The
reason is probably a reduction of the first term by the presence of ions.
- 8-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
From the equation of the Hamaker constant a useful approximation follows:
A A A
123 131 232

If we know the Hamaker constants of the single materials, one can estimate the
Hamaker constant for the interaction between the different substances. This relation
will be important later. With new spectroscopic data one can examine the accuracy of
the approximation. A comparison between precisely calculated Hamaker constants
and estimated values is listed in the following table. The data refer to ref.17. In most
cases the estimation fits quite well. In some cases, however, there are drastic
deviations.
Medium 1 Medium 3 Medium 2 Exactly
10
-20
J
Estimated
10
-20
J
SiO
2
(Silica) Vacuum Mica 8,01 9,35
SiO
2
(Silica) Water Mica 0,69 1,17
SiO
2
(Silica) Vacuum TiO
2
9,46 11,6
SiO
2
(Silica) Water TiO
2
0,69 2,37
SiO
2
(Silica) Vacuum KCl 5,94 6,99
SiO
2
(Silica) Vacuum Diamond 13,7 16,2
SiO
2
(Silica) Water Diamond 1,71 3,75
Al
2
O
3
Vacuum Mica 12,2 12,2
Al
2
O
3
Water Mica 2,15 2,22
Al
2
O
3
Vacuum Si
3
N
4
16,5 16,5
Al
2
O
3
Water Si
3
N
4
4,43 4,48
7.3.3 Surface energy and Hamaker constant
In order to calculate the surface energy of molecular crystals, one can imagine the
following experiment: A crystal is split and the two parts are separated to infinite
distance:
D
0

The work per surface necessary for it is W A D = 12
0
2
. D
0
is the distance between
two atoms. Thereby two surfaces are formed. The surface work necessary for it
amounts to 2
S
. Equating the results leads to

S
A
D
=
24
0
2

Examples:
- 9-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
Helium.
1
As an interatomic distance often a value of 1.6 is used. If one calculates
the surface energy from the Hamaker constant of 5.710
-22
J, then a surface energy of
0.29 mJ/m
2
results. This value is in good agreement with measured values for liquid
helium of 0.12-0.35 mJ/m
2
.
Teflon.
20
Atomic spacing is 1.7 , Hamaker constant in vacuum is 3.4-3.910
-20
J,
surface energy is 15-16 mJ/m
2
.
7.4 Concepts for the description of surface forces
7.4.1 The Derjaguin-approximation
Often it is easy to calculate the energy per unit area V
A
between two infinitely
extended solids which are separated by a gap of width x. From this it is possible to
approximate the potential energy V between solids of any arbitrary form and which
are at the distance d:
21

= dA x V d V
A
) ( ) (
The integration takes place over the entire surface of the solid. Often one has to deal
with rotational-symmetric configurations. Then it is reasonable to integrate in cylin-
drical coordinates:
( )

=
0
) ( 2 ) ( dr r r x V d V
A

The approximation is good if the characteristic decay length of the surface force is
small in comparison to the curvature of the surfaces.
Derjaguin has calculated the interaction between two spheres with radius R.
1
In that
case x and r are related by
dx r R rdr dr
r R
r
dx r R R d r x =

= + =
2 2
2 2
2 2
2
2
2 2 ) (
d
r
R
x(r)
r
R

If the range of the interaction is substantially smaller than R, then one needs to
consider only the outer caps of the two spheres, and only the contributions with a
small r are effective. We can simplify
Rdx dx r R rdr =
2 2
2
and the integral becomes
- 10-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
( )

=
d
A
dx x V R d V ) (
Example: For two spheres with radii 1 m which attract each other due to the vdW
force the approximation is quite good. In contact the solids have a typical distance of
3 (double atomic radius). After 3 nm the vdW force has already droped to 1/100
(supposing that V d 1
2
).
From the potential energy one can calculate the force between two spheres:
( ) ) (d V R dx x V
dd
d
R
dd
dV
F
A
d
A
=
|
|
.
|

\
|
= =


since 0 ) ( =
A
V . In summary: Assuming that the range of the interaction is
substantially smaller than R, there is a simple relationship between the potential
energy per surface V
A
and the force between two spheres.
Example: Calculate the vdW-force between two spheres from the vdW energy per
unit area
2
12 x A V
A
= . From the Derjaguin approximation we can directly write
2
12d
AR
F =
One can compare this with the results from the last chapter. There the vdW energy
between two spheres was approximated with d AR V 12 = . Setting dd dV F =
one immediately gets the same result.
Derjaguins approximation (in the strict sense) refers to the interaction between two
spheres with the same shape. That is an important case, because in this way the
behaviour of many monodisperse dispersions can be treated. White
22
extended the
approximation to solids with an arbitrary shape. For a sphere with radius R which is at
the distance d from a planar surface one gets
) ( 2 d V R F
A
=
The approximation has a fundamental consequence. Now it is possible to divide the
force (or energy) between two solids (which generally depends on the form, on the
materials and on the distance) into a pure geometrical factor (R or 2R, respectively)
and into a material and distance dependent term V
A
(x). Thus it is possible to describe
the interaction independent of geometry!
In the following only V
A
is considered, i.e. the interaction between two parallel
surfaces which limit infinitely extended solids.
7.4.2 The "disjoining pressure"
The term "disjoining pressure" was introduced in 1936 by Derjaguin. The disjoining
pressure is equal to the difference between the pressure within a film between two
surfaces and the pressure in the bulk phase. It is defined as:
=
1
A
G
x
A T V

, ,

- 11-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001

x
Pressure p
0
of
the bulk phase
Pressure in the
film
p p = +
0


The film is in equilibrium with the bulk phase. The disjoing pressure is not in
contradiction to the formalism of the surface forces. It is sometimes more useful to
think in terms of disjoining pressure If for instance the stability of a liquid film is
examined it is more appropriate to think in terms of an increased pressure in the film
rather than of the fact that this pressure is caused by the force between two water-air
interfaces.
7.5 Measurement of surface forces
In direct methods two solids are brought close together and the force between them
is measured as function of the distance. A new overview is ref.23. Problems in the
direct measurement of surface forces are:
The roughness of the surfaces must be small. The roughness limits the space
resolution. In particular this puts limits for small distances. Thereby is of interest
the roughness of surfaces which interact with each other.
The surfaces must be clean and free from dust.
Retarded vdW forces were measured for the first time between two quartz plates and
later between two metal wires.
A relatively universal method is the surface force appartus (SFA).
24,25
The
development of the surface forces apparatus was a big step forward because it
allowed to measure directly the force law in liquids and vapors at ngstrm resolution
level.
26
The SFA contains two crossed atomically smooth mica cylinders of roughly 1
cm radius between which the interaction forces are measured. One mica cylinder is
mounted to a piezoelectric translator. With this translator the distance is adjusted.
The other mica surface is mounted to a spring of known and adjustable spring
constant. The separation between the two surfaces is measured by use of an optical
technique using multiple beam interference fringes. Knowing the position of one
cylinder and the separation to the surface of the second cylinder the deflection of the
spring and the force can be calculated.

