Sie sind auf Seite 1von 110

CEDRIC J.

GOMMES

P H Y S I C A L C H E M I S T RY
O F I N T E R FA C E S

UNIVERSITY OF LIGE
Copyright 2014 Cedric J. Gommes

First printing, December 2014


T H E R E I S N O E A S Y W AY T O T R A I N A N A P P R E N T I C E . M Y T W O T O O L S A R E

E X A M P L E A N D N A G G I N G . I W I L L S H O W Y O U W H AT I T I S I D O , A N D T H E N I

W I L L T E L L Y O U T O D O O T H E R T H I N G S Y O U R S E L F. D O Y O U U N D E R S TA N D ?

L E M O N Y S N I C K E T, A L L T H E W R O N G Q U E S T I O N S

N O U S P R E N O N S E N G A R D E L E S O P I N I O N S E T L E S AVO I R D AU T RU I , E T P U I S

C E S T T O U T : I L L E S FAU T FA I R E N T R E S . N O U S S E M B L O N S P R O P R E M E N T

C E L U I , Q U I AYA N T B E S O I N D E F E U , E N I R A I T Q U R I R C H E Z S O N V O I S I N , E T

Y E N AYA N T T R O U V U N B E A U E T G R A N D , S A R R T E R A I T L S E C H A U F -

FER, SANS PLUS SE SOUVENIR DEN RAPPORTER CHEZ SOI. QUE NOUS

S E R T- I L D A V O I R L A PA N S E P L E I N E D E V I A N D E , S I E L L E N E S E D I G R E , S I

E L L E N E S E T R A N S F O R M E E N N O U S ? S I E L L E N E N O U S AU G M E N T E E T F O R -

TIFIE ?

M I C H E L D E M O N TA I G N E , E S S A I S , I , 2 5 , D U P D A N T I S M E
Contents

Introduction 0-1

Surface energies and Laplaces pressure I1

Wetting, a short introduction II1

Solved problems involving wetting III1

Intermolecular forces IV1

Gibbs-Thomson Effects V1

Thermal fluctuations VI1

Metastability and nucleation VII1

Adsorption on surfaces VIII1


Introduction

Gold is known as a shiny, yellow noble metal that does not


tarnish, has a face centred cubic structure, is non-magnetic
and melts at 1336 K. However, a small sample of the same
gold is quite different, providing it is tiny enough: 10 nm
particles absorb green light and thus appear red. The melting
temperature decreases dramatically as the size goes down.
Moreover, gold ceases to be noble, and 2-3 nm nanoparticles
are excellent catalysts which also exhibit considerable
magnetism. At this size they are still metallic, but smaller
ones turn into insulators. Their equilibrium structure changes
to icosahedral symmetry, or they are even hollow or planar,
depending on size.

Emil Roduner, Size matters: why nano materials are different.

Everything goes nano these days. Examples from the marketing


world are myriads, from cars to watches, shoes and toy helicopters.
I personally own a music player bearing that name, to which I am
listening at the very moment where I am writing these lines. Poli-
cymakers even seem to be anticipating a nano-revolution. So what
exactly is this hype all about? Figure 01: Thomas Graham (1805-
1869)

Let us not fool ourselves: nanomaterials arent anything


new. The oldest known example of a nanotechnology is probably
the Lycurgus Cup, which is a 4th century roman cup currently ex-
posed at the British Museum. The cup exhibits dichroism due to the
presence of colloidal silver within the glass. Another spectacular
example is that of so-called Damascus blades, which are made up
of steel with specific inclusions, which some researchers clim to be
carbon nanotubes. Closer to us, the entire field of colloids, born at
the end of the nineteenth century under the influence of people such
as Thomas Graham and Michael Faraday, among others, is about
nanometer-sized objects. The fact is that nanometer-sized objects
Figure 02: Michael Faraday (1791-1867)
have properties that are distinctly different from chemically identical,
but larger objects. The quote at the beginning of the chapter is the
spectacular example of gold.
0-2 c.j. gommes, chim0698

A first difference between today and say, twenty years ago is the
fantastic progress made in the field of electron microscopy, which
enables us to visualise nanometer-sized objects with unprecedented
accuracy. Advanced techniques such as electron tomography even Bottom-up and top-down methods
enable one a obtain 3D reconstructions of structures with nanome- refer to whether the larger structures
are created last or first. Typically,
ter resolution. The second difference is the progress in the synthesis lithography is top-down. Chemical
of nano structured materials by a variety of techniques, both phys- methods, by which molecules assemble
into nano structures is bottom-up.
ical and chemical, top-down and bottom-up. The fields of synthe-
sis of nano materials, and that of fundamental understanding of
nanometer-scale phenomena are clearly feeding on each other. It
is because we understand better nanometer-scale phenomena that
we can create structures on that scale in a more controlled way. In
turn, those better-defined structures enable us to better understand
nanometer-scale phenomena. This is incidentally the reason why sci-
entific or technological progresses are exponential: because they feed
on themselves.
There are, however, still some problems on the horizon. For exam-
ple, in the field of imaging of nanostructures one is often limited to
studying materials in vacuum, which is a very different environment
from that in which most nanomaterials are generally used. The de-
velopment of so-called in situ imaging techniques at the nanometer- Figure 03: One of the colloidal gold
suspension prepared by Faraday in the
scale, to visualise nano structures in working conditions, is still a 1850s, which is still kept by the Royal
challenge. One therefore often relies on indirect methods, such as Insitution. Read more about it here
small-angle X-ray scattering. In the field of synthesis too, our current
skills are limited. Although the word nano fabrication is sometime
used, our current abilities are far from the self-multiplying robots
that terrorised Great Britain in 2001 under the ominous name of
"Grey Goo". A positive way to look at this is to think that there are
many things left to be developed by new scientists and engineers in
that field. The word colloidal was coined by
Graham to characterise "glue-like" sub-
stances. The official IUPAC definition
When people express their enthusiasm about nanoma- of that word is as follows: "State of
subdivision such that the molecules or
terials, they often say that it is fantastic that we can control the polymolecular particles dispersed in a
shape of materials on a scale as small as a few nanometers. Before medium have at least one dimension
we comment on this, we should first emphasise that there is already between approximately 1 nm and 1
m, or that in a system discontinuities
an existing and successful science, the purpose of which is to control are found at distances of that order."
matter over a still smaller scale, that of the Angstrm. That science In other words, colloidal is almost
synonymous with nano.
is naturally chemistry. So size per se is not the issue. The real diffi-
culty with nano structures is that they at the interstice of two worlds: It is important that you become familiar
with a few orders of magnitude. For
the microscopic and the macroscopic. Systems in the former world example: the number of molecules
are generally made of a few molecules, interacting with very strong or atoms in a macroscopic amount of
matter is 1023 , the size of an atom is
covalent bonds, and subject to thermal fluctuations. By contrast, in 1 , the energy involved in chemical
the latter world systems comprise a huge number of molecules (say, bonding is 1 eV, the thermal energy at
one mole) interacting through very weak forces, and thermal fluctu- room temperature is 25 meV, etc.
introduction 0-3

ation is generally negligible. What is difficult and interesting about


naometer-sized systems is not really that they are small, it is that they
are intermediate between two worlds. It is often unclear which type Because nano materials are squeezed
of physics applies: would a 5 nm particle better be thought of as a between the macroscopic and the
microscopic worlds, one often refers to
small crystal or a large molecule? The answer is obvious "it depends". them as being mesoscopic. The greek
The purpose of CHIM0698-1 is to guide you into mesoscopic physics. prefix "meso" means "in between".

Figure 04: When a given object (here a


cube) is cut into objects of half the ini-
tial size, the total surface area exposed
is multiplied by two.

At this stage, you may wonder why the title of the course
is "Physical Chemistry of Interfaces" and not, say "Intro-
duction to nano sciences". The reason is that all many properties of
nanostructured materials result from their having a large surface-to-
volume ratios. In other words, interfaces play an important role at the
nanometer-scale. This is illustrated in Fig. 04 with the unfolding of
a paper cube, compared to that of eight smaller cubes. The surface is
two times larger in the latter case.
To put this in a more quantitative way, try to estimate the specific
surface area of a material made up of silica particles with a diameter
of 10 nm. Obviously, if you double the mass of material, you double
the surface. So the only sensible quantity to evaluate is expressed in
m2 /g, for example. The word "specific" means "per unit of mass".
Now, when I wrote "try to estimate.." I really meant it: before reading
what follows try to do it!
A possible way to calculate this is to estimate first the area of a sin-
gle nanoparticle, and then to calculate how many nanoparticles there
are in one unit of mass. The area of one spherical particle of radius R
is 4R2 so that the total area of the N particles in the material is

A = N 4R2 (01)

The mass of a single particle is 4/3R3 where is the density


(the specific mass) of the considered material. The number of parti- It is useful to know the order of mag-
cles in a mass m of material is therefore nitude of the density of a few common
materials. The density of water, and of
  most ordinary solvents is 1 g/cm2 , that
4 3 of noble metals is around 20 g/cm3 .
N = m/ R (02)
3 The density of silica (glass) and of
graphite are both close to 2 g/cm2 .
0-4 c.j. gommes, chim0698

The specific surface area of the material is finally


A 3
= (03)
m R
In the particular case of R = 5 nm, and = 2 g/cm3 , one finds A/m =
300 m2 /g. This is approximately the area of a tennis court! Because This is yet again an order of magnitude.
the density of packings of spheres are approximately 50 %, one gram The exact densities for the packing of
spheres range from 0.6 for extremely
of that powder occupies about one cubic centimeter. In other words, loose random packings to 0.7405
it it the total area of a tennis court that fits in a cubic centimetre. for dense packings. The order of
magnitude is 0.5.
The inverse relation that exists between the surface area and the
size of the structures is very general. It is independent of whether
the structures are spherical, laminar, or cylindrical. Exercises 1 and 2
should convince you of this. Going back to the tennis court analogy,
imagine that you have covered a tennis court with, say, a paper sheet
with infinitely small thickness so that you can fold it infinitely easily.
You then try to fold or crease the entire sheet into a tiny cube with
a size of 1 cm. The first fold will get you about half the size of the
court, 10 m., etc. To have the entire sheet fit in one cubic centimetre,
you will have to fold the sheet a very large number of times, with
increasingly smaller folds. What Eq. 03 tells you is that the average
distance between each fold once it fits in the cube is about 5-10 nm.

Why is the surface area relevant at all? For two reasons.


For specific applications, it is often desirable to have materials with a Figure 05: Tranmission Electron Mi-
croscopy (TEM) image of a Pt/Carbon
large surface area. For example, in the case of heterogeneous cataly- heterogeneous catalyst. The black dots
sis, the active sites are generally the surface atoms of metal particles. are the Pt nanoparticles. The scale bar is
20 nm.
As a rule of thumb: the large the surface area, the more active the
catalyst. Another example is the case of separation processes based
on adsorption. In that case too, the larger the surface area the more
useful the material, imply because there is more room on the surface
to capture contaminant. Another, less direct, reason is that the spe-
cific surface area of a material is roughly proportional to the fraction
of atoms or molecules that are on a surface, as opposed to the bulk
of the material. Atoms or molecules in the bulk of the material have
more neighbours than those close to the surface, so they are more
tightly bounded to each other. This means that the overall cohesion
of a material should decrease when it is divided into small particles,
because the fraction of atoms at the surface is larger. If you back to
the epigraph at the beginning of the chapter, you can already un-
derstand why gold nanoparticles have a lower melting temperature.
The modification of thermodynamic equilibria at the nanometer scale
as opposed to chemically identical macroscopic systems is a very
general phenomenon know as Gibbs-Thomson effects.

Let us estimate the fraction of atoms at the surface of a


introduction 0-5

material. The situation is sketched in Fig. 06 in the case of an ob-


ject having a deliberately complex shape. The volume of that object is
V and its surface area is A. The atoms that are at the surface occupy
a layer of thickness d all over the surface, where d is an equivalent
atomic diameter. The volume of that layer is therefore A d, so the
fraction of atoms on the surface is

Fraction of atoms Ad A
' = d (04)
at the surface V m

where the equality results from expressing the volume as m/. The
fraction in the Right-Hand Side (RHS) is nothing but the specific sur-
face area of the material. This equation has a very intuitive meaning:
V/A has the dimension of a length, and it is the order of magnitude
of the size L of the object (check for a sphere). What Eq. 04 tells you L d
is simply that the fraction of atoms at the surface is proportional to
d/L. If the object if of a size comparable to that of the atoms, then
there are may atoms at the surface. On the other hand, if the object is
macrosopic, there is a negligible fraction of atoms at the surface.
In practice, consider for example the fraction of atoms at the sur- Figure 06: A complex-shaped object,
with characteristic size L. All atoms
face of an ice cube. The ice cube is about 1 cm across so L ' 102
in a layer a thickness d, comparable to
m, and the diameter of the atoms is about d ' 1010 m. So the frac- a molecular diameter, can be consid-
tion is d/L ' 108 ; only ten atoms out of a billion are at the surface, ered to be at the surface. The fraction
of molecules at the surface is approx-
which isa negligible quantity. Note in passing that it does not matter imately equal to the fraction of the
whether the answer had been 20 108 or 0.2 108 . If the answer is volumes of the red and grey regions of
the drawing.
100 times larger or smaller, it is negligible anyway. This is the reason
why we do not have to care too much about the exact size of atoms
and of the ice cube for this type of calculation. This is the essence of
order-of-magnitude calculations.
If the result had not been negligibly small, say if we had found a
fraction of 0.01 then we would have needed to pay more attention
to the details. Because it makes a lot of difference whether 1 % or
100 % of atoms are at the surface. As an exercise, let us calculate the
fraction of water molecules that are at the surface of a 10 nm droplet.
We first have to find a reasonable value for d. This can be done by
expressing that the density of single molecules is the same as that of
macroscopic water. Assuming a cubic packing of molecules, which is
naturally a rough approximation, one has

M/NA
= (05)
d3

where M is the molar mass, NA is Avogadros number, and is


the macroscopic density. Using the values M = 18 g/mol, NA =
6.02 1023 , and = 1 g/cm3 , one finds d ' 3.1 108 cm. Using that
0-6 c.j. gommes, chim0698

value, the fraction of molecules on the surface is found to be


Ad 3d 3 3.1 108
= ' ' 0.2 (06)
V R 5 107
In a 10 nm droplet of water, twenty percent of molecules are at the
surface! This is a huge fraction.
If you compare the local environment of atoms or molecules at the
surface and in the bulk of a material, you will find that the molecules
at the surface are necessarily lacking neighbours with which to link.
The overall cohesion of water molecules in a 10 nm droplet must
therefore be significantly lower than bulk water, which must have a
dramatic influence on the equilibrium vapour pressure at any given
temperature. Do you expect the equilibrium vapour pressure over a
10 nm water droplet to be larger or smaller than over a macroscopic
water surface? If you prefer you can think of it in terms of a boiling
point: what is the boiling point of a 10 nm water droplet? This is
one of the many question of this type that we will be addressing in
CHIM0698.

Exercises

The main take-away message from this chapter is the relation be-
tween size and surface area. It is important that you feel comfortable
with this before studying the following chapters. The following exer-
cises should help you.

1. We obtained Eq. 03 by assuming spherical particles. How is


that relation changed if the nanomaterial consists of elongated Figure 07: A carbon nanotube sample,
needles of radius R instead of spheres? The fact that the needles imaged by transmission electron mi-
are elongated means that end surfacer of the needles is negligible croscopy (TEM). The thickest nanotubes
in the figure have an approximate
compared to the lateral surface. diameter of 10 nm.

2. Repeat the exercise in the case of a material made up of thin layers


of thickness t.

3. In a later chapter, we shall discuss nitrogen adsorption, which is


an experimental method to determine the specific surface areas
of materials. Using that technique on a given carbon nanotube
sample (see e.g. Fig. 07), one found A/m = 230 m2 /g. Can you
use this data to estimate the approximate radius of the tubes in
that sample?
Do the calculation twice, first assuming that the tubes are closed
so that the measured surface is the outer surface, and then by
assuming that the tubes are open so that the area is the sum of the
inner and outer surfaces. In both cases, you may assume that the
inner radius is about half the outer radius.
I0 c.j. gommes, chim0698

Further reading

The present lecture notes are supposed to be self-contained. These


further-reading suggestions are mostly meant to broaden your per-
spective on nano materials and nanotechnologies. Some links are
with restricted access, so you need to connect via your University
account to read them.

On the web
The NANOYOU (Nano for Youth) project is funded by the Euro-
pean Commission to increase young peoples understanding of
nanotechnologies. You can find many useful resources on their
website here.

The Royal Institution has a beautiful website on Faradays work,


and in particular about his celebrated colloidal gold suspensions.
You can find it here.

Here is a general overview to nano sciences by Paul Alivisatos, on


Youtube, who is a prominent researcher in the field and a profes-
sor at the University of California at Berkeley.

Regular papers
Emil Roduner, Size matters: why nanomaterials are different.
Chem. Soc. Rev., 2006, 35, 583-592
This tutorial review by Emil Roduner (cited as an epigraph of the
chapter) is a nice overview. You should not be able to understand
all of it at this stage. However, I encourage to try reading regularly
Figure 08: Richard Feynman (1918-
this paper along this semester to see your progress in that field. 1988).

R. Feynman, There is plenty of room at the bottom


This paper is a transcript of a historical conference that physics
genius and Nobel Laureate Richard Feynman gave in Pasadena in
1959. This conference is considered by many as a kick-off event in
the field of nanotechnologies.
Surface energies and Laplaces pressure

With respect to the cohesion and capillary action of liquids, I


have had the good fortune to anticipate Mr. Laplace in his
late researches, and I have endeavoured to show that my
assumptions are more universally applicable to the facts, than
those which that justly celebrated mathematician has
employed.

Thomas Young, A course of lectures on natural philosophy


and the mechanical arts, preface.

We have highlighted in the introduction that a specificity of


nanometer-sized systems is to have a large specific surface area,
which results in a large fraction of their atoms or molecules being po-
sitioned at interfaces. Because the local environment of molecules at
interfaces is different from the bulk, their binding energy is different, Figure I1: Thomas Young (1773-1829)
which results in an energy cost of the surface. This is sketched in Fig.
I8. A convenient way to think about this is in terms of missing bonds.
Form the point of view of the total system, the extra energy is pro-
portional to the number of molecules at the interface, i.e. to the area
of the interface. The extra energy, per unit area of the interface, is
the surface energy . The dimension of is an energy per unit area,
e.g. J/m2 , which can also be expressed as a force per unit length, e.g.
N/m. This is the reason for which is also referred to as the surface
tension, particularly when dealing with liquids.
To understand why the equivalence of N/m with J/m2 is not
merely dimensional but also physical, consider the situation in Fig.
I3. A thin film of water is held in a wire, one side of which is free
Figure I2: Pierre Simon de Laplace
to slide in such a way as to increase or decrease the area of the film (1749-1827)
A. The mechanical equilibrium of the system can be analysed by
applying the principle of virtual work. Because the natural tendency
for the film is to reduce its area as much as possible, you have to
apply a force F in order to prevent the area from decreasing.
If you moved the wire to the right by a small distance dx, the
mechanical work would be

W = Fdx (I1)
I2 c.j. gommes, chim0698

On the other hand, the change in energy resulting from the displace-
ment dx is dU = 2Ldx. The factor Ldx is the variation of the area
enclosed by the wire; is the energy per unit area of the free surface;
the factor 2 accounts for the fact that the film has two free surfaces,
one on top and the other on the bottom. When equating W with dU,
one finds
F/L = 2 (I2)
top view
which is indeed proportional to . The factor 2 proves that if is
to be interpreted as a force per unit length, then the force has to be F
considered as being exerted on the surface itself and not to the film A L
as a whole. In the present case, that force applies on both surfaces of
the film.
In principle, the wired frame of Fig. I3 could provide a means to side view
measure the surface tension of liquids. In practice, however, liquids F
films are generally too fragile for this is to be possible. Another ex-
perimental configuration is used, which is referred to as Wilhelmys Figure I3: Mechanical equilibrium of a
plate. It consists in dripping a plate of roughened platinum1 into the liquid film on a wire.
liquid and in measuring the force required for pulling it out of the 1
You will have to wait until we have
studied wetting phenomena to under-
liquid. stand why this is important.
The situation is sketched in Fig. I4; the thickness of the plate is t
and its length (in the direction orthogonal to the figure) is L, and we
call P0 the ambient (e.g. atmospheric) pressure. Let us write all the
forces exerted on the plate. The forces oriented downwards are

Forcesdown = mg + P0 L t + 2L (I3)

The first term is the weight of the plate, the second is the pressure
that is exerted on top of the plate by the ambient gas (applied on a
total area L t) and the last contribution is the same 2L as in Fig.
I3. On the other hand, the forces upwards are

Forcesup = F + ( P0 gz) L t (I4)


F
where the first term is the force exerted by the measuring device,
which are eventually interested in, and the second term is the pres-
sure in the liquid just underneath the plate. The latter is simply the
z
ambient pressure minus the hydrostatic head.
When balancing the forces up and down, one finds

F = mg + Ltzg + 2L (I5) Figure I4: A sketch of Wilhelmys


plate while it is being pulled out of the
liquid.
The first term is a known constant (independent of the liquid), the
second term is the weight of the small part of the liquid being pulled
out under the plate, and the last term is the one we are interested in.
Using a plate that is sufficiently thin, one can make the second term
surface energies and laplaces pressure I3

as small as desired. In this case the force needed to pull the plate out
of the liquid is proportional to .
In practice, a common experimental setup is a so-called Wil-
helmys ring (as opposed to plate), in which it is a horizontal ring
made up of a thin Pt wire that is pulled out of the liquid. The anal-
ysis is similar to that in Eq. I5, only L has to be understood as the
length of the ring. A few experimental values of obtained using
this type of setup are gathered in Tab. I1.

Substance Surface energy (mJ m2 )


Nitrogen (77K) 8.85
Ethanol (25C) 23.2
Acetone (25C) 23.5
Cyclohexane (25C) 25 Table I1: The surface energies of a few
usual and not so usual substances. Note
Formamide (25C) 57.0
that two substances in the table are
Water (10C) 74.2 solid.
Water (25C) 72.0
Water (50C) 67.9
Mercury (25C) 485.5
Aluminum (25 C) 35
Nickel (300C) 2450
Glass 2000-4000

With the values from Tab. I1 in mind, we are in a better situation


to examine the relative importance of the second and third terms in
Eq. I5. We shall for now keep the discussion as general as possible
and consider a system subject to it surface energy and gravity. The
total energy may be written as

U = A + mgh (I6)

where A is the total area of the liquid free surface, and the second
contribution is the gravitational energy, with m the mass, g the accel-
eration of gravity, and h the height of the centre of gravity. Calling
l a characteristic size of the system, the order of magnitude of each
contribution is the following. The surface contribution is You may consider that "a characteristic
size" is synonymous to "the order of
magnitude of the size". In the case of a
Usur f ace l 2 (I7) sphere, you may chose the radius R or
the diameter 2R. The type of mathemat-
and the the gravity contribution is ical relations involving characteristic
sizes, orders of magnitude, and the like,
are referred to as "scaling relations".
Ugravity gl 4 (I8) The symbol means that a dimen-
sionless factor is omitted, which is the
where is the density. To obtain that estimation, we have taken ac- reason why we should not care too
much about the accurate definition of
count that the mass scales as m l 3 and the height as h l. the characteristic length.
I4 c.j. gommes, chim0698

Comparing the two contributions to the energy, one sees that the
surface is dominant whenever the length satisfies
r

l (I9)
g

The right-hand side is referred to as the capillary length lc . In the


case of water, with ' 0.1 N/m, = 103 kg/m3 , and g ' 10 m.s2 ,
one finds lc ' 3 mm. This corresponds to the common observation
that objects smaller than that value, such as dew drops, tend indeed
to adopt a spherical shape. By contrast, larger droplets have a flat-
tened shape as a result of the gravitational energy.
It is convenient to summarise the present analysis by introducing
a dimensionless number. In the present case, a natural choice is the
ratio of a systems characteristic length and its capillary length. For
historical reason, it is the square of that ratio that have received a Figure I5: Lornd Etvs (1848-1919)
name
gl 2
Bo = Eo = (I10)

Actually, it has received two names: it is sometimes referred to as the
Bond number, and sometimes as the Etvs number.
Going back to the initial question, about the ratio of the second
and third terms in Eq. I5, it can be written as
gtz
(I11)
2
Figure I6: Wilfrid Noel Bond (?-1937)
Because the typical height of the liquid z is expected to be compa- Etvs and Bond have independently
rable to the thickness of the film, t ' z, the ratio is about one half given their names to (roughly) the same
dimensionless number. The Bond num-
of Bonds number. Experimentally, in Wilhelmys experiment we ber is used mostly in English-speaking
want to measure the surface tension so that any other effect would countries and the Etvs number is
used in continental Europe. Besides
be parasitic. In otegr words, we want Bonds number to be as small his work on capillarity, Etvs was
as possible, which can be achieved by making the wire be as thin as also a pioneer in the discussion of the
possible. Assuming a wire that would be about half a mm thick, and equivalence of inertial and gravitational
masses, which is considered now a
that the liquid is water, one finds that the weight of the liquid con- foundation of general relativity.
tributes just a few percent to the force in the Wilhemys experiment.
Most of the force is due to surface tension, which is good.

