Vous êtes sur la page 1sur 49

Surface Science Reports 14 (1992) 109-159 i:!:::::: . . . . . . . . ":~::!

:~i

North-Holland surface science


reports

Adhesion forces between surfaces in liquids


and condensable vapours
Jacob N. Israelachvili
Department of Chemical and Nuclear Engineering, and Materials" Department, Unicersity of California, Santa Barbara,
CA 93106, USA

Manuscript received in final form 10 October 1991

When a liquid is confined within a highly restricted space its properties become quantitatively and
qualitatively different from the bulk (continuum) values: for example, when confined between two
surfaces the molecules of a liquid may structure into quasi-discrete layers whose properties become
"quantized" with the number of layers, and such ultra-thin films often behave more like a solid or a
liquid crystal than a normal liquid, for example, withstanding finite compressive and shear stresses. The
forces between surfaces or particles in liquids can also be very complex once the surfaces approach
closer than 5-10 molecular solvent diameters. This is the regime where continuum and mean field
theories (e.g., of monotonically attractive van der Waals forces) break down and where the interactions
become sensitive to such fine details as the molecular structure of the liquid molecules and surfaces.
These short-range forces determine the adhesion and friction between surfaces and the properties of
colloidal dispersions. During the last ten years experiments using force balance techniques such as the
surface forces apparatus, and theories employing computer simulations, have totally changed our
conception of the short-range forces in liquids. It is now known that the force-laws are rarely
monotonically attractive or repulsive: the force can change sign at some small but finite separation or it
can be an oscillatory function of separation. Some of these forces are now well understood, but the more
important ones are not. These include the repulsive "hydration" and attractive "hydrophobic" forces
between hydrophilic or hydrophobic surfaces in water, the forces between metallic surfaces, and the role
of static and dynamic roughness on adhesion. An understanding of these interactions is both of
fundamental interest as well as having immediate practical applications in controlling the properties of
materials and complex macromolecular (e.g., colloidal) systems. Here we shall review the current state
of our knowledge in this area, where to a large extent, experiment is ahead of theory.

0167-5729/92/$15.00 © 1992 - Elsevier Science Publishers B.V. All rights reserved


112 J.N. Israelachl,ili

I. Introduction

It is now known that a number of quite diverse interactions occur between two surfaces
depending on whether the interaction occurs in vacuum, in vapour or in a liquid. In practice,
it is also important to distinguish between static (i.e., equilibrium) forces and dynamic (e.g.,
frictional and other energy-dissipating) forces.
In c a c u u m there are the long-range van der Waals and electrostatic (Coulombic) forces,
while at smaller surface separations - corresponding to molecular contacts ( D ~ 0.1-0.2 nm)
- there are additional forces such as covalent, hydrogen bonding and metallic bonding forces.
All these forces determine the adhesion between bodies of different geometries, the surface
and interracial energies of planar surfaces, and the strengths of materials, grain boundaries,
cracks, and other adhesive junctions. These adhesive forces are often strong enough to
elastically or plastically deform the shapes of two bodies or particles when they come into
contact.
When exposed to lJapours (e.g., atmospheric air) two solid surfaces in or close to contact
can now have a surface layer of chemisorbed or physisorbed molecules, or a capillary
condensed liquid bridge, between them. Each of these effects can drastically modify their
adhesion. The adhesion usually falls, but in the case of capillary condensation the additional
Laplace pressure, or attractive "capillary" force, between the surfaces may make the adhesion
stronger than in inert gas or vacuum.
When totally immersed in a liquid the force between two surfaces is once again completely
modified from that in vacuum or air (vapour). The van der Waals attraction is generally
reduced, but other forces now come into play which can qualitatively change both the range
and even the sign of the interaction. Again, the attractive forces can be either stronger or
weaker than in the absence of the intervening liquid medium, e.g., stronger in the case of two
hydrophobic surfaces, but weaker for two hydrophilic surfaces interacting in water. Further-
more, the force may no longer be purely attractive. It can be repulsive, or the force can
change sign at some finite surface separation, D, and the potential energy minimum may now
occur not at molecular contact but at some small distance farther out.
Until only a few years ago it was believed that only two forces operated between two
surfaces in a liquid such as water - the attractive van der Waals force and the repulsive
electrostatic "double-layer" force. These two forces together form the basis of the well known
D e r j a g u i n - L a n d a u - V e r w e y - O v e r b e e k , or D L V O theory [1], and are depicted schematically
in fig. 1.
More recent experiments have revealed that other types of attractive (and repulsive)
short-range forces can also arise in liquids, especially at short-range, i.e., at surface separa-
tions below a few nanometers or a few molecular diameters. These forces can be extremely
varied and complex, much more so than was imagined only a few years ago. This realization is
partly due to the ever increasing complexity of the systems being studied; for example, the
liquids are no longer simple one-component liquids but can consist of a polydisperse mixture
of anisotropic polar, amphiphilic or polymeric molecules. In addition, the two surfaces
themselves can be amorphous or crystalline, crystallographically matched or not, rough or
smooth, rigid or fluid-like (soft), hydrophilic or hydrophobic. All these factors are now
recognized as being critically important in determining the strength of the adhesion in
different systems, depending both on the physical and chemical properties of the surfaces as
well as on the nature of the bathing liquid or condensable vapours in the atmosphere.
In particular, both experiments and theory have shown that conventional continuum or
"jellium" theories are generally inadequate for describing the short-range interactions be-
Adhesion forces between surfaces in liquids and condensable capours 113

ES /Force or EnergyBarrier
Repulsion [ - - ~
g o
0,)
0
0
LI_

~ - - - ~ - - ~ Net DLVO Interaction

~o
B
t-
LIJ
t-
..
Mini m
:u
-g
O

.-'~ VDWAttraction
g
¢.-

Adhesive Contact
(Primary Minimum)
I I I
0 5 10 15 20
Distance, D (nm)
Fig. 1. Classical DLVO interaction potential energy as a function of surface separation between two flat surfaces
interacting in an aqueous electrolyte (salt) solution via an attractive van der Waals force and a repulsive screened
electrostatic "double-layer" force. The double-layer potential (or force) is repulsive and roughly exponential in
distance dependence, The attractive van der Waals potential has an inverse power-law distance dependence (for
example, Wct - 1 / D between two spheres; W c x - 1 / D 2 between two flat surfaces) and it therefore "wins out" at
small separations, resulting in strong adhesion in a primary minimum contact. The inset shows a typical interaction
potential between surfaces of high charge density in dilute electrolyte solution. All curves are schematic. It is
noteworthy that the interaction energy W between two fiat surfaces is directly proportional to the force F between
two curved surfaces of radius R according to the "Derjaguin approximation": F / R = 2~-W [eq. (6)].

tween two real surfaces (i.e., atomically " s t r u c t u r e d " surfaces which are not m a t h e m a t i c a l l y
smooth) across a real liquid (which c a n n o t be t r e a t e d as a structureless c o n t i n u u m at
distances below 5 - 1 0 m o l e c u l a r diameters).
A n o t h e r i m p o r t a n t d e v e l o p m e n t in r e c e n t years is the a p p r e c i a t i o n that in m a n y practical
situations o n e is often d e a l i n g with n o n - e q u i l i b r i u m forces. For example, certain organic
liquid films c o n f i n e d b e t w e e n two a p p r o a c h i n g surfaces may take a surprisingly long time to
equilibrate, as may the surfaces themselves, so that the s h o r t - r a n g e a n d a d h e s i o n forces
a p p e a r to be t i m e - d e p e n d e n t (section 10). F u r t h e r m o r e , u n d e r dynamic, as o p p o s e d to static
or e q u i l i b r i u m , c o n d i t i o n s the c o n f o r m a t i o n or " s t r u c t u r e " of liquid molecules b e t w e e n two
solid surfaces moving relative to each o t h e r may be quite different from that w h e n the
surfaces are at rest, a n d this too can modify the a d h e s i o n force b e t w e e n them. This, of course,
is in a d d i t i o n to any purely viscous or h y d r o d y n a m i c interaction, though in molecularly thin
liquid films these two effects can be difficult to distinguish from each o t h e r since they are
often related.

2. Experimental techniques for measuring intersurface forces

T h e simplest a n d most direct way to m e a s u r e the a d h e s i o n of two solid surfaces, such as


two spheres or a sphere on a flat surface, is to s u s p e n d o n e o n a spring a n d m e a s u r e - from
114 J.N. Israelacht,ili

out
-~s i in
ID

Distance, D
Fig. 2. Schematic attractive force-law between two macroscopic objects, such as two magnets, or between two
microscopic objects such as the van der Waals force between a metal tip and a surface. On lowering the upper
support the spring will expand such that at any equilibrium separation D the attractive force balances the elastic
restoring force. However, once the gradient of the attractive force d F / d D exceeds the gradient of the spring's
restoring force, defined by the spring constant K~, the upper surface will jump into contact (point P -+ P'). On
separating, the two surfaces will jump apart from Q to Q'. The distance Q - Q ' multiplied by K~ gives the adhesion
force, i.e. the value of F at Q.

the d e f l e c t i o n of that spring - the a d h e s i o n or " p u l l - o f f " force n e e d e d to s e p a r a t e the two


bodies. Fig. 2 illustrates t h e p r i n c i p l e of this m e t h o d w h e n a p p l i e d to the i n t e r a c t i o n of two
m a g n e t s . H o w e v e r , the m e t h o d is a p p l i c a b l e even at t h e m i c r o s c o p i c or m o l e c u l a r level, and it
forms the basis of all d i r e c t force m e a s u r i n g a p p a r a t u s e s [1,2] such as the surface forces
a p p a r a t u s ( S F A ) or a t o m i c force m i c r o s c o p e ( A F M ) .
If K~ is the stiffness of t h e f o r c e - m e a s u r i n g spring a n d A D the d i s t a n c e the two surfaces
j u m p a p a r t w h e n they s e p a r a t e , t h e n the a d h e s i o n force F~ is given by

F~=Fm~x=K~ AD, (1)

w h e r e we n o t e that in liquids the m a x i m u m or m i n i m u m in the force m a y occur at s o m e


n o n - z e r o surface s e p a r a t i o n , e.g., at p o i n t Q in fig. 2 o r p o i n t s P a n d Q in fig. 8 (see later).
F r o m F~ we m a y c a l c u l a t e the surface or i n t e r f a c i a l e n e r g y y. H o w e v e r , this d e p e n d s on
the g e o m e t r y of the two b o d i e s . F o r a s p h e r e of r a d i u s R on a fiat surface or for two c r o s s e d
cylinders of r a d i u s R we have [1]

y = F,/3rrR in units of N / m or J / m 2, (2)

while for two s p h e r e s of radii R 1 a n d R 2

= + ' {3)

Figs. 3 - 6 show two types of surface forces a p p a r a t u s e s ( S F A ' s ) suitable for m a k i n g


a d h e s i o n a n d force-law m e a s u r e m e n t s b e t w e e n two curved m o l e c u l a r l y s m o o t h surfaces
i m m e r s e d in a liquid o r in v a p o u r s of c o n t r o l l a b l e v a p o u r p r e s s u r e or humidity. T h e optical
t e c h n i q u e u s e d in t h e s e m e a s u r e m e n t s e m p l o y s m u l t i p l e b e a m i n t e r f e r e n c e fringes which
allows for surface s e p a r a t i o n s D a n d A D to b e m e a s u r e d to ± 1 A. F r o m t h e s h a p e s of t h e
Adhesion forces between surfaces in liquids and condensable capours 115

~fferentofrce-measunrig
~!!
.'.specttoromet
:~ erlight spring~
rod
amp

;~ort

)le-
ring

"ing

rod
U 1,2.111 ::) I 1
I t white Light
Fig. 3. Version of the surface forces apparatus (SFA Mk II) for measuring the forces between two curved molecularly
smooth surfaces in liquids at the ~ngstr6m resolution level [2,3]. Forces are measured from the deflection of the
"variable-stiffness force-measuring spring", whose stiffness can be varied by shifting the position of the "movable
clamp" using the "clamp adjusting rod". A variety of interchangeable force-measuring springs (two shown on top) can
also be used to allow greater versatility in measuring both repulsive or attractive forces over a range of greater than
six orders of magnitude. During the past few years, this apparatus has been used to identify and quantify most of the
fundamental interactions occurring between surfaces in various liquids and vapours, such as van der Waals and
double-layer forces, solvation (hydration and hydrophobic) forces, adhesion and capillary forces and the interactions
between polymer-covered and surfactant-coated surfaces.

i n t e r f e r e n c e fringes o n e also obtains direct q u a n t i t a t i v e visualization of any surface d e f o r m a -


tions b r o u g h t a b o u t by the a d h e s i v e c o n t a c t o f the two initially c u r v e d surfaces. T h e distance
o
b e t w e e n t h e two surfaces can also be i n d e p e n d e n t l y c o n t r o l l e d to within 1 A, and the force
sensitivity is ab o u t 10 - s N (10 6 g). Thus, for the typical surface radii o f R ~ 1 cm used in
t h es e e x p e r i m e n t s , y values can be m e a s u r e d to an accuracy of a b o u t + 10 -3 m J / m 2 ( + 10 -3
erg/m2).
T h e full f o r c e - la w b e t w e e n two surfaces (i.e., the force F as a f u n c t i o n o f surface
s e p a r a t i o n D ) can be m e a s u r e d in a n u m b e r of ways. T h e simplest is to e x p a n d the
p i e z o e l e c t r i c crystal s u p p o r t i n g the u p p e r surface by a k n o w n a m o u n t , say, A D e z x. If t h e r e is
a fo rce b e t w e e n the two surfaces this will cause the f o r c e - m e a s u r i n g spring to d ef l ect by, say,
116 J.N. lsraelachHli

Fig. 4. Photo of SFA Mk If, described in fig. 3.

ADspring , while the (measured) surface separation changes by AD. These three displacements
are related by
A Dspring = A D p z T - AD. (4)
Since both ZiDez T and A D can be controlled a n d / o r directly measured, eq. (4) provides a
way of measuring the spring deflection ADspring.The force difference between the initial and
final separations is therefore given by
AF = K~ ADspring. (5)

The above method therefore allows for the force difference to be measured between any
two separations. Thus by starting at some large separation where the force is zero ( F = 0) and
working one's way in, the full force-law can be constructed by carefully measuring the force
difference between any initial (or reference) separation and the final (or desired) separation.
Various surface materials have been successfully used in such studies including mica [3,4],
silica [5] and sapphire [6]. It has also been possible to measure the forces between adsorbed
polymer layers [7], surfactant monolayers and bilayers [1,8,9], protein layers [10], and metallic
layers deposited on mica [11]. The range of liquids and vapours that can be used is almost
endless, and so far these have included aqueous solutions, organic liquids and solvents,
polymer melts, various petroleum oils and lubricant liquids, and liquid crystals.
A particularly useful aspect of making force measurements between two curved surfaces
(for example, two spheres, a sphere and a flat, or two crossed cylinders), is that the measured
Adhesion forces between surfaces in liquids and condensable capours 117

Differential Micrometer ~ Pore

5 cm Micrometer
Micrometer for
M3 Differential Spring

White Light
Microscope Tube

Clamp ~

Piezoelectric Upper
Tube - (Control)
Air Outlet Chamber

Shaft

Spring Mount
Lower
Side Port Chamber
Clamp Spring
Inlet Hole
~Hinge

Fig. 5. New surface forces apparatus (SFA Mk IIl) for measuring the forces between two molecularly smooth surfaces
[73]. Mk III employs four distance controls instead of three as in Mk II. The four controls are: micrometer (M1),
differential micrometer (M2), differential spring (M 3) and piezoelectric tube. The mica surfaces are glued to
cylindrical support disks of radius R and positioned in a crossed cylinder geometry. The lower surface is mounted on
a variable-stiffness double-cantilever force-measuring spring (S) within the lower chamber and is connected to the
upper (control) chamber via a Teflon bellows (B).

force F can be directly r e l a t e d to the e n e r g y p e r unit a r e a E (or W ) b e t w e e n two flat


surfaces. This is given by the so-called D e r j a g u i n a p p r o x i m a t i o n [1]:

E=F/27rR or F/R =2~-E, (6)

w h e r e R is t h e r a d i u s of the s p h e r e (for a s p h e r e a n d a flat) or the radii of the cylinders (for


two c r o s s e d cylinders). F o r f u r t h e r d e t a i l s o f S F A a n d r e l a t e d f o r c e - m e a s u r i n g t e c h n i q u e s , see
refs. [1-3] a n d [12].