The second device with which surface forces can be measured directly and relatively
universally is the atomic force microscope.
27
In the atomic force microscope one
measures the force between a sample surface and a micro-fabricated tip, placed at
the end of an about 100 m long and 0.4-10 m thick cantilever. Alternatively colloidal
particles are fixed on the cantilever. With the atomic force microscope the forces
- 12-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
between surfaces and colloidal particles can be directly measured in liquid.
28,29
The
practical advantage is that measurements are quick and simple. Even better, the
interacting surfaces are substantially smaller than in SFA. Thus the problem of
surface roughness and of contamination is reduced. This again allows to examine
surfaces of different materials.
An important, but not so direct procedure to measure surface forces in liquid is the
osmotic stress method.
30,31
From a dispersion, where the particles to be examined
are in, is extracted liquid. This is done either via osmotic pressure or mechanically, by
pressing the liquid through a semipermeable membrane. The average distance
between the particles is determined, e.g., by X-ray diffraction. Finally pressure-
distance curves are obtained. With this procedure, e.g., was measured the force
between DNA molecules.
7.6 The electrostatic force
The interaction between two charged surfaces in liquid (usually water) depends on
the surface charge. We consider only the linear case, that is we calculate V
A
(x). If the
geometry of the solids must explicitly be considered the calculations become
substantially more complex. In addition only monovalent salts are considered (such
as KCl, NaI, etc). An extension to other salts can easily be made.
7.6.1 General equations
When calculating V
A
(x) two approaches are used: Either the change of Gibbs free
energy gained during the formation of the electrical double layer, is calculated, or the
disjoining pressure in the gap is determined. Both conditions give the same result.
Depending upon conditions and on a personal preference, one or the other method is
more suitable. Thus Hogg, Healy and Fuerstenau used the first condition, in order to
calculate the energy between two spheres with constant, but different, surface
potentials.
32
Parsegian and Gingell chose the way of the osmotic pressure, in order to
determine V
A
for two different surfaces with different boundary conditions.
33

We start with the first approach and calculate the change in Gibbs free energy for two
approaching double-layers. Gibbs free energy of the electrical double layer is


0
0
0
' ' d
For two homogeneous double-layers which are infinitely separated the Gibbs free
energy per unit area is
Laser light
Photo
detector
Particle fixed on the cantilever
Sample
- 13-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001

=
0
0
0
' ' 2 d G
If the two surface approach each other up to a distance x the Gibbs free energy
changes. Surface charge and potential depend on distance:

=
) (
0
0
0
' )' ( 2 ) (
x
d x x G
Hence, the interaction potential per unit area is

= = G x G G V
A
) (
The force per unit area is
dx G d =
In order to get an intuitive understanding we considers what happens to the surface
charges when the two surfaces are brought closer, holding the potential constant.
Remember: Surface charge and potential are related by
0
0
=


d
d

according to Gau law of electrostatics
d
. If two surfaces approach each other the
potential gradient d d decreases. The surface charge density decreases
accordingly. If for instance AgI particles are brought together, then I
-
ions are
removed from the surface and the surface charge density decreases. During the
approach of SiO
2
particles previously dissociated H
+
ions bind again. Neutral
hydroxide groups are formed and the negative surface charge becomes weaker. This
reduction of the surface charge is energetically unfavourable. Consequence: A
repulsive force.
The interaction can alternatively be calculated over the osmotic pressure or over the
disjoining pressure:

=
x
A
dx x x V ' ) ' ( ) (
This way is simpler in most cases. In the context of the Poisson-Boltzmann theory
one can deduce a simple expression for the disjoining pressure. For the linear case
and for a monovalent salt, the Poisson-Boltzmann equation is:

d
At this place is probably necessary to give an explanation about the symboles d, x and . d is the
shortest distance between solids of arbitrary geometry. x is the thickness of the gap between two
infinitely extended solids. It makes sense to distinguish between the two symbols (see illustration in
the chapter Derjaguin-approximation). is a place parameter, which does not give a distance, but
the coordinate inside the gap. With constant gap thickness x, the potential changes with .

0

x

0

x
- 14-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
0
2
2
0
0
=


|
|
|
.
|

\
|

dx
d
e e en
T k
e
T k
e
B B

The first term can be also written as ( ) T k e Tn k
B B
sinh 2
0
. Integration of the
equation leads to
p
dx
d
e e Tn k
T k
e
T k
e
B
B B
= |
.
|

\
|

|
|
|
.
|

\
|
+

2
0 0
2


So far p is only an integration constant. It has a physical meaning: p corresponds to
the pressure in the solution. The first term describes the osmotic pressure caused by
the increased number of particles (ions). The second term corresponds to the
electrostatic force, caused by the electric field of one surface which affects charges
on the other surface. This is calculated by the Maxwell stress tensor.
As force-per-surface only the difference of the osmotic pressure between the two
levels and the osmotic pressure in the reservoir is effective. In principle one must
imagine the "infinitely" extended gap in contact with
an infinitely large reservoir:
Therefore the osmotic pressure in the reservoir
0
2 Tn k
B
must still be subtracted from p, in order to
get the disjoining pressure. Finally, for the force-per-
surface one obtains
2
0 0
2
2
|
|
.
|

\
|

|
|
|
.
|

\
|
+ =

d
d
e e Tn k
T k
e
T k
e
B
B B

In order to determine the force in a specific situation, first the potential must be
calculated. This is done with the Poisson-Boltzmann equation. In a second step the
force-per-surface is calculated. It does not matter in which point we calculate , the
value must be the same for every .
The potential curve is not fixed explicitely by the Poisson-Boltzmann equation.
Additionally to the differential equation, boundary conditions must be fulfilled. Two
often used boundary conditions are:
Constant potential : with the approach of the solids, the surface potentials are
constant, i.e.
1
) 0 ( = = and
2
) ( = = x .
Constant load : during the approach the surface charge densities are constant. The
surface charge density the increase of the potential at the surface:
0 1
) 0 ( = = d d and
0 2
) ( = = d x d .
For large distances, i.e. for d
D
>> , both boundary conditions lead to identical forces.
In the case of small distances, with the constant charge conditions leads to more
repulsive forces as the constant potential condition.
7.6.2 Electrostatic interaction between two identical surfaces
An important case is the interaction between two identical parallel surfaces of two
infinitely extended solids:
x
- 15-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001

m
x


The interaction of identical solids is important for the aggregation of colloids or the
coagulation of emulsions. For identical solids the surface potential
0
on both
surfaces is similar. In between the potential decreases. In the center the gradient
must be zero for symmetry reasons, i.e. ( ) 0 2 = = d x d . Therefore the disjoining
pressure in the center is given by the osmotic pressure. The Maxwell stress tensor
does not contribute. Towards the two surfaces the osmotic pressure increases. It is
however compensated by the Maxwell stress tensor. Since the pressure must be the
same everywhere at the equilibrium, it applies:
|
|
|
.
|

\
|
+ =

2 ) (
0 T k
e
T k
e
B
B
m
B
m
e e Tn k x

m
is the potential in the center. depends on the gap thickness x, since
m
changes
with x. One can thus determine the force from the potential in the center.
The exponential functions can be developed in a series. For small potentials we
neglect all term higher than the quadratic one:
2
2
0 2
2 0
2 2
0
2
2 ... 1 ... 1
m
D
m
B
B
m
B
m
B
m
B
m
B
T k
e n
T k
e
T k
e
T k
e
T k
e
Tn k
=
|
|
.
|

\
|

|
|
.
|

\
|
+

+ +
|
|
.
|

\
|
+

+ =


If the double-layers of the two opposing surfaces overlap only slightly (d
D
>> ), then
we can approximate
) 2 ( ' 2 x
m
=
' is the potential of an isolated double layer. Fr ' one can use varoius exact
functions. The simplest approximation, T k e
B
<<
0
( )
D
= exp
0
, leads
to a repulsive force per area of
D
x
D
e x


=
2
0
2
0
2
) (
In order to calculate the energy per unit area we still have to integrate:
D D
x
D
x
x
D
x
A
e dx e dx x x V

= =

2
0 0 '
2
2
0 0
2
'
2
' ) ' ( ) (
If we uses a ' wich applies also for larger potentials we get
D
B
B
x
T k e
T k e
D B
A
e
e
e
T k n x V

|
|
.
|

\
|
+

=
2
2
2
0
1
1
64 ) (
0
0

- 16-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
7.6.3 The DLVO theory
It is been known already for over 100 years that many aqueous dispersions
precipitate upon addition of salt. Schulze and Hardy observed that most dispersions
precipitate with a concentration of 25-150 mM of monovalent counterions.
34,35
For
divalent ions they found far smaller precipitation concentrations of 0.5-2 mM. Trivalent
counterions lead to precipitation at even lower concentrations of 0.01-0.1 mM. Gold
colloids are e.g. stable in NaCl solution, as long as the NaCl concentration does not
exceed 24 mM. If the solution contains more NaCl, then the gold particles aggregate
and precipitate. The appropriate concentrations for KNO
3
, CaCl
2
and BaCl
2
are 23 mM,
0.41 mM and 0.35 M.
36