Let us step back for a moment and consider the case of Fig. I3
again from a microscopic perspective. What opposes the extension
of the liquid film is the fact that more molecules are pulled towards
the surface if the area A of the film is increased. This may seem like
a crazy explanation, and yet it is in quantitative agreement with the
experimental values of in Tab. I1.
Let us try to make quantitative the sketch in Fig. I8 and relate
to some more familiar characteristic of a fluid. The surface energy
is the energy required to bring a molecule/atom to the surface and

Figure I7: Joef Stefan (1835-1893)


surface energies and laplaces pressure I5

break a few bonds. On the other hand, if you think of it, from the
point of view of a molecule, evaporation consists in breaking all the
bonds that maintain it in the liquid. One therefore expects the energy
relevant to to be a fraction of the enthalpy of evaporation he .
Let us see how quantitative we can make that analogy, by assum-
ing an oversimplified liquid structure where the molecules/atom
would be positioned according to a cubic lattice, with spacing d.2 In 2
Remember, this is equivalent to what
we did on page 0-5 when we were
that case, the number of bonds of a molecules in the bulk is 6, and
trying to guess molecular dimensions.
only 5 at the surface. The surface energy would then be written as

1 he
' (I12) surface
6 v2/3
m

because the area occupied by a molecule at the surface is d2 = v2/3 m , bulk


according to the cubic model where vm is the molar volume.
To put real numbers on this, consider the case of cyclohexane at
Figure I8: The physical origin of
25 C. Looking up in tables of physicochemical properties, or even
surface energies: the local environment
on Wikipedia, you find the following values. Evaporation enthalpy of the atoms/molecules is different in
he = 30.5 kJ/mol, density (specific mass) = 773 kg/m3 , and molar the bulk and at interfaces.

mass M = 84.16 g/mol. Using the latter two values, you can estimate
the molecular diameter d = 5.56 . From that value, Eq. I12 leads
to = 26 mJ/m2 . Compare that value to the actual experimental
value in Tab. I1. The agreement is excellent! Of course we must Josef Stefan was an Austrian physicist,
have been particularly lucky with this calculation because no one in born in todays Solvenia, who gave
his name to Eq. (I12), among many
his own mind would accept that molecules in a liquid would pack others. To understand the importance
in a cubic way. Moreover, we have assumed that molecules on the of Stefans contribution, you have to
consider that the very existence of
surface loose just one neighbour. Why not half of them? You see atoms and molecules was controversial
my point: this is yet again an order of magnitude calculation. The at the time. The final demonstration
clear message, however, is that the story of molecules preferring to for the existence of atoms is generally
considered to be the work of Jean Perrin
be in the bulk rather than at the surface is indeed responsible for on Brownian motion, for which he
macroscopic manifestation of surface energies. received the Nobel prize in 1926. That
is 33 years after the death of Stefan!
So far we have talked mostly of liquids, but the concept of surface
energy applies to solids as well. Of course, it is difficult to imagine a
Wilhelmy-like experiment with a solid, but there are other manifes-
tations of surface energies for solids. For example, if you think of it a
mechanical fracture in a solid creates newly exposed surfaces, so frac-
ture criteria must have something to to with surface energies. Solids
with higher surface energies ought to be tougher is some sense. Ac-
tually a fracture is a process by which elastic deformation energy is
transformed into surface energy. This simple observation by Alan
Griffith resulted in a major breakthrough in solid mechanics, but we
will not have time to discuss this further.
What shall keep us busy in the remainder of the chapter is an
effect that is important for both solids and fluids, which bears the
name of Laplace pressure. The discovery that Thomas Young is refer-
Figure I9: Alan Arnold Griffith (1893-
1963)
I6 c.j. gommes, chim0698

ring to in the epigraph of this chapter is the observation that there


has to be a pressure jump across any curved interface. That pressure
jump is Laplaces pressure.
To understand the reason for this, consider two phases with vol-
umes V1 and V2 within a fixed total volume V = V1 + V2 , as sketch
in Fig. I10. The area of the contact between the two phases is A.
According to the general laws of thermodynamics, if you let the sys-
tem evolve it will reach an equilibrium state where free energy G is
minimal. In other words, the following differential should vanish
identically
dG = P1 dV1 P2 dV2 + dA (I13)

This means that if you change the shape of the two phases in any
A
imaginable way, the volumes V1 and V2 , and the interface area A
will obviously change but the free energy will be left unchanged. V1
To convert this statement into something more intuitive, about the
pressure difference P1 P2 , we need some mathematics. V2
Consider a volume like shown in Fig. I10. Each point of the sur-
face has a normal vector ~n pointing towards phase 2. Any infinitesi- Figure I10: Two phases with volumes
mal deformation of the surface can be described by a function V1 and V2 fill space. At equilibrium,
the free energy is left unchanged if the
surface is deformed in any possible
~r = ( x )~n (I14) way. This will eventually lead us to
Laplaces law.
where ( x ) is a function defined over the surface. If is positive
at a given point, the the surface is deformed outwards; it is deformed
inwards if is negative. The total change in volume can be written
in terms of as
Z
dV1 = d ( x ) (I15)
A

where the integral is over the total interface. Naturally dV2 is simply
the opposite of dV1 .
The change in surface area dA also depends on ( x ). According
to a famous theorem from differential geometry, the relation can
be written in terms of the local curvature of the surface. The actual
relation is Z To gain some intuitive understanding of
Eq. I16 consider the case of a spherical
dA = d K ( x )( x ) (I16) surface with radius R, with = dR.
A
In that case the total curvature is 2/R
where K ( x ) is the local total curvature of the surface at point x. Ge- everywhere on the surface so that the
integration is simply a multiplication by
ometrically, the total curvature is the sum of the inverse of the two
the area 4R2 . On the other hand the
principal radii of curvature at point x change in surface area is dA = 8R dR,
so that Eq. I16 is indeed satisfied.
1 1
K(x) = + (I17)
R1 ( x ) R2 ( x )

This is all the mathematics we need to put Eq. I13 is a more under-
standable form.
surface energies and laplaces pressure I7

With the help of Eq. I16, Gibbs equation can be put under the
form Z
dG = d [ P1 + P2 + K ] (I18)
A

Expressing that this quantity should vanish for any possible leads
us to the important result

P1 P2 = K (I19)

which is known as Laplaces law. It relate the pressure difference


across an interface to the curvature. Note that all quantities in the
equation may depend on the point considered on the surface. The
pressures P1 and P2 may vary from point to point, but they difference
is always proportional to the local curvature of the interface. It is
important to understand that the curvature can be either positive or
negative, as illustrated in Fig. I11.

a b c Figure I11: A few menisci config-


uration to illustrate the concept of
total curvature, in relation to the local
radii of creature. In (a) the curvature
is K = 2/R, in (b) it is K = 1/R be-
cause one radius is infinite, in (c) it is
K = 1/R1 1/R2 because the two
curvatures have opposite signs.

As we already mentioned, Laplaces pressure is not typical of


liquids. It applies to solids as well. Fig. I12 illustrates that point.
Let us see if we can find back the right order of magnitude of the
observed deformation from the known surface energy of aluminium
in Tab. I1 and from the macroscopic compression modulus K = 76
GPa. The latter quantity relates the change of volume of the material
to the pressure via
dV P
= (I20)
V K
In our case, the pressure P results from the surface tension, which in
the case of a spherical particle is written as

2 Figure I12: Compressive strain e on


P= (I21) the crystalline lattice parameters of of
R Ag and Al nanoparticles, determined
from the shift of scattering peaks of
The final relation between the change in volume and the size should
X-ray diffraction patters. Taken from
therefore be J. Phys. Chem. B 105 (2001) 6275. This
dV 2 contraction is a direct consequence of
= (I22) the Laplace pressure.
V RK
if our analysis is correct. Assuming a particle with radius R = 5 nm,
the predicted shrinkage of the nanoparticles should be of the order of
dV/V ' 1/5, i.e. e ' 1/15.
I8 c.j. gommes, chim0698

The agreement is not quantitative for several reasons. First the


compression modulus need not be the same for nanometer parti-
cles and for macroscopic Aluminium, if only because of the Laplace
pressure. Second, the nanoparticles are not spherical, and it is an ap-
proximation to model them as spheres. Anyway, our oversimplified
analysis predicts correctly the overall inverse relation between e and
the particle size, and the order of magnitude is correct.
Figure I13: A corrugated cylinder.
We shall finish the present chapter with interesting consequences It is stable or unstable depending
of surface energies, which are referred to capillary instabilities. These on whether the Laplace pressure is
larger at point 1 than at point 2, or the
instabilities result from the tendency of any system with a larger opposite.
surface-to-volume ratio to try and minimise its surface area. Consider
for example the case of a cylinder with radius R. You can clearly
lower the surface area if you transform it into a single sphere with
the same volume. The same hold if you decompose it into a series
of spherical droplets, provided the droplets are sufficiently large3 . 3
Try to prove that as an exercise. What
So a column is in principle unstable. This is something you know by is the smallest size of the droplets for
this to be the case?
observing a jet breakup into smaller droplets in a sink.
To see how this happens imagine that there some infinitesimal
corrugation to the column, so that it can be described by

r ( x ) = R + sin(2x/) (I23)

where x is the position along the column, R is the average radius, is


the amplitude of the corrugation, and is its periodicity. The ques-
tion we ask is: "Do we expect to increase with time or to decrease?".
The situation is sketched in Fig. I13. The column is unstable (and
increases) if the Laplace pressures at points 1 and 2 satisfy P2 P1 .
Such a pressure difference would indeed induce a flow of matter
from the thinner to the thicker parts of the columns, thereby increas-
ing its corrugation.
Let us calculate those pressures. For that purpose, it is important
to note that there are two orthogonal radii of curvature: one along
the axes of the column Rk , and in the other is R . The easiest to
calculate is R :
(
R + at point 1 Figure I14: Joseph Plateau (1801-1883).
R = (I24)
R at point 2

To calculate Rk , you have to remember from your calculus classes


that the radius of curvature of a curve parameterised as y( x ) is re-
lated to the second derivative of y via

1 d2 y
' 2 (I25)
R( x ) dx

where the ' sign means that the relation is valid provided the curva-
surface energies and laplaces pressure I9

ture is small enough. Applying that formula to Eq. I23, one finds
(
1 (2/)2 at point 1
= 2 (I26)
Rk (2/) at point 2

Note the minus sign in the second equation, which means that the
center of the tangent circle points outside of the surface, so that its
effect on the pressure via Laplaces law is negative. Putting all that
together, one finds the following pressures
"  2 # "  #
1 2 2 2 1
P1 = + ' + 2 (I27)
R+ R R

" 2 # "  #
2 2

1 2 1
P2 = ' 2 (I28)
R R R

where the second equality results from a Taylor development to the Figure I15: John William Strutt, 3rd
Baron Rayleigh (1842-1919).
first order in /R, which is justified because we have assumed that
was small. The condition for instability P1 P2 takes the simple form

2R (I29)
Isnt that a beautifully simple result?
Equation I29 means that a column is naturally unstable towards
corrugation with long periodicity. The periodicity that is actually ob-
served results from a balance between thermodynamics and kinetics.
Thermodynamics favours large values of , but the growth of such
an instability would be extremely slow because its would require the
transport of atoms/molecules over distances comparable to . The
values observed in practice are close to = 2R. This is illustrated
in Fig. I16, which shows the decomposition of a Cu nanowire into a
series of spherical nanoparticles.
For the records, this type of instability is know as the Plateau-
Rayleigh instability, after the famous british physicist Lord Rayeigh,
and the belgian physicist Joseph Plateau.

Exercises

1. Calculate the capillary length of mercury, with ' 0.5 N/M and
' 13 103 kg.m3 , as well as that of water on the moon, where
gravity is three times less than on Earth. Figure I16: Example of nanometer-
scale Rayleigh instability: spontaneous
2. Repeat the scaling analysis that we did for Bonds number to fragmentation of a Cu nanowire. Taken
from Appl. Phys. Lett. 85 (2004) 6275.
compare the surface energy to the kinetic energy of a flowing
liquid. That number is referred to as the Weber number We: find
an expression for it in terms of a characteristic length l, as well as
on any other relevant quantities.
I10 c.j. gommes, chim0698

3. Same question as before, but for the relative importance of the


viscous and surface forces. That number is generally referred to as
the capillary number Ca.

4. The list is longer than Ca and We: dimensionless numbers can be


derived to assess the relative importance of any physical force to
the capillary force. Can you think of other examples?

5. Analyze the definition of Bonds number Eq. I10 in terms of


Laplaces pressure.

6. Check the validity of Eq. I25 in the particular case of the equa-
tion p
y ( x ) = R2 x 2 (I30)
corresponding to a circle of radius R.

7. Figure I17 displays a string of air bubbles rising in a viscous


liquid (shampoo). Because air is almost inviscid, the pressure in
the bubbles can be considered as homogeneous. This is not the
case for the pressure in the liquid, which varies from point to
point. The variability of the liquid pressure controls the shape of
the bubbles via Laplaces relation. Considering the shape topmost
bubble in the figure, determine where the pressure on its surface is
the largest and where it is the smallest.

8. By a direct calculation of the surface area and of the volume cor-


responding to Eq. I23, show that the instability condition Eq.
I29 is equivalent to stating that the area should decrease when
increases, at constant volume.

9. Looking at Fig. I17, would you say that the capillary number (Cf. Figure I17: Air bubbles rising in a
viscous liquid.
point 3) is larger or smaller than one?

10. When pouring water on a hydrophobic surface, the water forms


a layer with thickness h, as sketched in Fig. I18. Assuming that
the lateral ends of the layer are circular, calculate the thickness h. h
Based on experiments that you can do at home, do you think this
is a good approximation? Do you expect the ends of the layer to be
Figure I18: Water on a hydrophobic
flatter or shaper than a circle?
surface. This is related to a more
general problem wittily referred to as
the housewife problem: when you pour
Further reading a bucketful of water on a floor, what
surface does it cover.
The series of lectures given by Thomas Young at the Royal Institu-
tion, from which the epigraph of the chapter is taken, are available
on Google books here. Browsing the table of contents of that book
will give you a hint of the type of universal genius that Thomas
Young was.
II0 c.j. gommes, chim0698

There are many good books on surface tension and wetting. An


excellent one is the following

de Gennes, Brochard-Wyart, and Quer


Capillarity and Wetting Phenomena: Drops, Bubbles, Pearls, Waves
Springer, 2004

The Nobel Lecture of Pierre-Gilles de Gennes is also interesting to


read.

Here is a paper from Nature (487, 2012, 463) about how nanometer-
scale Plateau-Rayleigh instabilities can be exploited to produce
nanoparticles with sophisticated structure.

Figure I19: Pierre-Gilles de Gennes


(1932-2007).
Wetting, a short introduction

But words are things, and a small drop of ink,


Falling like dew, upon a thought, produces
That which makes thousands, perhaps millions, think.

Lord Byron

The situations in which liquids are put in contact with solid sur-
faces are too many to enumerate, and their technological importance
can hardly be overestimated. The examples from either technology or
the natural world that immediately come to mind involve the paint-
ing of a surface, printing, lubrication, wicking of porous media by
Figure II1: Leonardo da Vinci (1452 -
a liquid, the creating of water-tight yet air-permeable fabrics, rain 1419)
drops sliding down a window or rolling down a hydrophobic leaf,
the walking of a insect on a pond, the capillary ascent of sap in the
xylem vessels of a plant, etc.
The topic of wetting is not a new one. The first documented study
about capillarity seems to be the deed of Leonardo da Vinci, and it
has been followed by a long series of scientists. A name you might
be familiar with is that of James Jurin who gave his name to the law
relating the height of capillary rise to the diameter of the capillary.
It is also a poorly known fact that Albert Einsteins first published
paper was about capillarity, in year 1900. Capillarity is an extremely
active field of research, and the work of Pierre-Gilles de Gennes (Fig. Figure II2: James Jurin (1684-1750)
I19) was instrumental in the bringing that topic to the forefront of
research.

Figure II3: Examples of a liquid


deposited on a clean surface, displaying
a typical truncated sphere shape with
a characteristic contact angle. This is
the case of water on differently treated
glasses, taken from Phys. Chem. Chem.
Phys. 6 (2004) 604.
II2 c.j. gommes, chim0698

The simplest question that you may want to ask is whether


a liquid will spread uniformly on a solid or not. This is a ques-
tion that is loosely referred to as that of the hydrophilicity or hydropho-
bicity of the surface. It is central to the question of painting a surface,
for example. If you are practically minded you may start by doing
experiments and deposit drops of liquids on the surface and see
what happens. What you will probably obverse at your first trial is
something like in Fig. II3: the liquid will adopt a truncated spher-
ical shape. If you repeat the experiment with a variety of liquids on
the same surface, you will notice that the angle changes from one
liquid to the next.
After experimenting with a few liquids, you may start to ratio-
nalise your findings in terms of sound physicochemical quantities.
Because small values of go along with large areas of the free sur-
face of the liquid, it seems natural to see whether the liquid surface
tension explains the variability of . After playing a bit with the Figure II4: William Zisman (1905-
1986). Picture taken from here.
data you will probably end up plotting them as 1 cos( ) against ,
and you would observe something like Fig. II5: the points fall on a
single line, which you write empirically as

1
= (II1)
cos( ) c

The reason why it makes sense to call the denominator c is that it is


a critical surface tension. When = c , the contact angle reaches the
value = 0. From this, you may infer that any liquid with c
would spread perfectly on the solid. This type of characterisation of
wetting, in particular II1 is called Zismans law.
The values of c for a few solids are given in Tab. II1. In the case
of clean glass, the value is quite large (c ' 150 mN/) which shows
Figure II5: Zisman plot, by which a
that any liquid with surface tension lower that than will spread on it.
series of liquids are deposited on a
This is notably the case for water and of most organic solvents (see given surface and the contact angle
Tab. I1 on page I3). On the contrary, polyethylene will be wetted is measured. The extrapolation to
cos( ) = 1 yields the critical surface
by most organic solvents but not by water. This corresponds to the tension of the surface. In this particular
common observation that polyethylene is hydrophobic. You may also case, the surface is terphene, and the
critical surface tension is found to be
note from the values in Tab. II1 that surface treatment may have
close to 25 mN/m.
dramatic effect, as already suggested by Fig. II3. Almost no liquid
will spread on fluorinated glass, except perhaps liquid nitrogen.

If you are more theoretically minded you may want to think


thoroughly about the question of wetting before experimenting. You
may come up to the following. In a situation like that of Fig. II3
there are three different interfaces, between the solid and air, the
liquid and air, as well as between the solid and the liquid. Each in-
terface has its own surface energy. The condition for the spreading of
wetting, a short introduction II3

the liquid on the surface necessarily depends on these three surface Solid c (mN/m)
energies. Clean glass 150
Fluorinated glass 10
We use the notations S , L , SL for the surface energies of the Polyethylene 31
solid, of the liquid, and of the solid-liquid interface. Assuming that PVF4 18
you know all the relevant surface energies, you can calculate the Table II1: Critical surface tension of
between the dry solid and the wetted solid. Because that quantity is a few solids. PVF4 is a fluorinated
directly related to the tendency of the liquid to spread, it is referred polymer.

to as the spreading parameter S. It can be written in terms of surface


energies as follows
S = S (L + SL ) (II2)
where the term between brackets is the energy of the wetted solid.
If S > 0 it is energetically favourable for the liquid to spread, which
happens for large values of S and a lower value of L . It may happen that I sometimes forgot
the subscript L in L , in line with the
The case S < 0 leads to partial wetting, in which the contact angle
notation of Chapter 1.
takes a finite value. All examples of Fig. II3 correspond to that
situation. In order to relate the contact angle to the surface energies,
let us consider the sketch of Fig. II6, which shows a close-up view
of a triple line between a solid, a liquid, and air. The total free energy
of such a system has contributions proportional to the area of each
interface, which can be written as
G = . . . + S AS + L A L + SL ASL (II3)
In this equation AS , A L and ASL are the surface areas solid-air,
liquid-air, and solid-liquid interfaces. The shape of the liquid on the
solid, and in particular the value of , is such that the free energy is L
stationary, i.e. that G is left unchanged by slight deformations of the
dx cos()
drop. The dots in the equation stand for the volume contributions,
LS S
which are assumed to be constant and are irrelevant in the present dx
context.
If you have in mind how we obtained the expression of the Laplaces Figure II6: Close up view of a triple
pressure, you may remember that the free energy has to be stationary line between a solid S, a liquid L and
air, with definition of the contact angle
with respect to any possible deformation. Let us see, however, if we as well as of the surface energies S ,
can learn something useful by considering only stationarity with re- L and SL .
spect to displacements of the triple line. If the triple line moves by an
infinitesimal quantity dx as shown in Fig. II6, the areas are modified
according to
dASL = l dx
dAS = l dx
dA L = l dx cos( ) (II4)
where l is the length of the triple line, and is the contact angle.
Introducing these expressions in the variation of the free energy Eq.
II3 leads to the following relation
S SL = L cos( ) (II5)
II4 c.j. gommes, chim0698

which is usually known as the Young-Dupr equation. In the par-


ticular case where the spreading parameter is equal to zero the
Young-Dupr equation predicts = 0, as it should. For S > 0,
the equation has no solution because this would require cos( ) > 1.
As we shall see in the rest of the chapter and also in the next, the
contact angle is often sufficient to calculate many useful properties of
a system involving wetting. In other words, it is often unnecessary to
know S , SL and L in more detail than though their combination
(S SL )/L .

Figure II7: Example of nanometer-


scale wetting: environmental TEM
shows that Cu nanoparticles adopt
different shapes in presence of different
vapours. Figure taken from Science 295
(2002) 2053.

Examples of wetting with different contact angles are given in Fig.


II3 and in Fig. II7. In the former case, It is the surface that is modi-
fied via plasma treatment, which resulted in varying S and SL and
hence . The latter case is more interesting: it consists in the wetting
of solids at the nanometer scale. For example, the difference between
Fig. II7A and II7E is the presence of carbon monoxide (CO), which
adsorbs on the Cu surface and reduces its surface energy. As a conse-
quence, the nanoparticle is found to flatten, in qualitative agreement
with the Young-Dupr relation. In the case of crystalline solids, the sur-
face energy is generally dependent on
It is interesting to look back on Zismans law in the light of the
the particular facet considered. In that
Young-Dupr equation; it can be re-written as case, the lower energy configuration
is not strictly a sphere, but a faceted
1 L shape known as Wullfs shape (see Fig.
= (II6)
cos( ) S SL II7). Most results obtained with liquids
remain, however, qualitatively correct in
which suggest that c = S SL . One has to be careful, however, the case of solids.
that SL is liquid-dependent so this is not the same as Zismans law.
The empirical observation that Zismans law is correct seems to sug-
gest that SL is almost liquid-independent. This might be reasonable
in the case of homologous liquids, e.g. a series of oils with increasing
molar mass or surfactant solutions with increasing concentrations.
However, Zismans law is empirical: there is no reason to believe that
the critical surface tension of a given solid would be the same for oils
wetting, a short introduction II5

and for aqueous solutions of surfactants. You will have to wait until
we discuss the physical origin of intermolecular forces in Chapter IV
to understand this better.