3. Van der Waals forces

M a n y y e a r s ago T a b o r a n d W i n t e r t o n a n d o t h e r s [13] m e a s u r e d t h e attractive van d e r


W a a l s f o r c e - l a w b e t w e e n two glass o r mica surfaces down to surface s e p a r a t i o n s o f D = 15 A,
118 J.N. lsraelachvili

Fig. 6. Photo of SFA Mk III, described in fig. 5.

and confirmed the so-called "Lifshitz" theory of van der Waals forces. According to this
theory the van der Waals force between two curved surfaces of radius R is given by

F --AR/6D 2, (7)

where A is the " H a m a k e r " coefficient, which for non-metallic solids is usually of the order of
10 ~9 j. For inert non-polar surfaces, e.g., of h y d r o c a r b o n s or van der Waals solids and
liquids, the Lifshitz theory has b e e n f o u n d to apply even at molecular contact, where it can
predict the surface energies (or tensions) of these solids or liquids. Thus, for h y d r o c a r b o n
surfaces the H a m a k e r constant is typically A -- 0.5 × 10-19 j. Inserting this value into eq. (7)
and using a cut-off distance of D o = 0.2 nm when the two surfaces are in contact [1], we
obtain the following estimate for the adhesion force (for R = I cm)

FS =AR/6D 8 ..~ 2.1 × 10 - 3 N. (8)


Using the S F A with a spring constant of K S = 100 N / m , such an adhesive force will cause
the two surfaces to j u m p apart by [cf. eq. (1)]: A D = F J K S = 2.1 x 10 .5 m (21/xm), which can
be accurately measured. Further, using eq. (2), the above values give for the surface energy

y = F s / 3 ~ r R = 22 m J / m 2,

a value that is typical for h y d r o c a r b o n solids and liquids (when -/ is referred to as the surface
A d h e s i o n forces between surfaces in liquids a n d condensable capours 119

-. a::~zoo-o::D-

_ ° ~ me
~ -0.1
E
or
~. -0.2 -

LL
F/R = - A/6D 2
_ A = 2.2 x 10 .20 J
-~ -o.3
E
8
u~ -0.4

I II I I I I I I I I I I I
-0.5
5 10
D i s t a n c e , D (nm)
Fig. 7. Attractive van der Waals force F between two curved mica surfaces of radius R = 1 cm measured in water and
various aqueous electrolyte solutions. The measured nonretarded H a m a k e r constant is A = 2.2 × 10 -2° J. Retarda-
tion effects are apparent at distances above 5 nm, as expected theoretically. A g r e e m e n t with the continuum Lifshitz
theory of van der Waals forces is very good down to surface separations of 2 nm.

tension). We see, therefore, how the surface energies of van der Waals solids can be directly
measured with the SFA. The measured values are generally in very good agreement with
literature values and, furthermore, they appear to be well accounted for by the Lifshitz
theory. In section 8, experiments involving capillary condensed liquids are described that show
how the surface energies (tensions) of hydrocarbon liquids have likewise been measured; and
in section 10, the effects of time and motion on the adhesion of surfaces are described.
For such adhesion measurements to be successful, however, the surfaces must be both
atomically smooth and clean. This is not always easy to achieve, and for this reason only inert,
low-energy surfaces (such as surfactant-coated mica surfaces exposing only hydrocarbon
groups) have been used so far. Other surfaces have also been successfully used, such as bare
mica, metal, metal oxide and silica surfaces; but these are high energy surfaces, and it is
difficult to prevent them from adsorbing a monolayer of organic matter or water from the
atmosphere.
Fortunately, such contaminants that physisorb onto mica and other surfaces from the
ambient atmosphere usually dissolve away once the surfaces are immersed in a liquid, so that
the short-range forces between such surfaces can usually be measured with great reliability.
Fig. 7 shows results of measurements of the van der Waals forces between two crossed
cylindrical mica surfaces in water and various salt solutions, showing the good agreement
obtained between experiment and theory (cf., solid curve, corresponding to F = A R / 6 D z,
where A = 2.2 × 10 -2° J is the fitted value which is within about 15% of the theoretical
non-retarded H a m a k e r constant for the m i c a - w a t e r - m i c a system). Note how at larger surface
separations, above about 5 nm, the measured forces fall off faster than given by the inverse
square law. This, too, is predicted by Lifshitz theory and is known as the "retardation effect"
[1].
From fig. 7 we may conclude that at separations above about 2 nm, or 8 molecular
diameters of water, the continuum Lifshitz theory is valid. This can be interpreted to mean
that water films as thin as 2 nm may be expected to have bulk-like properties, at least as far as
their interaction forces are concerned. Similar results have been obtained with other liquids
where in general, for films thicker than 8 - 1 0 molecular diameters and sometimes only 3 - 4
diameters, their continuum properties are already manifest.
120 J.N. lsraelacht,ili

4. Effect of liquid (solvent) structure on short-range forces

When a liquid is confined within a restricted space, for example, a very thin film between
two surfaces, it ceases to behave as a structureless continuum. Likewise, the forces between
two surfaces close together in liquids can no longer be described by simple continuum
theories. Thus, at small surface separations - below about 10 molecular diameters - the van
der Waals force between two surfaces or even two solute molecules in a liquid (solvent) is no
longer a smoothly varying attraction. Instead, there now arises an additional "solvation" force
that generally oscillates with distance, varying between attraction and repulsion, with a
periodicity equal to some mean dimension of the liquid molecules. Figs. 8 and 9 show the
force-law between two mica surfaces across the silicone liquid, octamethylcyclotetrasiloxane
(OMCTS), whose inert spherical molecules have a diameter of o- = 0.85 nm [14].
The short-range oscillatory force-law, varying between attraction and repulsion with a
periodicity of G, is related to the "density distribution function" and "potential of mean
force" characteristic of intermolecular interactions in liquids. These forces arise from the
confining effect that two surfaces have on the liquid molecules between them, forcing them to
order into quasi-discrete layers which are favoured (and correspond to the energy minima)
while fractional layers are disfavoured (energy maxima). This is illustrated schematically in fig.
10. The effect is quite general and arises with all simple liquids when they are confined
between two smooth surfaces (both flat and curved) and even between two solvent molecules.

loo , , , ,
pk-pk
~:2 amplitude

F P~-Qn

5
6
0 ,,7,
1 2 3 & 5 ?7
D

~0
=Q5

~-1 ~'Q~

2L!,
~Qz

0 1 2 3 Z, 5 6 7 8 9 10
Distance, D (nm)
Fig. 8. Measured oscillatory force between two mica surfaces immersed in the liquid OMCTS, an inert liquid of
molecular diameter o- ~ 0.85 nm. The arrows indicate inward or outward jumps from unstable to stable positions: the
arrows pointing to the right indicate outward jumps from adhesive wells. The inset shows the peak-to-peak
amplitudes of the oscillations as a function of D, which have an exponential decay of decay length roughly equal to
the size of the molecules.
Adhesion forces between surfaces in liquids and condensable t,apours 121

31-rE 0.4

0.2

'FIIJ ARACT,ON -0.2


E
-0.4
L~ v

5
~
rr ~ /3 _REPULSION B
0 L£1

-5
-40
-10
0 1 2 3
-80
I I I I I I I I I
1 2 3 4 5 6 7 8 9 10
Distance, D/(~
Fig. 9. Full force-law for two mica surfaces across OMCTS liquid (see fig. 8). Also shown are the "continuum" van
der Waals force-law (dotted curves) and a theoretical computation of the same system by Henderson and
Lozada-Cassou [18] using a molecular theory (inset in B). The right-hand ordinate gives the corresponding interaction
energy per unit area for two flat surfaces, according to the Derjaguin approximation, eq. (6). It is noteworthy that for
such inert liquids the strength of the final adhesion energy (or force) at molecular contact is often accurately given by
the continuum Lifshitz theory of van der Waals forces, even though this theory fails to describe the force-law at larger
distances.

Fig. 10. Schematic illustration of the way two smooth surfaces induce or enhance the layering of liquid molecules
between them, both in the case of simple spherical molecules (left) and short linear chain molecules such as liquid
alkanes or polymer melts (right). If the surfaces are not assumed to be mathematically smooth but are also laterally
"structured" - as occurs in practice due to the atomic-scale corrugations of surfaces - then the liquid molecules may
also have lateral or "in-plane" order (in addition to the "out-of-plane" order). The properties of liquids in such
confined spaces can no longer be described in terms of their bulk/continuum properties such as their density or bulk
viscosity. Such molecularly thin films can behave more like a solid or a liquid crystal, for example, being able to
sustain a finite normal load and exhibiting a finite yield stress when sheared. To fully understand the static and
dynamic properties of such molecularly thin films one must take into account the atomic-scale structure of the liquid
molecules as well as the structure of the confining surfaces.
122 J.N. Israelachcili

Oscillatory forces do not require that there be any attractive liquid-liquid or liquid-wall
interaction. All one needs is two hard walls confining molecules whose shapes are not too
irregular and that are free to exchange with molecules in the bulk liquid reservoir. In the
absence of any attractive forces between the molecules, the bulk liquid density may be
maintained by an external hydrostatic pressure. In real liquids, attractive van der Waals forces
play the role of the external pressure, but the oscillatory forces are much the same.
Thus, oscillatory forces are now well understood theoretically, at least for simple liquids,
and a number of theoretical studies and computer simulations of various confined liquids,
including water, which interact via some form of the Lennard-Jones or Mie potentials have
invariably led to an oscillatory solvation force at surface separations below a few molecular
diameters [15-19]. For example, a theoretical computation by Henderson and Lozada-Cassou
[18] for two mica surfaces across liquid O M C T S is in good agreement with the measured force
profile, as shown in the inset to fig. 9B.
In a first approximation the oscillatory force laws may be described by an exponentially
decaying cosine function of the form
E ~ E~) c o s ( 2 ~ ' D / c r ) e -l)/~r, (9)
where both theory and experiments show that the oscillatory period and the characteristic
decay length of the envelope are close to cr [19].
It is important to note that once the solvation zones of two surfaces overlap, the mean
liquid density in the gap is no longer the same as that of the bulk liquid; and since the van der
Waals-Lifshitz interaction depends on both the refractive index and dielectric constant,
which in turn depend on the density, we must conclude that van der Waals and oscillatory
solvation forces are not strictly additive. Indeed, it is more correct to think of the solvation
force as the van der Waals force at small separations with the molecular properties and
density variations of the medium taken into account.
It is also important to appreciate that solvation forces do not arise simply because liquid
molecules tend to structure into semi-ordered layers at surfaces. They arise because of the
disruption or change of this ordering during the approach of a second surface. If there were
no change, there would be no solvation force. The two effects are of course related: the
greater the tendency towards structuring at an isolated surface, the greater the solvation force
between two such surfaces, but there is a real distinction between the two p h e n o m e n a that
should always be borne in mind.
There is a rapidly growing literature on experimental measurements and other phenomena
associated with short-range oscillatory solvation forces. The simplest systems so far investi-
gated have involved measurements of these forces between molecularly smooth surfaces in
organic liquids. Subsequent measurements of oscillatory forces between different surfaces
across both aqueous and non-aqueous liquids have revealed their subtle nature and richness
of properties [8,9,20], for example, their great sensitivity to the shape and rigidity of the
solvent molecules, to the presence of other components, and to surface structure. In
particular, the oscillations can be smeared out if the molecules are irregularly shaped, e.g.,
branched, and therefore unable to pack into ordered layers, or when surfaces are rough even
at the fingstr6m level. The main features of these short-range forces will now be summarized.

4.1. Range and magnitude of oscillatory forces in simple liquids

In simple liquids such as CCl4, benzene, toluene, cyclohexane and O M C T S whose


molecules are roughly spherical and fairly rigid, the oscillatory force dominates the interaction
Adhesion Jorces between surfaces in liquids and condensable vapours 123

3b I

:~

~
2tt
1

0
Linear alkanes

~ . . . ...........
-

~-2
2_ y

-4 i ' -
0 1 2 3 4 5 6 7
Distance, D (nm)
Fig. 11. Forces between two mica surfaces across saturated linear chain alkanes such as n-tetradecane [21]. The 0.4
nm periodicity of the oscillations indicates that tbe molecules align with their long axis preferentially parallel to the
surfaces, as shown schematically in the upper insert. In contrast, irregularly shaped alkanes such as branched
isoparaffins cannot order into well-defined layers (lower insert) and these exhibit a smooth, non-oscillatory force-law
(dashed line) [27]. Similar non-oscillatory forces are also observed between "rough" surfaces, even when these
interact across a saturated linear chain liquid. This is because the irregularly shaped surfaces now prevent the liquid
molecules from ordering in the gap. The theoretical continuum van der Waals force is shown by the dotted line.

at s e p a r a t i o n s b e l o w 5 to 10 m o l e c u l a r d i a m e t e r s and, for simple ( n o n p o l y m e r i c ) liquids,


m e r g e s into the c o n t i n u u m ( n o n - o s c i l l a t o r y ) van d e r W a a l s or D L V O force at l a r g e r
s e p a r a t i o n s . T h e p e r i o d i c i t y o f the oscillations is always close to the m e a n m o l e c u l a r d i a m e t e r
~r, a n d the oscillations d e c a y roughly e x p o n e n t i a l l y with distance, with a c h a r a c t e r i s t i c d e c a y
l e n g t h b e t w e e n 1.0~r a n d 1.5~r. N o n - s p h e r i c a l b u t inert m o l e c u l e s possessing an axis of
symmetry, such as n - a l k a n e s , exhibit similar oscillatory solvation force-laws (fig. 11). F o r such
liquids, the p e r i o d of the oscillations is a b o u t 0.4 nm which c o r r e s p o n d s to the m o l e c u l a r
width a n d i n d i c a t e s that the m o l e c u l a r axes a r e p r e f e r e n t i a l l y o r i e n t e d p a r a l l e l to the surfaces
(fig. 11, t o p inset). S i m i l a r results have b e e n o b t a i n e d with s h o r t - c h a i n e d p o l y m e r melts such
as p o l y d i m e t h y l s i l o x a n e s [21].

4.2. Effect o f oscillatory forces on adhesion energy

C o n c e r n i n g the " a d h e s i o n " of two s m o o t h surfaces in a simple liquid, a glance at fig. 8


shows t h a t the a d h e s i o n has m o r e t h a n o n e value which d e p e n d s on t h e p o t e n t i a l e n e r g y or
force m i n i m u m from which t h e surfaces a r e p u l l e d apart. W e m a y r e f e r to this as " q u a n t i z e d
a d h e s i o n " , which has b e e n o b s e r v e d in the i n t e r a c t i o n s of c e r t a i n fibres. T h e d e p t h of the
p o t e n t i a l well at c o n t a c t ( D = 0) c o r r e s p o n d s to an i n t e r a c t i o n e n e r g y that is o f t e n surpris-
ingly close to t h e value e x p e c t e d from the c o n t i n u u m Lifshitz t h e o r y of van d e r W a a l s forces.
F o r e x a m p l e , the t h e o r e t i c a l Lifshitz H a m a k e r " c o n s t a n t " for the m i c a - O M C T S - m i c a system
124 J.N. lsraelacht:ili

is about 1 . 3 5 x 1 0 -21~ J. Using eq. (8) we obtain F / R ~ A / 6 D ~ = 5 6 mN m -~ for the


adhesion energy at contact (again assuming a contact cut-off distance D 0 of.0.2 rim). This is
very close to the value of W ~ 60 mN m 1 measured for the adhesion force in the primary
adhesion minimum (fig. 9, left-hand ordinate). Note, however, that oscillatory forces generally
lead to multivalued, or "quantized", adhesion values, depending on which energy minimum
two surfaces are being separated from.