This aggregation behaviour can be understood as follows: The gold particles are
negatively charged and repell each other. With increasing salt concentration the
electrostatic repulsion decreases. The particles, which move around thermally, more
frequently have a chance to approach each other to few nm. Then the vdW attraction
causes them to aggregate. Since divalent counterions weaken the electrostatic
repulsion more effectively than monovalent counterions, only small concentrations of
CaCl
2
and BaCl
2
are necessary for precipitation.
Quantitatively the aggregation of colloids was explained by Derjaguin, Landau,
Verwey and Overbeek.
37,38
For this reason the theory is called DLVO theory. If we
take into account the vdW attraction and the electrostatic repulsion it is possible to
approximate the energy per surface of two infinitely extended solids which are
separated by a gap x by
2
2
2
2
0
12
1
1
1
64 ) (
0
0
x
A
e
e
e
kT n x V
D
x
kT e
kT e
D
A

|
|
.
|

\
|
+



The following illustration shows the energy per unit area plotted versus the gap
thickness. The gap is filled with aqueous solution. Both surfaces have the same
surface potential of 30 mV. We see that
with rising salt concentration the repulsion decreases
with small distances the vdW attraction always wins

0 2 4 6 8 10
-0,2
0,0
0,2
0,4
0,6
0,8
A=10
-26
J
30 mV
100 mM
30 mM
10 mM
3 mM
distance / nm
e
n
e
r
g
y

p
e
r

a
r
e
a
- 17-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
7.7 Short range forces in solution
So far all forces were calculated with the help of continuum theory. The individual
nature of the molecules involved, their discrete size and shape, in particular that of
the solvent, were neglected. Often the structure close to the interface is different from
that in the volume phase. This region typically extends over few molecular diameters.
The special structure of the water causes two effects:
Hydration repulsion between hydrophilic surfaces
39,40

Hydrophobic attraction between hydrophobic surfaces
Both decrease exponentially with the distance between the surfaces, with typical
decay lengths of 1-2 nm.
Between hydrophobic surfaces there is a further attraction with large range
(10-100 nm). The cause for this force is not clear. A suggestion is that during
approximation of the surfaces gas bubbles form spontaneously. Those cause
an attraction.
7.8 Sterical interaction
7.8.1 Observations
Many dispersions are stabilised by polymers. The underlying interaction is often
called steric force. For the understanding of steric interactions it is necessary to know
some fundamentals of polymer physics.
The T

-temperature
To describe the behaviour of linear polymers in solution often the polymer is
considered to consist of a chain of n chain links. Each link has a length l. This
parameter l can correspond to the length of a monomer, it can also be larger. The
effective radius of an undisturbed polymer in solution, that is the square root of the
average square distance between the the two ends, is two times
R
l n
l M M
g
= =
6 6
0

R
g
: gyration radius; M: mass of the polymer; M
0
: mass of a monomer. Typical
example: with M=10
6
, M
0
=100, l=1nm results a length of the stretched polymer of
10
6
/1001nm=10m while R
g
=41nm.
The polymer has this radius only if the individual links can move freely. This happens
only in an "ideal" solvent. In an "ideal" solvent there is no interaction of the single links
with each other. That is the case when the interaction of the links with the solvent is
as strong as the interaction beweetn twop links. In a real solvent the actual radius can
be larger or smaller than R
g
. The radius is called also Flory radius R
F
. Both are related
over that intermolecular expansion factor :
R R
F g
=
In a "good" solvent a repulsive force acts between the monomers. The polymer swells
and is larger than unity. In a "bad" solvent the monomers attract each other, the
polymer shrinks and is smaller than unity.
- 18-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
A bad solvent often becomes a good solvent, if the temperature rises. The
temperature, at which the polymer behaves ideally, i.e. =1, is called Theta-
temperature T

. The ideal solvent is called also Theta solvent.


7.8.2 Force between surfaces which are coated with polymers
An overview is ref. 41. Already the old Egyptians knew that one can keep soot
particles dispersed, if they are asperged with albumin. In this way ink was made.
Steric stabilisation of dispersions has nowadays an important technical meaning.
Direct quantitative measures were made with the SFA
42,43,44,45,46,47
and the atomic
force microscope.
48,49,50,51,52,53,54

The force between surfaces, which are coated with polymers, depends on many
factors:
Quality of the solvent. In good solvents it come repulsive, in bad solvants
attractive. Moreover, it has to be considered that in good solvents little
polymer adsorbs to surface.
It is important how the polymer is bound to the surface. Physisorbed
polymers can for instance diffuse laterally. They have possibly more
binding sites and above all homopolymers can form loops. The adsorption
time plays a crucial role. Often bound polymers stand in a flow equilibrium
with solved polymers. With chemisorbed (graftet) polymers this is not the
case.
The density of the polymers at the surface is an important factor. With very
close packing (brush) the polymers and the steric force act over lengths,
which are substantially larger than the Flory radius. Polymer brushes can
be realised practically only by synthesis of the polymer directly on the
surface.
R
F
L>R
F
Graftet Adsorbed Brush

There is no simple theory which describes the entire steric interaction. Different
components contribute to the force and depending upon the situation dominate the
total force.
The most important interaction is repulsive and of entropic origin. Reason is the
reduced configuration entropy of the polymer chains. If the thermal movement of
a polymer chain at a surface is limited by approach of another surface, then the
entropy of the individual polymer chain decreases.
The entropic repulsion was calculated by different authors.
55,56,57
A simple
approximation, which is based on considerations of de Gennes, is:
58,48,54

L
x
e kT
2 2
3


is the density of the polymers on the surface in number per m
2
. L is the equilibrium
thickness of the polymere layer.
- 19-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
Intersegment forces, caused by the direct interaction between segments of
polymers of the two apporaching surfaces can be attractive. This interaction depends
strongly on the solvent. Below T


the interaction among the
monomers is stronger than the inter-
action of the monomers with the
solvent.
Bridging forces. It is also possible
that a polymer binds to both
surfaces. This usually leads to an
attraction at large separations.
Bridging is only effective at low
surface coverage.
The resulting force is schematically
represented in the illustration. It is
stressed once more that everything
is strongly simplified. Steric forces
are complex and difficult to treat.
7.9 Spherical particles in contact
So far we assumed that the surfaces, which interact, are not deformable. In reality all
solids have a finite elasticity. They deform upon contact. This has important
consequences for the aggregation behaviour and the adhesion of particles because
the contact area is larger than one would expect from infinitely hard particles.
Heinrich Hertz
e
layed the basis for the treatment of elastic solids in contact. As a
modell we cosnider two spheres with radii R
1
and R
2
in contact. The two spheres are
from materials with Youngs moduli E
1
and E
2
and with Poisson ratios
1
and
2
. The
two spheres are pressed together with a force F. This could for instance be the
gravitational force. Hertz supposed that no surface forces act between the solids.
Hertz showed that the pressure between the spheres decreases as a quadratic
function with the distance to the contact center. At the outer rim of the contact, at the
distance a, the pressure is zero. If one integrates the pressure over the entire contact

e
Heinrich Hertz, 1857-1894, german physicist, professor in Karlsruhe and Bonn.
F
R
1

R
2
2a

T<T

Force
Distance

R
F
Light
coverage
Strong
coverage
T<T

T>T

T>T

- 20-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
area, then the force F results. On the other hand, one can calculate the contact radius
from the force:
F
E
R
a =
*
*
4
3
3

R* designates the effective particle radius and E* the reduced elasticity constant:
2 1
2 1
*
R R
R R
R
+

= and
2
2
2
1
2
1
1 1
*
1
E E E

+

=
The contact radius a determines at the same time the indentation
*
2
R
a
=
As force distance relation we thus get, with * R a = ,
3
* *
3
4