Before we investigate the consequences of the Young-


Dupr equation in some specific situations, it might be useful
Figure II8: Self-explanatory illustra-
to draw your attention to some limitations of that law, so that we can tion of contact angle hysteresis. The
afterwards knowingly forget about them. When playing around with advancing contact angle (top) is gen-
erally arguer than the receding angle
liquid droplets on a surface you immediately notice the phenomenon (bottom). Figure borrowed from here.
of triple-line pinning, by which the triple line seems to stick to the
surface. That phenomenon is illustrated in Fig.II8. In general, for
a given liquid on a given solid, the angle can take any value in an
interval r a , where the subscripts a and r stand for advancing
or receding. The triple line moves macroscopically only when these
angles are exceeded.
a)
Some observations as familiar as the sticking of a water droplet 1 2
on a vertical window pane would be impossible without triple-line
pinning. It is because the contact angles at the top and at the bottom
of the droplet are different that such a droplet can be in mechanical b)
equilibrium. See also exercise XY for that matter.
1 2
The origin of contact angle hysteresis can be found in the hetero-
geneity of the surface. This is expected notably if a surface is made
up of microscopically small patches with slightly different contact
Figure II9: Two type of surface hetero-
angles 1 and 2 , as illustrated in Fig. II9a. In that case, the contact
geneities that may lead to contact angle
angle passes from one value to the other when the triple line passes hysteresis: (a) chemical heterogeneity
from patch to the next. If the characteristic sizes of the patches are and (b) structure heterogeneity.

very small, from a macroscopic perspective, it seems that the triple


line has not moved. Another situation in which contact angle hystere-
sis is expected is the type of surface roughness shown in Fig. II9b.
In that case, the microscopic contact angle of the liquid with the sur-
face is uniquely defined. However, the macroscopic contact angle is
subject to hysteresis.
Another interesting consequence of surface roughness, besides
hysteresis, is the following. Surface roughness increases he actual Figure II10: The Westinghouse Elec-
contact area of the liquid and the solid, and this is expected to in- trical Company research labs in Pitts-
burgh, where Robert N. Wenzel was
fluence the contact angle. To see how, repeat the derivation of the employed when he published his work
Young-Dupr relation, by replacing the areas dAS and dASL in Eq. on the wetting of rough surfaces. The
recognisable dome is a Van De Graaff
II4 by r dAS and r dASL , where r is the roughness factor of the
generator.
surface. This leads to the following relation for the contact angle
on the rough surface, compared to the contact angle on the chemi-
cally identical smooth surface.

cos( ) = r cos( ) (II7)

Because r is necessarily larger than one, contact angles larger than


II6 c.j. gommes, chim0698

90 are increased by surface roughness, and contact angles smaller The roughness factor of the surface r -
also called the rugosity - is defined by
than 90 are made even shallower. In other words, roughness en-
IUPAC as the ratio of the actual surface
hances both hydrophilicity and hydrophobicity. Eq. II7 is known as area to the apparent (projected) surface
Wenzels law. area.

A common illustration of wetting phenomena is the wicking


of a porous solid by a liquid, as illustrated in Fig. II11 which is
achtypica capillarity problem. It is an empirical observation by James
Jurin (Fig. II2) that a wetting liquid rises in a vertical capillary tube
in inverse proportion to the diameter of the tube. This situation can
be modelled by assuming that the pore space of the material (the
sugar cube) consists in cylindrical tubes. For methodological reasons,
we shall analyse that situation in two different manners: first based
on surface energies, and then on Laplaces pressure.
The energy of the liquid column has two contributions: the gravi-
tational and surface energies. The gravitational energy can be written
as
Figure II11: The wicking of a sugar
"   #
d 2 h cube by coffee is a common illustration
Ug = g h (II8)
2 2 of capillarity.

where h is the height of the column, d is the diameter of the capillary,


and is the liquid density. The term between square brackets is the
volume of the column and h/2 is the height of its center of mass. On
the other hand, the surface contribution is

Us = dh [SL S ] (II9)

It is the difference of surface energies that have to be considered


because the dry solid is replaced with wet solid when h increases.
Looking for the value of h that minimises Ug + Us yields immedi-
ately the value
4(S SL )
h= (II10)
gd
which is the celebrated Jurins law. It states that the liquid rises in a
capillary provided S > SL , and the thinner the capillary the larger
the rise. Try to think of the molecular origin of
Here is the second analysis, based on Laplaces pressure. If you Jurins law in relation to the surface-to-
volume ratio of the capillary.
look close to the surface of the liquid at the top of the capillary, you
will notice that the surface is curved. Provided the diameter of the
capillary is much smaller than the capillary length4 , the free surface 4
Be sure you understand this, and go
will have a spherical shape with radius of curvature given by back to chapter I if necessary.

d
R cos( ) = (II11)
2
As a consequence of that curvature, the pressure in the liquid just
wetting, a short introduction II7

h
Figure II12: Two different and equiva-
lent approaches to analyse the rising of
a liquid in a capillary, based on energy
minimisation or on Laplaces pressure.
d

below the meniscus is at a pressure

2L
P = P0 (II12)
R
where P0 is the ambient pressure. On the other hand, that pressure
can also be calculated from the height of the liquid as

P = P0 gh (II13)

Equating these to independent estimations of P brings you to

4L cos( )
h= (II14)
gd

which is exactly identical to Eq. II10 is the Young-Dupr equation


holds.
In our discussion of Jurins law so far, and particularly in Fig. II
12, we have implicitly assumed 0 or equivalently SL S . All
the results that we have obtained are still valid in the case of > 90,
but the formulae predict a negative capillary height. This means that
when a capillary is dipped into a non-wetting liquid, the latter does
invade the capillary until the point where the hydrostatic pressure
balances exactly the Laplace pressure, which is in this case positive.
An interesting way to look at it is to consider that you have to
exert a given pressure to push a non-wetting liquid in the capillary.
In a Jurin-like experiment that pressure is provided by gravity, but
you may also imagine immersing the capillary in liquid and increas- Figure II13: Example of two mer-
ing mechanically the pressure. The pressure that you would have to cury (intrusion) porosimetry curves
measured on two porous materials
reach to push the liquid in is with different pore sizes (courtesy of
Nathalie Job).
4L cos( )
Pi = (II15)
d
where the subscript i stands for intrusion. Note that Pi can be directly
obtained by applying Laplaces law at the meniscus. It is interesting
to note that the pressure that you have to exert to push a non-wetting
liquid in a capillary is inversely proportional to its diameter. The
II8 c.j. gommes, chim0698

smaller the diameter, the larger the pressure. The deep reason for this
is that wetting phenomena in pores scale like the surface-to-volume
ratio of the pores. You may at first think of mercury
That observation can be used practically measure the size of pores porosimetry as a technique from an
older age. After all, why not simply do
it porous materials, by a technique called mercury porosimetry, or electron microscopy to measure pore
sometimes intrusion porosimetry. The technique consists in immersing sizes? The reason is that the amount of
material that you actually characterise
the unknown porous material in mercury, and increasing progres- when doing microscopy is extremely
sively the pressure. At each particular pressure the volume of mer- small. A fragment 100 nm across, which
cury that has entered the pore space is measured. Mercury is used is the largest you can analyse in a
microscope with nanometer resolution,
for those experiments because it does not wet any known solid. All has a mass of 1015 g. Compared to
solids are mercuro-phobic, so to speak. the kg or even ton of material that is
used in any industrial context, this is an
A typical mercury intrusion curve is shown in Fig. II13: the in- extremely poor sampling. The strength
truded volume is plotted as a function of the pressure. One observes of techniques like mercury porosimetry
that intrusion occurs at a specific pressure that can be related to the is that they enable you to characterise
a large quantity of material, with a
pore size via Eq. II15. In practice, the accepted value for the contact nanometer-scale resolution.
angle of mercury with any material is 142 , which together with
the value of L given in Tab. I1 leads to the following quantitative
relation
1500
d[nm] (II16)
Pi [ MPa]
between the intrusion pressure and the pore size. In the case of the
figure, with intrusion pressures Pi = 20 MPa and Pi = 50 MPa, the
pore sizes are found to be 75 nm and 30 nm, respectively. In typical
mercury porosimeters, the pressure can reach 200 MPa, which means
that pores as small as 7 nm can be detected. In case you do not remember the
There is an interesting aspect to intrusion porosimetry that is not Hagen-Poiseuille, you can always find
it back using dimensional analysis. In
often discussed, which concerns the kinetics of fluid intrusion. When dimensional form: the volume flow
the pressure in the fluid is suddenly brought to a value larger than Q [m3 /s] depends on the pressure
gradient dP/dx [Pa/m], on the viscosity
the intrusion pressure Pi , how fast does intrusion take place? This [Pa.s] and on the tube diameter d [m].
question can be answered by noting that when a liquid invades a Only a single dimensionless number
porous material, there is a pressure drop in the liquid dP per unit can be formed, e.g. d4 dP/dx/(Q),
which can only be equal to a constant.
length dx of a pore, which can be approximately described by the This is exactly Eq. II17, except for the
Hagen-Poiseuille law, namely factor 128.

dP 128Q
= (II17)
dx d4
where is the viscosity of the fluid, Q is the volumetric rate, and d is
the pore diameter. Imagine that at given time t, the liquid has entered
the pore over a length L(t), from the outer surface of the material to
the meniscus. Over that distance, the pressure drops from P at the
outer surface to Pi just behind the meniscus. This enables us to write
the Hagen-Poiseuille law again as

d4
Q= ( P Pi ) (II18)
128L(t)
wetting, a short introduction II9

Because Q = (d/2)2 dL/dt, this can be expressed as a differential


equation, the solution of which is
s
L(t) =
d t
[ P Pi ] (II19) a)
4

The intrusion is proportional to t: it is initially very rapid, and then
very slow. This is the typical dynamics of many wetting phenomena.
b)
Although the fundamentals of wetting seem to be rela-
tively well understood, the way in which they apply in specific
natural or technological situations can be quite subtle. For example, Figure II14: Cassie-Baxter type of
wetting, in the case of (a) hydrophobic
consider again Wenzels law for the wetting of a rough surface (Eq. and (b) hydrophilic surfaces.
II7). When experiments are done on very rough surfaces, significant
deviation from Wenzels law is often detected. Wenzels law is valid
only for surfaces with small rugosity.
To understand why this happens, remember that when deriving
Wenzels law we have assumed that the entire surface was either dry
or wetted by the liquid. But this need not be the case. Based on what
we said about Jurins law in the case of a non-wetting liquid (Cf.
mercury porosimetry), you may expect something like Fig. II14a
for a drop deposited on a rough surface. In that case, the apparent
energy of the wetted surface SL could be modelled like


SL ' SL + (1 )L (II20)

where is the fraction of the surface where the two phases are actu-
ally in contact. On the remainder fraction of the surface, the interface
energy is simply L because the crevices of the surface are filled with
air. You may try to incorporate that type of thinking into our demon-
stration of the Young-Dupr equation, and this will lead you to a
different relation between surface roughness and contact angle.
Similarly, in the case of a wetting liquid you may think of the
crevices on the surface as of a porous material, which may be in- Figure II15: The Cassie-Baxter-like
super-hydrophobicity of the leafs
truded by the liquid following the usual laws of capillarity. In that of some plants is often referred to
case, on a microscopic scale things may look like in Fig. II142, which as the lotus effect. Picture from from
Wikipedia.
is also different from the Wenzels model. In that case too, you may
use an approach similar to Eq. II20 to derive the apparent energy of
the dry surface S .
The model corresponding to the wetting mechanism of Fig. II14
is known as the Cassie-Baxter model. In the particular case of hy-
drophobic surfaces it is sometimes referred to as the Lotus effect. The
surface of the lotus leaf is covered by a waxy substance that confers it
with hydrophocity; in addition to that, the surface comprises numer-
ous micrometer-sized posts that make it similar to Fig. II14a, and
confers super-hydrophobicity to the surface.
II10 c.j. gommes, chim0698

In myriads of biological contexts, natural selection has lead to


structured surfaces having wetting properties desirable to the species.
This is the case of super-hydrophobic leafs, which help keeping the
leafs clean from any macroscopic impurity. This is also the case of the
legs of insects walking on water, the surface of which is covered by
hydrophobic hair. The development of artificial surfaces with engi-
neered wetting properties is an extremely active field of research. If
that topic is of interest to you, you will find some suggested readings
hereunder.

Exercises
Figure II16: The legs of water strid-
1. Capillary action is one of the two mechanisms responsible for ers are covered with hydrophobic
hair, which confers them super-
the ascent of sap in plants from the roots to the leafs, in addition hydrophobicity. Picture from from
to osmotic pressure. Try to guess the order of magnitude of the Wikipedia.
diameter of xylem vessels, and then compare its with actual data.

2. A bucket of plastic is filled with water, and you punch with a a) b)


needle a hole about half a millimetre wide twenty centimetres
below the surface. Does water flow through that hole? You may
assume the contact angle of water with the plastic making up the
bucket is about 140 .
Figure II17: (a) The so-called capillary
3. The drawing in Fig. II17a is a historical proposal for a perpetual cup perpetual motion setup, and (b)
motion machine. I have proposed as simplified version of it in a simpler version of it, whereby a
Fig. II17b: it basically consists in a Jurin-like experiment with the capillary tube shorter than the capillary
height is dipped in water.
height of the tube shorter than the height predicted by Eq. II10.
Debunk this.

4. Think about Jurins law in the case where the spreading parameter
S is positive. Does Eq. II14 hold?

5. Next time you use a straw for drinking, pay attention to the small
drops of water that may remain stuck in the straw, while the larger
ones flow under the effect of gravity. The reason for this is contact
angle hysteresis. Write down a balance of force to determine the
maximum length of a column of water that may be blocked in a
straw, in terms of the advancing and receding contact angles a
and r . Do some experiments at home with water and a straw to
try learning something quantitative about a and r .

6. Think about the double-sided floater shown in Fig. II18. Doesnt Figure II18: A double-sided floater
with each side made up of a specific
something puzzle you about the horizontal equilibrium of such material, having different contact angles
an object? The force to the right has to be equal to the force to with water.
the left, otherwise it could serve as a perpetual motion machine.
Enumerate all the forces involved.
III0 c.j. gommes, chim0698

7. In Wilhemys plate experiment for measuring (see Fig. I4), the


plate is made of roughened platinum. Explain why this is important.

Further reading

Here is a website hosted at MIT, on the general subject of wetting


phenomena, with many videos;

David Qur gave two lectures about the effect of surface texture
on wetting, which you can find on Youtube the first here and the
second here. This is a fascinating and very timely field of research;

Here is a paper published in Science magazine, about Surface


Tension Transport of Prey by Feeding Shorebirds: The Capillary Ratchet
(Science, 320 (2008), 931). It is about how a seabird exploits contact
angle hysteresis;

Here is another paper about a desert beetle having hydrophobic


and hydrophilic patches on its back, which it uses to collect water
from fog-laden wind (Nature 414 (2001) 33).
Solved problems involving wetting

And so castles made of sand, melts into the sea eventually.

Jimi Hendrix

The present chapter is about some aspects of wetting phenomena


that would not justify a theoretical analysis in themselves, but which
are interesting to be aware of. They are treated here in the form of
problems. Most of them touch on the mechanical aspects of wetting.
An illustrative example is the cohesion of wet sand, which justifies
the epigraph of the present chapter and is treated in Problem 3. An-
other one is capillary stresses in partially saturated porous solids.
Figure III1: Jimi Hendrix (1942-1970)
That situation is presented in Problem 4 in the medical context of
infant pulmonary distress but, the physical phenomenon is quite
general.
On an almost philosophical note: you may have noticed from
your students life so far, that in many instances the creative part of
problem-solving is taken out of your hands. The following example
is borrowed from Eric Mazur5 : You drive to the supermarket and you 5
I am pretty sure of it but I cannot find
find no free parking spot; would you better wait until a car leaves and frees a the exact reference back. Eric Mazur is
a physics professor at Harvard, who is
spot, or try to park your car elsewhere? This is real-life decision that you famous notably for the development of
have to take, ideally based on some engineering-like rational analysis. Peer Instruction.
However, if that question was posed in a teaching context it would
probably be stated as follows. Given that there are 120 cars on the park-
ing lot, and that every customer stays about 20 min in the store, how long
will you have to wait on average for a free spot? In that form, the answer
is straightforward: you can expect to wait about 10 s. You should
definitely wait!
Note the leap between the actual problem and the way in which
the question is formulated for students. Of course the question is
easier, and I understand why you would rather have that type of
question to solve at the exam. However, all the creative and interest-
ing part has been taken out of it. Dont fool yourself: what will be
socially expected of you as a scientist is to answer the first type of
question, not the second!
III2 c.j. gommes, chim0698

In each problem hereafter, I state clearly the question that you


should try to answer. Finding the answer to that question is what
is really expected of you. You should try to answer it with no addi-
tional hint. If only to understand what is really expected of you, and
to measure its difficulty. It is generally acknowledged that
However, because the purpose of the present notes is teaching, it personal problem solving is central
to the development of mathematical
realise I have to guide you somehow in that process. This is the rea- skills. See the role of mathematical
son for which you will also find a series of subquestions that should puzzles. For some reason people
seem to think that this does not apply
help you get the final answer, if you cannot find one by yourself. That to other sciences... I really do not
list of subquestions is by no means the only way to get the answer. It understand why.
is just one possible way, that is either the only I found or the easiest
for me to explain.
There is room is science and engineering for personality. If you
find your own idiosyncratic approach to solving scientific problems,
wouldnt that make it more interesting?
solved problems involving wetting III3

A. Wetting in zero gravity


You can find on Youtube some movies
It is often reported in popular culture that a glass would not be able involving liquids in zero gravity. I
to hold any liquid, should gravity be absent. A famous illustration found this one particularly interesting.

is in Fig. III2. This is, however, not accurate because there is an


adhesion energy between water and glass (SL L ) that is expected
to keep the liquid inside the glass. Of course, the shape of the free
surface has no reason to be flat.

What is the equilibrium shape of a liquid in a glass in absence


of gravity, in relation to the surface energy of all relevant inter-
faces?

The unimaginative steps that you may follow if you feel com- Figure III2: This is roughly what this
pletely lost are the following exercise is about

1. Prove that the shape of the free surface of water is a spherical cap.

2. Assume a cylindrical glass with radius r. Given that the free sur-
h
face is a spherical cap and that the total volume of liquid V is a
r
given, express the total surface energy of the system in terms of a
single well-chosen geometrical parameter. And find the equation R
that the parameter has to satisfy at equilibrium.

3. Express the single geometrical parameter in terms of the contact


angle and express everything in a simple way in terms of the Figure III3: Definition of a spherical
cap. The area is A = r2 + h2 and its

Young-Dupr relation. volume is V = 6 h 3r2 + h2 .
III4 c.j. gommes, chim0698

Solution

1. Of course, the sphere is the geometrical shape that minimises the


area for a given volume. Note, however, that this is not directly rel-
evant here because there are other contributions to the energy than
the water/air interface. The simplest way to prove that the surface
has to be spherical is via Laplaces law: in absence of gravity the
pressure has to be is uniform in the drop, so that the curvature at
each point of the surface has to be the same. The only possibility is
a spherical surface.

2. The overall configuration is shown in Fig. III4, where some nota-


tions are also introduced. The free surface is a spherical cap, which
is uniquely specified through its radius R for a given radius r and
height h2 . Equivalently, the radius R can be replaced by the height
h1 . For a given volume of water V, the height h1 and h2 are not
independent because they are related through
 
V = r2 h2 h1 3r2 + h21 (III1)
6
the second term in this equation is the classical expression for the
volume of a spherical cap.
In absence of gravity the only contributions to the energy are those
from the various surfaces involved. Those are (1) the free surface
of water having surface tension and area A, (2) the surface of
the wet glass with surface energy SL and area Aw , and (3) the
Figure III4: Shape of the free surface
surface of the dry glass with surface energy S and area Ad . The and geometrical meaning of symbols.
water spontaneously adopts the configuration that minimises the
following energy

U = A + SL Aw + S Ad (III2)

The areas of the wet and dry parts of the glass are simply ex-
pressed as

Aw = 2rh2 and Ad = 2 ( H h2 ) (III3)

with the notations of Fig. III4.


The area of the cap is calculated by the classical formula
 
A = r2 + h21 (III4)

Using Eq. III1 to express h2 in terms of h1 , the expression of the


energy takes the form
  2 
U = r2 + h21 + V + h1 (3r2 + h21 ) (SL S ) + 2rHS
r 6
(III5)
solved problems involving wetting III5

The equilibrium condition dU/dh1 = 0 leads to


 2  
h1 h1
2 +1 = 0 (III6)
r S SL r

from which h1 can be calculated.

3. Instead of calculating h1 , it is more intuitive to calculate the con-


tact angle between the water and the glass that corresponds to Eq.
III6. Simple trigonometric calculations lead to the relation

2h1 /r
cos( ) = (III7)
1 + (h1 /r )2

This can be put as


 2  
h1 1 h1
2 +1 = 0 (III8)
r cos( ) r

Comparing Eqs. III6 and III8 shows that the value of h1 that
minimises the energy is exactly the one satisfying the Young-
Dupr equation. The practical problems of astronauts is
more often to get water out of the glass,
The answer is therefore simply that the configuration is such that than to put it back in as in Fig. III2.
Look here to see the contraption they
the free surface of the liquid is a spherical cap, and the contact angle use for drinking.
satisfies the Young-Dupr equation.

Thinking further

In absence of gravity, the pressure in the liquid is uniform. What is


the value of the pressure?

Equation III6 has two solutions. What do they correspond to?


III6 c.j. gommes, chim0698

B. The sticking together of two wetted plates


Look here for a Youtube movie of a
You have probably already played with microscope slides, which capillary bridge between two plates. Be
stick strongly together when they are wet. If you did so, you proba- careful, however, that their explanation
they give for the origin of the force is
bly also noticed that the further apart the two plates are, the easier it not accurate.
is to draw them separate them further. The attraction force between
the two plate is therefore inversely related to the separating distance.
The typical amplitude of roughness6 of glass is lower than 1 m, 6
See e.g. here for the roughness of
which means that this is about how close you can bring two glass some common materials.

plates together. Imagine two microscopy slides with area 2 cm2 and
water in between.

What is the value of the force that pushes them together assum- r
R h
ing = 75 mN/m and = 60 ?

The unimaginative steps that you may follow if you feel com-
pletely lost are the following Figure III5: Two flat plates, with a thin
liquid layer in between. In this view,
1. Looking at Fig. III5, what is the pressure in the liquid relative to the liquid layer is assumed to have
the shape of a disk with radius R and
the outer pressure P0 , as a function of h? Simplify that expression
thickness h.
for the case R  h;

2. You are pulling the two plates apart with a force F. In order to
determine the value of F that can overcome the capillary forces,
imagine cutting the liquid layer horizontally and account for all
forces exerted on, say the upper half of the system. Simplify again
that expression for R  h.
solved problems involving wetting III7

Solution

1. The two radii of curvature of the interface are r and R. The pres-
sure in the film is therefore
 
1 1
P = P0 + (III9)
R r

where we have assigned a different sign to each curvature, as we


should. From elementary trigonometry, the radius of curvature r is
related to h by
h
r= (III10)
2 cos( )
The final expression is therefore

2 cos( )
Pc = P0 (III11)
h
where we have taken account of R  h. The subscript c stands for
capillary pressure.