4.3. Effect of soh,ent molecular polarity (dipole moment and H-bonds)

The measured oscillatory solvation force laws for polar liquids such as acetone (dipole
moment: u = 2.85 D) are not very different from those of nonpolar liquids of similar
molecular size and shape. Similarly, in polar and hydrogen-bonding liquids of high dielectric
constant such as propylene carbonate (e = 65, u = 4.9 D), methanol (e = 33, u = 1.7 D), and
ethylene glycol (e = 41, u = 1,9 D) containing dissolved ions, the force is well described by the
continuum D L V O theory at large distances, but at smaller distances the oscillatory solvation
force dominates the interaction. It appears, therefore, that dipoles and H-bonds do not have a
large effect on the magnitude and range of oscillatory forces (though H-bonds may introduce
additional monotonic forces, discussed below).

4.4. Oscillatory forces in liquid mixtures

Christenson [22] found that the forces between two mica surfaces across a mixture of
OMCTS (o- -- 0.85 nm) and cyclohexane (or ~ 0.55 nm) are essentially the same as that of the
dominant component if its volume fraction in the mixture exceeds 90%, However, for a 50-50
mixture the oscillations are not well defined and their range is now less than for either of the
pure liquids. It appears that a mixture of differently shaped molecules cannot order into
coherent layers so that the range of the short-range structure becomes even shorter (note that
this is not the case for mixtures of homologous molecules, e.g., a mixture of alkanes or
short-chained polymers of different lengths but the same molecular width).
The case of immiscible liquid mixtures is quite different, for now one of the components is
usually preferentially attracted to the surfaces where they collect or adsorb, and this can have
a drastic effect on the force between the (now highly modified) surfaces. For example, the
presence of even trace amounts of water can have a dramatic effect on the solvation force
between two hydrophilic surfaces across a nonpolar liquid. This is because of the preferential
adsorption of water onto such surfaces, which disrupts the molecular ordering in the first few
layers. This effect usually leads to a shift of the oscillatory force curve to lower, more
adhesive, energies.

4.5. Asymmetrical and branched chain molecules

Irregularly shaped chain molecules with side-groups or branching lack a symmetry axis and
so cannot easily order into discrete layers or other ordered structure within a confined space.
In such cases the liquid film remains disordered or amorphous and the force-law is not
oscillatory but monotonic. An example of this is shown in fig. 11 for iso-octadecane where we
see how a single methyl side-group on an otherwise linear 18-carbon chain has totally
eliminated the oscillations. Similar effects occur with branched polymer melts, such as
polybutadienes [21].
Adhesion forces between surfaces in liquids and condensable t,apours 125

4.6. Liquid-solid phase transitions in thin liquid films

Fig. 11 shows the forces measured across liquid alkanes such as n-tetradecane. At large
separations ( D > 4 nm) the force is monotonically attractive and is reasonably well described
by the continuum theory of Lifshitz. But below 3 nm the oscillatory force of 0.4 nm periodicity
indicates that the molecules become ordered with their long axes preferentially aligned/ori-
ented parallel to the surfaces. Similar results have been obtained with low molecular weight
polymer melts [21] and show how force measurements can provide information on the
conformations of liquid molecules at and between two surfaces. Computer simulations of such
systems [23,24] indicate that in films thinner than 5 molecular diameters, the liquid films
actually freeze - becoming solid-like due to the applied pressure and surface force fields
acting on them from either side. Further, during transitions from, say, three layers to two
layers the film melts before it solidifies again. The knowledge that molecularly thin films
between two solid surfaces become solid-like has helped our understanding of why "liquids"
in confined spaces can sustain (withstand) finite compressit~e stresses. In section 10 we shall
see how the freezing of a thin film also leads to a finite shear stress, giving rise to frictional
forces, and also how freezing-melting transitions in thin liquid films have recently explained
the p h e n o m e n o n of stick-slip at interfaces.
As shown in fig. 11, it is often not possible to press two surfaces closer together beyond the
first or second oscillatory barrier or wail, i.e., to squeeze out the last layer or two of trapped
molecules. This is partly because the molecules have frozen, but mainly because at this stage
the compressed molecules are strongly interdigitated between the atomic corrugations (hills
and valleys) of the two confining surfaces. This locks the molecules in place and prevents
them from being pushed out regardless of how high the externally applied pressure. Indeed,
one may consider such trapped molecular layers as becoming a part of the solid, like an
intercalated layer in a crystal. As regards the adhesion of two surfaces that have trapped such
a layer between them, this will be much reduced from the adhesion in vacuum (in the absence
of the physisorbed molecules). In addition to the reduced adhesion, such surfaces also have a
much reduced friction than in vacuum.

5. Effects of surface structure on short-range forces

5.1. Effect of surface lattice

It has recently been appreciated that the structure of the confining surfaces is just as
important as the nature of the liquid for determining the solvation forces [23-25]. As we have
seen, between two surfaces that are completely smooth (or "unstructured") the liquid
molecules will be induced to order into layers, but there will be no lateral ordering within the
layers. In other words, there will be positional ordering normal but not parallel to the
surfaces. However, if the surfaces have a crystalline (periodic) lattice, this will induce ordering
parallel to the surfaces as well (fig. 10), and the oscillatory force should therefore also depend
on the structure of the surface lattices. Further, if the lattices of two opposing structured
surfaces are not in register but are at some "twist angle" relative to each other, or if the two
lattices have different dimensions ( " m i s m a t c h e d " or " i n c o m m e n s u r a t e " lattices), the oscilla-
tory force-law may be expected to be further modified.
McGuiggan and Israelachvili [26] measured the adhesion forces and interaction potentials
between two mica surfaces as a function of the orientation (twist angle) of their surface
126 J.N. lsraelacht ili

o o°

I [ I I I I I l [
-3-2-~ 0 ~ 2 ~ 4 5
Angle (deg)
Fig. 12. A d h e s i o n e n e r g y for two mica surfaces in a p r i m a r y m i n i m u m contact in w a t e r as a function of the m i s m a t c h
angle 0 b e t w e e n the two c o n t a c t i n g surface lattices. Similar p e a k s are o b t a i n e d at the o t h e r " c o i n c i d e n c e " angles:
0 = +_60% + 120 °, and 180 ° (inset).

lattices. The forces were measured in air, in water, and in an aqueous KCI solution where
oscillatory structural forces were present. In air, the adhesion was found to be relatively
independent of the twist angle 0 in the range - 10° < 0 < + 10° due to the adsorption of a 0.4
nm thick amorphous layer (of organics and water) at the interface.
The adhesion in water is shown in fig. 12. Apart from a relatively angle-independent
"baseline" adhesion, sharp adhesion peaks (energy minima) occurred at 0 = 0 °, _+60 °, _+ 120°,
and 180 °, corresponding to the "coincidence" angles of the surface lattices. As little as + 1°
away from these peaks, the energy decreases by 50%.
In aqueous KCI solution, due to potassium ion adsorption the water between the surfaces
becomes ordered, resulting in an oscillatory force profile where the adhesive minima occur at
discrete separations of about 0.25 am, corresponding to integral numbers of water layers. The
whole interaction potential was now found to depend on orientation of the surface lattices
(fig. 13), and the effect extended at least four molecular layers.
Although oscillatory forces are predicted from Monte Carlo and molecular dynamic
simulations [15-19], no theory has yet taken into account the effect of surface structure, or
atomic "corrugations", on these forces, nor any lattice mismatching effects. As shown by these
experiments, within the last one or two nanometers, these effects can alter the adhesive
minima at a given separation by a factor of two. Our results show that the force barriers, or
maxima, also depend on orientation. This could be even more important than the effects on
the minima. A high barrier could prevent two surfaces from coming closer together into a
much deeper adhesive well. Thus, the maxima can effectively contribute to determining not
only the final separation of two surfaces, but also their final adhesion. Such considerations
should be particularly important for determining the thickness and strength of intergranular
spaces in ceramics, the adhesion forces between colloidal particles in concentrated electrolyte
solutions, and the forces between two surfaces in a crack containing capillary condensed
water.
Adhesion forces between surfaces in liquids and condensable vupours 127

lO.O-

0
En

0123456
(D-W/o
“E -
2 6.0 -
ui

Angle, ~3(deg)
Fig. 13. Measured adhesion energies E,,, E,, E,, and E, versus twist angle 0 in 1mM KCI solution where the
force-law is oscillatory (inset). E, corresponds to the “primary minimum” (fig. 11, E, to the second minimum where
the surfaces are separated by an additional layer of water molecules, E, to the third minimum where the surfaces are
separated by two more water layers, etc. D, is the thickness ( - 0.3 nm) of the adsorbed layer of adsorbed potassium
ions relative to the contact position in distilled water (II = 0). The mean periodicity in the oscillations is 0.27iO.05
nm which corresponds to the diameter ( - 0.28 nm) of the water molecule.

The intervening medium profoundly influences how one surface interacts with the other.
As the results show, when the surfaces are separated by as little as 0.4 nm of an amorphous
material, such as adsorbed organics from air, then the surface granularity can be completely
masked and there is no mismatch effect on the adhesion. However, with another medium,
such as pure water which is presumably well ordered when confined between two mica
lattices, the atomic granularity is apparent and alters the adhesion forces and whole interac-
tion potential out to D > 1 nm. Thus, it is not only the surface structure but also the liquid
structure, or that of the intervening film material, that together determine the short-range
interaction and adhesion.

5.2. Effect of surface roughness

On the other hand, for surfaces that are randomly rough, the oscillatory force becomes
smoothed out and disappears altogether, to be replaced by a purely monotonic solvation
force. This occurs even if the liquid molecules themselves are perfectly capable of ordering
into layers. The situation of symmetric liquid molecules confined between rough surfaces, is
therefore not unlike that of asymmetric molecules between smooth surfaces (cf. fig. 11).
128 J.N. lsraelachuili

To summarize some of the above points, for there to be an oscillatory solvation force, the
liquid molecules must be able to be correlated over a reasonably long range. This requires
that both the liquid molecules and the surfaces have a high degree of order or symmetry. If
either is missing, so will the oscillations. A roughness of only a few ~ngstrams is often
sufficient to eliminate any oscillatory component of a force-law [27].

5.3. Effect of surface curuature and geometry

It is easy to understand how oscillatory forces arise between two flat, plane parallel
surfaces (fig. 10). Between two curved surfaces, e.g., two spheres, one might imagine the
molecular ordering and oscillatory forces to be smeared out in the same way that they are
smeared out between two randomly rough surfaces. However, this is not the case. Ordering
can occur so long as the curvature or roughness is itself regular or uniform, i.e., not random.
This interesting matter is due to the Derjaguin approximation, eq. (6), which relates the force
between two curved surfaces to the energy between two flat surfaces. If the latter is given by a
decaying oscillatory function, as in eq. (9), then the energy between two curved surfaces will
simply be the intergral of that function, and since the integral of a cosine function is another
cosine function (with some appropriate phase shift), we see why periodic oscillations will not
be smeared out simply by changing the surface curvature. Likewise, two surfaces with
regularly curved regions will also retain their oscillatory force profile, albeit modified, so long
as the corrugations are truly regular, i.e., periodic. On the other hand, surface roughness, even
on the nanometer scale, can smear out any oscillations if the roughness is random and the
liquid molecules are smaller than the size of the surface asperities.

6. Entropic (steric) forces between diffuse surfaces and interfaces

Dynamic or "fluid-like" surfaces and interfaces

If a surface or interface is not rigid but very soft or even fluid-like, this too can act to smear
out any oscillations. This is because the thermal fluctuations of such interfaces make them
dynamically " r o u g h " at any instant, even though they may be perfectly smooth on a time
average. Two common types of surfaces fall into this category: (i) solid surfaces coated with
thin organic monolayers or polymer layers, and (ii) amphiphilic m e m b r a n e surfaces in water.
Examples of the first include colloidal particles coated with surfactant monolayers, as occur
in lubricating oils, paints, toners, etc. The surfaces of these particles are typically covered with
a 1.5-2.5 nm layer of surfactant molecules. These molecules have a long hydrocarbon chain
with some functional anchoring group at one end. When this group becomes attached to a
surface, then depending on the surface coverage and temperature, the hydrocarbon chain
region will form either a solid crystalline layer, an amorphous layer, or a fluid layer. In the
latter case the forces between two such surfaces immersed in simple organic liquids are not
oscillatory - as expected for two smooth surfaces - but monotonically attractive down to
surface separations of 1-2 nm, then monotonically repulsive at smaller distances. Indeed, the
forces look very similar to those between two smooth, rigid surfaces across branched or
irregularly shaped organic liquids (fig. 11).
Examples of the second type of dynamically rough surface include free surfactant and lipid
bilayers in aqueous solutions, biological membranes, micelle and microemulsion droplets, and
o i l / w a t e r interfaces stabilized by adsorbed amphiphilic molecules or polymers, gas bubbles
Adhesion forces between surfaces in liquids and condensable z'apours 129

II i I
- Steric-Hydration ~ Frozen chains
- Repulsion ~ \~ Fluid chains
\
E -,,
0 DGDG " . . , ~.,-*~7-
v - ], ....
=o
- I" ...... " " " - 5
I VDW Attraction
LU I I
- | I
- '..,'.E i
E - It ! Water :.:::::::::::::::.::::

- ~ / Bilayer --~I~Iii ,l ~ -
-1.0 I I I
0 1 2 3 4
Interbilayer Separation (nm)
Fig. 14. Measured attractive van der Waals and repulsive steric-hydration forces in water between adsorbed bilayers
of the most common uncharged lipids of biological membranes: (i) phosphatidylcholine or "lecithin" (PC), showing
the effect of increased bilayer fluidity in enhancing the steric repulsion, (ii) phosphatidylethanolamine (PE) whose
head groups are smaller, less hydrated and less mobile than those of PC, resulting in a much reduced steric-hydration
repulsion and increased adhesion and (iii) digalactosyldiglyceride (DGDG), one of the most common lipids of plant
membranes. Note the absence of any oscillatory components in the measured forces.

stabilized by surfactant molecules, soap films and foams. T h e short-range forces between such
surfaces in water have also been found to be monotonically repulsive at small surface
separations (see fig. 14). In contrast, the short-range solvation forces in water between smooth
rigid surfaces of clays, mica, etc., are always oscillatory (see fig. 13 above, and fig. 16 below).
These very short-range repulsive forces are very effective at stabilizing the attractive van
der Waals forces at some small but finite separation, D, between 1 and 2 nm, instead of at
D = 0.2 nm as occurs when surfaces come into true molecular contact. But this small shift in
the energy m i n i m u m can have a dramatic effect on reducing the adhesion energy or force by
up to three orders of m a g n i t u d e (recall that the van der Waals force between two surfaces
varies as 1/D3). It is for this reason that fluid-like bilayers, biological membranes, emulsion
droplets in a salad dressing or gas bubbles in beer a d h e r e to each o t h e r only very weakly.
T h e r e are currently no agreed u p o n theories that account for these very short-range
repulsive forces. Because of their short-range it was, and still is, c o m m o n l y believed that they
arise from water ordering or "structuring" effects at surfaces, and that they reflected some
unique or characteristic property of water. However, it is now known that these repulsive
forces also exist in o t h e r liquids. Moreover, they a p p e a r to b e c o m e stronger with increasing
temperature, which seems unlikely for a force that originates from molecular ordering effects
at surfaces which should b e c o m e diminished with increasing temperature. It is m o r e likely
that they have an entropic origin, arising from the osmotic repulsion between exposed
thermally mobile surface groups once these overlap in a liquid. Such forces are often referred
to as " t h e r m a l fluctuation" forces, in contrast to those that arise from solvent structuring
effects which are known as "structural" forces or - if the solvent is water - " h y d r a t i o n "
forces. If these short-range repulsive forces do indeed have an entropic origin [28,29] they
would be m o r e akin to the "steric" forces associated with the interactions of polymer-covered
surfaces in liquids than to any solvation-type interaction.
131) J.N. Israelachcili