R E F =
If no external force ( 0 = F ) act the indentation is zero and the contact radius is also
zero. Since no attractive surface forces were considered, there is also no adhesion in
the Hertz model.
This case, 0 = = F , does not occur practically. Reason: Surface forces, e.g. the
van der Waals force, attract the two solids even for 0 = F . Attractive surface forces
were considered by Johnson, Kendall and Robert;
59
their model is called JKR model.
Theiy considered the following: If two solid surfaces come into contact then free
surface disappears. Let us call the associated energy gain per unit area cohesion
work W. The work of cohesion is correlated with a energy loss of the system. Hence,
it would happen spontaneously. At the same time work must be done for the elastic
deformation of the solids. Johnson, Kendall and Robert calculated this elastic
deformation energy with the help of the Hertz theory. The enlargement of the contact
area is limited by the additional necessary elastic deformation energy. The real
contact radius is given by
|
.
|

\
|
+ + + =
2 3
*) 3 ( * 6 * 3
*
*
4
3
WR F WR WR F
E
R
a
For 0 = W one immediately obtains the Hertz result.
In the JKR model a force is necessary to separate two solids. This force is called
adhesive force. The adhesive force amounts to * 2 / 3 WR F
adh
= .
Usually the cohesion work is expressed in terms of the surface energy of the solid
S

by
S
W 2 =
This is a critical assumtpion and could better be consdered a definition of the surface
energy of a solid. Thus we obtain
* 3 R F
S adh
=

The adhesive force increases linearly with the particle radius. It is surprisingly
independent on the elasticity of the materials. Explanation: In a hard material the
deformation of the solid is small. Thus, on one side the contact area and the
- 21-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
attraction are small. On the other hand the repulsive elastic component is small. Soft
materials are strongly deformed. Thus both the attractive surface energy term and the
repulsive elastic term are high.
In the JKR theory it is accepted that surface forces are active only in contact areas. In
the reality surface forces are active also outside of the direct contact. This fact was
considered by Derjaguin, Muller and Toporov in the so-called DMT theory.
60
A
consequence is that a kind of neck or meniscus forms at the contact line. The case of
a hard sphere on a soft surface is shown in the following illustration.
Unfortunately substantial results of the DMT theory cannot be written as convenient
analytic expressions. There is, however, one simple result: For the adhesive force
they obtain:
* 4 R F
S adh
=

Exact analyses show that the two models represent two extrems of the real
situation.
61,62,63
For large, soft solids the JKR model describes the situation
realistically. For small, hard solids it is appropriate to use the DMT model. A criterion,
which model is to be used, results from the height of the neck
3 1
2
2
*
*
|
|
.
|

\
|

E
R
h
S
n


If the neck height is larger than some atomic distances, then the JKR model is more
favourable. With shorter neck heights the DMT model is more suitable.
Example: A SiO
2
-sphere with 20 m diameter sits on a silicon oxide surface. 1.
Estimate the contact radius at negligible external force 2. Calculate the adhesive
force. 3. How much does the contact radius increase if we take the weight of the
sphere into account? E=5,410
10
N/m
2
, =0,17,
S
=50 mN/m, density =3000 kg/m
3
.
R
1
=10 m and R
2
= it follows m R R 10 *
1
= = .
2
10
2 10
3
10 8 , 2 *
/ 10 4 , 5
17 , 0 1
2
*
1
m
N
E
m N
E
=

=
The contact radius is estimated with the JKR model. Without external force:
m a
m
m
N
m N
m
a
7
20 5
2 10
5
3
10 51 , 2
10 59 , 1 10 05 , 0 2 6
/ 10 8 , 2
10
4
3

=
= |
.
|

\
|

=

The weight amounts to
h
n

- 22-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001
N
s
m
m
kg
m g R F
10
2 3
3 15 3
1
10 23 , 1 81 , 9 3000 10
3
4
3
4

= = =
The contact radius is enlarged by the weight to
3 20
11 15 5 10 3 16
2 5 10 5
5 10
10
3 5
3
10 58 , 1
10 88 , 8 10 32 , 2 10 42 , 9 10 23 , 1 10 415 , 8
) 10 05 , 0 6 ( 10 23 , 1 10 05 , 0 12
10 05 , 0 6 10 23 , 1
10 8 , 2
10
4
3
m
m
m
a

=
|
.
|

\
|
+ + + =
|
|
.
|

\
|
+ +
+


Hence, practically the increase due to the weight is negligible.
The calculation of the contact radius was made with the help of the JKR theory. Exact
calculations by Maugis show that the JKR theory can be used for ( ) 10 3 2
3 1
>
S
RE
to calculate the contact radius (not the adhesion!).
64

The neck height amounts to
m
m N
m m N
h
n
10
3 1
2 2 10
5 2 2 2
10 2 , 3
) / 10 8 , 2 (
10 / 05 , 0

=
|
|
.
|

\
|


The neck height is about as large as an atomic diameter. Therefore the DMT model is
suitable. The adhesion therefore is
N m m N F
Adh
3 , 6 10 / 05 , 0 4
5
= =


Experimentally the predictions of the theories for the behaviour of two spheres
in contact can be measured in different ways. Adhesive forces are for in-
stance measured in a centrifuge. The centrifugal force, which is necessary to
loosen particles from a planar surface, is determined.
65,66,67,68,69
The procedure
is, however, limited to particles with R larger than about 5 m. An overview of
older procedures is contained in ref. 70. The JKR theory predicts correct con-
tact radii for relative soft surfaces with effective radii larger than 100 m. This
was shown in direct force measurements with the SFA
71,72
or specifically de-
signed systems. For smaller spheres it was verified with the atomic force
microscope.
73



1
J.N. Israelachvili, Intermolecular and Surface Forces, Academic Press, London, 1992.
2
R.J. Hunter, Foundations of Colloid Science, Vol I+II, Clarendon Press, Oxford, 1995.
3
P. Debye, Physik. Zeitschr. 1920, 21, 178.
4
W.H. Keesom, Physik. Zeitschr. 1921, 22, 129.
5
F. London, Zeitschr. f. Physik 1930, 63, 17.
6
J. Mahanty, B.W. Ninham, Dispersion Forces, New York, Academic Press, 1976.
7
H.C. Hamaker, Physica 1937, 4, 1058.
8
D. Henderson, D.-M. Duh, X. Chu, D. Wasan, J. Colloid Interface Sci. 1997, 185, 265.
9
I.E. Dzyaloshinskii, E.M. Lifshitz, L.P. Pitaevskii, Adv. Phys. 1991, 10, 165.
10
A. Meurk, P.W. Luckham, L. Bergstrm, Langmuir 1997, 13, 3896.
11
L.B. Boinovich, Adv. Colloid Interf. Sci. 1992, 37, 177.
12
V.A. Parsegian, G.H. Weiss, J. Colloid Interface Sci. 1981, 81, 285.
13
J. Visser, Adv. Colloid & Interface Sci. 1972, 3, 331.
14
H.D. Ackler, R.H. French, Y.M. Chiang, J. Colloid Interface Sci. 1996, 179, 460.
15
C.J. Drummond, G. Georgaklis, D.Y.C. Chan, Langmuir 1996, 12, 2617.
16
C.J. Drummond, D.Y.C. Chan, Langmuir 1996, 12, 3356.
- 23-
PCSURF07. DOC BUTT PC2 UNI SI EGEN 10. 01. 2001