2. If we imagine cutting the system horizontally in two identical


parts through the centre of the layer, the forces pushing the system
up are
Fup = F + Pc R2 (III12)

The forces pulling down are

Fdown = P0 R2 + 2R (III13)

where the second term is the direct contribution of the surface. 7 7


If this is not clear to you, go and
Equating both terms leads to read again our analysis of the forces in
Wilhemys plate experiment on page
I2.
2 cos( )
F = R2 + 2R (III14)
h
Note that the second term is approximately R/h times smaller
than the first one, so that it is of the same order of magnitude as
the term that we neglected in Laplaces pressure. To be consistent,
the second contribution has to be neglected.
The rest is just a numerical application. The capillary pressure
is Pc = 75 kPa. Multiplying this by the area leads to a force of 15
N. That is about the weight of 1.5 kg. I am not surprised by that
figure. Are you?

Thinking further

1. If the spacing h is increased indefinitely, while preserving the


volume of the liquid, how does the force F depend on h?
III8 c.j. gommes, chim0698

2. The general expression of the force that you found in point 2


changes sign for a particular value of the contact angle . Explain
why.

3. The calculation of F that we discussed was based on Laplaces


pressure. Redo the same calculation based on energies. It may
help you to go back to chapter II and read again the two deriva-
tions of Jurins law.
solved problems involving wetting III9

C. The capillary force between two spheres


A Youtube movie showing a capillary
It is a fact known to all kids that wet sand is better for making cas- bridge between two glass beads is
tles than dry sand. The reason is that water binds the sand grains visible here.

together.
At the level of the sand grains, the situation is similar to that
shown in Fig. III7. For the purpose of the present problem you
may assume that the contact angle between the water and the sand is
= 0.

What is the force that pulls the two sand grains together in the
limit of small amounts of water? Figure III6: Wet sand is stickier than
dry sand because of Laplaces pressure
in the liquid bridge that forms at the
We shouldnt give you any hint for the present problem, except contact between two grains.
that it looks very much like the previous one, doesnt it?

r Figure III7: Two touching spheres


R with a liquid meniscus at their contact
point.
x
III10 c.j. gommes, chim0698

Solution

The quantity x in Fig. III7 is a useful measure for the amount of


liquid in the bridge. The limit of very small amounts of water is
equivalent to the limit x 0. We shall therefore first express all
relevant quantities in terms of x.
The smallest radius of curvature in the system in r, which we may
express as a function of x using the Pythagorean theorem as

( x + r )2 + R2 = ( R + r )2
x2 + 2xr = 2Rr (III15)

Because the point of contact between the two spheres is locally flat,
r decreases more rapidly than x for x 0. In the case where r  x,
the relation between r and x simplifies to

1 2R
= 2 (III16)
r x

The pressure in the liquid is then


 
1 1 2R
Pc = P0 + ' P0 = P0 2 (III17)
x r r x

The equilibrium is analysed in the usual way: cut mentally the


system by a vertical plane passing between the two spheres and
express that each sphere is mechanically equilibrated. This leads to
 
2 2R
P0 x + 2x = P0 2 x2 + F (III18)
x

The first terms on the left- and right-hand sides (proportional to x2 )


are the integrals over the corresponding surface of the sphere of the
pressure force. This is not as simple as it may seem: it is the result
of an integral over a spherical surface of a projected force, which
happens to be proportional to the projected area of the surface. Be
sure you understand it.
The final result is then simply

F = 2x + 2R ' 2R (III19)

When the quantity of water in the meniscus becomes vanishingly


small (i.e. for x 0), the force converges to a constant value 2R,
which only depends on the radius of the particle. The reason for this
is that the capillary pressure happens to be inversely proportional to
the surface over which it is exerted, so that the force is constant. This
is is a geometrical accident.
solved problems involving wetting III11

Thinking further

1. In practice, a surface is never so smooth enough for the asymp-


totic limit of the previous paragraph to hold. At a very small scale,
the surface of contact looks more like many small spheres touch-
ing each other than two large spheres. How does that affect the
force?

2. Generalise the analysis for arbitrary values of .

3. Try to apply an analysis based on energies to calculate the force


between the two spheres.

4. Do the same analysis for the capillary force between two cylin-
ders. This might be relevant to the sticking of your hair together
when you get out of a shower.
III12 c.j. gommes, chim0698

D. Capillary stresses and surfactant deficiency disorder

The final structure in the respiratory tracts of mammals is the pul-


monary alveolus III8, which is a locally spherical cavity, with ap-
proximate diameter d = 100 m. The inner surface of alveoli are
coated with a liquid having a finite surface tension, which would
tend to make the structure of the lung collapse. To prevent this from
happening, the alveolar cells produce a lipoprotein complex that acts
as a surfactant, and reduces the surface tension of the liquid.
Figure III8: Alveoli are the final and
In premature newborns, the alveolar cells are unable to produce
smallest structures in the lung, at which
that protein. This is the origin of so-called Surfactant Deficiency scale the exchange between between the
Disorder, which may be fatal. air and the blood takes place (Image
from Wikipedia).

What is the order of magnitude of the (negative) pressure that


has to be exerted on a lung to prevent its alveoli from collapsing?
For a fuller account of SDD, including
The unimaginative steps that you may follow if you feel com- some historical aspects, see e.g. S.
Wrobels Bubbles, Babies and Biology:
pletely lost are the following The Story of Surfactant published in
the FASEB Journal, 18 (Oct. 2004).
1. The free surface of the liquid in the lungs has a given total area
Av per unit volume. Imagine cutting off a cubic piece of lung with
side l, and deforming it by a small quantity dl in all directions.
How does the surface energy change? What is the energy cost of
that change?

2. Once you have solved point 1, all you have to do is propose a


crude structural model of a lung, to estimate the area AV . For
example, you may assume spherical alveoli with diameter d filling
a fraction e of the total volume of the lung. The value e = 0.5 is
probably close to reality.
solved problems involving wetting III13

Solution

1. For small linear deformations dl/l  1, the relative change in


the area of individual alveoli satisfies da/a = 2dl/l. This implies
in turn that the change in the total area of the all alveoli in the
volume is also dA/A = 2dl/l. As a consequence, the energy cost
of dilating a lung is
dl
dA = A2 (III20)
l
For that deformation to be possible, the change in surface energy
has to be equal to the mechanical work of the forces F exerted on
the faces of the cube, namely

dW = 3Fdl (III21)

This leads to the following simple relation

F 2 A
= 3 (III22)
l2 3 l

where A/l 3 is noting but what we have called AV . Note also that
F/l 2 is the tensile stress, equivalent to a negative pressure.

2. All is left to do is to estimate the surface area AV of a lung. If the


number of alveoli per unit volume of the lung, nV , were known,
you would calculate the area as

AV = nV 4R2 (III23)

The number of alveoli can be determined from the porosity e via

4
e = nV R3 (III24)
3
This results in an area per unit volume

3e
AV = (III25)
R
Note that you may have used another
With the values R = 50 m and e = 0.5, one finds AV ' 3 104 model for the lung, for example a cubic
m1 . If the liquid was pure water, with surface tension = 75 packing of alveoli. In that case you
would have a surface 4R2 in a volume
mN/m, the capillary stress would be of (2R)3 . That would be AV = /(2R).
Compare that with Eq. III25 with
F e = 0.5.
= 1500 Pa (III26)
l3
That pressure is equivalent to about 1.5 kg exerted on a surface of
10 10 cm2 . Imagine a newborns lung having to exert that at each
breath!
III14 c.j. gommes, chim0698

Thinking further

1. The purpose of the alveoli is to enable mass transfer between the


air and blood. What is the total area of the alveoli in an adults
lung?

2. Estimate the order of magnitude of the capillary stress for a


nanoporous material with, say AV ' 300 m2 /cm3 .

3. In the analysis here above, we have neglected the mechanical


stiffness of the material. How can that be taken into account?
solved problems involving wetting III15

E. The mechanical equilibrium of a meniscus

The present problem is motivated by the double-sided floater case


presented in Fig. II18, the mechanical equilibrium of which may
seem paradoxical.
We shall consider the simpler case of just a wall wetted by a liquid,
as in Fig. III9, and ask the question

What is the total force exerted by the liquid on the wall? And R(z)
how does it depend on the various surface energies involved? h
z
There is a straightforward way to estimate that force, which does
require any calculation. Try to find it. If you do find it, use the result
to calculate the height h of the triple line above the surface of the
liquid. Figure III9: Shape of the meniscus
On the other hand, if you are running out of imagination, you may close to a flat vertical wall.
also go through the following steps

1. As a starting point in any analysis, it is often advisable to do a


dimensional analysis. Use Buckinghams theorem to find the
general expression for the height h of the meniscus as a function
of the contact angle , the surface tension , the density , and the
acceleration of gravity g.

2. Expressing that the pressure in the liquid is hydrostatic, calculate


the local radius of curvature of the meniscus R(z) as a function of
the height above the free surface.

3. The surface can be parameterized as z() where is the local


slope of the surface, as shown in Fig. III9. The angle increases
from 0 far from the wall to /2 on the wall. Assume for now
that /2. Use the answer to point 2 to find the differential
equation that governs dz/d and solve it. What is the height h that
the liquid climbs on the wall? Compare your answer to that of
point 1.

4. What is the total force that the liquid exerts on the wall?
III16 c.j. gommes, chim0698

Solution

The straightforward way consists in noting that the force ex-


erted on the wall by the liquid is equal to the force exerted on the
liquid by the wall. Because any vertical slab of liquid has to be equili-
brated in its own right. Consider then a thick slab of liquid extending
from the wall to far from it, where the free surface is practically hori-
zontal. The force pulling that slab to the right is simply . Therefore,
the horizontal force exerted by the liquid on the wall is also .
You can use that result to find the height h by expressing explicitly
the equilibrium of horizontal forces on the liquid slab as
Z h
sin( ) = + gz dz (III27)
0

In the left-hand side are the force pulling towards the wall, and in In this analysis, we have assumed that
the right-hand side the forces pulling/pushing away from the wall. the ambient pressure is P0 = 0, but
this does not matter. If you keep P0 in
This leads immediately to the calculation, you will find that its
q effect cancels out in the end. With some
experience, you should progressively
h = lc 2(1 sin( )) (III28) feel comfortable with this.
p
where lc = /(g) is the capillary length.

As for the unimaginative way, here it is.

1. There are four dimensional variables h, , g, , and a dimension-


less variable . The dimensional variables are based in three di-
mensions: length, time, and mass. You can therefore only form one
dimensionless variable (4 3 = 1). Let 0 be that variable. The
general relation that you look for can only be of the type

0 = f ( ) (III29)

where f (.) is can a priori be any function.


As for 0 , any suitable combination would do. Because h is what
you are interested in, and you know that the capillary length
p
lc = /(g) has the dimension of a length, it is natural to
chose 0 = h/lc . The mathematical expression for the height
can therefore only be of the form

h = lc f ( ) (III30)

where f ( ) is a yet unknown function. It is important that you un-


derstand that, whatever the final answer you find for this problem,
it has to be compatible with Eq. III30.
solved problems involving wetting III17

2. The pressure in the liquid obeys

P(z) = P0 gz (III31)

and the pressure in the gas phase is simply P0 (it is much lighter
than the liquid so the vertical pressure gradient in it can be ne-
glected). The pressure discontinuity at the interface is accounted
for by Laplaces law, which leads to

R(z) = (III32)
gz
Just a check: R for z 0, as it should.

3. Draw an infinitesimal part of the surface with height dz and hy-


pothenuse dl (along the surface). The length of the hypothenuse is
related to the local radius of curvature by Rd = dl, and we also
have the relation dz = dl sin(). The relation between dz and d is
therefore
dz = R sin()d (III33)
Combining this with Eq. III32 leads to the following differential
equation
dz sin()
= (III34)
d gz
the solution of which is
q
z() = lc 2(1 cos()) (III35)

where we chosen the integration constant in such a way that z( =


0) = 0. The final answer is
q
h = lc 2(1 sin( )) (III36)

which is indeed compatible with Eq. III30.

4. There are two component to the horizontal force exerted on the


wall by the liquid: the direct term, equal to sin( ) per unit length
of the triple line, and the pressure term. Because the ambient
pressure is irrelevant, we may set it equal to 0, in which case the
latter contribution is equal to
Z h
1
(0 gz)dz = + gh2 (III37)
0 2
Pay attention to the signs: this is a negative pressure on the right-
side on the wall, so the net effect is a positive force in the right
direction. The total force is finally
1
F= gh2 + sin( ) (III38)
2
Using Eq. III36, one finds F = independently of the contact
angle . The fact that the total force is indepen-
dent of the contact angle solves the
apparent paradox raised by Fig. II18.
III18 c.j. gommes, chim0698

Further thinking

1. Look back to our analysis of Wilhelmys plate (Fig. I4) with the
present problem in mind.
IV0 c.j. gommes, chim0698

As a general conclusion for these few problems I would


like to insist again that the skill that you should develop is to find
you own way to answer the questions in boldface characters in each
problem. Answering the subquestions is not what is expected of you,
this is just useful for learning.
Reflecting on the general way in which every problem here was
solved, it seems that Plyas general steps of problem-solving (see
Fig. III10) can be followed. The steps are
1. Understand the problem. That seems obvious, but this means
here understanding the physics of the problem. In the case of
problem D on capillary stresses, that step was realising that any
macroscopic change of volume of the lung is accompanied by
a change in the surface area of liquid free surface, and that this
change is surface area required some energy;

2. Devise a plan. In that case, the plan is (i) calculate the change in
surface area resulting from a macroscopic change of volume, and
(ii) equate the resulting change of energy to the mechanical work
of the macrosopic forces; Note that step 3 (carry out the plan)
is the only one that you do if you
3. Carry out the plan. That step is self-explanatory. This is the only try to apply formulae in a rote way.
one where some mathematics is involved; Most of the time, that step is the least
interesting one, but for some reason it
4. Look back. Have you answered the question? Does the result seems to be the one on students spend
make sense? In the case of the lung, estimating the total area of the more time.

lung, and relating the pressure to a force is a way to look back.


The solution of each problem is ended with a section I have en-
titled "Further thinking". Some of the questions raised there are
amenable to a Plya-like analysis. See if you can answer those ques-
tions.

Further reading

Many interesting applications of fluids involve some type of capil-


lary instability, which is the general term for Plateau-Rayleigh-like
type of phenomenon, by which a system evolves under the influ-
ence of capillary forces. There is collection of solved problems on
capillary instabilities in the book Hydrodynamique physique: Prob-
lmes rsolus avec rappels de cours by Marc Fermigier, Dunod (1999).

The hungarian mathematician George Polya published in 1945 a


book entitled How To Solve It which sold over one million copies,
in more than 15 languages. The topic of the book is to propose
Figure III10: George Plya (1887-1985).
general principles for problem solving. If you get a chance, it is
Picture from Wikipedia.
interesting to read. For a glimpse of what is in the book, you can
look here.
Intermolecular forces

The itsy bitsy spider climbed up the waterspout.


Down came the rain and washed the spider out.
Out came the sun and dried up all the rain
And the itsy bitsy spider climbed up the spout again.

Folk song, ca 1910

Intermolecular forces are forces exerted between molecules, as


opposed to intramolecular forces which exert within individual
molecules. All the forces and energies that we have mentioned when
discussing surface energies and surface tension in terms of missing
bonds are intermolecular forces. There are entire books written about Figure IV1: Johannes Diderik van
intermolecular forces. The present chapter is just a very basic intro- der Waals (1837-1923) . Picture from
Wikipedia.
duction that should provide you with a minimal working knowledge.
The intermolecular forces are often referred to as van der Waals
forces, after the Dutch chemist who hypothesised them to explain the
properties of non-ideal gases though his celebrated equation of state
 
a
P + 2 (Vm b) = k B T (IV1)
Vm
where the constants a and b account for the interaction between
molecules and their finite volume, respectively. To make a long story
short, when vapours are compressed close to their dew point, the
pressure is lower than predicted by the ideal gas law, which hints at
attractive forces between molecules.
Although intermolecular forces were discovered in the context
of non-ideal gases, their effects extend in many different areas. The
epigraph of the present chapter is a reminder that van der Waals
forces are strong enough to enable insects to walk on vertical sur-
faces, apparently defying the laws of gravity. In the case of geckos,
the contact of the animals feet with the surface is maximised thanks
to micrometer-sized setae.
The intermolecular forces are electromagnetic in nature. If molecules
carry an electric charge, they are expected to attract or repel each Figure IV2: The feet of the gecko are
covered with innumerable setae that
other in agreement with Coulombs law. Less intuitively, even elec- maximise the contact area with the
trically neutral molecules influence each other; unlike Coulombic underlying surface. This the animal
to climb on vertical walls, despite
his relatively large size. Picture from
Wikipedia.
IV2 c.j. gommes, chim0698

interactions, however, the forces between neutral molecules are al-


ways attractive, as we now discuss.
Consider the case of two polar molecules, carrying an electrical
dipole 1 and 2 . If the molecules are free to rotate, they will prefer-
entially orient themselves in such a way that their parts with opposite
charges will face each other. This will result in an average attractive
energy given by
21 22 C
wK = = K6 (IV2)
3(4e0 )2 k B Td6 d
where e0 is the vacuum electrical permittivity, k B is Boltzmanns
Figure IV3: Willem Hendrik Keesom
constant, T is the temperature, and d is the distance between the (1876 -1956). Picture from Wikipedia.
two molecules. The interaction energy depends on the temperature Keesom had been a student of van der
because the lower the temperature is, the more likely the two dipoles Waals and of Kamerlingh-Onnes. Does
the name Kamerlingh-Onnes ring a bell
will be aligned. The interaction described by Eq. IV2 is called the to you?
Keesom interaction, after the Dutch physicist who first calculated
them in 1921.
Another case is the situation where a molecule carrying dipole
interacts with a non-polar, yet polarisable molecule with polaris-
ability . If the molecule is free to rotate, the interaction energy is

2 C
wD = = D (IV3)
(4e0 )2 d6 d6
which bears the name of Peter Debye. Note that here too, the depen-
dence in the distance is proportional to d6 .
Yet another situation is that of two neutral molecules bearing no
electrical dipole whatever. A handwaving explanation for the at-
traction between two such molecules is the following. Electrons in
each molecule move around very rapidly with a frequency compara-
ble to the ionisation frequency .8 As a consequence, the molecules Figure IV4: Peter Debye (1884 -1966).
are non-polar only on average: at every moment they carry a small Picture from Wikipedia.
8
Ionisation frequencies are of the
dipole that fluctuates very rapidly. If the fluctuations in each molecules
order of 1015 1016 Hz. The ionisation
were independent of one another, this would not result in any force process by which an electron is stripped
between them. The force would be now attractive now repulsive and out of a molecule a photon, i.e. by an
oscillating electric field, can be thought
the two would cancel out exactly. However, when the two molecules of as a resonance phenomenon. The
are brought close together, the fluctuations of the dipoles in each one frequency of the field has to match
the natural frequency of the electron.
of them become anti-correlated: if one points, say up the other will
be more likely to point down. This results in a net attraction between
the two molecules. Imagine compressing gaseous nitrogen
The quantitative analysis of the attraction between two non-polar at a constant temperature. When you
reach a given volume the gas will
molecules requires quantum-mechanical calculations. The calcula- start to condense, which means that
tions were done for the first time by Fritz London in 1921, and the molecules have come sufficiently close
to each other to attract each other.
result can be written as Have you ever thought about why two
3 1 2 h1 2 C nitrogen molecules attract each other at
wL = = 6L (IV4) all?
2 (4e0 )2 d6 (1 + 2 ) d
intermolecular forces IV3

where 1 and 2 are the polarisabilities of the molecules, and 1 and


2 are their ionisation frequencies. Once again, the distance depen-
dency is d6 . This contribution to van der Waals forces is called the
London force, and sometimes the dispersive force.
The three contributions - Keesom, Debye, and London - are ad-
ditive. Imagine approaching two polar molecules A and B. The
molecules will naturally undergo the Keesom attraction. However,
the molecules having a permanent dipole does not prevent them
from being polarisable too. This means that if the molecules were
put in an electrical field E, their dipoles would have a contribution
There is a factor 2 in Eq. IV5 in front
proportional to E in addition to their permanent dipoles . The of the Debye contribution. This is not a
molecules will therefore undergo the Debye and London attraction typo. Think about it.
too. The resulting energy is therefore
CK + 2 CD + CL Ctot
w= = 6 (IV5)
d6 d
where we have introduced the obvious notation Ctot .
The values of the various contributions to the van der Waals forces
are given in tab. IV1 for a few common molecules. In the particular
case of non-polar molecules such as He or N2 the Keesom and Debye
contributions are absent, and one is left only with the London force.
In the table, the value of the constant Ctot is compared to an experi-
mental value derived from the van der Waals coefficients a and b via

9ab
Cexp = (IV6)
4
which results from assuming that the molecules behave like hard
spheres at short distances. Globally the agreement between the elec- Figure IV5: Fritz London (1900 - 1954).
Picture from Wikipedia.
tromagnetic calculation of the actual intermolecular forces is fair. It
particularly interesting to compare the various contributions to the
intermolecular forces in the case of polar molecules, such as HCl. The
largest contribution by far is generally the dispersive London force!
A notable exception is water, for which the Keesom contribution is
dominant.

/(4e0 ) h CK CD CL Ctot Cexp Table IV1: The values of the Keesom,


Debye, and London contributions to
(D) (1030 m3 ) (eV) van der Waals forces between a few
He 0 0.2 24.6 0 0 1.2 1.2 0.86 usual molecules, in units of 1079 Jm6 .
CH4 0 2.59 12.5 0 0 101.1 101.1 103.3 Adapted from Butt, Graf and Kappa,
Physics and Chemistry of Interfaces,
HCl 1.04 2.7 12.8 9.5 5.8 111.7 127.0 156.8 Wiley: 2005.
CH3 OH 1.69 3.2 10.9 66.2 18.3 133.5 217.9 651.0
H2 O 1.85 1.46 12.6 95.8 10.0 32.3 138.2 176.2
N2 0 1.74 15.6 0 0 56.7 56.7 55.3
Equation IV5 describes the attraction between two individual
molecules. To see how this converts to attraction between two macro-
IV4 c.j. gommes, chim0698

scopic bodies, say A and B, one has to sum over all pairs of molecules
with one in A and the other in B. Mathematically, this means that we
have to calculate a six-dimensional integral. Let us write the Eq. IV5
as w = C AB /d6 where C AB is specific to the types of molecules in A
and B. The macroscopic interaction would be
C AB A dVA B dVB
Z Z
WAB = (IV7)
VA VB [( x A x B )2 + ( y A y B )2 + ( z A z B )2 ]3
where A and B are the densities of media A and B expressed as a
number of molecules per unit volume, in such a way that A dVA and
B dVB are the number of molecules in the elementary volumes.

a) b) Figure IV6: a) Van der Waals interac-


D tions between two semi-infinite bodies
y x d
r A and B separated by a distance D,
z z
z
and b) interaction between a single
y D
x molecules A at a distance D from a
A B B semi-infinite body B.