7. Forces in water and aqueous solutions

7.1. Repulsive hydration forces

The forces occurring in water and in electrolytc (salt) solutions are more complex than
those occurring in nonpolar liquids. At long-range, in addition to the attractive van der Waals
forces we now also have repulsive electrostatic "double-layer" forces (since most surfaces are
charged in water). However, according to continuum theories the attractive van der Waals
force is always expected to ultimately win out at small surface separations (fig. 1). However,
certain surfaces (usually oxide or hydroxide surfaces such as clays and silica) swell sponta-
neously or repel each other in aqueous solutions even in very high salt. Yet in all these
systems one would expect the surfaces or particles to remain in strong adhesive contact or
coagulate in a primary minimum if the only forces operating were D L V O and oscillatory
solvation forces.
There are many other aqueous systems where D L V O theory fails and where there is an
additional short-range force that is not oscillatory but smoothly varying, i.e., monotonic.
Between hydrophilic surfaces this force is exponentially repulsive and is commonly referred to
as the hydration or structural force. The origin and nature of this force has long been
controversial especially in the colloidal and biological literature. Repulsive hydration forces
are believed to arise from strongly H-bonding surface groups, such as hydrated ions or
hydroxyl ( - O H ) groups, which modify the H-bonding network of liquid water adjacent to
them. Since this network is quite extensive in range [30] the resulting interaction force is also
of relatively long range.
Repulsive hydration forces were first extensively studied between clay surfaces [31]. More
recently they have been measured in detail between mica and silica surfaces [3-5] where they
have been found to decay exponentially with decay lengths of about 1 nm. Their effective
range is about 3 - 5 nm, which is about twice the range of the oscillatory solvation force in
water. Empirically, therefore, the hydration repulsion between two hydrophilic surfaces
appears to follow the simple equation
W = + W 0 c -~/A'', (10)
where A0 = 0.6-1.1 nm for 1 : 1 electrolytes [3,4], and where W0 depends on the hydration of
the surfaces but is usually below 3 - 3 0 mJ m 2 _ higher W0 values generally being associated
with lower A0 values.
In a series of experiments to identify the factors that regulate hydration forces, Pashley [4]
found that the interaction between molecularly smooth mica surfaces in dilute electrolyte
solutions obeys the D L V O theory. However, at higher salt concentrations, specific to each
electrolyte, hydrated cations bind to the negatively charged surfaces and give rise to a
repulsive hydration force (fig. 15). This is believed to be due to the energy needed to
dehydrate the bound cations, which presumably retain some of their water of hydration on
binding. This conclusion was arrived at after noting that the strength and range of the
hydration forces increase with the known hydration numbers of the cations in the order
Mg2+> Ca2+> L i + ~ N a + > K + > Cs +.
While the hydration force between two mica surfaces is overall repulsive below about 4 nm,
it is not always monotonic below about 1.5 nm but exhibits oscillations of mean periodicity
0.25 + 0.03 nm, roughly equal to the diameter of the water molecule. This is shown in figs. 15
and 16, where we may note that the first three minima at D = 0, 0.28, and 0.56 nm occur at
negative energies, a result that rationalizes observations on clay systems: Clay platelets such
Adhesion forces between surfaces in liquids and condensable capours 131

I i 1 I I I I [ I I I I
15°1 ",I ' ' '1
1 M KCI
.._.- HydrationForce ~ lOO~ ~ -~-
13- "~ 50110-3 M l¢oo,L?] j
>~ FA ~. Y'¢.. 7
I
E
lO-1

¢-
10- 2 Lu
>~ o.1I 10-4 M
E .

II I I~ I I I ~ I I I I I I
O,Ol0 50 100
Distance, D (rim)
Fig. 15. Measured forces between charged mica surfaces in various dilute and concentrated KCI solutions. In dilute
solutions (10-SM and 10-4M) the repulsion reaches a maximum and the surfaces "jump" into molecular contact
from the tops of the "force barriers" (see fig. 1). In dilute solutions the measured forces are excellently described by
the DLVO theory, based on exact numerical solutions to the nonlinear Poisson-Boltzmann equation for the
electrostatic forces and the Lifshitz theory for the van der Waals forces (using a Hamaker constant of A = 2.2 × 10 2o
J). At higher electrolyte concentrations, as more hydrated cations adsorb onto the negatively charged surfaces, an
additional hydration force appears superimposed on the DLVO interaction. This has both an oscillatory and a
monotonic component and is shown in more detail in the inset and in fig. 16. Inset: Short-range hydration forces
between mica surfaces plotted as pressure against distance. Lower curve: force measured in 10-3M KCI solution
where there is one K + ion adsorbed per 1.0 nm2 (surfaces 40% saturated with K + ). Upper curve: force measured in
IM KCI where there is one K + ion adsorbed per 0.5 nm2 (surfaces 95% saturated with K +). At larger separations
the forces are in good agreement with the DLVO theory.

as m o t o m o r i l l o n i t e often repel each o t h e r increasingly strongly down to s e p a r a t i o n s of ~ 2


nm, b u t they also stack into stable aggregates with w a t e r interlayers of typical thickness 0.25
a n d 0.55 n m b e t w e e n t h e m [32,33]. In chemistry we would refer to such structures as stable
hydrates of fixed stoichiometry, while in physics we may think of t h e m as e x p e r i e n c i n g an
oscillatory force.
Such e x p e r i m e n t s showed that h y d r a t i o n forces can be modified or r e g u l a t e d by exchanging
ions of different hydrations on surfaces, an effect that has i m p o r t a n t practical applications in
controlling the stability of colloidal dispersions. It has long b e e n k n o w n that colloidal particles
can be p r e c i p i t a t e d ( c o a g u l a t e d or flocculated) by i n c r e a s i n g the electrolyte c o n c e n t r a t i o n -
an effect that was traditionally a t t r i b u t e d to the d e c r e a s e d s c r e e n i n g of the electrostatic
" d o u b l e - l a y e r " r e p u l s i o n b e t w e e n the particles due to the d e c r e a s e d D e b y e s c r e e n i n g length
in the solution. However, there are m a n y examples where colloids are stable, n o t at lower salt
c o n c e n t r a t i o n s , b u t at high c o n c e n t r a t i o n s . This effect is now recognized as b e i n g d u e to the
i n c r e a s e d h y d r a t i o n r e p u l s i o n e x p e r i e n c e d by c e r t a i n surfaces w h e n they b i n d highly hydrated
ions at higher salt c o n c e n t r a t i o n s . " H y d r a t i o n r e g u l a t i o n " of a d h e s i o n a n d i n t e r p a r t i c l e forces
promises to b e c o m e a n i m p o r t a n t tool for controlling various technological processes such as
clay swelling [32,34], ceramic processing a n d rheology [12,35], c o n t r o l l i n g fracture [12], a n d
c o n t r o l l i n g colloidal particle a n d b u b b l e coalescence [36].

7.2. Attractiue hydrophobic forces

W a t e r a p p e a r s to be u n i q u e in having such " h y d r a t i o n " or " s t r u c t u r a l " forces that exhibit


both an oscillatory a n d a m o n o t o n i c c o m p o n e n t . B e t w e e n hydrophilic surfaces the m o n o t o n i c
132 J.N. &raelachl'ili

• 100 'Th;or
F/R10
1O0

LI-

0.1
0 1 2 3 4
Distance, D (nm)
Fig. 16. Experimental and theoretical interaction potentials between two mica surfaces in 10 3M KCI solution, where
the concentration of K + ions bound to the surfaces is about one ion per 1 nm 2. In more dilute electrolyte solutions,
the interaction is pure D L V O (cf. fig. 15). At 10 3 M KCI and higher electrolyte concentrations more cations adsorb
(bind) to the surfaces along with their water of hydration, which causes an additional hydration force characterized by
short-range oscillations (of periodicity 0.22 to 0.26 nm, about the diameter of the water molecule) superimposed on a
longer-ranged monotonically repulsive tail. Similar results have been obtained with other electrolytes. The main
figure shows the measured force-law; the inset is a theoretical computation for the same system by Henderson and
Lozada-Cassou [18].

component is repulsive, but between hydrophobic surfaces it is attractive and the final
adhesion in water is much greater than expected from the Lifshitz theory (fig. 17).
A hydrophobic surface is one that is inert to water in the sense that it cannot bind to water
molecules via ionic or hydrogen bonds. Hydrocarbons and fluorocarbons are hydrophobic, as
is air, and the strongly attractive hydrophobic force has many important manifestations and
consequences, some of which are illustrated in fig. 18.
In recent years there has been a steady accumulation of experimental data on the
force-laws between various hydrophobic surfaces in aqueous solutions. These surfaces include
mica surfaces coated with surfactant monolayers exposing hydrocarbon or fluorocarbon
groups, or silica and mica surfaces that had been rendered hydrophobic by chemical methyla-
tion or plasma etching [37,38]. These studies have found that the hydrophobic force-law
between two macroscopic surfaces is of surprisingly long range, decaying exponentially with a
characteristic decay length of 1-2 nm in the range 0-10 nm, and then more gradually farther
out. The hydrophobic force can be far stronger than the van der Waals attraction, especially
between hydrocarbon surfaces for which the Hamaker constant is quite small.
As might be expected, the magnitude of the hydrophobic attraction falls with the decreas-
ing hydrophobicity (increasing hydrophilicity) of surfaces. Thus, Helm et al. [39] measured the
forces between uncharged but hydrated lecithin bilayers in water as a function of increasing
hydrophobicity of the bilayer surfaces. This was achieved by progressively increasing the
head-group area per amphiphilic molecule exposed to the aqueous phase, i.e., by progres-
sively exposing more of the hydrocarbon chains. The results (fig. 19) showed that with
Adhesion forces between surfaces in liquids and condensable t,apours 133

t
Force

Adhesive
Contact

0 1 3 2 4 5 6
D/(~
Fig. 17. Typical short-range solvation (hydration) forces in water as a function of distance, D, normalized by the
diameter of the water molecule, tr (about 0.25 nm). The hydration forces in water differ from those in other liquids in
that there is a monotonic component in addition to the normal purely oscillatory component. Depending on the local
density and orientation of the water molecules (or the hydrogen-bonding network) at the surfaces the monotonic
component can be attractive or repulsive and thus dominate the oscillatory component. For hydrophilic surfaces the
monotonic component is repulsive (upper dashed curve), whereas for hydrophobic surfaces it is attractive (lower
dashed curve). For simpler liquids there are no such monotonic components and both theory and experiments show
that the oscillations simply decay with distance with the maxima and minima, respectively, above and below the
baseline of the van der Waals force (middle dashed curve) or superimposed on the net DLVO interaction.

i n c r e a s i n g h y d r o p h o b i c area the forces b e c a m e progressively more attractive at longer range,


that the a d h e s i o n increased, a n d that the stabilizing repulsive s h o r t - r a n g e steric-hydration
forces decreased. This shows how the overall force curve c h a n g e s w h e n an initially hydrophilic
surface b e c o m e s progressively m o r e hydrophobic, viz., as the hydrophilic h e a d - g r o u p s b e c o m e
r e p l a c e d by the h y d r o p h o b i c h y d r o c a r b o n chains.
F o r two surfaces in water their purely h y d r o p h o b i c i n t e r a c t i o n energy (i.e., i g n o r i n g D L V O
a n d oscillatory forces) in the r a n g e 0 - 1 0 n m is given by

W= -2y i e -°/a,,, (11)

where, typically, Yi = 1 0 - 5 0 mJ m 2, a n d Ao = 1 - 2 nm.


A t a s e p a r a t i o n below 10 n m the h y d r o p h o b i c force a p p e a r s to be insensitive or only
weakly sensitive to c h a n g e s in the type a n d c o n c e n t r a t i o n of electrolyte ions in the solution.
T h e a b s e n c e of a " s c r e e n i n g " effect by ions attests to the n o n - e l e c t r o s t a t i c origin of this
i n t e r a c t i o n . In contrast, some e x p e r i m e n t s have shown that at s e p a r a t i o n s g r e a t e r t h a n 10 n m
the a t t r a c t i o n does d e p e n d o n the i n t e r v e n i n g electrolyte, a n d that in dilute solutions, or
solutions c o n t a i n i n g divalent ions, it can c o n t i n u e to exceed the van der Waals a t t r a c t i o n out
to s e p a r a t i o n s of 80 n m [37,40].
134 .L N. Israelachvili

. :.. iiili. iiii! iiji

Fig. 18. Examples of attractive hydrophobic interactions in aqueous solutions. (a) Low solubility/immiscibility; (b)
micellization; (c) dimerization and association of hydrocarbon chains; (d) protein folding; (e) strong adhesion; (f)
non-wetting of water on hydrophobic surfaces; (g) rapid coagulation of hydrophobic or surfactant-coated surfaces; (h)
hydrophobic particle attachment to rising air bubbles (basic mechanism of "froth flotation" used to separate
hydrophobic and hydrophilic particles).

The long-range nature of the hydrophobic interaction has a number of important conse-
quences. It accounts for the rapid coagulation of hydrophobic particles in water, and may also
account for the rapid folding of proteins. It also explains the ease with which water films
rupture on hydrophobic surfaces. In this the van der Waals force across the water film is
repulsive and therefore favours wetting, but this is more than offset by the attractive
hydrophobic interaction acting between the two hydrophobic phases across water. Finally,
hydrophobic forces are increasingly being implicated in the adhesion and fusion of biological
membranes (fig. 20). It is known that both osmotic and electric field stresses enhance
membrane fusion, an effect that may be due to the resulting increase in the hydrophobic area
exposed between two adjacent surfaces (figs. 20C and 20D).

7.3. Origin of hydration forces

From the previous discussions we can infer that the hydration force is not of a simple
nature, and it may be fair to say that it is probably the most important yet the least
understood of all the forces in liquids. Clearly, the very unusual properties of water are
implicated, but the nature of the surfaces is equally important. Some particle surfaces can
have their hydration forces regulated, for example, by ion exchange. Other surfaces appear to
be intrinsically hydrophilic (e.g., silica) and cannot be coagulated by changing the ionic
conditions. However, such surfaces can often be rendered hydrophobic by chemically modify-
Adhesion forces betweensurfaces'in liquids and condensable t,apours 135

+4 ~..lOO

+3 -~ lo

1 I
+2 ~ o I 2
"T
E +1 I1 D(nm)
ZE ?#It j Steric-hydration
cc I ~ '~ Repulsion
I #,l~. J L J

--1 Full / I /
Bilayer ~ iI -200[, I
h
~ -2 i i1~11 ~ -I00
I|| ...,.,,
Partia y ~ / " ~.- _l I
Depleted ~ 0 [- , , , "Y, ~I
- Bilayer
-6 -4 -2 0
Thinning of Hydrated Bilayer (rim)
I r 1 I I I I
1 2 3 4 5 6 7 8
Interbilayer Separation, D (nm)
Fig. 19. Induction of fusion between two supported lecithin bilayers in the fluid state (see inset to fig. 14) by
increasing the hydrophobic attraction between them. (*) Forces between two unstressed bilayers, showing a van der
Waals attraction beyond 2.5 nm and steric-hydration repulsion below 2.5 nm. The van der Waals attraction causes the
two surfaces to jump into "contact" from the point J at D = 4.2 nm. No fusion is observed even up to very high
compressive pressures. (©) Forces between two depleted bilayers (i.e., under tension) where the bilayers are about
15% thinner than in their unstressed equilibrium state (i.e., where each lipid exposes an additional 0.1 nm 2 of its
hydrophobic chains to the aqueous phase). The two surfaces now jump into contact from farther out: point J at
D = 6.2 nm and at F the bilayers spontaneously fuse into one bilayer (upper inset). For even thinner bilayers, the
adhesion progressively increases (see lower inset), the range and magnitude of the attractive hydrophobic force
increases and fusion now occurs more or less spontaneously as soon as the bilayers come within 1.0-2.0 nrrl of
contact.

c .... " ....