17
L. Bergstrm, Adv. Colloid Interface Sci. 1997, 70, 125.
18
C.M. Roth, B.L. Neal, A.M. Lenhoff, Biophys. J. 1996, 70, 977.
19
J.N. Israelachvili, Langmuir 1994, 10, 3369.
20
C.J. Drummond, G. Georgaklis, D.Y.C. Chan, Langmuir 1996, 12, 2617.
21
B.V. Derjaguin, Kolloid Zeitschrift 1934, 69, 155.
22
L.R. White, J. Colloid Interface Sci. 1983, 95, 286.
23
P.M. Claesson, T. Ederth, V. Bergeron, M.W. Rutland, Advances in Colloid and Interface Science
1996, 67, 119.
24
D. Tabor, F.R.S. Winterton, R.H.S. Winterton, Proc. Roy. Soc. A 1969, 312, 435.
25
J.N. Israelachvili, D. Tabor, Proc. Royal Soc. London A 1972, 31, 19.
26
J.N. Israelachvili, G.E. Adams, J. Chem. Soc. Faraday Trans. I 1978, 74, 975.
27
G. Binnig, C.F. Quate, C. Gerber, Phys. Rev Lett. 1986, 56, 930.
28
W.A. Ducker, T.J. Senden, R.M. Pashley, Nature 1991, 353, 239.
29
H.J. Butt, Biophys. J. 1991, 60, 1438.
30
D.M. LeNeveu, R.P. Rand, V.A. Parsegian, Nature 1976, 256, 601.
31
V.A. Parsegian, R.P. Rand, N.L. Fuller, D.C. Rau, Methods in Enzymology 1986, 127, 400.
32
R. Hogg, T.W. Healy, D.W. Fuerstenau, Trans. Faraday Soc. 1966, 62, 1638.
33
V.A. Parsegian, D. Gingell, Biophys. J. 1972, 12, 1192.
34
H. Schulze, J. prakt. Chemie 1882, 25, 431 und 1883, 27, 320.
35
W.B. Hardy, Proc. Roy. Soc. London 1990, 66, 110; Zeitschr. physikal. Chemie 1990, 33, 385.
36
J.T.G. Overbeek, in: Colloid Science, Vol. 1, ed. H.R. Kruyt, Elsevier, Amsterdam, 1952, S. 82.
37
B. Derjaguin, Acta Physicochimica URSS 1939, 10, 333.
38
E.J.W. Verwey, J.T. Overbeek, Theory of the Stability of Lyophobic Colloids, Elsevier, Amsterdam,
1948.
39
J. Israelachvili, H. Wennerstrm, Nature 1996, 379, 219.
40
S. Leikin, V.A. Parsegian, D.C. Rau, R.P. Rand, Ann. Rev. Phys. Chem. 1993, 44, 369.
41
S.S. Patel, M.Tirrell, Ann. Rev. Phys. Chem. 1998, 40, 597.
42
J. Klein, Nature 1980, 288, 248.
43
J. Klein, P.F. Luckham, Macromolecules 1984, 17, 1041.
44
J.N. Israelachvili, M. Tirrel, J. Klein, Y Almog, Macromolecules 1984, 17, 204.
45
Y. Almog, J. Klein, J. Colloid Interface Sci. 1985, 106, 33.
46
J. Marra, M.L. Hair, Macromolecules 1988, 21, 2356.
47
Marra, J., Christenson, H.K. J. Phys. Chem. 1989 93, 7180.
48
S.J. OShea, M.E. Welland, T. Rayment, Langmuir 1993, 9, 1826.
49
A.S. Lea, J.D. Andrade, V. Hlady, Colloids and Surfaces A 1994, 93, 349.
50
G.J.C. Braithwaite, A. Howe, P.F. Luckham, Langmuir 1996, 12, 4224.
51
H.G. Brown, J.H. Hoh, Biochemistry 1997, 49, 15035.
52
X. Chatellier, T.J. Senden, J.F. Joanny, J.M. di Meglio, Europhys. Lett. 1998, 41, 303.
53
T.J. Senden, J.M. di Meglio, P. Auroy, Eur. Phys. J. B 1998, 3, 211.
54
H.-J. Butt, M. Kappl, H. Mller, W. Meyer, J. Rhe, Langmuir, eingereicht.
55
Dolan, A.K.; Edwards, S.F. Proc. Roy. Soc. London A 1974, 337, 509.
56
de Gennes, P.G. Adv. Colloid Interface Sci. 1987, 27, 189.
57
Milner, S.T.; Witten, T.A.; Cates, M.E. Macromolecules 1988, 21, 2610.
58
Israelachvili, J.N. Intermolecular and Surface Forces, 2nd. Ed., Academic Press, London, 1992, p.
295.
59
Johnson, K.L.; Kendall, K.; Roberts, A.D. Proc. Roy. Soc. London A 1971, 324, 301.
60
Derjaguin, B.V.; Muller, V.M.; Toporov, Y.P. J. Colloid Interface Sci. 1975, 53, 314.
61
Muller, V.M.; Yushchenko, V.S.; Derjaguin, B.V. J. Colloid Interface Sci. 1983, 92, 92.
62
Maugis, D. J. Colloid Interface Sci. 1992, 150, 243.
63
Pashley, M.D.; Pethica, J.B.; Tabor, D. Wear 1984, 100, 7.
64
Maugis, D. Langmuir 1995, 11, 679.
65
Larsen, R.I. Amer. Ind. Hyg. Assoc. J. 1958, 19, 256.
66
Zimon, A.D. Kolloid Zh. 1963, 25, 317.
67
Bhme, G.; Kling, W.; Krupp, H.; Lange, H.; Sandstede, G. Zeitschr. Angew. Physik 1964, 16, 486.
68
Booth, S.W.; Newton, J.M. J. Pharm. Pharmacol. 1987, 39, 679.
69
Podczeck, F.; Newton, J.M. J. Pharmaceutical Sci. 1995, 84, 1067.
70
Corn, M. in Aerosol Science (ed. Davies, C.N.) 359-392 (Academic Press, New York, 1966).
71
Israelachvili, J.N.;Perez, E.; Tandon, R.K. J. Colloid Interface Sci. 1980, 78, 260.
72
Horn, R.G.; Israelachvili, J.N.; Pribac, F. J. Colloid Interface Sci. 1987, 115, 480.
73
Heim, L.O.; Blum, J.; Preuss, M.; Butt, H.-J. Phys. Rev. Lett. 1999, 83, 3328.
- 1-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
8 CONTACT ANGLE PHENOMENA AND WETTING
This section is about contact angles or wetting phenomena, that is about the behavior
of a liquid on a solid surface. Wetting phenomena are technically of great relevance.
For the flotation or in washing processes contact angles play a decisive role. In many
applications one would like to achieve complete wetting. Examples are the spreading
of ink on paper, coatings and paints, the spreading of liquids for the treatment of
textiles, the distribution of insecticide sprays on the surface of leaves or on the
epidermis of insects. In other cases one would like to avoid wetting especially of
water, e.g. in order to make clothes rain-proof or to prevent the destruction of road
pavements. A good introduction to the topic is ref. 1. Contact angles are treated in
more detail in the book by Neumann und Spelt.
2

8.1 The Young equation
8.1.1 The macroscopic contact angle
Youngs equation is the basis for a quantitative description of wetting phenomena. If a
drop of a liquid is put on a solid surface there are two possibilities: The liquid wets the
surface completely (contact angle =0) or a finite contact angle is established.
a
In
the second case a three phase contact line is formed. At this line three phases are in
contact: The solid, the liquid, and the vapour. The free energies per unit area of the
interfaces are given by
S
,
L
and
SL
.
The parameters
S
,
L
,
SL
and describe the situation on the macroscopic scale, at a
distance of about 0.1-1 m from the contact line. Close to the contact line, in the core
region, the situation is more complicated. Here, surface forces can change the shape
of the drop and we have to distinguish between the microscopic contact angle (in
the region few nm from the three phase
contact line) and the macroscopic
contact angle (which can be observed
with optical methods including optical
microscopy).
It is possible to calculate the macroscopic
contact angle from the interfacial energies
without knowing what happens in the core
region. This relation between the inter-
facial energies and the contact angle is
the Young equation:
3

SL S L
= cos

8.1.2 Derivation
We derive the Young equation for a typical example, that of a circular drop on a
planar solid surface. Therefore we consider the change of free enhalpie dG when the
drop spreads on the surface. While spreading the radius of the contact zone

a
In a strict sense the terms are used as follows: completely wetting: =0; "wetting: <90
0
; "not
wetting: >90
o
. Like De Gennes
1
, Id like to distinguish only two cases: "wetting for =0 und "partial
wetting for >0.