We shall calculate explicitly the integral in the simple case of two


semi-infinite media separated by a spacing D, as sketched in Fig.
IV6. Let us calculate that integral in two steps, first by calculat-
ing the interaction between a single A-type molecule and a semi-
infinite medium of type B, and then by integrating that energy over
all molecules in medium A. The first step is equivalent to calculating
a three-dimensional integral on volume B, which is best calculated in
cylindrical coordinates as
Z Z
C AB B
WA | B = dz 2rdr 3
(IV8)
0 0 [(z + D )2 + r2 ]
where D is the distance between the A-like molecule and the semi-
infinite medium B. Note that we have used the notation WA| B be-
cause this is not the same as WAB . This leads to the following expres-
To get easily from Eq. IV8 to Eq. IV9
sion make a change of variable = r2 , and
C AB B
WA | B = (IV9) integrate first with respect to then to
6 D3 z.
Note that the very strong d6 dependence becomes here a weaker
d3 dependence.
To obtain the van der Waals energy between two semi-infinite me-
dia, Eq. IV9 has to be integrated over all the molecules in the semi-
infinite volume A. Of course, in the strict case of two (semi-)infinite
media the interaction energy is infinite, unless it is expressed per unit
area of the two bodies. This leads to the following expression
Z
A dz
WAB = B C AB (IV10)
6 0 [ D + z ]3
intermolecular forces IV5

Two flat surfaces Two spheres Two parallel cylinders Two crossed cylinders

R1
D
R1 R1
R2 R2
D R2

A A R1 R2
q
R1 R2
W = 12D 6D A A
6D R1 R2
2 R1 + R2 12 2D3/2 R1 + R2

Figure IV8: Van der Waals interaction


energies between bodies of various
and therefore to shapes. In all cases A is the Hamaker
A AB constant, and D is the spacing.
WAB = (IV11)
12D2
where A AB is the so-called Hamaker constant, defined as

A AB = 2 C AB a B (IV12)

The Hamaker constant characterises the two interacting media. Note that D does not have the same
meaning in Eq. IV9 and in Eq. IV
The expression of WAB in Eq. IV7 is quite general; you can apply
10. In the former equation, it is the
it in principle to calculate the van der Waals interactions between distance between a molecule of A and
bodies of any size and shape. Depending on the shapes of the two the surface of B; in the latter it is the
distance between the surfaces of A and
bodies, the calculation can be cumbersome, but the calculations are B.
conceptually as simple as those leading to Eq. IV11. A few expres-
sions of the energies are gathered in Fig. IV8 for interactions be-
tween spheres and cylinders. Note that these are expression for the
energy. The corresponding forces are calculated as

dW
F= (IV13)
dD
We shall come back to these expressions later.
Before we proceed to see what Eq. IV11 can be good for, it might
be useful to elaborate a bit on Hamakers constant. You may have
noticed that the relevant constant in Eq. IV11 characterises the inter- Figure IV7: Hugo Christian Hamaker
(1905-1993). Picture from Langmuir 7
action of two different media A and B, while we have only discussed (1991) 209. In that same paper the au-
the forces between identical molecules so far in Eqs. IV2, IV3 and thors mention the following anecdote.
Hamaker left the field of colloids in
IV4. The fact that the dispersion forces is the largest of the three the forties to focus on the development
contributions for non-polar (or weakly polar) molecules simplifies the of statistical methods for the Phillips
analysis. From Eq. IV4, the following approximation is seen to hold company, and he stopped following
the developments in that field. In 1965
p one of his sons who was studying soil
A AB ' A B A B ' A A A B (IV14) science asked him: "Dad, there is some-
thing called a Hamaker constant. Is it
In other words, all you need to know to estimate the interaction named after some relative of ours?"
between two bodies via van der Waals forces is the Hamaker constant
of individual bodies, and to compose then according to Eq. IV14.
Eq. IV14 is sometimes referred to as the Berthelot composition rule.
IV6 c.j. gommes, chim0698

A (1020 J) (mJ m2 )
from Eq. IV19
Liquid He 0.057
PTFE (teflon) 3.8 18.5
Pentane 3.8-4 Table IV2: Hamaker constants of a
few materials (both liquid and solid),
Ethanol 4.2 20.5
sorted in increasing order, and inferred
Acetone 4.1 20 value of the surface energy. Compare
Hexane 4.32 those values to those from Tab. I1 on
page I3.
Octane 4.5-5.02
Cyclohexane 4.8 - 5.2
Decane 5.45
Water 3.7 - 5.5 18-27
Toluene 5.4
Formamide 6.1 30
Fused Quartz 6.3
Silica (SiO2 ) 6.5 24-29
Diiodomethane 7.2 - 7.8
Polyvinyl chloride (PVC) 7.5 7.8 38.0
Quartz 7.93
Polyethylene (PE) 10
Mica 10-11
Aluminium 15
ZnS 15-17 73 - 83
Iron oxide (Fe3 O4 ) 21
Silicon 25.5
Zirconia (ZrO2 ) 27
Diamond 28.4
Copper 28.4
Germanium 30
Silver (Ag) 40
Rutile (TiO2 ) 43
Gold 45
Graphite 47
intermolecular forces IV7

The order of magnitude of Hamaker constant in air is 1019 J. It


is interesting to use that value to see how large van der Waals forces
can be. For that purpose, consider the force between two solids with
It is a better strategy for the gecko to
a contact area of 1 cm2 . The force per unit area is calculated from the minimise the distance between his feet
derivative of Eq.IV11, which leads to and the surface than to maximise the
contact area, because the dependence is
A D 3 .
F= (IV15)
6D3
The typical distance D between solids in contact is D ' 0.2 nm.
With these values, the van der Waals force across a surface of 1cm2 is
found to be of the order of 7 104 N. This corresponds to a weight of 7
tons! This means that the contact area between a geckos feet (see Fig.
IV2) and the surface on which it is walking does not have to be large
to stand the geckos weight.
Let us make another estimation of van der Waals forces, this time
from the nanometer world. In many practical situations, one has to
deal with metal nanoparticles dispersed on a support (see e.g. Fig.
05 on page 0-4). A legitimate question is how strong is the force
that keeps the particle on the surface. This can be calculated from
the energy given in Fig. IV8 for two interacting spheres, in the limit
where one sphere becomes infinitely larger. In that case the energy
becomes
AR
W= (IV16)
6D
and the force is
AR
F= (IV17)
6D2
In the case of metal on silica, the Hamaker constant can be calculated
Figure IV9: A metal nanoparticle, say,
more accurately than in the case of the gecko. Using the geometrical 4 nm across sits on a silica surface: how
mean as in Eq. IV14 and the values from Tab. IV2, one finds strong is the force keeping the particle
on the surface ?

Ametal/SiO2 ' 6.5 35 = 15 1020 J (IV18)

With a radius R = 2 nm, and a distance D = 0.2 nm, the force is


found to be F = 1.25 nN.
The only way to make sense of a force of the order of 109 N is
to compare it to another force. For example, one may estimate the
weight of the 4 nm nanoparticle. Assuming a density of = 20
g/cm3 , the weight of the nanoparticle is found to be of the order of
1020 N. In other words, the van der Waals force is about 1011 times
stronger than that. Another way to get a feeling for this would be to
consider that you would need to centrifuge with an acceleration of a
hundred billion gs if you wanted to detach the nanoparticle from the
support.
Let us push this type calculations to still smaller sizes. If you think Figure IV10: Marcelin Berthelot (1827-
1907). Picture from Wikipedia.
of it, the van der Waals forces are responsible for what we loosely
IV8 c.j. gommes, chim0698

referred to as the missing bonds when discussing the physical origin


of surface tension. There should therefore be a relation between the
surface tension/energy A of a given substance and its Hamaker
constant A A . This is illustrated in Fig. IV11. Consider for now the
case where A and B are identical media, say, the same liquid. In B
B
that case, the energy it takes to cut the liquid into two halves is 2 A
per unit area, because two new surfaces are created. On the other A
hand, this energy can also be calculated as the van der Waals energy A
between the two halves when they were in tight contact. Assuming
an initial spacing D0 of molecular dimensions, and using Eq. IV11 Figure IV11: The energy cost of mov-
ing a liquid B away from a solid A,
the relation we find is therefore
A + B AB , is the van der Waals in-
AA teraction between the two macroscopic
2 A = (IV19) liquid bodies.
12D02
The value of D0 that is universally accepted for most liquids and
solids alike for this type of calculation is D0 ' 1.65 . Some values
of Hamaker s constants are gathered in Tab. IV2, together with
the values of inferred from Eq. IV19. Comparing those values
with Tab. I1, one sees that the calculations are reasonably accurate
in the case of weakly-polar substances. This is, however, not the
That a single value can be used for D0
case for polar molecules and for metals, for which dispersive forces
for a broad variety of substances is a
responsible for only a fraction of the cohesion. For example, it the bit of a mystery. You can read some
case of water disperse forces account for only about 25 % of all the explanations in the excellent book of J.
Israelachvili on Intermolecular Forces,
intermolecular forces. in section 13.13.
The reasoning leading to Eq. IV19 based on Fig. IV11 can be
repeated in the case of two distinct media. In that case, equating the
change in surface energy to the energy that is needed to overcome
the van der Waals attraction between the two media leads to
A AB
A + B AB = (IV20)
12D02

Expressing that A AB ' A A A B leads to the following relation

AB ' A + B 2 A B (IV21)
You can find in the literature expres-
sions similar to Eq. IV21 with a semi-
which can be used to estimate interface energies, when only the
empirical factor in front of A B to
surface energies are known. Of course you do not expect this relation account for the polarity of the sub-
to be very accurate in the case of polar substances, but if you dont stance. That is of course numerically
more accurate... but also less general.
know better, this relation can be a good starting point.
So far, we have considered only the interaction between two bodies
in vacuum. In that case, the van der Waals forces are always attrac-
tive. This is not necessarily the case if the two interacting bodies
are immersed in a liquid. To make things more practical, consider
two bodies A and B and a liquid L. Imagine the situation where
the Hamaker constant between A and B would be weaker than be-
tween A and L. In that case, you may expect that the liquid will try
intermolecular forces IV9

to squeeze between A and B, thereby leading to an effective repul-


sion. This situation is described using the concept of disjoining pres-
sure, introduced by Boris Derjaguin. In the case of two flat plates A
and B facing each other at distance D, the disjoining pressure ( D )
is the pressure that has to be exerted on the plates to prevent a spe-
cific liquid to squeeze in the interstice and increase D.
To determine the energy related to the insertion of a liquid layer
of thickness D between two media A and B, consider the various
steps sketched in Fig. IV14, from 1 to 4. We refer to that energy as
Wwetting . If Wwetting is positive, the process will not spontaneously
take place. On the contrary, if it is negative a positive pressure will Figure IV12: Boris Vladimirovich
need to be exerted on A and B to prevent the liquid from pushing Derjaguin (1902-1994). Picture from
them apart. Colloid Journal 64 (2002) 648.

Before we proceed with that particular calculation, it is necessary


to slightly generalise Eq. IV11, in order to calculate the interaction
energy between a semi-infinite medium B and a slab of A of finite
thickness extending, say, from distances z0 to z1 from the surface of
B. Starting from Eq. IV9 the calculation is straightforward:
Z z
1 C
AB B
W= A dz (IV22)
z0 6 z3 z1
B
z0
which leads to !
A 1 1
W = AB 2
2 (IV23)
12 z0 z1 A
Note that this energy reduces to Eq. IV11 for z1 , as it should.
We can now turn to Fig. IV14. The change in energy from 1 to 2 Figure IV13: The interaction energy
between a slab of B extending from z0
corresponds to increasing the distance between A and B from D0 to
to z1 above a semi-infinite medium A is
D, i.e. ! calculated as Eq. IV23.
A AB 1 1
W12 = 2 (IV24)
12 D02 D

according to Eq. IV11. Note that this energy is positive. The energy
change from 2 to 3 can be calculated using Eq. IV23 with z0 = D0
and z1 = D. The result is
!
AL 1 1
W23 = 2 2
12 D02 D
! !
A LA 1 1 A LB 1 1
2 2 (IV25)
12 D02 D 12 D02 D

The positive contribution on the first line is the energy it takes to


remove the slab of liquid aways from the two semi-infinite L-like
media, one on the top and the other on the bottom. The two negative
contributions result from the slab being put in contact with the semi-
infinite A and B. The last contribution is the putting back together
IV10 c.j. gommes, chim0698

A L A L A L A
L
1) 2) 3) 4)
B
B B B

of the two semi-infinite L-like media, from 3 to 4. This results in a Figure IV14: The insertion of a liquid
change in energy layer L between two solids A and B has
an energy associated with it, which can
! be calculated via the following steps.
AL 1 1
W34 = 2 (IV26)
12 D02 D

which is negative as it should.


Putting all contributions together results in the following expres-
sion for Wwetting = W12 + W23 + W34 :
!
A ALB 1 1
Wwetting = 2 (IV27)
12 D02 D

where the effective Hamaker constant is given by

A ALB = A AB + A L A AL A BL (IV28)

The subscript highlights that it corresponds to the interaction of


media A and B, with medium L in between. Whether the liquid
layer L will tend to thicken and move the two media A and B apart
depends on the sign of the effective Hamaker constant. If A ALB is
negative, the film tends to increase in thickness.
The disjoining pressure is defined as the pressure that has to be
exerted on the film to prevent it from thickening further. It is there-
fore positive when A ALB is negative. It is calculated as the derivative
of the wetting energy

dW A
( D ) = = ALB3 (IV29)
dD 6D
If A ALB 0 then the disjoining pressure is negative, which means
that a tension (negative pressure) has to be exerted to prevent the
film for becoming thinner.
It is interesting to relate this analysis to the macroscopic approach
of Chapter II via the definition of a spreading parameter S as in Eq.
II2. Clearly the two approaches should lead to the same result in
the macroscopic limit, i.e. for large values of D. In other words, one
should have
A ALB
lim Wwetting = = AL + BL AB (IV30)
D 12D02
intermolecular forces IV11

I will let you find out on your own9 that this is indeed the case. 9
Hint: use Eqs. IV19 and IV20.
Therefore, the present analysis enables us to find the correction to
the macroscopic wetting energy AL + BL AB in the case where
the layer is not macroscopically thick. The result can be written as
"  #
D0 2

W ( D ) = ( AL + BL AB ) 1 (IV31)
D

W /( AL + B L AB)
1

The value obtained with the generally accepted values D0 = 1.65 is 0.8
plotted in Fig. IV15. The figure shows that the macroscopic values 0.6
of the energies are reached for thicknesses of about 2 nm. For objects
0.4
smaller than about that size, you expect significant deviations from
the predictions of the macroscopic laws of wetting. 0.2

0
0 5 10 15 20
Let us now discuss the sign of A A LB and of the disjoining pres-
D (
A)
sure in a few characteristic situations. Using the geometric-mean rule
Figure IV15: Normalized wetting
Eq. IV14 for composing Hamaker constants, enables one to rewrite energy, as a function of the thickness D
A e f f in the following way as a function of the Hamaker constants of of the liquid layer. This was plotted by
A, B and L: assuming D0 = 1.65 .
p p  p p 
A A LB = AA AL AB AL (IV32)

Based on that general equation, a few interesting situations can be


discussed.
In the case where media A and B are identical, the effective
Hamaker constant is
p p 2
A ALA = AA AL (IV33)

which can only be positive. Two identical media can only attract each
other.
When one of the two media, say, B is vacuum, the situation is The rule of thumb for spreading is
simply that of a liquid spreading over a surface. The question is that a liquid spreads on any solid
more polarisable than itself. This is the
about whether the liquid will spontaneously thicken or not. In that
reason why oil spreads on metals, and
case, the effective Hamaker constant is obtained by setting A B = 0, water does not spread on most plastics.
i.e. If you look at the Hamakers constant
of liquid helium in Tab. IV2, you will
p p p 
A A LO = A L AA AL (IV34) also understand why liquid helium is a
universally wetting fluid.
According to this analysis, the liquid spreads on the solid A if the
Hamaker constants satisfy A L A A . If you recall that the ori-
gin of A is the London dispersive forces described by Eq. IV4, this
criterion is equivalent to stating that the liquid is electrically less
polarisable than the solid. The criterion for wettability based on
In the general situation, the effective Hamaker constant is positive Hamakers constant looks very much
like Zismans criterion based on a
except when A L has a value intermediate between A A and A B . In critical surface tension C discussed in
that case, it is energetically favourable for the liquid to enter in the in- Chapter II.
terstice between A and B, thereby exerting a pressure that pushes A
IV12 c.j. gommes, chim0698

Figure IV16: Direct van der Waals


force measurements using an Atomic
Force Microscope (AFM), published
in Measured long-range repulsive
Casimir-Lifshitz forces, Nature, 457
(2009) 170. The voltage is a measure
of the force. On top: attractive force
between a gold particle and a gold
substrate in bromobenzene, and bot-
tom: repulsive force between the same
particle and silica in the same solvent.

and B apart. This is illustrate on Fig. IV16. Using Atomic Force Mi-
croscopy (AFM) the authors of that study measured directly the force
between a gold nanoparticle and a support, in bromobenzene. When
the support is also made of gold, the force is attractive as expected
based on Eq. IV33. By contrast, when silica is used as a support
a repulsive force is observed. Silica and gold were not chosen ran-
domly for that experience. For a repulsive force to be observed, the
Hamaker constant of bromobenzene has to be intermediate between
silica and gold. From Tab. IV2, the Hamaker constants of silica and
gold are very different (ASiO2 ' 6.5 1020 J and A Au ' 45 1020 J),
which is a favourable condition to observe repulsion.
Figure IV17: Hendrick Casimir (1909-
The calculation of the effective Hamaker constant via Eqs. IV28 2000). Picture from Wikipedia.
or is interesting conceptually because it enables one to understand
physically the sign of A ALB , but is not necessarily practical because it
requires the three Hamaker constants A A , A B and A L to be known.
Another way has been developed by Lifshitz based on electromag-
netic calculations. The following formula is a good approximation for
A ALB

3he (n2A n2B )(n2L n2B )


A ALB = +
8 2(n2A + n2B n2L + n2B ) (n2A + n2B )1/2 + (n2L + n2B )1/2
 
)(
3k B T e A eB eL eB
+ (IV35)
4 e A + eB eL + eB
where k B is Boltzmanns constant, h is Plancks constant, e is the Figure IV18: Evgeny Mikhailovich
Lifshitz (1915-1985). Picture from
main electronic absorption frequency, and ei and ni are the static
Wikipedia.
dielectric constant and refractive index of phase i.
Remember that the dielectric constant e and the refractive index
n both characterise the electrical polarisability of a medium. There-
intermolecular forces IV13

fore, from a physical perspective, Eqs. IV35 and Eq. IV4 are not so
different after all.
A practically important situation where repulsive van der Waals
forces play a role is the freezing a water. It is well known that grow-
ing ice crystals are able to to push away solid material initially in
contact with water. This is notably striking in frost heave, by which
entire layers of soil are deformed by growing ice crystals (see Fig. IV
19). If you think of it, this is possible only if a repulsive force exists
between the growing ice and the suspended material. In absence of
Figure IV19: Stone rings is an example
any repulsive force, the growing ice crystal would simply engulf the of structures that can form further to
solid and grow around it rather than push it. frost heaving. Picture from Wikipedia.
To analyse phenomena like frost heave, the Hamaker constant
between ice and a few oxides across water are reported in Tab. IV3.
These values have been calculated using the general relation given
Eq. IV35, which is the reason for which the dielectric constants and
refractive indices are reported in the table. For the calculations, the
value e ' 3 1015 Hz was used for the ionisation frequency. The
values in the table show that ice attracts ice across water, as testified Table IV3: Van der Waals interaction
by a positive value of the Hamaker constant A. This was expected energies between ice and various
from our previous analysis based on Eq. IV33 because two identical oxides. Taken from G.W. Scherer,
Cement and Concrete Research, 30
media can only attract each other, and it is also confermed by the (2000), 673.
Lifshitz expression.
Substance e n A (1021 J)
More interestingly, the values in the table also show that ice repels Water 80 1.33 -
all the considered oxides (quartz , mica, and alumina) as testified by Ice 110 1.309 1
Quartz 3.8 1.448 -1.20
the negative values of A. The van der Waals repulsion between ice Mica 7.0 1.60 -1.211
and these oxides across water is the nanometer-scale phenomenon Alumina 11.6 1.75 -2.94
that is responsible for such spectacular effects as frost heave (see e.g.
Fig. IV19).
An interesting technical application of repulsive van der Waals
forces between a solidifying solvent and suspended particles is the
so-called freeze-casting methods for shaping ceramic and metallic
materials. This is illustrated in Fig. IV20. With this type of process,
one can chose appropriately the solvent to have a selective repulsion
of some suspended particles (not necessarily all of them), to design
porous materials with specific morphologies.
IV14 c.j. gommes, chim0698

Figure IV20: Examples of microstruc-


tures formed by freeze-casting, i.e. by
the repulsion of suspended material
by a solidifying solvent. The pores in
those figures were initially occupied
by the solvent crystals, which were
subsequently freeze-dried (sublimated).
Picture taken from S. Deville, Freeze-
Casting of Porous Ceramics: A Review
of Current Achievements and Issues,
Advanced Engineering Materials, 10
(2008) 155.

Further reading

The comprehensive reference for many topics covered in this chap-


ter is Jacob N. Israelachvili Intermolecular and Surface Forces, 2011,
of which you can find an online version here. To have full access,
however, you will have to connect from your university account.

The mechanism by which the gecko climbs on vertical surfaces is


fraught with beautiful physics, some related to wetting phenom-
ena and van der Waals forces, some to the mechanics of the setae
which are responsible for the close contact between the animals
feet and the surface. One of the first papers on the subject is by
K. Autumn et al. Evidence for van der Waals adhesion in gecko setae
Proc Nat Acad. Sci. USA 99 (2002) 12252. A more comprehensive You will find many things to read on
the Internet about geckos. I love the
study was published by the same authors under the title Properties,
title of this blog post How Geckos Stick
Principles, and Parameters of the Gecko Adhesive System in Biological on der Waals!
Adhesives (Smith and Callow Ed.), 2006, Springer, of which you
can find a preprint here.
Repulsive van der Waals forces play an important role in the dam-
aging of porous rocks by the crystallisation of salts in their pores.
You can have a look at this lecture by George Scherer on that topic,
which he gave on the occasion of his receiving the Onsager medal
in 2011.

Exercises

1. Use dimensional analysis to find the D-dependence of the van der


Waals interaction energy between two semi-infinite bodies.

2. Compare the van der Waals forces between two spheres with the
capillary force calculated on Page III9.

3. Define a dimensionless number to characterise the relative impor-


tance of van der Waals forces compared to gravity.
V0 c.j. gommes, chim0698

4. According to Eq. IV33, for very large values of the Hamaker


constant A L , the effective attraction of the A-like media is larger
than in vacuum. How is this possible physically?

5. Carbon nanotubes synthesised by the chemical vapour decompo-


sition of gaseous hydrocarbons often come in the form of bundles
comprising several thousand nanotubes arranged in a hexagonal
packing. Assuming that the tubes are about 10 nm in diameter and
a few microns in length, calculate the energy needed to disperse
the nanotubes (i.e. to destroy the bundle).
Gibbs-Thomson Effects

Definition - If to any homogeneous mass we suppose an


infinitesimal quantity of any substance to be added, the mass
remaining homogeneous and its entropy and volume
remaining unchanged, the increase of the energy of the mass
divided by the quantity of the substance added is the
potential for that substance in the mass considered.

J. W. Gibbs On the Equilibrium of Heterogeneous Substances,


Transactions of the Connecticut Academy of Arts and
Sciences 111 (1878) 108 .

This chapter is an incomplete draft, which I included here for


your convenience. Please rely on your own notes to study this
chapter. Figure V1: Josiah Willard Gibbs (1839-
1903) . Picture from Wikipedia.

In technical terms, the subject of the present chapter can be put


as that of the thermodynamical consequences of dispersion. In plain
english, this is equivalent to determining how the size of an object
influences its physicochemical properties such as its vapour pressure,
its solubility, its melting point, etc.
It so happens that the size-dependence is identical for phenomena
apparently as different as vaporisation, dissolution, and melting,
among others. That particular dependence, independently of the
specific phenomenon it is applied to, is known as the Gibbs-Thomson
relation.
A spectacular example of size-dependent property is provided
by the melting of nanometre-sized crystallites, which occurs at a
temperature significantly lower than the melting temperature of the
bulk material. This effect is illustrated in Fig. V2 in the case of gold
nanoparticles, which was already touched on in the epigraph of the
introduction section. For very large particles the melting temperature
approaches that of the bulk (Tmb ' 1330 C), and it decreases dramati-
See in particular how it is used in Fig.
V2.