Sferic stresses

Caz Electric field stresses

Fig. 20. Effects of various stresses on bilayers and membranes which expose hydrophobic regions (dashed) that act as
adhesion or fusion sites. (A) steric stresses due to packing mismatches with embedded proteins or other lipids, (B)
ionic stresses due to asymmetric ion binding of divalent ions, (C) osmotic stresses due to water diffusion and (D)
electric field stresses.
136 J.N. lsraelachuili

ing their surface groups. For example, by heating silica to above 600°C, two surface s i l a n o l - O H
groups release a water molecule and combine to form a hydrophobic siloxane - O - group,
whence the exponentially repulsive hydration force changes into an exponentially attractive
hydrophobic force of similar decay length.
How do these exponentially decaying repulsive or attractive forces arise? Theoretical work
and computer simulations [17,18,41,42] suggest that the solvation forces in water should be
purely oscillatory, while other theoretical studies [43-48] suggest a monotonic exponential
repulsion or attraction, possibly superimposed on an oscillatory profile. The latter is consis-
tent with experimental findings, as shown in the inset to fig. 15 where it appears that the
oscillatory force is simply additive with the monotonic hydration and D L V O forces, suggesting
that these arise from essentially different mechanisms.
It is probable that the short-range hydration force between all smooth, rigid or crystalline
surfaces (e.g., mineral surfaces such as mica) has an oscillatory component. This may or may
not be superimposed on a monotonically repulsive profile due to image interactions [46]
a n d / o r to structual or H-bonding polarization interactions [43-45].
It also appears that between rough surfaces (e.g., of silica) and especially between fluid
surfaces (e.g., of lipid bilayers), the oscillations are smeared out and that any longer-ranged
structural force collapses. What remains is a much shorter-ranged steric-type repulsion, as has
so far always been observed between such fluid-like interfaces (fig. 14).
Like the repulsive hydration force, the origin of the hydrophobic force is still unknown.
Luzar et al. [48] carried out a Monte Carlo simulation of the interaction between two
hydrophobic surfaces across water at separations below 1.5 nm. They obtained a decaying
oscillatory force superimposed on a monotonically attractive curve, i.e., similar to fig. 17.
It is questionable whether the hydration or hydrophobic force should be viewed as an
ordinary type of solvation or structural force - simply reflecting the packing of the water
molecules. It is important to note that for any given positional arrangement of water
molecules, whether in the liquid or solid state, there is an almost infinite variety of ways the
H-bonds can be interconnected over three-dimensional space while satisfying the " B e r n a l -
Fowler" rules requiring two donors and two acceptors per water molecule. In other words, the
H-bonding "structure" is actually quite distinct from the "molecular" structure. It is the
energy (or entropy) associated with the H-bonding network, which extends over a much larger
region of space than the molecular correlations, that is probably at the root of the long-range
solvation interactions of water. It is clear that the situation in water is governed by much more
than the simple molecular packing effects that seem to dominate the interactions in non-
aqueous liquids.

8. Capillary forces: effect of ambient conditions

When considering the adhesion of two solid surfaces or particles in air or in a liquid, it is
easy to overlook or underestimate the important role of capillary forces, i.e., forces arising
from the Laplace pressure of curved menisci which have formed as a consequence of the
capillary condensation of a liquid in the small gap between the surfaces.
For the case of a spherical particle of radius R in contact with a flat surface the adhesion
force in an inert atmosphere is

F~ = 47rR-/sv, (12)
~
Adhesion forces between surfaces in liquids and condensable capours 137

force)

Fs = 4 = R ~ s v

!!ii!iii: ::i F~= 4,m [~',v co~e. ~s,]

iiiii! , ii,
Fig. 21. Sphere on flat in an inert atmosphere (top) and in an atmosphere containing vapour that can "capillary
condense" around the contact zone (bottom). At equilibrium the concave radius, r, of the liquid meniscus is given by
the Kelvin equation. The radius r increases with the relative vapour pressure, but for condensation to occur the
contact angle 0 must be less than 90 ° or else a concave meniscus cannot form. The presence of capillary condensed
liquid changes the adhesion force F~, as given by the two equations in the figure. Note that this change is
independent of r so long as the surfaces are perfectly smooth. Experimentally, it is found that for simple inert liquids
such as cyclohexane, these equations are valid already at Kelvin radii as small as 1 nm - about the size of the
molecules themselves. Capillary condensation also occurs in binary liquid systems, e.g., when small amounts of water
dissolved in hydrocarbon liquids condense around two contacting hydrophilic surfaces, or when a vapour cavity forms
in water around two hydrophobic surfaces.

but in an atmosphere containing a condensable vapour (fig. 21), the above becomes replaced
by

F~ = 47rR[YLV cOS 0 + YSL], (13)

where the first term is due to the Laplace pressure of the miniscus and the second is due to
the direct adhesion of the two contacting solid surfaces within the liquid.
Note that the above equation does not contain the radius of curvature, r, of the liquid
meniscus (fig. 21). This is because for smaller r the Laplace pressure YLv/r increases, but the
area over which it acts decreases by the same amount, so the two effects cancel out. A natural
question arises as to the smallest value of r for which eq. (13) will apply. Experiments with
inert liquids, such as hydrocarbons, condensing between two mica surfaces indicate that eq.
(13) is valid for values of r as small as 1-2 nm, corresponding to vapour pressures as low as
40% of saturation [49]. With water condensing from vapour or from oil it appears that the
bulk value of YLV is also applicable for meniscus radii as small as 2 nm.
However, unlike with inert liquids, the condensation of water can have very dramatic
effects on the whole physical state of the contact zone. For example, if the surfaces contain
138 J.N. lsraelachcili

A B

C D
Fig. 22. (A, B) Most likely conformation of single-chained surfactant monolayers (for which the head-group area is
larger than the hydrocarbon chain area) under dry conditions. (C) Hydrated monolayer, i.e., monolayer exposed to
humid atmosphere, showing that water molecules (black dots) have penetrated into the head-group/interracial
region. The surfactant molecules are now highly mobile and can diffuse laterally as well as flip-flop across the
monolayer, (D) Penetration of organic molecules (e.g. alkanes) into monolayer chain region.

i o n s t h e s e will d i f f u s e a n d b u i l d up w i t h i n t h e l i q u i d b r i d g e , t h e r e b y c h a n g i n g t h e c h e m i c a l
c o m p o s i t i o n o f t h e c o n t a c t z o n e as w e l l as i n f l u e n c i n g t h e a d h e s i o n , M o r e d r a m a t i c e f f e c t s
c a n o c c u r if t h e s u r f a c e s a r e c o v e r e d by a s u r f a c t a n t m o n o l a y e r . O n e x p o s u r e to h u m i d air this
c a n b e c o m e h y d r a t e d , swell a n d b e c o m e m o r e f l u i d - l i k e (fig. 22C), a n d w h e n t w o such
surfaces come into contact some of the molecules can turn over rendering the once hydropho-
bic s u r f a c e p a r t i a l l y h y d r o p h i l i c (fig. 23). W a t e r c a n c o n d e n s e a r o u n d t h e c o n t a c t z o n e a n d
t h e a d h e s i o n f o r c e will also b e a f f e c t e d - g e n e r a l l y i n c r e a s i n g w e l l a b o v e t h e v a l u e e x p e c t e d
for t w o i n e r t h y d r o p h o b i c s u r f a c e s for w h i c h YLv = 25 m N / m .

Fig. 23. Capillary condensed water at bifurcation of two monolayer-coated surfaces exposed to humid air near
saturation [74]. Aqueous regions are shown by the shaded patches. Compare the states of the hydrated monolayers in
the contact zone with their states on the isolated surfaces (fig, 22C).
Adhesion forces between surfaces in liquids and condensable t,apours 139

It is clear that the adhesion of two surfaces in vapour or a solvent can often be largely
determined by capillary forces arising from the condensation of liquid that may be present
only in very small quantities, e.g., 10% of saturation in the vapour, or 20 ppm in the solvent.

9. Other types of adhesion forces: specific interactions

9.1. Bridging forces

This review is by no means exhaustive, and in particular it does not touch upon attractive
forces occurring between polymer-covered surfaces in liquids [7]. Such forces can be particu-
larly strong if the surface coverage is below saturation whence polymer bridging between two
surfaces gives rise to an attractive "bridging" force. This can be of long range, depending on
the molecular weight of the polymer, and the attraction appears to decay exponentially with
distance. (Strongly adhesive bridging forces of a different type also arise between charged
surfaces that are "bridged" or cross-linked by divalent counterions, such as when Ca 2+
bridges two negatively charged surfaces together.) Another type of attractive force between
polymer-covered surfaces arises in " b a d solvents", when the polymer layers attract each other
and in this way pull the two surfaces or colloidal particles together, but these are rarely as
strong as bridging forces.

9.2. Ligand-receptor interactions

Van der Waals, electrostatic and hydrophobic forces are not system-specific in the sense
that the interaction potential is a known function of distance, such as a power law or
exponential function. Less studied is a specific binding mechanism between certain molecules
that have a perfect geometrical fit. Such "lock and key" or " l i g a n d - r e c e p t o r " interactions [50]
give rise to very strong physical - as opposed to covalent - bonds with minimal expenditure of
energy. Such noncovalent bonds are central to a molecular understanding of biological
recognition and molecular engineering applications. Helm et al. [51] recently studied the
interactions between Biotin ligands and Streptavidin receptors (fig. 24). This is one of the
most thoroughly studied l i g a n d - r e c e p t o r systems whose binding energy of 88 k J / m o l (about
3 5 k T per bond) is also one of the highest known. The measured force-law is also shown in fig.
24.
It is interesting to note that the Avidin-Biotin force-law is particularly featureless, viz.,
there is essentially zero force until the surfaces are within less than 0.5 nm of contact, then an
extremely strong attractive force, then a "hard-wall" repulsion at "contact". In other words,
in a first approximation, the interaction potential may be represented by a delta function
where all that matters is the strength and position of the adhesive peak, just as for a covalent
bond. Finer details, such as the weak electrostatic double-layer repulsion and finite compress-
ibility of the "hard-wall" are but small perturbations on this overall potential. It is clear that
nature has here developed an extremely efficient mechanism whereby noncovalent adhesive
junctions having the effective strength of covalent bonds can be switched on, or mechanically
"locked", quickly and with minimal expenditure of energy (no energy barrier of formation).
While the exact biological function of the Avidin-Biotin system is still not fully known, it is
likely that these types of bonds can be " u n l o c k e d " equally easily, e.g., by a change in the pH
or by light stimulation. There is no reason why similar molecular mechanisms could not be
developed for a wide range of technological applications.
140 J.N. lsraelacht,ili

INN+N t
~
: Receptor ~

REPULSION
Fp~G,/'-"AIUI[I~
J u m p / in
Ill
z
E
0

-I0
ATTRACTION
ew -20
u
L. -30
o

Jump out
-40
0 5 I0 15
Distance, D (am)
Fig. 24. Two initially asymmetric surfaces of Avidin and Biotin before and after they have locked together to form a
symmetrical l i g a n d - p r o t e i n - l i g a n d junction that is strongly and irreversibly adhesive. Dower part: Force-distance
profile between an Avidin and a Biotin surface. The adhesion force of 35 m N / m corresponds in an effective surface
energy of E ~ 7.4 m J / m 2. Given that only 5% of the area is actually involved in the binding, this value implies a local
interfacial energy of at least 150 m J / m 2.

10. Effect of time and motion on adhesion: adhesion dynamics

T h e r e has been much recent activity devoted to probing dynamic (non-equilibrium)


interactions of surfaces. These studies have provided significant new insights into how energy
is dissipated during real adhesion and debonding processes (loading-unloading cycles), and
also to our understanding of friction at the molecular level. Central to these studies is an
increased appreciation of how both the static properties (e.g., the equilibrium structure) and
the dynamic properties (e.g., the viscosity and rheology) of liquids in ultra-thin films differ
from the bulk liquid properties, and how these are related to each other. These advances will
now be reviewed - with the emphasis being on "ideal" surfaces and interfaces, i.e., surfaces
that are molecularly smooth, and interfacial films that are no more than a few molecular
layers thick.

10.1. Non-equilibrium adhesion processes

Under ideal conditions the adhesion energy is considered to be a well-defined thermody-


namic quantity. It is normally denoted by y or W, and it gives the work done on bringing two
Adhesion forces between surfaces in liquids and condensable ~,apours 141

surfaces together or the work needed to separate two surfaces from contact. Under ideal,
equilibrium conditions these two quantities are the same, but under most realistic conditions
they are not: the work needed to separate two surfaces is always greater than that originally
gained on bringing them together. An understanding of the molecular mechanisms underlying
this phenomenon is essential for understanding many adhesion phenomena, energy dissipa-
tion during loading-unloading cycles, contact angle hysteresis, and - ultimately - the
molecular mechanisms associated with many frictional processes. We start by describing both
the theoretical and experimental basis of adhesion hysteresis, and how it arises even between
perfectly smooth and chemically homogeneous surfaces.
Most real processes involving adhesion are hysteretic or energy-dissipating even though
they are usually described in terms of (ideally) reversible thermodynamic functions such as
surface energy, adhesion free energy, reversible work of adhesion, etc. For example, the
energy change, or work done, on separating two surfaces from adhesive contact is generally
not fully recoverable by bringing the two surfaces back into contact again. This may be
referred to as adhesion hysteresis, and expressed as
wR > wA
receding advancing
(separating) (approaching)

or

A W = ( W R - WA) > 0, (14)


where WR and Wa are the adhesion energies for receding (separating) and advancing
(approaching) two solid surfaces, respectively. Adhesion hysteresis is responsible for such
phenomena as "rolling" friction and elastoplastic adhesive contacts [52-54] during loading-
unloading and bonding-debonding cycles.
Hysteresis effects are also commonly observed in wetting-dewetting phenomena [55]. For
example, when a liquid spreads and then retracts from a surface the advancing contact angle
0a is generally larger than the receding angle 0 a (cf. fig. 25A). Since the contact angle, 0, is
related to the liquid-vapour surface tension, y, and the solid-liquid adhesion energy, W, by
the Dupr6 equation (fig. 25B):
(1 + cos 0)-/L = W, (15)
we may conclude that wetting hysteresis or contact angle hysteresis (0A > OR) actually implies
adhesion hysteresis, WR > WA, as given by eq. (14).
In all the above cases at least one of the surfaces is always a solid. In the case of solid-solid
contacts, the hysteresis has generally been attributed to viscoelastic bulk deformations of the
contacting materials or to plastic deformations of locally contacting asperities [52,53]. In the
case of solid-liquid contacts, hysteresis has usually been attributed to surface roughness or to
chemical heterogeneity [55] as illustrated in figs. 25C and 25D, though there have been
reports of significant hysteresis on molecularly smooth and chemically homogeneous surfaces
[561.
Here we shall focus on two more fundamental mechanisms that can give rise to hysteresis.
These may be conveniently referred to as (i) mechanical hysteresis, arising from intrinsic
mechanical irreversibility of many adhesion-decohesion processes (see fig. 25B inset, and fig.
26), and (ii) chemical hysteresis, arising from the intrinsic chemical irreversibility at the
contacting surfaces associated with the necessarily finite time it takes to go through any
adhesion-decohesion or wetting-dewetting process (fig. 25E). Henceforth we shall use the
142 .I.N. lsraelacht'ili

A .. '!