L
Contact angle
Contact line

SL

- 2-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
increases from a to a. Its
height decreases from h to
h. Is the change in free
enthalpie negative the
process is going to occurr
spontaneously. For positive
dG the drop will contract
spontaneously. In equili-
brium, which is the energe-
tically most favorable
situation, we have dG=0.
To simplify the derivation we assume that the drop is so large that contributions
! of the Laplace pressure
! an elongation of the three-phase contact line
do not lead to a significant change in the Gibbs free energy. For very small drops we
have to introduce correction factors. On the other hand we further assume that
the drop is small enough, that gravitation is negligible. As a consequence the
surface of the drop is spherical.
In summary: We only consider surface contributions to the Gibbs free energy.
When the drop sreads free solid surfaces disappears and is replaced by solid-liquid-
interface. The change in area da a dA
SL
= 2 is accompanied by a change in energy
( )
SL S SL
dA . In addition the surface area of the liquid-gas interface changes.
Elementary geometry tells us that the surface of a spherical cap is
( )
2 2
h a A
L
+ =

A small change of a when the drop spreads is given by
dh h da a dh
h
A
da
a
A
dA
L L
L
+ =

= 2 2

Unfortunately the change in surface area depends on two variables: a and h (we do
not have to consider a change of . It is of second order). However, these two
variables are not independent. In most cases we can safely assume that the volume
of the drop is constant. The volume of a spherical cap is
( ) ( ) dh h a da ah dV h h a V + + = + =
2 2 3 2
2
3
6


Since the volume is supposed to be constant, 0 = dV , we have
( )
2 2
2 2
2
2
h a
ah
da
dh
dh h a da ah
+
= + =


Using Pythagoras law ( )
2 2 2 2 2
2 h Rh a h R a R = + = we get
R
a
h h Rh
ah
da
dh
=
+
=
2 2
2
2

and
R
h

a
R
h
a
a
h
- 3-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
da a da
R
h R
a da
R
h
a da
R
a
h da a dA
L
=

= = cos 2 2 1 2 2 2

Now we can write the total change in Gibbs free energy:
( ) ( ) da a da a dA dA dG
L S SL L L SL S SL
+ = + = cos 2 2

At equilibrium 0 = da dG . From this Youngs equation follows directly.
8.1.3 Additions to Youngs equation
At this point we consider two of the above assumptions.
For small drops the bending of the surface might be significant and the contribution of
the Laplace pressure might become important. Then we have to add an additional
term: ( )
2 1
1 1 R R d V Vdp
L
+ = . In most practical cases however, we can neglect it.
The three-pahse contact line might be curved. For drops smaller than 1 mm a
term has to be added to the Young equation which contains the line tension
(Linienspannung) :
4

a
SL S L

= cos

The line tension can be interpreted as the energy per unit length required to create a
three phase-contact line.
One should keep in mind that Youngs equation is only valid in thermodynamic
equilibrium, hence in the presence of saturated vapor of the liquid.
8.1.4 Surface roughness and inhomogeneity of the surface
Surface roughness. With small contact angles and not too rough surfaces one can
use the approximation
= cos cos R
rough

Here R is the ratio between actual and projected surface. Since R 1 a surface
roughness for <90 leads to
smaller contact angles, whilst
for >90 the contact angle
increases! That is pointed out
in illustration 2. To a certain
contact angle belongs a cos-
value (1). Multiplication with R
results in a new cos-value (2)
which leads to a new contact
angle (3). One thus finds the
largest contact angles on
rough surfaces
5
.
This gives rise to the so called
Lotus effect: On some plant
leaves the contact angle of water is very large and water drops easily go away. The
surface is hydrophobic and rough. For some technical applications this is used to
create surfaces which stay clean.
1
90 180
0
-1
1
2
3
1
2
3
- 4-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
Inhomogeneity of the surface of the solid. There are different ways to consider
inhomogeneities.
6
If there are two different components on a surface, a start is to
consider a kind of average contact angle according to
7

2 2 1 1
cos cos cos + = f f

f
1
and f
2
are the surface ratios of the two components,
1
and
2
are the contact
angles on the pure components.
8.1.5 Complete wetting
The case
L S SL
= look at first glance like an exception. That, however, is not
the case. In equilibrium, in particular in the presence of saturated vapor,
S
can never
become larger than
L SL
+ . If this were possible, then the solid could reduce its
surface tension forming a liquid film on ist surface. Hence, is the system is in
equilibrium and we have complete wetting then
L S SL
= .
In non-equilibrium situations the so-called spreading coefficient
S
S SL L
=

can be positive.
8.2 Rise of a liquid in a capillary
An important application of Youngs equation and a procedure for measuring the
contact angle at the same time, is the rise of a liquid in a tube (capillary). If a capillary
is lowered into a liquid, ofte the liquid rises in
the capillary. The height is given by
h
r g
L
K
=
2

cos

is the density of the liquid, g is the
acceleration due to gravity (9,81 m/s
2
).
The liquid rises only for for <90. For >90
the liquid is pressed out of the capillary!
8.2.1 Derivation
We assume that the capillary is thin and the liquid surface is spherical. To derive the
equation we consider the change of Gibbs free energy upon an infinitesimal rise of
the liquid dh. The shape of the liquid-vapor-interface does not to change. The
accompanying change in Gibbs free energy is:
( ) dG r dh r gh dh
K S SL K
= + 2
2

The first term represents the surface work. During the rise free solid surface is lost,
while solid-liquid interface is formed. This is the actual drive for the liquid rise: The
liquid ascends because the high-energy solid surface is replaced by a low-energy
solid-liquid interface. The second term corresponds to the work required to lift the
liquid with the weight r h
K
2
by a step dh against the gravity. In equilibrium the
hydrostatic pressure of the liquid column is equal to the contribution of the interfacial
tensions and we have
h

2r
K
- 5-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
( ) ( )
dG
dh
r r gh r gh
K S SL K S SL K
= + = = 2 0 2
2

Replacing
S SL
by
L
cos immediately leads to the equation above.
8.3 Measurement of the contact angle
8.3.1 Methods
The most usual method to measure the contact angle is to observe a sessile drop
with a telescope or microscope and measure the
contact angle with a goniometer. With small
drops, where hydrostatic effects are negligible,
one gets the contact angle also from the height of
the drop (or from other easily measurable
parameters).
8
As predicted by the Laplace equa-
tion small drops have a circular cross section. For
geometrical reasons we have ( ) tan 2 = h r . In
more exact procedures the cross section of the
drop is recorded and fitted with a solution of the
Laplace equation. From the fit the contact angle
follows.
Alternatively one can measure the contact angle at the edge of a bubble (captive or
sessile bubble). The method is less sensitive to pollution of the interface. In addition,
the vapor phase automatically is 100% saturated.
A widely used technique is the Wilhelmy plate method (chapter 2). If the contact
angle is larger than zero, then the force, with which the plate is pulled into the liquid,
is 2
L
l cos. Here l is the width of the plate.
One problem is to quantify the wetting behavior of powders. The usual way to do this
is the capillary rise method. Solids, which are available only as powders, can be
pressed into a capillary of typically 1 cm diameter. This porous material is then
treated as a bundle of thin capillaries with a certain effective radius r. In order to
measure this effective radius, first a completely wetting liquid is used. Either the
speed of the liquid rise is measured or the pressure required to keep the liquid out of
the porous material. This counterpressure is equal to the Laplace pressure of the
liquid, given by p r
w L
w
= 2 . Then one measures the Laplace pressure with the
liquid of interest. It amounts to p r
L
= 2 cos . A direct comparison gives the
contact angle.
8.3.2 Hysteresis in contact angle measurements
Practically one often observes that in contact angle measurements the advancing
contact angle
adv
is larger than the receding contact angle
rec
. There are at least
three reasons:
Surface roughness.
Inhomogeneity or pollution of the solid surface.
9