V2 c.j. gommes, chim0698

cally for very small particles. For particle of 4 nm in diameter (similar


to that we considered in Fig. IV9) the melting temperature is as low
a 1000 K.

Figure V2: The melting temperature


of gold nanoparticles measured as a
function of their diameter. The figure is
taken from Phys. Rev. A 13 (1976) 2287.

The most obvious characteristic of nano-materials is their large


surface-to-volume ratios, which scale like 1/R (see the Introductory
Chapter). The contribution interfaces to the thermodynamic proper-
ties, compared to the volume, is therefore also expected to scale in
that way. Can this be the origin of the trends in Fig. V2? It could
be, if only because the divergent behaviour at small sizes is remi-
niscent of 1/R. In order to analyse this in detail, i.e. to eventually
have a quantitative relation describing Fig. V2, we will apply Polyas
method, namely: (i) Understand the physics; (ii) Devise a plan; (iii)
Carry out the plan; (iv) Look back.

To try solving a new problem, it is always convenient to


start solving a related but simpler problem. What other
situation do you know, in which the melting temperature changes?
The first example to pop in your mind might be the effect of salt on
ice, but this is a complicated situation involving two substances. In
Fig. V2 we just have pure gold. With a pure substance, the simplest
example that I have of a changing melting temperature is the sheer
effect of pressure. In general, the melting temperature of solids in-
creases when the pressure is increased. The case of gold is shown
in the phase diagram of Fig. V3. You most probably studied this in
an introductory thermodynamics course: the mathematical relation
relevant here is called the Clausius-Clapeyron equation, but do you
remember what it meant physically?

Here is a simple way to look at it. Melting consists in the pass- Figure V3: Phase diagram of gold,
ing of the substance from a crystalline state with low energy uC and from D.A. Young, Phase diagrams of
the elements, technical report from the
low entropy sC to a liquid state with higher energy u L and larger US Department of Energy (1975).
gibbs-thomson effects V3

entropy s L . The energy uC u L because the atoms in the crystal sit


in fixed positions that minimise the energy, while in the liquid they
constantly move around each other. On the other hand, because of
that motion there are more degrees of freedom in the liquid state to
store thermal energy than in the crystal state, i.e. s L sC . As a first
approximation, we may assume that the energies and entropies uC/L
and sC/L are independent of the temperature. With that particular
model in mind, the transition from C L is possible only if the
temperature T is sufficiently large, so that the heat taken up from the
environment Q = T (s L sC ) is enough to account for the change in
energy u L uC . In other words, the melting temperature is given by

0 u L uC
Tm = (V1)
s L sC

To be fully accurate, we would have to account for the fact that in


general the volume is larger in the liquid state v L than in the crys-
talline state vC . In this case, there is an addition energy cost to melt-
ing, that is the mechanical work exerted by the melting substance on
its environment, namely W = P(v L vs ). In that case, the heat taken
from the environment is needed for u L uC as well as for W. This The lower-case letters in the notations
uC/L , sC/L and vC/L refer to molar
eventually leads to quantities. Implicitly, we are making
here energy balances on one mole of
u L uC v vC
Tm = +P L (V2) substance.
s L sC s L sC

This simple physical picture is all is needed to understand the data in


Fig. V3.
In other word, the melting temperature increases with the pressure
because the larger the pressure, the more energy is required by the
melting-induced swelling of the particle. With that in mind, the data
in Fig. V2 suggests that the energy cost for melting a nanoparticle is
lower than u L uC . This would explain why the melting temperature
is lower, because less heat has to be taken from the environment.
Moreover, the smaller the particles are the lower that energy is. This
clearly hints at a surface energy contributions.
Before we proceed to explain the thermodynamic origin of Fig.
V2, it is useful to put this analysis in a more abstract form. In terms
of chemical potential, the condition for the equilibrium of the crystal
and liquid state is
C = L (V3)

It might not strike you, but Eq V3 is an equation describing the


shape of the transition region between the crystal and the liquid in
the phase diagram, because the chemical potentials C/P are func-
tions of P and T. Eq. V3 is not true for all values of P and T. It is
only true at equilibrium, i.e. when T = Tm ( P) where Tm ( P) is the
Eq. V3 is something you knew: at
an equilibrium phase transition, the
V4 c.j. gommes, chim0698 chemical potentials are the same in
the two phases. However, do you
remember where that comes from?

melting temperature at pressure P. In order to actually use that equa-


tion to practically determine how Tm depends on P, however, one has
to know exactly how C/P depends on P and T.

The equation that tells you how the chemical potential


depends on P and T is a cornerstone of thermodynamics.
It is called the Gibbs-Duhem equation. For a pure substance, it states

d = vdP sdT (V4)

where v and s are the molar volume and entropy, respectively. Imag-
ine we knew the chemical potential, say, of the liquid state L at a
given pressure P0 and temperature T0 . Using a Taylor development,
one could estimate the value of L for a value of P and T not very far
from P0 and T0 using the following Taylor development
   
L L
L ' L ( P0 , T0 ) + [ P P0 ] + [ T T0 ]
P T T P
 2   2 
1 L 1 L
+ 2
[ P P0 ] 2 + [ T T0 ] 2
2 P T 2 T 2 P
 2 
L
+ [ T T0 ] [ P P0 ] + (V5)
TP

Using that general development and limiting yourself to the first-


order accuracy, the relation can be written as follows in the particular
case of the crystal and liquids C and L

C/ L ( P, T ) ' C/ L ( P0 , T0 ) + v C/ L [ P P0 ] s C/ L [ T T0 ] (V6)

where we have used the Gibbs-Duhem relation Eq. V4 to express


( C/ L /P ) T = v C/ L and ( C/ L /T ) P = s C/ L . Putting that
particular dependence in Eq. V3 leads to

C ( P0 , T0 ) + v C [ P P0 ] s C [ T T0 ]
= L ( P0 , T0 ) + v L [ P P0 ] s L [ T T0 ] (V7)

Of course, you can chose in here any value of P0 and T0 that would
make your life easier. A convenient choice here is to take P0 = 0 and
for T0 the melting temperature at that particular pressure Tm 0 . With

that particular choice, one has C ( P0 , T0 ) = L ( P0 , T0 ) so that Eq.


V7 becomes significantly simpler. One finds

0 v L vC
Tm = Tm +P (V8)
s L sC
which is identical to Eq. V2, as it should.

We are now facing a typical situation: how do we generalise


this type analysis to determine the melting point of nanoparticles?
Figure V4: Pierre Duhem (1861-1916) .
Picture from Wikipedia.
gibbs-thomson effects V5

Does Eq. V3 still apply? There is no way you can answer that ques-
tion unless you know exactly where that equation comes from. Think
about this before proceeding with the studying/reading of this chap-
ter: where does it come from?
As a matter of fact, the equality of the chemical potentials of two
phases at equilibrium is not a first principle of thermodynamics. The
general law of thermodynamics is that the free energy of a system
is minimal, subject to whatever external constraint is imposed on it.
The particular form of the free energy that is relevant depends on
the variables that are used to characterise the system. For example,
we are currently interested in analysing the melting/freezing of
a substance at a given temperature T and pressure P. With these
variables, its is the Gibbs free energy G that is relevant.
The general differential form of Gibbs equation in the case of a
system comprising a crystalline phase C and a liquid phase L can be
written as follows
Figure V5: William Thomson, first
dG = VC dP S C dT + C d NC + C d A C Baron Kelvin (1824-1907). Picture from
+ VL dP S L dT + L d N L + L d A L (V9) Wikipedia.

where P is the pressure, T is the temperature, and VC/ L , S C/ L , C/ L


and NC/ L are the volume, entropy, chemical potential and number
of molecules in phases C and L, respectively. Compared to the Gibbs
equation that you encountered in introductory thermodynamics,
there are here two new contributions in the form of d A accounting
for the changes in surface energies. Note that the other terms in the
Gibbs equation, such as V dP and SdT are proportional to volumes,
so that the relative contribution of d A compared to the latter scale
like area-over-volume ratios. In other words, the d A contributions
are important only when the systems are small; when they reach
nanometer sizes they are crucial.
Figure V6: Joseph John (JJ) Thomson
The particular process that we are interested in now is the melt- (1856-1940). Picture from Wikipedia.
ing/freezing of nanoparticles, which takes place at constant tem-
perature and pressure. The only variable left are NC and N L and
A C and A L . Because any nanoparticle in the system is either frozen
or molten, the total number of atoms is a constant: d NC = d N L .
Moreover, if we assume that the particles are spherical, their area and
the number of atoms they contain are not independent quantities.
They are related by

dA 2v
dA = vd N = dN (V10)
dV R
where v is the molecular volume, and the second equality results
from assuming that the particles are spherical. With that in mind, the
equilibrium state of the system comprise molten and frozen nanopar-
V6 c.j. gommes, chim0698

ticles, expressed as dG = 0, becomes simply

2 2
L ( P, Tm ) + L v L = C ( P, Tm ) + C v C (V11)
R R
This equation generalises Eq. V3 to dispersed system, i.e. in which
surface energies are not negligibly small.
In order to convert Eq. V11 into a more usable equation relat-
ing the melting temperature to the particle size, all we need to do is
express the chemical potentials explicitly in terms of the tempera-
ture. This can be done using Eq. V6 with suitable values of P0 and
T0 . A convenient choice is to take P0 as the ambient pressure and
T0 = Tmb as the melting temperature of the bulk material at that par-
ticular pressure, i.e. such that C ( P0 , Tmb ) = L ( P0 , Tmb ) . With that
particular choice, Eq. V11 takes the following simple form

Tm 2 C v C L v L
= 1 (V12) The choice we made of the reference
Tmb R h m temperature and pressure T0 and
P0 has no consequence on Eq. V12.
where we have used the notation h m = Tmb ( s L s C ) for the en- It simply enabled us to express the
thalpy of melting of the substance. Eq. V12 is the Gibbs-Thomson thermodynamic quantities in terms
of the bulk melting temperature Tmb .
equation. It relates the melting temperature Tm of particles to their As an exercice, repeat the analysis
size. Note that the dependence is in 1/R, and that Tm converges to without making any explicit choice
for P0 and T0 . You will then obtain an
the bulk melting temperature Tmb for R . expression similar to Eq. V12 with a
With respect to the Polyas steps, we now reached the point where cryptic combination of C/L ( T0 , P0 ). If
we are supposed to look back, and see if we achieved what we were you express that for large particles, the
melting temperature should coincide
aiming at. The purpose of our analysis, which eventually lead to Eq. with Tmb , it will all simplify into Eq.
V12, was to describe the data in Fig. V2. Does it make sense? In V12.
many respects it does. However, we still have to be sure we under-
stand the sign of Tm Tmb . As a first approximation, consider the
simple case where vC ' v L . In that case, the numerator in the Gibbs-
Thomson equation is proportional to C L , which determines
whether the small particles melt at a higher or lower temperature.
To understand the sign of C L , I encourage you to go back to
Chapter 1 and be sure you are comfortable with Stefans interpre-
tation of surface energy (see Eq. I12) before you continue reading.
For the sake of a back-of-the-envelope calculation, consider again a
6-coordinated cubic arrangement of atoms, both in the crystalline
and in the liquid states. Because we assumed vC ' v L , we may also
assume that the spacing between atoms d is the same in both phases.
The order of magnitude of d is the Angstroem. In the crystalline (the
liquid) phases, each atom has a binding energy eC (e L ) with each of
its six neighbours, except at the surface where one bond is missing.
The surface energies of the crystalline and liquid phases are therefore
something like
e
C/L ' C/L (V13)
d2
gibbs-thomson effects V7

where d2 is the area occupied by an atom at the surface. On the other


hand, you know that it takes some energy to melt a solid. In the
context of the simple cubic model, the enthalpy of melting hm is
related to the binding energies via

hm = 6 (eC e L ) (V14)

The direct consequence of this analysis is that C can only be larger


than L . Accordingly, Eq. V12 does indeed predict that small parti-
cles melt at a lower temperature.
We can push the back-of-the-envelope calculation further and
estimate how small the particle ought to be for their melting point to
be significantly impacted. For that purpose, we may use Eqs. V13
and V14 to rewrite the Gibbs-Thomson relation in the form

Tm 1 d
b
' 1 (V15)
Tm 3 R

where we have used vC = v L ' d3 . Of course this equation is a


crude approximation, if only because of the ridiculous cubic model.
However, its physical insight is interesting: the melting point depres-
sion is significant when the size of the atoms is not negligibly small
compared to that of the particles.
There is an alternative way to understand the thermodynamics of
Gibbs-Thomson phenomena based on Laplaces pressure inside the
nanoparticles. As we saw in Chapter 1, Laplaces pressure applies to
liquids as well as to solids (see e.g. Fig. I12). For a nanoparticle with
diameter 4 nm and a surface energy of 1 J/m2 the Laplace pressure is
as large as 104 atmospheres. Instead of taking explicitly the interface
energy into account in the Gibbs relation (Eq. V9), you may alter-
natively take the Laplace pressure into account when calculating the
chemical potentials via Eq. V6. This results in additional terms of
the type
2
vC/L C/L (V16)
R
in the chemical potentials of the crystalline and liquid phases. If you
do that, of course you should not put the dA terms in the Gibbs
equation, so that the equilibrium condition is Eq. V3. Basically, in
order to account for surface energies you have two options:

Either, you write Gibbs equation explicitly with the dA terms as


in Eq. V9;

Or you use the usual thermodynamic relations, e.g. Eq. V3, but
you account for the change in chemical potentials resulting from
Laplaces pressure.
V8 c.j. gommes, chim0698

Sometimes the second option is more intuitive, but it is also less gen-
eral. For example, Exercice XX at the end of the chapter is devoted to
the calculation of melting point depression in lamellar crystals, which
have no curvature. In that case the Laplace pressure is irrelevant but
the experimental melting point is nevertheless found to depend on
the thickness of the lamellae.
The rest of the chapter is a generalisation of Gibbs-Thomson equa-
tions to

other phase equilibria than melting-freezing: solubility and liquid-


vapor (Kelvin equation);

phases confined inside porous materials (role of contact angle,


capillary condensation and evaporation);

I had no time to type this down; you will have to rely on your own
notes to study that. Next years students will be luckier than you,
perhaps.
VI0 c.j. gommes, chim0698

Further reading

Exercices

1. Use the chemical potential formalism to explain the phase dia-


gram in Fig. V3. In other words, start from Eq. V9 with no dA
term and derive Eq. V2.

2. Similarly, use the same type of basic energy balance approach as


in Eq. V2 to derive the Gibbs-Thomson relation Eq. V12.

3. In semi-crystalline polymers, crystalline regions have the shape


of lamellae with a thickness of a few nanometers surrounded by
amorphous (molten) regions. In that case, the observation is that
the melting temperature depends on the thickness of the lamellae.
Derive the expression for the melting point depression, assuming
that the crystals grow or melt laterally, while keeping a constant
thickness.

4. Use the Gibbs-Duhem equation to find a Taylor development of Figure V7: Example of lamellar struc-
similar to Eq. V6 but valid to the second order in P and T. tures in semi-crystalline polymers: in
that case the melting temperature de-
pends on the thickness of the lamellae.
Picture from Macromolecules 42 (2009)
2135.
Thermal fluctuations

If, in some cataclysm, all of scientific knowledge were to be


destroyed, and only one sentence passed on to the next
generation of creatures, what statement would contain the
most information in the fewest words? I believe it is the
atomic hypothesis (or the atomic fact, or whatever you wish
to call it) that all things are made of atoms: little particles that
move around in perpetual motion, attracting each other when
they are a little distance apart, but repelling on being
squeezed into one another. In that one sentence, you will see,
there is an enormous amount of information about the world,
if just a little imagination and thinking are applied.

Richard Feynman, Lectures on Physics

Figure VI1: Ludwig Boltzmann (1844-


1906). Picture from Wikipedia.
This chapter is an incomplete draft, which I included here for
your convenience. Please rely on your own notes to study this
chapter.

Very few scientific items, let alone equations, have the privilege of
becoming cultural symbols. Einsteins E = mc2 is definitely one such
item. The second equation on that list is probably Boltzmanns def-
inition of entropy S = k ln[W ], which can be read on his tombstone
in Vienna. To the present day, Boltzmanns definition of entropy is
the best link we have to relate the microscopic and the macroscopic
worlds, i.e. the motion of atoms and molecules to the property of
matter at our scale.
Richard Feynman necessarily had that equation somewhere in his
mind when he wrote that sentence in the epigraph. The goal of the
present chapter is to provide you with a crash course on a few topics
that are usually part of a broader course in statistical thermodynamics.
In other words, this chapter is about the type of imagination a thinking
that Feynman refers to.
VI2 c.j. gommes, chim0698

Microstates and the ergodic hypothesis

The distinction between microstate and macrostate is central to sta-


tistical mechanics. A macrostate is typically what is studied in ther-
modynamics. In this context, the state of a system is defined by its
macroscopic properties such as volume its V, its total number of
molecules N, the pressure, the chemical potential, etc. By contrast,
a microstate is a complete microscopic description of the system, at
the level of the molecules or atoms, or of whatever building block is
necessary for the type of system considered. For example, describing
the microstate of a monoatomic gas consists in specifying the 3 coor-
dinates and the 3 velocities of each atom. Depending on the degree
of accuracy of the description, the state of the electrons in each atom
might also have to be specified, etc.
When molecules collide with each other, or simply interact through
van der Waals-like forces, they constantly exchange energy. During
any evolution of a system, its total energy is a constant. What fluctu-
ates, however, is the way in which the energy is split among all the
degrees of freedom of the system. When two molecules collide, the
translation kinetic energy of the two molecules may decrease, if one
molecules starts spinning thereby increasing its rotational energy.
Similarly, some of the energy may be transferred to the electrons of a
molecule, which may later result in a photon emission, etc. Through
any of these processes, the energy is constantly transformed from one
form to another.
Keeping track of the microstate of a system comprising of the or-
der of 1023 molecules, i.e. checking whether any particular molecule
is spinning rather than another one is vibrating, is a daunting task
beyond any possible analysis. A breakthrough we owe to the genius
of Boltzmann consists in assuming that what we call thermal equi-
librium is a situation where the energy of a system is shared evenly
over all its possible degrees of freedom. Put differently: at thermo-
dynamic equilibrium, all the microstates with the same energy E are
equiprobable. This is sometimes referred to as the ergodic hypothesis. You can find on the Internet many Java
A central characteristic of a macrostate is the number of mi- applets dealing with simple molecular
dynamic simulations. See e.g. this one,
crostates compatible with it. Given the ergodic hypothesis, a sys- and use toi check that each degree of
tem is more likely to be found in macrostate that corresponds to freedom.
One Helium atom (atomic weight 4)
numerous microstates than the other way around. This is exactly the and one Neon atom (atomic weight
meaning of Boltzmanns equation; on Bolztmanns grave, it is writ- 20) are trapped inside a small box
ten as S = k B ln(W ), where the W stands for Wahrscheinlichkeit, i.e. and they collide repeatedly with each
other. Assuming ergodicity, can you
probability. determine the average kinetic energy of
A more common notation for the number of microstates having each atom?
energy E is ( E). Boltzmann equation for the entropy is

S = k B ln [] (VI1)
thermal fluctuations VI3

which can also be written as


S = k B ln [ p] (VI2)
where p = 1/ is the probability of any microstate. V,N
Es Sub-system

Boltzmanns Distribution heat

Heat Bath
Imagine a small system with a given volume and a given number of
T, Eb
molecules, which is in contact with a much larger system with which
it can exchange energy under the form of heat. For example: a tiny Figure VI2: A subsystem in con-
dust particle floating in a large room filled with air. The particle is tact with a heat bath, with which it
constantly hit by air molecules, which makes its every atom jiggle exchanges energy.

around some equilibrium position. Each time an air molecule hits the
dust particle, it may loose some of its energy in favor of the particle.
The opposite may also happen, an atom of the particle may hit an
air molecules in its vicinity, thereby loosing some of its energy. The
dust particle is therefore constantly fluctuating between different
microstates. We want to characterize these fluctuations.
In the following, we refer to the dust particle as the subsystem s
and the air filling the room as the heat bath b. The system comprising
the subsystem and the heat bath is the composite system. Although
the energies Es and Eb do fluctuate in a complex way, the energy
of the composite system E = Es + Eb is a constant. Whenever the
subsystem has energy Ei , the heat bath has energy E Ei .
The clever trick for analyzing the fluctuations of s consists in not-
ing that the ergodic hypothesis does apply to the composite system.
Accordingly, the probability that the subsystem be in state s is
1 b ( E Es )
ps = (VI3)
i b ( E Ei )
where the denominator accounts for the fact that the probabilities
should add up to 1.
One interesting observation that can be done at this stage is that ps
does only depend on the energy Es . This is naturally a consequence
of the ergodicity of the composite system. To express the dependence
of ps on Es in a more practical form, one can write
Sb ( E Es )
ln[ ps ] = ln[b ( E Es )] C = C (VI4)
kB
where the constant C is simply the logarithm of normalizing con-
stant, and the second equality results from Boltzmanns equation. The thermodynamic relation T1 =
 
Remember that the subsystem is much smaller than the heat bath, S
E V,N
can be considered as the
so that Es  E. This enables to write a Taylor development, limited definition of the temperature.
to the first term as
S ( E) 1 dSb
ln[ ps ] ' b Es C (VI5)
kB T k B dE
VI4 c.j. gommes, chim0698

Remembering the thermodynamic definition of the temperature, the


result takes the final form
1 Es
ps = exp[ ] (VI6)
Z kB T
where we have incorporated all the constants in the single constant
Z. In many statistical mechanics texts you
The seemingly irrelevant normalization constant Z happens to be a will find the notation = 1/(k B T ).

central quantity in statistical physics. It is called the partition function

E
Z= exp[ k B iT ] (VI7)
i

Knowing the dependence of the partition function on the macro-


scopic variables V, N, T, is equivalent to knowing all the thermody-
namical properties of the system, and even slightly more.

Relation to thermodynamics

The average energy h Ei can be calculated as i Ei pi , which can be


expressed in terms on the derivatives of Z as

1 E ln[ Z ]
h Ei =
Z Ei exp[ k B iT ] =
(VI8)
i

The average entropy is calculated in the same way, starting from Eq.
??, i.e. as
hSi = k B pi ln[ pi ] (VI9)
i
Note that this definition coincides with the microcanonical entropy
(Eq. ??) in the particular case where pi = 1/ for all states.
Using Boltzmanns distribution (Eq. VI6) for pi , the definition of
the average entropy can be put under the form

h Ei
hSi = + k B ln[ Z ] (VI10)
T
This expression enables us to identify Helmholtzs free energy F =
h Ei T hSi with
F = k B T ln[ Z ] (VI11)
Therefore, knowing the partition function is equivalent to knowing a
thermodynamic potential of the system.
All the relevant thermodynamic properties can then be calculated
in the usual way, as
     
F F F
S= P= = (VI12)
T V,N V T,N N T,V
The derivatives of the free energy result
from the differentiation of F = E TS,
when using the general relation dE =
TdS pdV + dN.
thermal fluctuations VI5

The statistical mechanical approach enables to go further than the


classical thermodynamics. For instances, the energy fluctuations of
the system can be estimated as
D E D E2
E2 = E2 E (VI13)

Quantifying the amplitude of the thermal fluctuations as


D E 1 E 1 2 Z
E2 =
Z Ei2 exp[ k B iT ] = Z 2
(VI14)
i

and using the fact that ln[ Z ]/ = h Ei, one reaches the final
result p
E = k B T 2 CV (VI15)
where CV = E/T is the heat capacity of the subsystem. Can you imaging an experimental
An interesting observation in Eq. VI15 is that E increases only setup to put statistical fluctuations in
evidence?
as N 1/2

A statistical derivation of the ideal gas law

We consider here the behavior of a large collection of N molecules in


a box of volume V. We know that this system should obey the ideal
gas law P = Nk B T/V. The purpose of this section is to illustrate the
method of statistical mechanics on a well known physical system. Gaussian integrals are integrals of the
The states of the system are defined here over a continuous space, type
Z + r
both for the 3N positions and the 3N momentum coordinates. The exp[ ax2 ]dx =

a
partition function is calculated by replacing the sum by an integral.
" !# Could you guess its a1/2 dependence
N
1 1 1 k p n k2 from dimensional considerations?
Z Z

k B T n
3N 3N
Z= d x d p exp (VI16)
N! h 3N =1 2m

The first factor accounts for the fact that interchanging any two parti-
cles does not change the state. The volume integral is simply equal to
V N . The momentum integral can be calculated as
3N
1 p2
Z 
dp exp (VI17)
k B T 2m

which can be calculated by Gaussian integration.