Initial equilibrium .B "- " iI


ii;ii!iiiiiiiii:~;~!i;i!!ii!ii;iiiiiii
::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::

:i:i:~:i:~:i:i:i:i:i:i:~:?i:i:i:~:~:i:i:i:i:i:~:i:i:i:i:i:i:i:~:i:i:i:.....
- " " " . . • . . " , " " " ' - " " I

iiii iiiiiiiiiiiiiiiiiiiiii!
iiiii!!!iii!!i!ili!i!iiiiiii!iiiiiiiiiiiiiiiiiiii!iiii
iiii
Advancing interface Advancing (growing) droplet
!Ci iii!iiiiiiiiiiii::iiiiiii
ii!iii!iiiii!i!iiii
ii,

•2.2.2.2 (2 2 2 2 2 ~ -_ ~.2.2
:i:::::::i:::: :' :: ' :::::::i

Receding interface
Receding (contracting) droplet
...`...'`...•'...•............................`...••......•'.•........'`.•'''.....'...'.......
........:::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::

Fig. 25. Examples of wetting and contact-angle hysteresis. (A) Solid surface in equilibrium with vapour. On wetting
the surface, the adt,ancing contact angle 0A is observed, on dewetting it decreases to the receding angle, OR. This is
an example of adhesion hysteresis during wetting-dewetting that is analogous to that occurring during the
Mading-unloading of two solid surfaces. (B) Liquid droplet resting on a flat solid surface. This is not a true
equilibrium situation: at the three-phase contact boundary the normal liquid stress, 7~. sin O, is balanced by high local
stresses on the solid which induce elastic or plastic deformations (inset) a n d / o r chemical rearrangements to relax
these stresses. (C, D) Contact angle hysteresis is usually explained by the inherent roughness (left side) or chemical
heterogeneity (right side) of surfaces. (E) Interdiffusion, interdigitation, molecular reorientations and exchange
processes at an interface may induce roughness and chemical heterogeneity even though initially both surfaces are
perfectly smooth and homogeneous.

term approach-separation t o r e f e r q u i t e g e n e r a l l y t o a n y cyclic p r o c e s s , s u c h as a d h e s i o n - d e -


c o h e s i o n , l o a d i n g - u n l o a d i n g , a d v a n c i n g - r e c e d i n g a n d w e t t i n g - d e w e t t i n g cycles.
Because of natural constraints of finite time and the finite elasticity of materials most
approach-separation cycles are thermodynamically irreversible, and therefore energy dissipat-
ing. B y t h e r m o d y n a m i c i r r e v e r s i b i l i t y w e s i m p l y m e a n t h a t o n e c a n n o t go t h r o u g h t h e
approach-separation cycle v i a a c o n t i n u o u s s e r i e s o f e q u i l i b r i u m s t a t e s b e c a u s e s o m e o f t h e s e
a r e c o n n e c t e d via s p o n t a n e o u s - a n d t h e r e f o r e t h e r m o d y n a m i c a l l y i r r e v e r s i b l e - i n s t a b i l i t i e s
Adhesion forces between surfaces in liquids" and condensable t'apours 143

~ Approach Separation~
s v~ ~ ~-~
IIIIIIIgllllllMlll IIIIIIHIIIIIIIIlU IIIIIIIIIIIIIIIIgll Ul l l
S' Adhesive
contact

Macroscopic jumps Molecular jumps (peeling)

Hi m It,o:,
5 D
IIIIIIIIIIIIIIIIIIIII *

I IHI H IIII

Repulsion

Approach
FORCE O- ~_ c
Distance, D

'y
Jump

Attraction ," ,~

lump out
o

Fig. 26. Origin of mechanical adhesion hysteresis during the approach and separation of two solid surfaces. Top: In
all realistic situations the force between two solid surfaces is never measured at the surfaces themselves, S, but at
some other point, say S', to which the force is elastically transmitted via the backing material supporting the surfaces.
Centre (left): "Magnet" analogy of two approaching surfaces, where the lower is fixed and where the other is
supported at the end of a spring of stiffness K s. Bottom: Force-distance curve for two surfaces interacting via an
attractive van der Waals-type force-law, showing the path taken by the upper surface on approach and separation. On
approach, an instability occurs at D = DA, where the surfaces spontaneously jump into "contact" at D ~- D o. On
separation, another instability occurs where the surfaces jump apart from ~ D o to D R. Centre (right): On the
molecular or atomic revel, the separation of two surfaces is accompanied by the spontaneous breaking of bonds,
which is analogous to the jump apart of two macroscopic surfaces.

o r t r a n s i t i o n s . D u r i n g s u c h t r a n s i t i o n s t h e r e is a n a b s e n c e o f m e c h a n i c a l a n d / o r c h e m i c a l
e q u i l i b r i u m . I n m a n y c a s e s t h e t w o will b e i n t i m a t e l y r e l a t e d a n d o c c u r at t h e s a m e t i m e ( a n d
t h e r e is u s u a l l y a l s o a n a b s e n c e o f t h e r m a l e q u i l i b r i u m ) , b u t t h e a b o v e d i s t i n c t i o n is
n e v e r t h e l e s s a u s e f u l o n e s i n c e t h e r e a p p e a r t o b e t w o fairly d i s t i n c t m o l e c u l a r p r o c e s s e s t h a t
give r i s e t o t h e m . T h e s e will n o w b e c o n s i d e r e d in t u r n .
144 J~N. lsraelachcili

10.2. Mechanical hysteresis

Consider two solid surfaces a distance D apart (fig. 26) interacting with each other via an
attractive potential and a hard-wall repulsion at some cut-off separation, D 0. Let the
materials of the surfaces have a bulk elastic modulus K, so that depending on the system
geometry the surfaces may be considered to be supported by a simple spring of effective
spring constant K s. When the surfaces are brought towards each other a mechanical
instability occurs at some finite separation, D A, from which the two surfaces jump sponta-
neously into contact (cf. lower part of fig. 26). This instability occurs when the gradient of the
attractive force, d F / d D , exceeds K S. Likewise, on separation from adhesive contact, there
will be a spontaneous jump apart from D 0 to D R. Separation jumps are generally greater
than approaching jumps.
Such spontaneous jumps occur at both the macroscopic and atomic levels. For example,
they ocur when two macroscopic (R = 1 cm) surfaces are brought together in surface forces
experiments; they occur when STM or A F M tips approach a flat surface [57], and they occur
when individual bonds are broken during fracture and crack propagation in solids [58,59]. But
such mechanical instabilities will not occur if the attractive forces are weak or if the backing
material supporting the surfaces is very rigid (high K or Ks). However, in many practical
cases these conditions are not met and the adhesion-decohesion cycle is inherently hysteretic
regardless of how smooth the surfaces, of how perfectly elastic the materials, and of how
slowly one surface is made to approach the other (via the supporting material).
Thus, the adhesion energy on separation from contact will generally be greater than that
on approach, and the process is unavoidably energy dissipative. It is important to note that
this irreversibility does not mean that the surfaces must become damaged or even changed in
any way, or that the molecular configuration is different at the end from what it was at the
beginning of the cycle. Energy can always be dissipated in the form of heat whenever two
surfaces or molecules impact each other.

10.3. Chemical hysteresis'

When two surfaces come into contact, the molecules at the interfaces relax a n d / o r
rearrange to a new equilibrium configuration that is different from that when the surfaces
were isolated (fig. 25E). These rearrangements may involve simple positional and orienta-
tional changes of the surface molecules, as occurs when the molecules of two homopolymer
surfaces slowly intermix by diffusion [60-62]. In more complex situations, new molecular
groups that were previously buried below the surfaces may appear and intermix at the
interface. This commonly occurs with surfaces whose molecules have both polar and nonpolar
groups, for example, copolymer surfaces, protein surfaces, and surfactant surfaces [63]. All
these effects act to enhance the adhesion or cohesion of the contacting surfaces.
What distinguishes chemical hysteresis from mechanical hysteresis is that during chemical
hysteresis the chemical groups at the surfaces are different on separation from on approach.
However, as with mechanical hysteresis, if the cycle were to be carried out infinitely slowly, it
should be reversible.

10.4. Adhesion mechanics

Modern theories of the adhesion mechanics of two contacting solid surfaces are based on
the J o h n s o n - K e n d a l l - R o b e r t s (JKR) theory [64,65]. In the J K R theory two spheres of radii
Adhesion forces between surfaces in liquids and condensable vapours 145

R 1 and R 2, bulk elastic moduli K, and surface energy y per unit area, will flatten when in
contact. The contact area will increase under an external load or force F, such that at
mechanical equilibrium the contact radius a is given by (cf. fig. 27)

a 3 = -- F + 6 ~ R y + ~ / 1 2 r r R y F + (6~-Ry) 2 (16)
K
where R = R 1 R 2 / ( R 1 + R;). Another important result of the J K R theory gives the adhesion
force or "pull off" force:
f S= - 3 ~ - R y ~ , (17)
where, by definition, the surface energy "/R, is ideally related to the reversible work of
adhesion W, by W = 2y R = 2y. Note that according to the J K R theory a finite elastic
modulus, K, while having an effect on the l o a d - a r e a curve, has no effect on the adhesion
force - an interesting and unexpected result that has nevertheless been verified experimen-
tally [1,64].
Eq. (16) is the basic equation of the J K R theory and provides the framework for analysing
the results of adhesion measurements of contacting solids ("adhesion mechanics" or "contact
mechanics" [65]) and for studying the effects of surface conditions and time on adhesion
energy hysteresis. This can be done in two ways: first, by measuring how a varies with load (cf.
fig. 27) and comparing this with eq. (16), and second, by measuring the "pull off" force and
comparing this with eq. (17).

10.5. Measurements o f adhesion hysteresis

The surface forces apparatus technique was used for measuring the adhesion of "pull off"
forces Fs, as well as the loading-unloading a - F curves of a variety of differently prepared
surfaces, and surface combinations (figs. 28 and 29) under different experimental conditions.
The pull-off method allows a m e a s u r e m e n t only of YR, while the a - F curves give both 7A and
YR- We may note that if all these processes were occurring at thermodynamic equilibrium,
then YA and "/R should be the same and equal to the well-known literature values of solid or
liquid hydrocarbon surfaces, viz. 3/= 23-31 m J / m 2 [66].
Fig. 30 shows the measured a - F curves obtained for a variety of surface-surface combina-
tions. Both the advancing (open circles) and receding (black circles) points were fitted to eq.
(16). These fits are shown by the continuous solid lines in fig. 30, and the corresponding fitted
values of 7A and YR are also shown.
As already mentioned, under ideal (thermodynamically reversible) conditions 3, should be
the same regardless of whether one is going up or down the J K R curve, as was shown in fig.
27A. This was found to be the case for two solid crystalline monolayers (fig. 30A) and almost
so for the two fluid monolayers (fig. 30B). The greatest hysteresis was found for two
amorphous monolayers (fig. 30C). However, no hysteresis was measured when an amorphous
or a fluid monolayer was brought together with a solid crystalline monolayer (fig. 30D).
Additionally, in all cases where "/R was also determined from the measured pull-off force, eq.
(17), it was found to be the same as the value determined independently from the receding
branch of the a - F curve (see legend to fig. 30).
These results provide a convincing test of the validity and inherent consistency of the two
basic J R K equations, eqs. (16) and (17). They also show that experimental pull-off forces
should, in general, be higher than given by the J K R theory, eq. (17), unless the system is truly
close to equilibrium conditions.
146 J,N. lsraelachrili

gA) R E V E I ~ S I B L E 'I Advancino


Contact I K R ('i t > O)
radius
a

Separation
(pull-off) Spontaneous

Adhesionforce: ~ 0 F

4--- Negative loads Positive loads--~


(tension) (compression)

{B]IRREVERSIBLE i ~G E
/0ranch <UR)I'I

. / rs, con,a¢,
Separation~ 4 ~ C (
(pull-off)

Fig. 27. (A) Reversible contact-radius versus load curve of nonadhesive (Hertzian) contact and adhesive (JKR)
contact under ideal conditions. No hysteresis. (B) Irreversible a - F curves and the hysteresis loops they give rise to
during an advancing-receding cycle (also commonly referred to as compression-decompression, loading-unloading
and b o n d i n g - d e b o n d i n g cycles).

The data also indicate that chain interdiffusion, interdigitation or some other molecular-
scale rearrangement occurs after two amorphous or fluid surfaces are brought into contact,
which enhances their adhesion during separation. The observation that two solid-crystalline or
a crystalline and an amorphous surface do not exhibit hysteresis is consistent with this
scenario, since only one surface needs to be frozen to prevent interdigitation from occurring
with the other. All this is illustrated schematically in fig. 29. The much reduced hysteresis
between two fluid-like monolayers probably arises from the rapidity with which the molecules
at these surfaces can disentangle (equilibrate) even as the two surfaces are being separated
(peeled apart).
It appears, therefore, that the ability of molecules or molecular groups to interdiffuse,
interdigitate a n d / o r reorient at surfaces, and especially the relaxation times of these pro-
cesses, determine the extent of adhesion hysteresis (chemical hysteresis). Little or no hystere-
sis arises between frozen, rigid surfaces since no rearrangements occur during the time course
Adhesion forces between surfaces in liquids and condensable uapours 147

Solid
Crystalline ~
(condensed) i l ~ \ \ \ \ \ \ \ \ \ \ \ \ \ - ~

DPPE 42.~2
DMPE 43,~,2
so,,0
Crystalline
(expanded) ~.\\\\\\\\\\\\\'~

Amorphous
DHDAA 75~,2
CTAB 6o#

(liquid-like) CaABS 59,~2

Fig. 28. Molecularly smooth mica surfaces onto which well-characterized surfactant monolayers were adsorbed, either
by adsorption from solution ("self-assembly") or by the Langmuir-Blodgett deposition technique. Different types of
surfactants and deposition techniques were used to provide surface-adsorbed monolayers with a wide variety of
different properties such as surface coverage and phase state (solid, liquid or amorphous). The figure shows the likely
chain configurations for monolayers in the crystalline, amorphous and fluid states (schematic). The first two phases
shown are solid, the third is glassy or amorphous and the last is liquid-like. The full n a m e s of the surfactants are:
D P P E (di-palmitoyl-phosphatidyl-ethanolamine), D M P E (di-myristoyl-phosphatidyl-ethanolamine), D H D A A (di-
hexadecyl-dimethyl-ammonium-acetate), C T A B (cetyl-dimethyl-ammonium-bromide), CaABS (calcium-alkyl-ben-
zene-sulphonate). The values next to each surfactant gives its molecular area in the monolayer.

(A) (8)

(C)

Fig. 29. Schematics of likely chain interdigitations occurring after two surfaces have been brought into contact. (A)
Both surfaces in the solid crystalline state - no interdigitation. (B) Both surfaces amorphous or fluid - interdigitation
(entanglements) and disentanglements occur slowly for two amorphous surfaces and rapidly for two fluid surfaces. If
the surfaces are separated sufficiently quickly, the effective molecular areas being separated from each other will be
greater than the " a p p a r e n t " area, and the receding adhesion will be greater than the advancing adhesion. (C) One
surface solid crystalline, the other amorphous or fluid - no interdigitation.
148 J.N. lsraelachrili

DMPE (43A2) CaABS (59,~,2)


2 "- DMPE (43,~2) CaABS (59~.2) '
E
v
::::L • " ~ ~
O

E9
£3 A B
,<
n" 0 I~ I I I I I I 0 I I I I I I I I
F-- -5 0 5 10 15 -10 10 20 30 40
O F(mN) F(mN)
'<
I-- 6
Z CTAB (60~,2)
O 5 CTAB (60,~2) ~ O ~ 1 CTAB(60A2) ~
E . F- ' fX v

~a
e~'2
JJ + G

1 C
0 I L I I I 0 I I L I I I I
-20 -10 0 10 20 30 40 -10 0 10 20 30
F(mN) F(mN)
APPLIED LOAD
Fig. 30. Measured advancing and receding a - F curves at 25°C for tour surface combinations. The solid lines are
based on fitting the advancing and receding branches to the JKR theory, eq. (16), from which the indicated values of
3,A and 3,R were determined. At the end of each cycle the pull-off force was measured, For the four cases shown here
the following adhesion energy values were obtained based on eq. (17): (A) crystalline on crystalline: 3,R = 28 m J / m ~,
(B) fluid on fluid: 3,R = 36 m J / m 2, (C) amorphous on amorphous: 3,R = 44-76 m J / m 2, (D) amorphous on crystalline:
3,R = 32 m J / m = . The equilibrium (literature) values for 3' are in the range 23-31 m J / m 2.

of a typical loading-unloading cycle. Liquid-like surfaces are likewise not hysteretic, but now
because the molecular rearrangements can occur faster than the loading-unloading rates.
Amorphous surfaces, being somewhat in between these two extremes are particularly prone to
being hysteretic because their molecular relaxation times can be comparable to loading-un-
loading times (presumably the time for the bifurcation front to traverse some molecular scale
length).
If this interpretation is correct it shows that very significant hysteresis effects can arise
purely from surface effects, which would be in addition to any contribution from bulk
viscoelastic effects. The former involves molecular interdigitations that need not go much
deeper than a few hngstr6ms from an interface.