Disolved substances in the liquid are deposited at the solid

r
h

- 6-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
On rough surfaces with very large contact angle bubbles can be traped:
Liquid
Solid
Vapour


8.3.3 Results
Contact angles can have quite different values. Examples of advancing contact
angles are tabulated here.
Liquid
( / mJ/m
2
)
Solid
adv

Water (72) Paraffin 110
Teflon 98-112
Polypropylene 108
Polyethylene 88-103
Graphite 86
Platinum 40
Glass small
Gold 0
Rough alkyl 174
n-Decan (23) Teflon 32-40
Graphite 120
n-Octane (21,6) Teflon 26-30
n-Propanol (23) Teflon 43
Paraffin 22
Polyethylene 7
Mercury (484) Teflon 150
Glass 128-148
8.4 Theoretical aspects of contact angle phenomena
8.4.1 Problems
The Young equation is extensively used to quantify the wetting properties of liquids.
This is misleading because it has not been really verified empirically . A major
problem lies in the value for the surface tension of solids. The surface tension of a
solid is not a real thermodynamic quantity because it depends on the way the surface
has been created. Solid surfaces can be formed plastically or elastically. Depending
on the way of formation, the surface tension or energy can be different.
8.4.2 Effect of surface forces
Within the core region the profile of a drop is modified by surface forces. As long-
range forces vdW and electrostatic forces are important.
1
These forces affect the
- 7-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
profile in a range of 2-100 nm. They can cause a difference between the microscopic
contact angle and the macroscopic one (which enters into Youngs equation).
8.4.3 The model of Girifalco, Good and Fowkes
In many applications it is possible to determine
L
and
S
but not
SL
. Then it would be
helpful to express
SL
through
L
and
S
. Therefore Girifalco, Good, and Fowkes con-
sidered molecular solids and liquids, in which the molecules are held together by vdW
forces. Then, in a Gedanken experiment they seperated two solids at the interface:
1
2
Vacuum

The necessary work per cross-sectional area is
W = +
1 2 12

Two new surfaces are formed while interface disappears. Now we remember chapter
7. The work to separate two solids to infinite distance against the vdW attraction is
given by
W A D
V
=
1 2 0
2
12
The Hamaker constant can approximately be expressed by
A A A
V V V 1 2 1 1 2 2

The Hamaker constants of the single materials are related to the surface energies in
accordance with
2
0 1 1 1
24 D A
V
= and
2
0 2 2 2
24 D A
V
=
Substituting results in
10,11

2 1 2 1 12
2 + =

A similar formula is used not only for the vdW force but also for other force
components. Example: VdW (dispersive) and polar interactions
12


SL S L S
d
L
d
S
p
L
p
+ 2 2

8.4.4 Coexistence of the drop with a liquid film
The Young equation is only valid in equilibrium, i.e. only if the gas phase is saturated
with the vapour of the liquid (p/p
0
=1). A possible adsorption isotherm is shown in
curve A. The larger the vapour pressure, the thicker the adsorbed liquid film. The

mac

mic

- 8-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
problem is that for p/p
0
1 the liquid film becomes infinitely thick, i.e. it becomes a
drop. In order to have a drop in the equilibrium on the surface, the drop must be able
to coexist somehow with the liquid film. This is possible for adsorption isotherms like
curve B. Imlication: The presence of the solid surface changes the state of the liquid.
0,0 0,2 0,4 0,6 0,8 1,0 1,2
0
5
10
15
20
25
30
35
40
p/p
0

nm
-2
A
B

8.5 Kinetics of wetting
Essentially experiments were performed using two arrangements:
A liquid is pressed into a capillary. Here the capillary number w v = ,
where v is the flow velocity and is the viscosity of the liquid. With w<<1 the
contact angle grows in accordance to w
3
. At w=1 it is approximately 90.
As a second method, the propagation of drops on surfaces was examined. An
empirical law, valid for small drops for which the gravity is negligible and they
therefore have a circular cross section, is
R t V
n p 2

R is the radius of the contact line, V is the volume of the drop, assumed to be
constant, t is the time, n and p are empirically determined coefficients with the
values n0,16...0,32 and p2/3.
It is well-known that a spreading drop forms a thin (< 1m) film. This film is called
precursor film. Its thickness and extension are determined by surface forces.
During the investigation of wetting it makes a difference whether one works in
saturated vapour or "dry". In the first case the drop spreads on a thin liquid film. In the
latter case the solid surface is not covered by molecules of the liquid.
Dry wetting
The schematic structure of an advancing drop is shown in the figure. One can classify
the wetting processes according to the dissipated energy and we can distinguish
three regions:
- 9-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
1. The movement in the thick, wedge-shaped part of the drop is like in a caterpillar:
Small eddies cause a flow. Heat is released due to viscous friction.
2. In the precursor film energy is likewise dissipated by viscous friction. The liquid
transport in the precursor film is driven by differences in the disjoining pressure,
thus by surface forces.
With classical continuum theory there is a problem.
13
One normally accepts that
the liquid directly on the solid surface does not move. This leads to an infinitely
large loss of energy. A possibility to remove this energy divergence from the
theory consists in permitting to the liquid to slide along the solid surface. Another
possibility is that the drop spreads on the precursor film.
3. At the real contact line on the front end of the precursor film liquid molecules bind
to the solid surface. During this binding normally heat is released.
In soft solid surfaces one more effect is added: the solid is deformed at the contact
line and additional energy is dissipated.
14

The adsorption-desorption model
15

How does wetting of a dry solid surface at the end of the real contact line take place?
The advancing of the real contact line is probably determined by statistical kinetics on
molecular level. Liquid molecules bind to adsorption sites on the solid surface. They
must in part displace adsorbed gas molecules. These binding and dissociating
processes take place statistically, however with higher probability in the advancing
direction. Behind the contact line the molecules can still rearrange and reorganize.
The spreding velocity of a liquid depends thus on the disturbance of the adsorption
equilibria.
Free adsorption
sites
Liquid
molecules
Gas
molecule

The adsorption-desorption model predicts a maximum wetting speed v
max
and a
minimum dewetting speed v
min
. At speeds larger than v
max
gas bubbles form. This was
indeed observed. For water v
max
5-10 m/s.
The wetting speed is determined by the interaction between solid and liquid. One can
therefore expect that for strong interactions the adsorption and desorption procedures
are slower and therefore the speed decreases. Strong interaction is present with
precursor
film

Real contact line
- 10-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
small contact angles. I.e. small contact angles should be correlated with a small v
max
.
This was observed.

w
180

Dewetting
Wetting
v
max
v
min


8.6 Wetting and flotation
In this chapter application-oriented aspects of contact angle phenomena are treated.
8.6.1 Good wetting
In many applications one would like to achieve a good wetting. In addition
SL
and
L

should be as small as possible because then the spreading coefficient
S
L S S SL L /
= is large. Therefore surfactants are usually added to the liquid. It
is not enough to reduce
L
. Often the reduction of
SL
is just as, or even more,
important. The selection of a suitable surfactant depends on the solid.
Wetting as a capillary phenomenon
In many cases, e.g. for the wetting of textiles or the use of cleaing paper to pick up
fluids, the liquid should penetrate into interspaces. This procedure is similar to the rise
of a liquid in a capillary. The pressure p, which is needed for preventing liquid to
penetrate into a capillary with radius r is p r
L
= 2 cos . The larger this pressure,
the better the wetting characteristics. We distinguish between two cases:
( )
> =