The partition function of the ideal gas is therefore

1 VN
Z= 3N
(2mk B T )3N/2 (VI18)
h N!
from which the free energy is found to be
 
3
F = k B T N ln[V ] ln[ N!] + N ln[2mk B T ] 3N ln[h ] (VI19)
2
VI6 c.j. gommes, chim0698

To estimate the chemical potential,


All the properties of the ideal gas can then be obtained by simple
we have used Stirlings approxima-
differentiation, following Eq. VI12. In particular, one finds immedi- tion ln[ N!] ' N ln[ N ] N, which
ately holds for large values of N.
Nk B T
P= (VI20)
V
as one should. The chemical potential can be put under the form

= k B T ln[ P/P0 ] (VI21)

where P0 ( T ) is a function that depends only on the temperature

1
P0 = (2m)3/2 (k B T )5/2 (VI22)
h 3
VII0 c.j. gommes, chim0698

Further reading

1. There are many good textbooks on statistical mechanics. An excel-


lent starting point is the relevant chapters in the Feynman lectures
on Physics;

2. If you like online courses, a few good lectures by Leonard Susskind


are available on Youtube. These lectures are more general, and also
more theoretically-oriented, than what I saw in this chapter.

Exercices

1. About Macrostrate and Microstates: estimate the memory space


required to store on a computer the microstate of 1 mg of oxygen
(in double precision variables). Compare that figure to the total
computer memory space available on Earth.

2. Determine the vertical density profile of an isothermal ideal gas in


a gravitational field. Do you find some similarity with Boltzmanns
distribution ?
Metastability and nucleation

A lot of people like snow. I find it to be an unnecessary


freezing of water.

Carl Reiner

This chapter is an incomplete draft, which I included here for


your convenience. Please rely on your own notes to study this
chapter.

When studying macroscopic equilibria the specifics of how a phase


is created are generally irrelevant, as far as its equilibrium proper-
ties are concerned. This is because the state of a system is uniquely
defined by its thermodynamic variables. This is not necessarily the
case for nanometer-sized systems, for which the size and shape of
the phases matter, in addition to the macroscopic variables. This is
incidentally the reason why nanometer-sized systems often exhibit
hysteresis: several states are possible for a given set of thermody-
namic variables. We shall come back to this later.
For the moment, consider a supersaturated system free to evolve
isothermally. It can a priori be any type of system: a gas compressed
isothermally above its saturation pressure, a vapour cooled below
its dew point, a melt quenched into the unstable region of it phase
diagram, a salt solution exceeding its solubility product, whatever.
What all these systems have in common is that they are unstable:
a phase transition is possible according to macroscopic thermody-
namics. The reason why it does not necessarily take place is that the
formation of germs of the new phase would necessarily come with
an interface having a thermodynamic cost ( per unit area of the
interface). Because the new germs have to be small initially, the rela-
tive importance of surface tension is initially very large. Although a
phase transition may be possible macroscopically, the thermodynamic
cost of creating the interface may be prohibitive, so that the transition
VII2 c.j. gommes, chim0698

does not necessarily take place. An example of such a metastable sys-


tem is freezing rain: it consists of undercooled rain drops that remain
liquid while in the air, and that freeze upon touching a preexisting
icy surface.

Supersaturation

Let us see how this can be analysed thermodynamically. We need a


thermodynamic measure of how far we are from equilibrium. The
natural measure would be , the difference in chemical potential
between the metastable and the equilibrium phases. Note that, by
definition 0. It is customary to define the supersaturation as


ln( ) = (VII1)
kB T
We now consider a few particular cases, and explicit the values of
and in these cases.

Vapour compressed above its saturating pressure The chemical potential


of a vapour at a given temperature is
 
P
= 0 ( T ) + k B T ln (VII2)
P0 ( T )

where P0 ( T ) is the saturating pressure and 0 ( T ) is the chemical


potential of the liquid 10 . In this case = 0 ( T ), and the super- 10
Note that this relation satisfies = 0
saturation is simply for P = P0 , as it should.
P
= (VII3)
P0

Undercooled liquid The chemical potential of the liquid and solid


phases are equal at the meting temperature Tm . To calculate their val-
ues at other temperature, we can use the general relation (/T ) P =
s and write

l = l ( Tm ) sl ( Tm )( T Tm )
s = s ( Tm ) ss ( Tm )( T Tm ) (VII4)

The subscripts stand for solid and liquid. These expressions are ex-
pected to hold for temperature not too far from Tm . Expressing that
l ( Tm ) = s ( Tm ), we find

l s = ( T Tm )(ss sl ) (VII5)

Taking account of the general relation

hm
sl sl = (VII6)
Tm
metastability and nucleation VII3

where hm is the enthalpy of melting, one has the final relation

hm
 
T
ln( ) = 1 (VII7)
kB T Tm

Supersaturated salt solution Consider a salt solution with the follow-


ing simple equilibrium

A+ + B AB(s) (VII8)

where AB is a solid. The difference in chemical potential is

= AB A B (VII9)

where A and B depend on the concentrations C A and CB through

A = 0A + k B T ln(C A )
B = 0B + k B T ln(CB ) (VII10)

and AB is a constant. The equilibrium corresponds to the situation


where = 0, which can be put in the familiar form

C A CB = K S (VII11)

where the solubility product KS is given by

k B T ln(KS ) = AB 0A 0B (VII12)

The supersaturation can therefore be written in this case

C A CB
= (VII13)
KS
which can be easily generalised to more complex equilibria than Eq.
VII8.

Nucleation barrier and critical radius

We consider now a supersaturated system. It can be any type of


system, but it might be convenient at this stage to think of it as com-
pressed vapour.
Because 0 the vapour should spontaneously condense
into liquid, according to macroscopic thermodynamics. If one looks
carefully into this, however, one realises that in order to create a
liquid region one also has to create a liquid/vapour interface with
a given thermodynamic cost per unit area. In other words, the
extra free energy associated with the creating of a liquid droplet in a
vapour at pressure P and temperature P0 is

G = N + A (VII14)
VII4 c.j. gommes, chim0698

where N is the number of molecules in the droplet, which is also the


number of molecules taken out of the vapour, and A is the surface
area of the droplet. Particularising this to a spherical drop, with the
vapour chemical potential given by Eq. VII2, one has

4R3
G = + 4R2 (VII15)
3vm
where R is the radius of the droplet, and vm is the molecular volume.
The first term in Eq. VII17 is the one we had in mind when think-
ing of macroscopic thermodynamics: according to this term the free
energy is decreased when R is made as large as possible, i.e. when
as much vapour as possible is condensed (as expected for P P0 ).
However, the second term balances this effect by assigning a positive
thermodynamic cost to the creating of a droplet. What you should
expect to happen therefore depends on more than merely the com-
parison of P and P0 .
When you have a situation such as described by Eq. VII17, it
is always useful to put it first in a dimensionless form, in order to
understand the role of each of the dimensional parameters vm , , and
k B T (which we take as an order of magnitude for = K B T ln( )).
For that purpose, on may first notice that the first term is of the order
k B TR3 /vc ; it is balanced by the second term, which is of order R2 .
This suggests using
vm
= (VII16)
kB T
as a unit for R and, say 42 as the unit of G.
Using these units, the extra free energy associated with the cre-
ation of a droplet of radius R can be written in the simpler way
 2  3
G R 1 R
= ln ( ) (VII17) 10
42 3
The situations sketched in Fig. VII1. For < 1 the volume contribu-
G/(4 2 )

tion is positive, so that the minimum of G is reached for R = 0, i.e. 5

no condensation is expected whatever. By contrast, in the case of su-


persaturated vapour ( > 1), the state with R = 0 is only metastable.
0
A state with a lower fee energy can be reached for large values of R,
but this requires passing a finite free-energy barrier.
The critical radius of the droplet that has to be formed for conden-
5
0 1 2 3 4 5
sation to proceed is obtained by expression that dG/dR = 0, which R/
leads to
2 Figure VII1: Free energy correspond-
R = (VII18)
ln( ) ing to the creation of a spherical object
with radius R for various supersatu-
and the associated free energy barrier is rations: = 1/2 (red), = 1 (green),
= 2 (blue), and = 4 (black). The
4 42 1
G = = 4R2 (VII19) length unit is = vc /(k B T ).
3 [ln( )]2 3
metastability and nucleation VII5

It is easy to remember that the free energy barrier is simply one third
of the critical nucleus surface energy.
To make these results more "real", let us consider the practical case
of undercooled water. In this case, the critical radius can be written
as
vm 1
R = 2 (VII20)
hm 1 T/Tm
In the case of water ice with vm ' 1.09 106 m3 /g, hm = 333 J/g,
and ' 30 103 J/m2 , one has
vm
R (0) = 2 ' 2 (VII21)
hm

This is the critical radius extrapolated to T 0 K. Remember,


however, that the assumptions underlying Eq. (VII7) hold only for
T ' Tm . Anticipating on the next section, the relevant quantity for
nucleation is G /(k B T ). In the case of undercooled water, it can be
written as Reaching 50 C of supercooling is not
rare. A famous case is sodium acetate
used in hand warmers. The melting
G 1 4R (0)2 1
= (VII22) temperature of sodium acetate is 58 C,
kB T 3 k B Tm T/Tm (1 T/Tm )2 but it can remain in a metastable liquid
state down to about 0 C.
with
4R (0)2
'4 (VII23)
k B Tm
The critical nucleus size and the corresponding free energy barrier
are plotted in Fig. VII2. A typical value for the supercooling of
water is ' 220 K. This is e.g. the temperature of freezing rain. This
corresponds to T/Tm ' 0.8; The critical nucleus is about 2 nm across.

2 4
10 10 Figure VII2: Size of the critical ice ra-
dius and the corresponding free energy
3
barrier in the case of supercooled water
10
1 (Tm = 273 K).
G /(k B T )

10
R (nm)

2
10

0
10
1
10

1 0
10 10
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
T /T m T /T m

Let us consider now the case of a water vapour at T = 373 K


compressed beyond P0 = 1 atm. The molecular volume of liquid
water is vm ' 2.99 1029 m3 , the surface tension is ' 75 103 N/m,
which leads to ' 4.3 . The free energy barrier is therefore

G 1 4 (2)2 1
= (VII24)
kB T 3 k B T [ln( P/P0 )]2
VII6 c.j. gommes, chim0698

with
4 (2)2
' 139 (VII25)
kB T
The critical nucleus size and the corresponding free energy barrier
are plotted in Fig. VII3.

2 4
10 10 Figure VII3: Size of the critical ice
radius and the corresponding free
3
energy barrier in the case of water
10
1 vapour at 373 K compressed above

G /(k B T )
10
P0 = 1 atm.
R (nm)

2
10

0
10
1
10

1 0
10 10
0 0.5 1 1.5 2 0 0.5 1 1.5 2
ln(P /P 0 ) ln(P /P 0 )

The free energy barrier is directly related to the nucleation rate.


Imagine the case of nucleation from a supersaturated solution of a
simple molecule, say sugar. Sugar molecules in the solution are con-
stantly assembling into fleeting clusters. If these clusters are smaller
than the critical size R , they can only dissolve back into the solution.
If one cluster forms larger than the critical size, then it is unstable
towards growth and it eventually leads to the crystallisation of a
macroscopic quantity of sugar.
The probability of seeing a fluctuation is directly related to the cor-
responding change in the free energy G. In fact, it is proportional
to11 11
This is Einsteins thermodynamic
Prob( f luctuation) ' eG/k B T (VII26) approach to fluctuations. We shall come
back to it in details in a further chapter.
The ' sign means that many factor have not been written. In partic-
ular, it is clear that the probability has to depend on the total volume
that is being considered, as well as of the observation time. Clearly,
the probability of seeing a fluctuation of a given amplitude has to
be two times larger is the volume of the solution is two times larger.
Also, if you wait two times longer, you expect to see two times as
many critical nuclei forming. What we are interested in is therefore
the nucleation rate, i.e. the number of critical nuclei forming per unit
volume of the solution and per unit time. The missing factor in Eq.
VII26 has the dimensions of m3 s1 .
Physically, one expects the nucleation rate (in m3 s1 ) to depend
on the mobility of the molecules in the solution. If the molecules
were very slow, the statistics of the fluctuations would still be de-
scribed by eq. VII26 but the time it would take to see the clusters
appear and disappear would be very long. A convenient quantity
to use is therefore the viscosity of the solution , the unit of which
metastability and nucleation VII7

is Pa.s = J m3 s. The other dimensional quantities that may enter


the expression are the molecular volume vm and the thermal energy
k B T. The only combination of these quantities with the correct di-
mension is k B T/(v2m ). One therefore reaches the conclusion that the
homogeneous nucleation rate has to be of the form

2
 
k T 16
I ' 2B exp (VII27)
vm 3 k B T [ln( )]2

where the missing factor is now dimensionless and of the order of


one.

Heterogeneous nucleation

The situation we have implicitly analysed in the the previous section


is that of homogeneous nucleation. In case there is a solid surface in
the system, the metastability can be significantly reduced as we now
show.
Imagine we are condensing, say water vapour. If the solid surface
is hydrophilic, clearly the thermodynamic cost of the free surface is
reduced if the vapour condenses on the surface, rather than as a free
droplet. In this case the shape of the nucleus is like in Fig. VII4, and
the term A in Eq. VII14 would have to be replaced by
h
FA = A L + (SL S ) ASL = ( A L ASL cos( )) (VII28)
r
where A L is the surface area of the liquid, ASL is the surface area of
the wetted solid and the s are the corresponding surface energies.
R
For the RHS, we have used the Young-Dupr relation.
Let us do some geometrical calculation to express the volume and
the relevant surface areas in terms of the radius of curvature R of the
droplet12 . The volume of the droplet is
Figure VII4: Drop on a surface, with
contact angle . Cf. the first exercise we
V = R3 (1 cos( ))2 (2 + cos( )) (VII29) did on surface tension.
3 12
The calculations are straightforward.
and the areas are They are based on the formulae V =
h(3r2 + h2 )/6 and A = (r2 + h2 ) for
AL = 2R2 (1 cos( )) the volume and area of a spherical cap.

ASL = R2 (1 cos( )2 ) (VII30)

Using these expressions, the free energy of formation of a nucleus on


a surface can be put as

4R3
 
G = + 4R2 f ( ) (VII31)
3vm

with
1 3 1
f ( ) = cos( ) + cos3 ( ) (VII32)
2 4 4
VIII0 c.j. gommes, chim0698

Eq. VII31 is identical to Eq. VII17 except for the factor f ( ). This
means that Eq. VII18 remains valid in the case of heterogeneous nu-
cleation, but the radius has to be understood as a radius of curvature

2
R = (VII33)
ln( )
On the other hand, the free energy barrier has to be multiplied by a
factor f ( ), the dependence of which on is plotted in Fig. VII5. As 1

you may have expected, f ( ) is close to 1 for close to 180 , because


0.8
in this case one has a spherical nucleus sitting on a surface, with a
vanishingly small contact area. By contrast, f ( ) is close to 0 for 0.6

f ()
close to 0. In this case case, the change of surface energy SL S is
0.4
so negative that it almost balances the L contribution.
Remember that the very reason why there is a nucleation barrier is 0.2

the finite thermodynamic cost of the interface between the medium 0


0 45 90 135 180
and the new phase being created. In the case of heterogeneous nucle- ( )
ation on a surface with ' 0, the cost of the liquid/vapor interface
Figure VII5: Function f ( ) defined in
L is almost exactly compensated by the reduction of the solid sur- Eq. VII32.
face energy, when passing from S to SL . This is the reason why
the nucleation barrier vanishes for 0. Condensation occurs then
exactly at saturation.
It is interesting to analyse also the nucleation rate in the case of
heterogeneous nucleation. In this case, the quantity you are inter-
ested in is no longer the number of nucleation events per unit time
and volume, but per unit time and area of the solid surface. So the
unit of I has to be s1 m2 . The same dimensional analysis as in the
case of homogeneous nucleation leads now to

2
 
kB T 16
I ' 5/3 exp f ( ) (VII34)
vm 3 k B T [ln( )]2

which differs from Eq. VII27 though the function f ( ) and through
the exponent of vm in the pre-exponential factor.
Another thing that I encourage you to think about is the case
of Fig. VII6 which is more realistic that a flat surface. For a given
contact angle and a given volume of the droplet, the ratio of the
liquid/solid to the free liquid surface area is larger than on a flat
surface. You expect this to decrease still the nucleation barrier.
Figure VII6: Surfaces are seldom
perfectly flat. Remember that the length
scales we are talking about are just a
few nanometers. On that scale, any
surface has defects, which may be
modelled as ditches: you may therefore
expect this configuration to be more
realistic than Fig. VII4.
Adsorption on surfaces

There are no beautiful surfaces without a terrible depth.

Friedrich Nietzsche

This chapter is an incomplete draft, which I included here for


your convenience. Please rely on your own notes to study this
chapter.

An adsorption isotherm describes the mass-transfer equilibrium Figure VIII1: Irving Langmuir (1881-
(a given temperature) between a reservoir and a surface on which 1957) is the first industrial researcher to
receive a Nobel Prize (in 1932).
molecules can adsorb. The general relation takes therefore the form

n = f () (VIII1)

where n is the amount of molecules adsorbed (e.g., in moles of


molecules per gram of adsorbent), and is the chemical potential of
the adsorbate molecules in the reservoir. Just a summary of standard thermody-
If the reservoir is gaseous, it can often be considered as made up namics of open system. Gibbs relation

of ideal gases with chemical potential


dG = SdT + Vdp + dN (VIII2)

= 0 ( T ) + k B T ln( P/P0 ) (VIII6) which notably enable one to write


 
G
where P is the partial vapour pressure, P0 is the saturating pressure = (VIII3)
N T,p
and 0 ( T ) is the chemical potential at saturation. Because, at a given One of Maxwells relations is
temperature the chemical potential only depends on P, the adsorp- V
= (VIII4)
tion isotherms are often expressed as n = f ( P). In case the reservoir p N
is a dilute solution, one can approximate its properties by those of an In the case of an ideal gas, this leads to
ideal solution. In particular, the chemical potential is k T
= B (VIII5)
p p
= 0 ( T ) + k B T ln(C/C0 ) (VIII7) which leads directly to Eq. VIII6.

where C is the solution and C0 is the saturating solution. In this case,


the adsorption isotherm is expressed as n = f ( P). In both cases, the
VIII2 c.j. gommes, chim0698

chemical potential is of the form

= B( T ) + k B T ln( x ) (VIII8)

where B is a temperature-dependent constant, and x is an intensive


variable that controls the composition.
Note that Eqs. VIII6 and VIII7 are the chemical potentials per
gas phase
molecule. The same expressions are valid for the chemical potentials
per mole if Boltzmanns constant k B is replaced by the ideal gas con-
stant R. This is merely a question of units. adsorbed
phase

Langmuirs adsorption Figure VIII2: Adsorption is a phase


equilibrium between a gaseous phase
(or a solution) and an adsorbed phase.
Langmuirs model of adsorption consists in assuming that the sur-
face of the adsorbent is equivalent to a checkerboard, on each square
of which a single molecule can come and stick with a specific adsorp-
tion energy u a . Because the molecules are attached to the surface, the
corresponding energy u a is negative. A possible origin of that energy
is dispersive forces, but the nature of the forces is irrelevant here.
Let us assume that the adsorbent has a total of A adsorption sites
on which the molecules can adsorb. If we want to describe quan-
titatively adsorption in this model system, we have to answer the
following question. When the adsorbent is put in contact with a
reservoir with chemical potential and temperature T, how many
sites become occupied by a molecule?
The answer to that question is a famous result in surface science
and bears the name of Irving Langmuir. If you Google the words
"Langmuirs isotherm" you will find many different mathematical
derivations of it. This is the characteristic of important results: there
are many different ways in which they can be understood.
The derivation that we present here is not the most common, but it
is probably the most general. Its structure is the following

1. We first use the fundamental principles of statistical mechanics to


calculate the chemical potential ads of molecules adsorbed on the
surface, as a function of the temperature T and of the total number
of molecules adsorbed N. This will lead to a general expression of
the form
ads = Function( N, T ) (VIII9)

The fact that ads depends on N is very natural: the more molecules
you have in a system, the larger their propensity to leave that sys-
tem, and vice versa.

2. We will then express that the surface with N adsorbed molecules


is in mass-transfer equilibrium with a reservoir with chemical
adsorption on surfaces VIII3

potential res . The equilibrium condition is expressed as res = a) b) c)


ads , which will eventually enable us to find an expression for N
vs. res , at a given T.

To calculate the chemical potential ads one has to calculate first


the Helmholtz free energy
Figure VIII3: On a given surface with
A adsorption sites, there are a) 1 = A
Fads ( N, T ) = Uads ( N, T ) TSads ( N, T ) (VIII10) different configurations with N = 1
molecules, b) 2 = A ( A 1)/2
where U is the energy and S is the entropy, and using the general configurations with 2 molecules, and
A
thermodynamic relation c) in general = ( N ) for N adsorbed
molecules.
 
F
= (VIII11)
N T

to calculate ads .
The energy is easy to calculate: it is simply given by

Uads = Nu a (VIII12)

The entropy is barely more difficult to calculate. All you have to re-
member is the microscopic definition of the entropy in terms of the The difference between the Helmnholtz
number of configurations compatible with whatever macroscopic free energy F and Gibbs free energy
G is important only for systems that
variables define your system. This is the celebrated Bolzmanns equa- can undergo large changes of volume.
tion In the present context, the difference is
irrelevant.
S = k B ln() (VIII13)
In our case, is the number of ways in which N molecules (macro-
scopic variable) can be arranged on a surface with A adsorption sites.
As we show in Fig. VIII3, the total number of different configura-
tions with N adsorbed molecules is
 
A A!
= = (VIII14)
N N!( A N )!

In principle, the problem is now solved. In practice, however, the


derivative of F is going to be difficult to calculate if we keep the fac-
torials. Luckily enough, any macroscopic number of molecules, such
as N is extremely large, so that the factorial can be approximated
with a very good accuracy using Stirlings formula

ln( N!) ' N ln( N ) N (VIII15)

Using this approximation, the number of configurations can be writ-


ten as
      
N N N N
ln() = A ln + 1 ln 1 (VIII16)
A A A A

which is plotted in Fig. VIII4. Note that N/A is nothing but the
fraction of the sites that are occupied by a molecule, which we shall
VIII4 c.j. gommes, chim0698

refer to as . The quantity ln() is sometimes referred to as the con-


figurational entropy. The latter quantity is found here to be maximal 0.8

for = 1/2, and to vanish for = 1 and = 0. These two values 0.6

ln()/A
simply mean that there is a single way of having all sites occupied or
0.4
empty, respectively.
The chemical potential of the adsorbed phase is then obtained by 0.2

taking the derivative of F with respect to N, which leads to 0


0 0.2 0.4 0.6 0.8 1
  N/A
N
ads = u a + k B T ln (VIII17)
AN Figure VIII4: Number of configura-
tions of N molecules adsorbed on a
It is plotted in Fig. VIII5. The chemical potential has two contribu- surface with A sites, as calculated from
tions: the first (u a ) is energetic and the second is entropic. Because of Eq. VIII16 .

the latter, the chemical potential diverges for N A and for N 0.