10.6. Effects of contact time, loading-unloading rates, temperature and capillary condensation
on adhesion hysteresis

The adhesion energy as determined from the pull-off force generally increased with the
contact time for all the surfaces studied. This is shown in fig. 31 for two amorphous
monolayers of CTAB for which the effect was most pronounced. Notice how the hysteresis
increases as the more the monolayer goes into the amorphous, glassy state (15°C) and
disappears once it is heated to above its chain-melting temperature (35°C).
Adhesion/orces between surfaces in liquids and condensable t'apours 149

2.0 i i i i i i i i i

CTAB
1.8-
LLI D
r-
._o 1.8 D-- ~ 150C
d~
t.-

1.4 25oc

1.2
E
0 1.0 F "
t". I -" 35°C
z
0.8 r I I I I I I I I I
0 2 4 6 8
C o n t a c t T i m e (min)
Fig. 31. Effect of contact time on the normalized adhesion energy of two CTAB (amorphous) monolayers at different
temperatures (at 35°C CTAB is in the liquid state - cf. fig. 28).

Similarly, the rate at which two surfaces were loaded or unloaded also affected their
adhesion energy. Again, for two amorphous CTAB surfaces, fig. 32 shows that on slowing
down the loading-unloading rate, the hysteresis loop becomes smaller and that both the
advancing and receding energies, YA and YR, approach the equilibrium value.
One should note that decreasing the unloading or peeling rate may sometimes act to
increase the adhesion, since by decreasing the peeling rate one also allows the surfaces to
remain longer in contact. In fig. 32, the surfaces were first allowed to remain in contact for
longer than was needed for the interdigitation processes to be complete (as ascertained from
the contact time measurements of fig. 31).
Fig. 33 shows that when liquid hydrocarbon vapour is introduced into the chamber and
allowed to capillary condense around the contact zone, all hysteresis effects disappear. This
again shows that by fluidizing the monolayers they can now equilibrate sufficiently fast to be

I I I I I I I I I l I

5
CTAB

3 L~IOOP
~ / ~ -P)
2-
O
O
1- (1 min p - p )

0 I I I I I I I I I
-20 -10 0 10 2O 3O 40
Applied Load, F (mN)
Fig. 32. Effect of advancing-receding rates on a - F curves for two CTAB monolayers at 25°C. By fitting the data
points to eq. (16) the following values were obtained. For the fast loop (1 rain between data points): YA= 20 mJ/m 2,
YR = 50 mJ/m 2. For the slow loop (5 min between points): YA= 24 mJ/m 2, YR = 44 mJ/m 2.
150 J.N. lsraelachHli

Surfaces exposed Surfaces exposed to


to inert dry air saturated hydrocarbon vapour
6 A
5
CTAB (60~2) Dodecane
2
E4 E
=L
"*O 3
2-
"- 1

1 F-
0
-20
I
-10 0
I
10
I I
20
I
30 40
0
--10
I I 1 I I I 1 i~
0 10 20 30
F (mN) F(mN)

_ DHDAA (75,~2) Dodecane

f
co
DHDAA (75,~2) "5~
F,
< ~'2
rc

I- m 1
Z
0 e~'l f
(..)

I I I I I ~0 ~ 1 I I I I I
-10 10 20 30 -10 0 10 20 30
F(mN) F (mN)

CaABS (59A2)
o
C Decane
3 3-
CaABS (~9~2) ~
Ezt E
v 2 ~. 2v
o o
0
,r-

03
4,: %1

-10
l I
10
I
20
I I I
30
I
40 -10
!
0 10 20 30
I
40
F(mN) F(mN)
APPLIED LOAD
Fig. 33. Disappearance of adhesion hysteresis on exposure of monolayers to various organic vapours. The adhesion
energies as measured by the pull-off forces were between 18 and 21 m J / m 2 for the three systems shown.

considered always at equilibrium (like a true liquid). Such effects may be expected to occur
with other surfaces as well, so long as the vapour condenses as a liquid that also wets the
surfaces.
The above results show that the adhesion of two molecularly smooth and chemically
homogeneous surfaces can be hysteretic due to structural and chemical changes occurring at
the molecular (or even ~ngstr6m) level. Adhesion hysteresis increases with:
(i) the ability of the molecular groups at the surfaces to reorient and interdiffuse across the
contact interface, which is often determined by the phase state of the surface molecules;
(ii) the time two surfaces remain in contact and the externally applied load during this time;
and
(iii) the rate of approach and separation (or peeling) of the surfaces.
Adhesion forces between surfaces in liquids and condensable aapours 151

For the molecularly smooth surfaces studied here, it appears that chemical hysteresis is far
more important than mechanical hysteresis. The results also question traditional explanations
of hysteresis based purely on the static surface roughness and chemical heterogeneity of
surfaces (cf. fig. 25), and focus more on the dynamics of these effects.

11. Adhesion and friction

11.1. Properties of liquids in very thin interfacial films between two surfaces

When a liquid is confined between two surfaces or within any narrow space whose
dimensions are less than 5 to 10 molecular diameters, not only the static but also the dynamic
properties of the liquid can no longer be described even qualitatively in terms of the bulk
properties. We have already seen that molecules confined within such molecularly thin liquid
films become ordered into layers ("out-of-plane" ordering), and they can also have lateral
order within each layer ("in-plane" ordering), as was illustrated in fig. 10. Such films appear
to be more solid-like than liquid-like, at least when the surfaces are separated by discrete
multiples of the diameter of the confined molecules. However, during transitions from n
layers to n + 1 layers the films are believed to undergo a transient melting transition
[23,24,67], viz. there is an overall solid to liquid to solid transition when the surface separation
changes by one molecular layer. This scenario helps understand what happens when two
surfaces move laterally (or slide) past each other, as occurs during shear and frictional motion.
Work has also recently been done on the dynamic, e.g., viscous or shear, forces associated
with molecularly thin films. Both experiments [67-71] and theory [23,24] indicate that even
when two surfaces are in steady-state sliding they still prefer to remain in one of their stable
potential energy minima, i.e., a sheared film of liquid can retain its basic layered structure,
though the time scales of molecular hops and the in-plane ordering are modified. Thus, even
during motion the film does not regain its totally liquid-like state. Indeed, such films exhibit
yield-points before they begin to flow. They can therefore sustain a finite shear stress, in
addition to a finite normal stress. The value of the yield stress depends on the number of
layers comprising the film and represents another "quantized" property of molecularly thin
films.
The dynamic properties of a liquid film undergoing shear are very complex. Depending on
whether the film is more liquid-like or solid-like, the motion will be smooth or of the
"stick-slip" type - the latter exhibiting yield-points a n d / o r periodic "serrations" characteris-
tic of the stress-strain behaviour of ductile solids. During sliding, transitions can occur
between n solid-like layers and n - 1 layers or n + 1 layers, and the details of the motion
depend critically on the externally applied load, the temperature, the sliding velocity, the twist
angle, and the sliding direction relative to the surface lattices.

11.2. Molecular events within a thin interracial film during shear

H e r e we briefly review recent results on the shear properties of simple molecules in thin
films and how these are related to changes in their molecular configurations. These have been
studied using the A F M technique [72] or a sliding attachment for use with the surface forces
apparatus technique (fig. 34). Fig. 35 shows typical results for the friction measured as a
function of time (after c o m m e n c e m e n t of sliding) between two mica surfaces separated by
152 J.N. lsraelachcili

Driving Velocity, v

~ Translation
Stage
To chart
Springs recorder

Strain Gauges
~

Cylindrical
Disks
IL

S = F/A
K
F ----~ P=L/A
x----
:.:.:.'.:. area A
!!!~!~ ~ii:!~!: ~i~i!~!~ ~I"..:E~!ii!i:iiii E!!!I E{ •

i
Fig. 34. Schematic drawing of the sliding attachment for use with the surface forces apparatus (fig. 3),

60~ h I b ?,~ I I V._.~____~.__, I I

50
n=l
40 / V, , A/I ~T~" FS

3ol- v r~-:.i~ .o.f


o 20 n:2 #

10 k .4,,.kq)'~'7Sc, ,~l'~,-n,i~tt~ j "


^ '/ equal
o/ ~ i ;~l i i i i
o I 2
Time (minutes)
Fig. 35. Measured change in friction during interlayer transitions of the silicone liquid octamethylcyclotetrasiloxane
(OMCTS, an inert liquid whose quasi-spherical molecules have a diameter of 0.85 nm). In this system, the shear
stress (friction force/contact area) was found to be constant so long as the number of layers n remained constant.
Qualitatively similar results have been obtained with other simple liquids such as cyclohexane. The shear stresses are
only weakly dependent on the sliding velocity c. However, for sliding velocities above some critical value co, the
stick-slip component disappears and the sliding becomes smooth or "steady" at the kinetic value, F k.
Adhesion forces between surfaces in liquids and condensable capours 153

Applied stress stress

(a) A T R E S T ~ (b) ST I C K I N G =:~ (c) SLIPPING


(whale film melts)
stress ]m stress ]p

IIIIIIIIIIm mlmIB

(c') SLIPPING ~ (c") SLIPPING c=~ (d) R E F R E E Z l N G


(one layer melts) (interlayerslip)
Fig. 36. Schematic illustration of molecular rearrangements occurring in a molecularly thin film of spherical or simple
chain molecules between two solid surfaces during shear. Note that, depending on the system, a number of different
molecular configurations within the film are possible during slipping and sliding, shown here as stages (c) total
disorder, whole film melts, (c') partial disorder, part of film melts and (c") order persists even during sliding with slip
occurring at a single slip-plane either within the film or at the film/substrate interface (or even more locally within
the film). The configurations of branched-chained molecules is much less ordered (more entangled) and remains
amorphous during sliding, leading to reduced friction and smoother sliding with little or no stick-slip.

n = 3 molecular layers of the inert liquid O M C T S , and how the stick-slip friction increases to
higher values in a quantized way when the n u m b e r of layers falls to n = 2 and then to n = 1.
With the m u c h a d d e d insights provided by recent c o m p u t e r simulations of such systems
[24,25] a n u m b e r of distinct regimes can be identified during the stick-slip sliding that is
characteristic of such films, shown in fig. 36 (a) to (d).
Surfaces at rest - fig. 36 (a): With no externally applied shear force, s o l v e n t - s u r f a c e
interactions induce the liquid molecules in the film to adopt solid-like ordering. Thus at rest
the surfaces are stuck to each other t h r o u g h the film.
Sticking regime (frozen, solid-like film) - fig. 36 (b): A progressively increasing lateral shear
stress 7, is applied. This causes a small increase in the lateral displacement, x, and film
thickness D, but only by a small fraction of the lattice spacing or molecular dimension, or. In
this regime the film retains its solid-like " f r o z e n " state - all the strains are elastic and
reversible, and the surfaces remain effectively stuck to each other.
Slipping and sliding regimes (melted, liquid-like film) - fig. 36 (c), (c'), (c"): W h e n the
applied shear stress has r e a c h e d a certain critical value, the film suddenly melts (known as
" s h e a r melting") and the two surfaces begin to slip rapidly past each other as the "static shear
stress" (7 s) or "static friction force" ( F s) has been reached (in the language of materials
science the " u p p e r yield-point" has been reached).
If the applied stress ~- is kept at a constant value, the u p p e r surface will continue to slide
indefinitely once it has settled down to some constant velocity. Even if the shear stress is
r e d u c e d below 7s, steady-state sliding will continue so long as it remains above the kinetic
shear stress r k (or the kinetic friction force F k, or the " l o w e r yield-point"). T h e experimental
observation that the static and dynamic stresses are different suggests that during steady-state
sliding the configuration of the molecules within the film is almost certainly different from
that during the slip. Experiments with simple linear chain (alkane) molecules [70] show that
154 J.N, lsraelacht,ili

the film thickness remains quantized during sliding, so that the structure of such films is
probably more like that of a nematic liquid crystal where the liquid molecules have become
shear aligned in some direction enabling shear motion to occur while retaining some order
within the film.
Computer simulations for simple spherical molecules by Thompson and Robbins [24]
indicate that during the slip, the film thickness, D, is roughly 15% higher (i.e., the film density
falls), and the order p a r a m e t e r drops from 0.85 to about 0.25. Both of these are consistent
with a disorganized liquid-like state for the whole film during the slip, as illustrated
schematically in fig. 36 (c). At this stage, we can only speculate on other possible configura-
tions of molecules in the sliding and slipping regimes. This probably depends on the shapes of
the molecules (e.g., whether spherical or linear or branched), on the atomic structure of the
surfaces, on the sliding velocity, etc. Fig. 36 (c), (c') and (c") show three possible sliding
modes wherein the molecules within the shearing film either totally melt, or where movement
occurs only within one or two layers that have melted while the others remain frozen, or
where slip occurs between two or more totally frozen layers. Other sliding modes, for
example, involving the movement of dislocations or disclinations are also possible, and it is
unlikely that one single mechanism applies in all cases.
Freezing transitions - fig. 36 (d): The slipping or sliding regime ends once the applied shear
stress falls below ~'k, when the film freezes and the surfaces become stuck once again. The
freezing of a whole film can occur very rapidly. Depending on the system, freezing can occur
after prolonged sliding or immediately after the slip. The latter case is particularly common
whenever the stress is applied not directly at the two surfaces but transmitted through the
material on either side of the surfaces. In such cases the stress on the surfaces relaxes
elastically during the slip. If the slip is rapid enough and if the molecules in the film can
freeze quickly, the slip will be immediately followed by a stick; and if the externally applied
stress is maintained the system will go into a continuous "stick-slip" cycle. On the other
hand, if the slip mechanism is slow the surfaces will continue to slide smoothly and there will
be no stick-slip. However, unless the liquid molecules are highly entangled or irregular in
shape, there will always be a single stick-slip "spike" on starting. This is known as "stiction"
(fig. 37), and it can be a serious cause of damage when two surfaces start moving from rest.
A novel interpretation of the well known p h e n o m e n o n of decreasing coefficient of friction
with increasing sliding velocity has been proposed by Thompson and Robbins [24] based on
their computer simulation which essentially reproduced the above scenario. This postulates
that it is not the friction that changes with sliding speed ~, but rather the time various parts of
the system spend in the sticking and sliding modes. In other words, at any instant during
sliding, the friction at any local region is always F~ or F k, corresponding to the "static" or
"kinetic" values. The measured frictional force, however, is the sum of all these discrete
values averaged over the whole contact area. Since as t; increases each local region spends
more time in the sliding regime ( F k) and less in the sticking regime (F,) the overall friction
coefficient falls. Above a certain critical velocity L,c the stick-slip totally disappears and
sliding proceeds at the kinetic value.