0
2
p
r
S SL

, for good wetting
SL
should be small
= = 0
2
p
r
L

, for good wetting


L
should be large
The point is thus to find a surfactant, which reduces
S
without simultaneously
decreasing
L
too strongly.
8.6.2 Water repulsion
In many applications one would like to achieve a repulsion of water or other liquids
e.g. to prevent destruction of road pavements or to make clothes waterproof.
In this case ( ) p r
S SL
= 2 should be as negative as possible. This can be
achieved by reduction of the surface energy
S
of the solid.
- 11-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
The illustration shows
one important aspect
of liquid penetration
through fibrous
networks: In a capillary
one must apply a
counterpressure for
<90
o
in order to
prevent a rise of the
liquid. Nevertheless the
liquid does not pass spontaneously trough a fiber twine, one has to press it through.
8.6.3 Flotation
Flotation is a purification method. It is used to
process crushed ore on a large scale. Originally the procedure was applied only to
some sulfides and oxides. Meanwhile many minerals are separated by flotation.
Annually 10
9
tons ore are processed with the help of flotation (world annual
production of iron 1981: 491 million tons)
remove unwanted material during the water purification
clean industrial waste products
deinking of paper
In flotation of ore one first crushes the material to under 0,1 mm particle size. The
particles are mixed with water. Air bubble are passed through the container. The
mineral rich particles bind to the air bubbles by hydrophobic forces and are carried to
the surface. A foam forms. With the foam the mineral rich particles can be removed.
The used and surfactants are important. They contain:
Collectors adsorb selectively on the mineral and make its surface hydrophobic
activators support the collectors
depressants reduce the collectors effect
Frothing agents increase the stability of the foam
Liquid
Fibre
>90
o
<90
o

- 12-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
Foam with
hydrophobic
material
Suspension
Air

Role of the contact angle
If the contact angle is not zero, then a particle at the liquid surface is stable (see
illustration). To quantify this we calculate the work
required to separate a sphere from the liquid-gas
interface.
Sphere surface area in air:
( ) ( )
( )

r h R
R
2 2 2 2 2
2
1 2
2 1
+ = + +
=
sin cos cos
cos


with r R = sin and = cos R R h . How large is
the difference in Gibbs free energy between the
stable state of the sphere in the interface and the
situation, where the particle is completely in liquid?
( ) ( ) G R R
SL S L
= + 2 1
2 2 2
cos sin
With Youngs equation cos =
L S SL
we get
( )
( ) ( )
( )

G R R
R R
R
L L
L L
L
= +
= + = +
=
2 1
2 2 2 1
1
2 2 2
2 2 2 2 2
2 2



cos cos sin
sin cos cos cos cos
cos

Example: J G m R m J
o
L
12 2
10 5 . 5 5 , 90 , / 07 . 0

= = = = . One can estimate
the necessary force from F G R N = / . 11 .
One can also regard the case, in which the bubble is substantially smaller than the
particle. Before the adhesion of a bubble to a particle the bubble is normally
Water
Air

h
r
R

- 13-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
deformed. This is shown a little exaggerated in the following illustration. The surface
of the bubble becomes gradually larger. That leads to a brief increase of the energy -
a kind of activation energy.
approach deformation adhesion

8.7 Washing and cleaning
8.7.1 Definitions and introduction
It is about the theory and practice of the removal of foreign material from solids
through chemical substances. This definition excludes pure mechanical cleaning.
Also a pure chemical cleaning, e.g. by solvation of the foreign material, is not treated.
First some definitions:
Soap Sodium or potassium salts of a higher fatty acid
Detergent Soap-free wetting or cleaning agent
Surfactant Generic term: Strongly interface active substance
Table 1:Applications of surfactants
49% Detergent household and business
5% Industrial cleaning
11% Cosmetics and pharmaceutical industry
10% Textiles and fibres
8% Chemical industry
3% Food industry
3% Construction, metal processing
2% Mining industry, flotation, oil extraction
2% Colours and varnishes, plastics
2% Leather-, fur-, cellulose-, paper industry
2% Plant protection, pest control
3% Others
Surfactants are divided into anionic, cationic, non-ionic and zwitterionic (amphoteric)
surfactants.
Table 2: Surfactants consumption in the FRG 1985
Anionic 252.000 t
Not ionic 143.000 t
Cationic 30.000 t
Amphoteric 5.500 t

- 14-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
In textiles usually oily substances attach to
the fibres (animal fats, fatty acids,
hydrocarbons, etc.). Also dust, soot and
other solid particles have to be removed in a
washing process. In order to test the
effectiveness of a surfactant, textiles are
often polluted with standard dirt mixtures and
cleaned with a standard washing procedure
(launderometer). One usually measures the cleanliness on the basis of the optical
reflectivity of white textiles. For this purpose a calibration curve dirt quantity against
reflectivity is created before.
Attached dirt particles are separated spontaneously from a textile fibre if
( ) 0 + =
TD TW DW
A G ,
i.e. if
TW DW TD
+ .
With liquid dirt the contact angle must become zero,
i.e. the spreading coefficient should be zero or
positive. This leads to the same condition.
An effective surfactant should reduce
TW
and
DW
, without reducing
TD
too strongly. A
decrease of the surface tension of the water - visible by bubble formation - is not a
proof for an effective surfactant for washing.
8.7.2 Aqueous surfactant solutions
Surfactant solutions show certain distinct properties. Therefore they are called
"colloidal electro-
lytes". Many proper-
ties change
drastically if the con-
centration exceeds
the so-called critical
micellar concen-
tration (cmc). Light
scattering increases.
Hydrophobic sub-
stances are dis-
persed. Surface and
interface tension,
osmotic pressure
and the electrical
conductivity change
in a characteristic
way. Micells usually
show a relatively
homogeneous size distribution. They contain typically 50-100 monomers. Micells form
only above a certain temperature, the Krafft temperature. Below this temperature
the surfactant precipitates.
8.7.3 Factors, which affect the washing behaviour
Originally it was assumed that the cleaning effect of surfactants is caused by their
ability to accomodate hydrophobic substances inside micells. However, this does not

TD

DW

TW

dirt
water
textile
dirt
water

Concentration cmc
Osmotic
pressure
- 15-
PCSURF08. DOC BUTT PC2 UNI SI EGEN 31. 01. 2001
seem to be the dominating factor because the cleaning ability increases already with
surfactant concentration at concentrations below the cmc, thus before micells are
formed. Today one accepts rather that the reduction of the free interface enthalpies

TW
and
DW
is the main cause for cleaning.
An important characteristic of surfactants is their ability to keep dirt particles in
solution (suspending power). Without this ability a washing procedure would only lead
to a uniform distribution of the dirt. Obviously the surfactant is bound to the surface of
the dirt particles and keeps them dispersed. Aggregation and floculation is prevented
e.g. by electrostatic repulsion.

1
P.G. de Gennes, Reviews of Modern Physics 1985, 57, 827.
2
A.W. Neumann, J.K. Spelt, Applied Surface Thermodynamics, Surfactant Science Series 63, Marcel
Dekker, New York, 1996.
3
T. Young, Phil. Trans. Royal Soc. London 1805, 95, 65.
4
Pethica, B.A. J. Colloid Interface Sci. 1977, 62, 567.
5
T. Onda, S. Shibuichi, N. Satoh, K. Tsujii, Langmuir 1996, 12, 2125.
6
J. Drelich, J.L. Wilbur, J.D. Miller, G.M. Whitesides, Langmuir 1996, 12, 1913.
7
A.B.D. Cassie, Discuss. Faraday Soc. 1948, 3, 11.
8
A.S. Dimitrov, P.A. Kralchevsky, A.D. Nikolov, H. Noshi, M. Matsumoto, J. Colloid Interface Sci. 1991,
145, 279.
9
S. Brandon, A. Marmur, J. Colloid Interface Sci. 1996, 183, 351.
10
R.J. Good, L.A. Girifalco, J. Phys. Chem. 1960, 64, 561.
11
F.M. Fowkes, Ind. Eng. Chem. 1964, 56, 40.
12
D.K. Owend, R.D. Wendt, J. Appl. Polym. Sci. 1969, 13, 741.
13
C. Huh, L.E. Scriven, J. Colloid Interface Sci. 1971, 35, 85.
14
A. Carr, J.C. Gastel, M.E.R. Shanahan, Nature 1996, 379, 432.
15
T.D. Blake, Dynamic Contact Angles and Wetting Kinetics, in: Wettability ed. J.C. Berg, Marcel
Dekker, New York, 1993.

You might also like