This means that molecules would tend to desorb when N ' A and
tend to adsorb when N ' 0, even if it is energetically unfavorable. It
is interesting to note that the entropic term is proportional to k B T, the
thermal energy. At the absolute zero, the chemical potential would An order of magnitude that is useful
simply be u a because a molecule could only go to a state with lower to know is the following: at room
temperature, the thermal energy is 1/40
energy. At finite temperature, thermal fluctuations enable a molecule eV. Think of it. How does that compare
to move to a state of higher energy by taking the missing energy in a with the typical energy involved in a
chemical reaction?
reservoir under the form of heat. All this is in Eq. VIII17.

5 Figure VIII5: Chemical potential ads


( a d s u b )/(k B T )

as function of surface coverage N/A, as


calculated from Eq. VIII17 .

5
0 0.2 0.4 0.6 0.8 1
N/A

If the surface defined by Eq. VIII17 is put in contact with a reser-


voir defined by Eq. VIII8, molecules get transferred from the surface
to the reservoir and vice versa until the equilibrium is reached, which
corresponds to ads = res . This can be written as

N Kx
= (VIII18)
A 1 + Kx
where K is a constant, given by

B ua
 
K = exp (VIII19)
kB T
adsorption on surfaces VIII5

and x is a composition variable (e.g. concentration or partial vapour


pressure). Equation VIII18 is the celebrated Langmuirs adsorption
isotherm.

1 Figure VIII6: Langmuirs adsorption


isotherm for K = 1 (blue), 5 (green) and
20(red).
0.8

0.6
N/A

0.4

0.2

0
0 1 2 3 4 5
x

In the limit of small concentrations or vapour pressure, i.e. if


Kx  1, it reduces to a simple proportionality law

N
' Kx (VIII20)
A
which is sometimes referred to as Henrys law.
The two parameters of the Langmuirs isotherm are A and K. The
total number of adsorption sites A is proportional to the specific
surface area. When the concentration or the vapour pressure is very
large, all adsorption sites tend to be occupied. In this case N A.
The other parameter is K, which depends on the energetics of the
adsorption (relative to the thermal energy k B T.) The isotherms are
plotted in Fig. VIII6. Note the effect that a change in K has on the
overall shape of the adsorption isotherm . Think about how a change in temper-
ature affects K and the isotherm. Does
To obtain a more precise, even if qualitative, meaning of the con-
that make sense to you, based on your
stant K, one needs a better understanding of the various terms in intuitive understanding of thermal
the reservoir chemical potential, Eq. VIII8. Consider for instance fluctuations?

a solution. If the solution is dilute, the interactions between solute


molecules can be neglected. In this case, we could use a cartoon
model of a solution similar to the Langmuir model of a surface (see
Fig. VIII7). In this case the number of adsorption sites A is replaced
by the number of pseudo-sites in the solution V, the adsorption
energy is replaced y the solvation energy us , and N is replaced by
x V, where x is the molar fraction of the solutes13 . Remembering 13
We make here an implicit assumption
that the solution is diluted, i.e. that x  1 , Eq. VIII17 becomes about the sizes of the solute and solvent
molecules. Can you state it explicitly?

sol = us + k B T ln( x ) (VIII21)

which is exactly of the type of Eq. VIII8 , with B = us . In this case


VIII6 c.j. gommes, chim0698

the constant K takes a very explicit meaning

us u a
 
K = exp (VIII22)
kB T

K is therefore just a Boltzmann factor of the usual type exp(E/(k B T ))


.
In order not to let you think that Lagmuir-like adsorption is so
simple as what we have presented here, I would like to end the dis-
Figure VIII7: A cartoon model of
cussion with Fig. VIII8. In order to adsorb a molecule from a solu- a solution: the solute molecules are
tion, one has to remove solvent molecules adsorbed on the surface positioned in a 3D cube. This leads to
the same mathematics as Fig. VIII3.
to free some space for the solute, one also has to remove salvation
molecules from around the solute, etc. The expression of K that we
have obtained remains essentially valid, but the change of energy E
is naturally not simply us u a . Can you think of a more accurate
expression, and enumerate all the energies that should enter it?

a) b) c) Figure
500VIII8: The situation is actually
a bit more involved. For example, to
adsorb a solute molecule on a surface
(a), one
400has first to remove the solvent
molecules adsorbed on the surface,

V a d s (cm 3 /g )
as well as those solvating the solute
(b), and
300only then can adsorption take
place.

200
Multilayer adsorption
100
When nitrogen adsorption is measured on carbon nanotubes the
isotherms look like in Fig. VIII9, which is clearly different from the
0
Langmuir-type isotherm that we last considered. Isotherms of that 0 0.5 1
type are quite common, and they are referred to as "type II". P /P 0
Unlike for "type I" (Langmuir-like) isotherms, the quantity ad-
Figure VIII9: Example of nitrogen
sorbed does not saturate when the pressure is increased. The most adsorption at 77 K measured on a
obvious explanation for this type of behaviour is multilayer adsorp- carbon nanotubes sample. Note the
tion. different shape with respect to the type
I isotherms described by Langmuirs
To describe quantitatively this type of adsorption, the assump- equation.
tions of the Langmuir model have to be relaxed. The BET adsorption
model is identical to Langmuir in all respects, except that we now
allow several molecules to adsorb on top of each other in each ad-
sorption site.
adsorption on surfaces VIII7

Figure VIII10: The three authors of


the BET equation on the occasion of
a later reunion (picture taken from
I. Hargittai, The Martians of Science,
Oxford University Press). From left
to right: Paul Emmett (1900-1985)
was an American Chemical Engineer,
Stephen Brunauer (1903-1986) is a
Hungarian physicist who emigrated
to the USA, Edward Teller (1908-2003)
is also Hungarian. Besides the BET
theory, Teller was also the director
of the American fusion (H) bomb
program, as well as of the so-called
Starwars program. Teller regretted not
to receive a Nobel prize for the BET; he
was, however, the first recipient of the
Ig-Nobel prize for peace in 1991 for his
contribution to the hydrogen bomb.

In principle, the equilibrium properties of the model in Fig. VIII


11 can be calculated in the same statistical way that we used for
Langmuirs isotherm. In practice, however, the evaluation of is
extremely complicated when multi-layer adsorption is permitted.
Just to convince you that this is a difficult problem, we shall calcu-
late the first few values, up to N = 4. To make things even simpler,
we shall suppose that the adsorption energy is the same whether a Figure VIII11: In the BET model
several molecules may adsorb on top
molecule adsorbs on a surface or on another molecule. Therefore, we of each other on each adsorption site.
do not have to keep track of the energy of each configuration because This is the only difference with the
Langmuir model.
they all have the same energy. Even so, this is a difficult problem.

1. For N = 1, one has = A.

2. For N = 2, one has = A( A 1)/2 + A. The first term is the


number of configurations with a single molecule per site (as in
Langmuir), and the second term is the number of configurations
with 2 molecules on top of each other. So far so good.

3. For N = 3, one has = ( A3 ) + A + A( A 1). The first term is


the same as in Langmuir, the second is the configurations with
3 molecules piled up in a single site, and the last term is for the
configurations having one site with 1 molecule and another with
2 molecules. We call this the [1, 2] configurations. Note that [1, 2]
is distinguishable from [2, 1], so that term doesnt come with the
VIII8 c.j. gommes, chim0698

factor 1/2.

4. Things start to get tricker for N = 4. In that case you expect

A1
   
1 A
= A + A ( A 1) + A ( A 1) + A + (VIII23)
2 2 4

The first term is for 4 molecules in a site, the second is for [3, 1]
configurations, the third for [2, 2], the fourth for [2, 1, 1] and the
last one for the Langmuir-like configurations.

5. Etc.

As you can see, things become rapidly intractable. Remember that we


need a general relation, then to estimate its asymptotic approxima-
tion for large values of A and N, and finally to calculate its deriva-
tive! I give up.
There is fortunately another possible approach. The validity of the
calculations is limited to gas phase adsorption. But this is anyway
the conditions in which multi-layer adsorption is generally observed.
Here is how it works.

Langmuir again, but differently


We start by considering again the Langmuir adsorption scenario, to
build some confidence in the approach. The approach here is dy-
namical. Each adsorbed molecule desorbs at a given rate that we can
calculate. At the same time, the surface is constantly hit by molecules
from the gas phase, and we can also calculate the adsorption rate. We
shall then calculate the equilibrium as the condition under which the
adsorption rate exactly balances the desorption rate.
The surface has A sites, of which a fraction 0 is free and a fraction
1 = 1 0 is occupied by one molecule. At any finite temperature,
each molecule adsorbed on the surface vibrates as a result of all
the collisions with other molecules as well as of the vibration of the
surface. Occasionally one collision may bring an energy larger than
the adsorption energy u a , and the molecule desorbs (or evaporates).
The rate is expected to be proportional to the number of adsorbed
molecules N = 1 A and we may write the proportionality constant as
1/a where a is a characteristic desorption time.

N e = A1 /a (VIII24)

We shall come back later to the temperature dependence of a .


During the same time, molecules come from the gas and hit the v dt
surface, and may therefore adsorb. The number of collisions per
Figure VIII12: Simplified calculation
unit time and unit area can be estimated in the following way. The of the collision frequency. Because the
molecules have an average thermal
velocity v, the number of collision in
a time laps dt is roughly equal to the
number of gas molecules in a layer of
thickness vdt over the surface.
adsorption on surfaces VIII9

thermal energy being k B T, the average velocity of a molecule in the


gas is p
v ' k B T/m (VIII25)
where m is the molecular mass14 . The number of collisions of the 14
This is obtained by equating the
molecules with any surface in a time laps dt is roughly equal to the thermal energy k B Twith the kinetic
energy of a molecule1/2mv2 . The
number of molecules in a layer of thickness vdt that covers the sur- symbol ' means that a factor of the
face. Therefore order of 1 has been
neglected. In this
case the factor is 2
P p
#collisions/time/area ' k B T/m (VIII26)
kB T
where the first factor is simply the concentration of the gas15 Of 15
From the perfect gas law: PV = nk B T
course there are many numerical factors missing in this estimation16 , and therefore n/V = P/(k B T ).
16
We could make the calculation more
but the important result is that the rate is proportional to the pres- accurate by noting that only half the
sure. In the following, we shall simply write molecules in the vdt-thick layer are
going towards the surface. Etc.
#collisions/time/area = P (VIII27)

where is a constant that depends only on the gas.


The rate of adsorption of molecules can now be written as

N a = ea 0 AaP (VIII28)

where a is the area occupied by a single adsorption site. In this equa-


tion, starting from the right: P is the collision rate per unit area,
Aa is the total surface area, 0 is the fraction of the area that is avail-
able for adsorption, and ea is a factor lower than one that we have
introduced to account for the fact that not every collision results in
adsorption. You may think of ea as the probability that a molecule
hitting the surfacer does adsorb. Some molecules may simply bounce
on the surface; they do so with probability 1 ea .
Equating the adsorption and desorption rates N a = N d one finds

1 = a ea aP0 (VIII29)

which eventually results in


Ka P
1 = (VIII30)
1 + Ka P
with
Ka = a ea a (VIII31)
Equation VIII30 is exactly Langmuirs adsorption isotherm.

Multilayer adsorption
The analysis leading to Eq. VIII30 assumed that only one molecule
at most could exist at a given site, so the system was completely char-
acterised by the fraction of occupied sites 1 . If multi-layer adsorption
VIII10 c.j. gommes, chim0698

takes place, we have to calculate the fraction of sites occupied by one


molecule 1 , by two molecules 2 , three molecules 3 , etc. Once all
theses quantities will be known to us, we will be able to calculate the
average number of molecules per site as

hni = 1 + 22 + 33 + . . . (VIII32)

from which the total number of molecules adsorbed will be calcu-


lated as
N = Ahni (VIII33)

Our aim now is therefore to calculate hni. Before we start the analy-
sis, note finally that the s satisfy

0 + 1 + 2 + 3 + . . . = 1 (VIII34)

because each site is either bare or occupied by some molecules.


To calculate the s, we will proceed as in Sec. , but we now have
to write for each layer a balance equation similar to Eq. VIII29 .
Because the parameters e and may vary from one layer to another,
we shall use the notation Ki = i ei a for the characteristics of the ith
layer, starting the numbering at 0. The equilibrium of each layer is
written as

K0 P0 = 1
K1 P1 = 2
...
Ki1 Pi1 = i
... (VIII35)

where K0 is the quantity we called Ka in Sec. . The second equation,


for example, means that the number of molecules condensing on
the sites that host already one molecule is equal to the number of
molecules evaporating from the sites that host two molecules. The
other lines have similar meanings.
A standard BET assumption is that only the first layer (in direct
contact with the solid) is different from the others. The properties of
the latter layers are assumed to be indistinguishable from those of
the bulk liquid. In the following, we use the index b (e.g. Kb ) for the
values of the bulk, and index a for the values of the first layer (as in
Ka ) for the. With that assumption in mind, it proves convenient to
use the notations
Ka
= Kb P and C = (VIII36)
Kb
With these notations the solution to Eq. VIII35 is simply i = Ci 0 .
Using that result, the average number of molecules per site is17 17
This part here is almost recreational
mathematics! The first step itself is
already nice: To estimate i=0 i for
any 1, you can write it as

(1 + + 2 + ...)(1 )
(1 )
Of course you havent changed any-
thing by multiplying and dividing by
(1 ). If you develop the numerator,
however, you will find that it is equal to
adsorption on surfaces VIII11


C
hni = ii = (1 )2
0 (VIII37)
i =1

This is not the solution yet because 0 is still unknown. The value
found by expressing that i i = 1, which leads to

(1 C )(1 ) + C
1= 0 (VIII38)
1

The final result therefore takes the form


C
hni = (VIII39)
(1 )(1 + (C 1) )
which is the BET equation, for the derivation of which Edward Teller
considered he deserved a Nobel Prize.
To put the result under a more familiar form, we have to under-
stand the physical meaning of . Remember that

= b eb aP (VIII40)

Because is dimensionless and proportional to the pressure, all the


factors in front of P can be written as 1/Pb where Pb is some yet to
be specified pressure that characterises the bulk adsorbate at the
considered temperature. To gain some insight into this, consider for
instance what happens for P = Pb , which leads to = 1. This can be
written as
1
= eb aPb (VIII41)
b
This equation means that, at the pressure Pb , the number of molecules
that evaporate from the surface of the bulk adsorbate (left-hand side)
is equal to the number of molecules that condense on it from the gas
phase (right-hand side). This is exactly the definition of the satura-
tion. In other words Pb = P0 ! The final form of the BET equation is
therefore the following

N CP/P0
= (VIII42)
A (1 P/P0 )(1 + (C 1) P/P0 )
which is the one that we shall use in Sec. to determine specific sur-
face areas from experimental adsorption data. We postpone to Sec.
the thorough discussion of the BET constant C.
Examples of BET isotherms are plotted in Fig. VIII13 for three
different values of C. At low relative pressure, the same Henry-like
linear adsorption as in the Langmuir isotherm is observed. When
the pressure approaches the saturation, the number of molecules
adsorbed per site diverges. This corresponds to the condensation
of a thick liquid layer on the surface. The effect of the constant C is
significant only when the thickness is relatively small, say hni 5.
VIII12 c.j. gommes, chim0698

5 Figure VIII13: The BET adsorption


20 isotherm, corresponding to Eq. VIII42,
for C = 1 (blue), 10(green) and 100
4 (red).
10

hni = N/A
3
0
0 0.5 1

0
0 0.2 0.4 0.6 0.8 1
P /P 0

Note in passing that no condensation at all would be observed for


P P0 in absence of a solid surface. What Fig. VIII13 is showing is
an example of how a nanostructure18 can shift a phase equilibrium. 18
Remember that a large surface
Imaging you had a porous solid with pores that would be just large area can only be achieved with nano-
structured materials.
enough for about 10 molecules to fit in it, 5 on each side. Looking at
Fig. VIII13 you can already guess that these pores will be filled with
condensate at P ' 0.8P0 . We will analyse this is a more quantitative
way when discussing the Gibbs-Thomson effect.

Discussion

The BET constant


The constant C in the BET equation is referred to as the BET constant.
Remember that it is defined as
   
K e
C= = (VIII43)
Kb eb b

where the es are the fraction of collisions from the gas phase that
result in condensation (on the surface or on the bulk liquid surface),
and the s are characteristic evaporation times. Evaporation is a
thermally activated process. A molecule close to the surface of a
bulk liquid (but inside the liquid) has a binding energy ub 0 that
keeps it in the liquid. Because of the thermal motion, the molecule
constantly receives energy from its neighbours or by being hit by
an incoming gas molecule, etc. As a consequence, the energy of the
molecule is fluctuating randomly in time, and the probability of
finding it in a state of energy E above its ground state is given by
adsorption on surfaces VIII13

Boltzmanns factor
E
 
Proba( E) ' exp (VIII44)
kB T
where the missing factor19 is a normalising constant. If that extra- 19
That is the meaning of the symbol '.
energy happens to be larger than the binding energy of the molecule,
i.e. if E + ub 0 then it can evaporate. One therefore expects the
following temperature-dependency for the evaporation rate 1/b
 
1 ub
= b exp (VIII45)
b kB T
where b is a typical vibration frequency. Be sure you understand this
equation, in particular the sign in the exponential: ub is negative so
that if the binding energy is very large (i.e. very negative). A similar
expression can of course be given for adsorbed molecules. This leads
to
ub u a
 
ea b
C= exp (VIII46)
eb a kB T
The most important factor in there is the exponential. The prefactor
involves b /a the ratio of the natural vibration frequencies of the
molecules in the bulk and on the solid surface, as well as ea /eb the
ratio of the "probabilities of capture" when a gas molecule hits the
solid surface or the surface of the bulk liquid. One can reasonably
expect that ratios are close to unity. Moreover, we do not expect them
to be temperature-dependent. The important physical quantity that
controls the value of the exponential factor is (ub u a )/(k B T ), having
a clear physical meaning. The constant C therefore characterizes the
binding of a molecule to the surface. It has exactly the same meaning
as the constant K in Langmuirs equation.
The relation can be made even more explicit by considering the
general expression that we have found earlier: K = exp(( B
u a )/k B T ) with B = 0 the chemical potential of the liquid, in the
case of vapour adsorption. Since the chemical potential is Gibbs
energy per molecule, one can write
0 = ub + P0 vb Tsb (VIII47)
where ub , vb and sb are the molar energy, volume, and entropy in
the bulk state. Taking this into account, the expression of K can be
written as
P0 vb Tsb ub u a
   
K = exp exp (VIII48)
kB T kB T
where the similarity with Eq. VIII46 is now obvious. The relation
between the two approaches (statistical and dynamical) could be
pushed even further. However, what we have done here should suf-
fice to convince you that the two approaches are consistent with each
other, and that they actually give the same result.
VIII14 c.j. gommes, chim0698

Measuring specific surface areas

The BET model is extremely famous, mostly because it provides an


very practical way of analysing adsorption data to determine specific
surface areas. Many people actually refer to nitrogen adsorption
measurement as a "BET measurement": "lets do a BET!" I wouldnt
be surprised if half of you would end up one day calling a nitrogen
adsorption measurement a "BET". Thats fine in the long run, but in
the coming year or so I will not let you do that. BET is a model, not
an experimental method.
Anyway, the idea is that you expect molecules to occupy a given
area on a surface, so that if you know the number of adsorption sites,
you can determine the total area of the surface. So determining the
surface area is tantamount to determining the number of adsorption
sites, which we have referred to so far as A. What you measure, on
the other side, in N as a function of P/P0 . Here is the usual way to
extract A from such a datatset. From Eq. VIII42, you can write

P/P0 1 C1 P
= + (VIII49)
(1 P/P0 ) N AC AC P0

The left-hand side is something you can calculate in a straightfor-


ward way from the measured data, e.g. with a spreadsheet; we refer
to this quantity as the BET function. Equation VIII49 tells you that if
you plot the BET function against the reduced pressure, you should
obtain a straight line. Calling the intercept and the slope , the BET
constant C and the number of adsorption sites A are then calculated
as
1
C = 1+ and A = (VIII50)
C
As simple as that.
The BET function corresponding to the data in Fig. VIII9 is
shown in Fig. VIII14. The values of and obtained from a least-
square fit from P/P0 = 0.05 to P/P0 = 0.35 are C ' 79 and
A ' 53cm3 /g. The latter value corresponds to a volume in STP
conditions20 which is just a weird unit for a number of moles. The 20
Standard temperature and pressure
value 53 STP cm3 /g converts to 2.37 103 mol/g. This is the number conditions, T = 300 K, and P = 1 atm.
One mole of gas occupies 22.4 liters in
of molecules needed to cover the entire surface of the nanotubes with STP conditions.
a layer that is exactly one molecule-thick. Looking on the isotherm it-
self (Fig.VIII9), one sees that this value is the position of the plateau
in the isotherm. In principle, the value could be read directly on the
isotherm, but the BET fit is a more accurate way of doing this. If we
accept the usual figure that one nitrogen molecule occupies about
16.2 2 on a surface. The value of A coverts to 230 m2 /g. This is the
specific surface area of the nanotubes in Fig. VIII9.
adsorption on surfaces VIII15

Figure VIII14: BET plot corresponding


0.03 to the nitrogen adsorption data of Fig.

(g /cm 3 )
VIII9. The red line is the least-square
fit of the data for 0.05 P/P0 0.35 .

BE T f unction 0.02

0.01

0
0 0.5 1
P /P 0

Exercices

1. The modelling of adsorption presented here is extremely simpli-


fied. Enumerate some ways in which adsorption could in principle
differ from what is presented here, and things that we might have
neglected.

2. We have often mentioned energies in this chapter: adsorption


energies, salvation energies, etc. What is the nature of the physical
forces responsible for those energies?

3. In principle, a nitrogen molecule can adsorb in two different ways


on a surface: perpendicular to it, or flat on it. How is the chemical
potential of an adsorbed molecule modified if it can adsorb in two
different configurations? Assume that the two configurations have
the same energy. How is the adsorption isotherm modified?

4. If a solution is not very diluted, deviations from ideality have


to be considered. How is the adsorption isotherm modified if a
Margules equation is used?

5. Same question for a Van der Waals gas.

6. One classically advocates that the BET model should only be used
from P/P0 = 0.05 to P/P0 = 0.35. Figure VIII15 shows how
the BET model actually compares with the data. To your opinion,
why is there any deviation between the data and the model? The
reasons need not be the same at low and high pressure.
IX0 c.j. gommes, chim0698

Figure VIII15: Real volumes calcu-


500 lated from the BET fit in Fig. VIII14.
Can you think of reasons (possibly
60 many) for the deviation between the
400 40 data and the BET model at low and at
large pressures ? The reasons need not
V a d s (cm 3 /g ) be the same in both cases.
20
300
0
0 0.05 0.1
200

100

0
0 0.2 0.4 0.6 0.8 1
P /P 0

7. When analysing the nitrogen adsorption data in Fig. VIII9 we


found a specific surface area of 230 m2 /g. Assuming that nan-
otubes are closed, so that adsorption only takes place on their
outer surface, make a rough estimate of their diameters. (Hint: the
density of the graphitic carbon that makes up the nanotubes walls
is ' 2 g.cm3 ).

8. Can you guess the order of magnitude of the area occupied by


water molecule adsorbed on a surface.

9. Typical values for the BET constant are C ' 20 for adsorption on
polymeric materials and around C ' 200 for silica. Can you make
sense of this? What does this mean?

Langmuir (statistical mechanical approach) BET


Broekhof-de Boer Metastability.

Further reading

Exercices

Das könnte Ihnen auch gefallen