11.3. Shearing experiments with different types of liquids

The above scenario is already quite complicated, and yet this is the situation for the
simplest type of experimental system. The factors that appear to determine the critical
velocity L,e depend on the type of liquid between the surfaces (as well as on the surface lattice
structure). Small spherical molecules such as cyclohexane and O M C T S have been found to
Adhesion forces between surfaces in liquids and condensable l,apours 155

Starting spike
(stiction)

stopping spike

orFricti°n ~!iiii~!iii~!iiii
forCeshear
force iiiiiiiiiiiii~iii!iii:ii!iii~iiiiliiiiii
:::5::::::5::::::::::::5::::::::::::::::::::;:5::::::::::::::::::::::5::::::::::5::::::::::::',:::::::::5::::::::::::::::::5:::::-:-:-.-.,.......-.......-,-,.......,,-.-
,.....-........,...
:-:-:.: ,.-.-.-.,...,.,...,..........-,........-.-.....,...-....~..,...-,-,.r
:.:.:.~.:.:.:.:.:.~.:.:.:.:.:.~.:.:.:~:.:.~.:`:.:.:.:.:.:.:.:.:.~.:.:c.:.:.:.~.:' ....,,,-.-,-.....-.-...,... ,-.-.-.,.........-......,-................r ...-.-.....,.-,.,
:~:.:.:.:.:.:.:.~+:.:.:.:.:.:.~.:.:.:.~.:~:.:.:.:.:.:.~.::~:.~.:.:.:~:.:.~.:.~.:.:.:.~.
;.:.:,:.:.:.:.;,:.:.:.:.:.: :.:.:.:.:.:.:.:.:.:.:.:.:.:.:.:.:.•.:.:.•.:.•.:.:.:•:.:.:.:.:.:.•.:.:.:.:•:.:.:.•.:.:.•.:.•.:.:.:.:.:.:.:.:.:.•.:.•.:.:•:.:.:.:.:.:.:':.:.•`:.:.:.:.:.:
•ii•i!i!!!!iiiii•!i!!i!iiii!•!•!!i!iiiiii!•i!i!iiiiii!•!!i!iiii!i•••!iiiiii!i!!i!i!iiiiii•!i!i!iiii!i•ii!i!iiii!i•!i!iiiiii!•••i•iiiiii!i•!i!iiiiii!ii•i!iiii!i#iiiiii!i•i•!iiii:
~ii!i~:~i~ii:!i~i~i:ii!i!:~!i!ii:!:~i!i:i:.i!i:~i~i:i:i!:.~!i:i:i~i~:i:i:ii:!:~i!i!i~i~:~!i!i:.~:.~:~i!i!ii:~i~:~!i!i~:~:~:~i!i!U:~!i!i!iU:i!i~:~!:~i~i:i:
start Time ~- stop
sliding sliding
Fig. 37. " S t i c t i o n " is t h e h i g h s t a r t i n g f r i c t i o n f o r c e e x p e r i e n c e d by t w o s u r f a c e s w h i c h c a u s e s t h e m to j e r k f o r w a r d
r a t h e r t h a n a c c e l e r a t e s m o o t h l y f r o m rest. It is a m a j o r c a u s e o f s u r f a c e d a m a g e a n d e r o s i o n . T h e f i g u r e s h o w s a
" s t i c t i o n s p i k e " o r " s t a r t i n g s p i k e " as w e l l as " s t o p p i n g s p i k e " . T h e l a t t e r o c c u r s w h e n t w o s l i d i n g s u r f a c e s a r e
b r o u g h t to r e s t o v e r a f i n i t e t i m e d u r i n g w h i c h t h e m o l e c u l e s in t h e film c a n f r e e z e a n d stick b e f o r e t h e s u r f a c e s h a v e
stopped moving.

have very high u~, which indicates that these molecules can rearrange relatively quickly in thin
films. Chain molecules and especially branched chain molecules have been found to have
much lower u~, which is to be expected, and such liquids tend to slide smoothly rather than in
a stick-slip fashion. However, the values of l'c also depend on the number of liquid layers
comprising the film, the structure and relative orientation of the two surface lattices, the
externally applied load, and of course on the stiffness of the spring (and in practice of the
material of the surfaces). With more asymmetric molecules, such as branched isoparaffins and
polymer melts, no regular spikes or stick-slip behaviour occurs at any speed since these
molecules can never order themselves sufficiently to "solidify". Examples of such liquids are
perfluoropolyethers and polydimethylsiloxanes (PDMS).
The shear properties of seven different types of organic and polymeric liquids are listed in
table 1, together with the type of sliding observed, the friction coefficient, and the bulk
viscosity of the liquids (given for reference purposes). From the data of table 1 (top part) it
appears that there is a direct correlation between the shapes of molecules and their
coefficient of friction. Small spherical or chain molecules have high friction with stick-slip
because they can pack into ordered solid-like layers, whereas longer chained and branched
molecules give low friction and smoother sliding.
It is interesting to note that the friction coefficient generally decreases as the bulk viscosity
of the liquids increases. This is because the factors that are conducive to low friction are
generally conducive to high viscosity. Thus, molecules with side-groups such as branched
alkanes and polymer melts usually have higher bulk viscosities than their linear homologues
for obvious reasons. However, in thin films the linear molecules have higher shear stresses. It
is probably for this reason that branched liquid molecules are better lubricants - being more
disordered in thin films because of this branching. In this respect it is important to note that if
an "effective" viscosity were to be calculated for the liquids of table 1, the values would be
106 to 100 times the bulk viscosities (106 for cyclohexane, 100 for PBD). This indicates that
the bulk viscosity plays no direct role in determining the frictional forces in such ultra-thin
films, at least at low shear rates. However, the bulk viscosity should give an indication of the
156 J.N. lsraelachvili

Table 1
Tribological characteristics of some liquids and polymer melts in molecularly thin films between two shearing mica
surfaces (note that a low friction coefficient is generally associated with a high bulk viscosity)
Liquid (dry) Shorbrange Type of Friction Bulk viscosity
force sliding coefficient (cP)
Spherical molecules a)
Cyclohexane (o- = 5/k) Adhesive Stick-slip >> 1 (quantized) 0.6
OMCTS (o- = 9 A) Adhesive Stick-slip >> 1 (quantized) 2.3
Chain molecules ~
Octane Adhesive Stick-slip 1.5 0.5
Tetradecane Adhesive Stick-slip 1.0 2.3
Octadecane (branched) Repulsive (Stick-slip) 0.35 5.5
PDMS (M = 3700, melt) Repulsive Smooth 0.42 50
PBD (M = 3500, branched) Repulsive Smooth 0.03 800
Water
Water (KCI solution) Repulsive Smooth 0.01-0.03 1
Hydrocarbon liquids ( w e t ) A d h e s i vhi
e Smooth 0.03 ~1
a) PDMS: polydimethylsiloxane;PBD: polybutadiene; OMCTS: octamethylcyclotetrasiloxane.
b) The strong adhesion between two hydrophilic mica surfaces in wet hydrocarbon liquid is due to capillary forces,
i.e., to the resolved Laplace pressure within the condensed water bridging the two surfaces. The direct force
between the two surfaces across the liquid (water) is actually repulsive.

lowest possible viscosity that might be a t t a i n e d in such films. Based o n this hypothesis we may
surmise that friction coefficients as low as 1 0 - 4 - 1 0 3 might be a t t a i n a b l e with the right
system.
T h e only exception to the above c o r r e l a t i o n s is water, which has b e e n f o u n d to exhibit both
low viscosity and low friction [68,71], yet water is essentially a small molecule. I n addition, the
p r e s e n c e of water can drastically lower the friction a n d e l i m i n a t e the s t i c k - s l i p of hydrocar-
b o n liquids w h e n the sliding surfaces are hydrophilic. O n the o t h e r h a n d , we have n o t e d that
with c e r t a i n (hydrophobic) s u r f a c t a n t - c o a t e d m o n o l a y e r surfaces and p o l y m e r melts the
p r e s e n c e of water can act very differently, e.g., e n h a n c i n g stick-slip. However, the results with
o t h e r surfaces are too few a n d too p r e l i m i n a r y to allow us to draw any g e n e r a l conclusions
a b o u t the tribological role of water at this stage.

12. C o n c l u d i n g remarks

T h e s h o r t - r a n g e ( < 2 n m ) forces b e t w e e n surfaces across liquids can be very complex.


Below a b o u t t e n m o l e c u l a r d i a m e t e r s c o n t i n u u m theories often b r e a k down a n d the forces,
such as the a d h e s i o n force, are d e t e r m i n e d by the m o l e c u l a r structure of the liquid molecules
a n d the s t r u c t u r e of the c o n f i n i n g surfaces. Such i m p o r t a n t forces as repulsive h y d r a t i o n and
attractive h y d r o p h o b i c forces are still not u n d e r s t o o d , while only recently has progress b e e n
m a d e at u n d e r s t a n d i n g how the simplest types of i n t e r a c t i o n s b e c o m e modified w h e n surfaces
are in shear m o t i o n relative to each other. W e may anticipate m a n y new advances in the next
few years in this area, b o t h e x p e r i m e n t a l (employing SFA, S T M a n d A F M m e a s u r e m e n t s )
a n d theoretical ( e m p l o y i n g c o m p u t e r e x p e r i m e n t s such as m o l e c u l a r dynamics simulations).
Adhesion forces between surfaces in liquids and condensable t,apours 157

Acknowledgment

I thank the Department of Energy for financial support under DOE grant number'
DE-FG03-87ER45331, though this support does not constitute an endorsement by DOE of
t h e v i e w s e x p r e s s e d in t h i s a r t i c l e .

References

[1] J.N. Israelachvili, Intermolecular and Surface Forces: With Applications to Colloidal and Biological Systems
(Academic Press, New York, 1985) (2nd ed., 1991).
[2] J.N. Israelachvili, Chemtracts - Anal. Phys. Chem. 1 (1989) 1.
[3] J.N. Israelachvili and G.E. Adams, J. Chem. Soc. Faraday Trans. I, 74 (1978) 975.
[4] R.M. Pashley, J. Colloid Interf. Sci. 80 (1981) 153; 83 (1981) 531;
R.M. Pashley, Adv. Colloid Interf. Sci. 16 (1982) 57;
R.M. Pashley, Chem. Scr. 25 (1985) 22.
[5] R.G. Horn, D.T. Smith and W. Hailer, Chem. Phys. Lett. 162 (1989) 404.
[6] R.G. Horn, D.R. Clarke and M.T. Clarkson, J. Mater. Res. 3 (1988) 413.
[7] J. Klein, J. Chem. Soc. Faraday Trans. I, 79 (1983) 99;
J. Klein, Macromol. Chem. Macromol. Symp. 1 (1986) 125;
H.J. Ploehn and W.B. Russel, Adv. Chem. Eng. 15 (1990) 137;
S.S. Patel and M. Tirrell, Annu. Rev. Phys. Chem. 40 (1989) 597.
[8] J.N. Israelachvili and P.M. McGuiggan, Science 241 (1988) 795;
J.N. Israelachvili, Acc. Chem. Res. 20 (1987) 415.
[9] H.K. Christenson, J. Disp. Sci. Technol. 9 (1988) 171.
[10] C.S. Lee and G. Belfort, Proc. Natl. Acad. Sci. USA 86 (1989) 8392.
[11] C.J. Coakley and D. Tabor, J. Phys. D 11 (1978) L773;
J.L. Parker and H.K. Christenson, J. Chem. Phys. 88 (1988) 8013;
C.P. Smith, M. Maeda, L. Atanasoska, H.S. White and D.J. McClure, J. Phys. Chem. 92 (1988) 199.
[12] R.G. Horn, Am. Ceram. Soc. 73 (1990) 1117.
[13] D. Tabor and R.H.S. Winterton, Proc. R. Soc. London Ser. A 312 (1969) 435
J.N. Israelachvili and D. Tabor, Proc. R. Soc. London Ser. A 331 (1972) 19.
[14] R.G. Horn and J.N. lsraelachvili, J. Chem. Phys. 75 (1981) 1400.
[15] I.K. Snook and W. van Megen, J. Chem. Phys. 70 (1979) 3099;
I.K. Snook and W. van Megen, J. Chem. Phys. 72 (1980) 2907;
I.K. Snook and W. van Megen, J. Chem. Soc. Faraday Trans. II, 77 (1981) 181;
W. van Megen and I.K. Snook, J. Chem. Soc. Faraday Trans. II, 75 (1979) 1095;
W. van Megen and I.K. Snook, J. Chem. Phys. 74 (1981) 1409.
[16] R. Evans and A.O. Parry, J. Phys. (Condens. Matter) 2 (1990) SA15.
[17] R. Kiellander and S. Marcelja, Chem. Phys. Lett. 120 (1985) 393;
R. Kjellander and S. Marcelja, Chem. Scr. 25 (1985) 73.
[18] D. Henderson and M.J. Lozada-Cassou, J. Colloid Interf. Sci. 114 (1986) 180.
[19] P. Tarazona and L. Vicente, Mol. Phys. 56 (1985) 557.
[20] H.K. Christenson and R.G. Horn, Chem. Scr. 25 (1985) 37.
[21] H.K. Christenson, D.W.R. Gruen, R.G. Horn and J.N. Israelachvili, J. Chem. Phys. 87 (1987) 1834;
R.G. Horn and J.N. Israelachvili, Macromolecules 21 (1988) 2836;
R.G. Horn, S.J. Hirz, G.H. Hadziioannou, C.W. Frank and J.M. Catala, J. Chem. Phys. 90 (1989) 6767;
J.N. Israelachvili and S.J. Kott, J. Chem. Phys. 88 (1988) 7162.
[22] H.K. Christenson, Chem. Phys. Lett. 118 (1985) 455.
[23] M. Schoen, C. Rhykerd, D. Diestler and J. Cushman, Science 245 (1989) 1223;
C. Rhykerd, M. Schoen, D. Diestler and J. Cushman, Nature 330 (1987) 461.
[24] P. Thompson and M. Robbins, Science 250 (1990) 792.
[25] U. Landman, W.D. Luedtke, N.A. Burnham and R.J. Colton, Science 248 (1990) 454.
[26] P. McGuiggan and J. Israelachvili, J. Mater. Res. 5 (1990) 2232.
[27] M.L. Gee and J.N. Israelachvili, J. Chem. Soc. Faraday Trans. 86 (1990) 4049.
[28] J.N lsraelachvili and H. Wennerstr6m, Langmuir 6 (1990) 873; J. Phys. Chem., in press.
158 J.N. lsraelach~,,ili

[29] M.K. Granfeldt and S.J. Miklavic, J. Chem. Phys. 95 (1991) 6351.
[30] H.E. Stanley and J. Teixeira, J. Chem. Phys. 73 (1980) 3404.
r~31] H. van Olphen, An Introduction to Clay Colloid Chemistry, 2rid ed. (Wiley, New York, 1977) ch. 10.
[32] B.E. Viani, P.F. Low and C.B. Roth, J. Colloid lnterf. Sci. 96 (1984) 229.
[33] U. Del Pennino, E. Mazzega, S. Valeri, A. Alietti, M.F. Brigatti and L. Poppi, J. Colloid Interf. Sci. 84 (1981)
301.
[34] J.P. Quirk, Israel J. Chem. 6 (1968) 213.
[35] B.V. Velamakanni, J.C. Chang, F.F. Lange and D.S. Pearson, Langmuir 6 (1990) 1323.
[36] M. Elimelech, J. Chem. Soc. Faraday Trans. 86 (1990) 1623;
R.R. Lessard and S.A. Zieminski, Ind. Eng. Chem. Fundam. 10 (1971) 260.
[37] H.K. Christenson, J. Fang, B.W. Ninham and J.L. Parker, J. Phys. Chem. 94 (1990) 8004.
[38] P.M. Claesson and H.K. Christenson, J. Phys. Chem. 92 (1988) 1650;
P.M. Claesson, C.E. Biota, P.C. Herder and B.W. Ninham, J. Colloid Interf. Sci. 114 (1986) 234;
J.L. Parker, D.L. Cho and P.M. Claesson, J. Phys. Chem. 93 (1989) 6121;
R.M. Pashley, P.M. McGuiggan, B.W, Ninham and D.F. Evans, Science 229 (1985) 1088;
Ya.I. Rabinovich and B.V. Derjaguin, Colloids Surf. 30 (1988) 243;
J.N. Israelachvili and R.M. Pashley, Nature 300 (1982) 341;
K. Kurihara, S. Kato and T. Kunitake, Chem. Lett. (Chem. Soc. Jpn.) 1990 (1990) 1555.
[39] C.A. Helm, J.N. Israelachvili and P.M. McGuiggan, Science 246 (1989) 919; Biochemistry, in press.
[40] H.K. Christenson, P.M. Claesson, J. Berg and P.C. Herder, J. Phys. Chem. 93 (1989) 1472.
[41] B. J6nsson, Chem. Phys. Lett. 82 (1981) 520.
[42] N.I. Christou, J.S. Whitehouse, D. Nicholson and N.G. Parsonage, Syrup. Faraday Soc. 16 (1981) 139.
[43] S. Marcelja and N. Radic, Chem. Phys. Lett. 42 (1976) 129.
[44] S. Marcelja, D.J. Mitchell, B.W. Ninham and M.J. Sculley, J. Chem. Soc. Faraday Trans. II, 73 (1977) 630.
[45] D.W.R. Gruen and S. Marcelja, J. Chem. Soc. Faraday Trans. II, 79 (1983) 225.
[46] B. J6nsson and H. Wennerstr6m, J. Chem. Soc, Faraday Trans. II, 79 (1983) 19.
[47] D. Schiby and E. Ruckenstein, Chem. Phys. Lett. 95 (1983) 435;
P. Attard and M.T. Batchelor, Chem, Phys. Lett. 149 (1988) 206.
[48] A. Luzar, D. Bratko and L.J. Blum, Chem. Phys. 86 (1987) 2955.
[49] H.K. Christenson, J. Colloid Interf. Sci. 121 (1988) 170;
L.R. Fisher and J.N. lsraelachvili, Colloids Surf. 3 (1981) 303.
[50] M. Wilchek and E.A. Bayer, Methods in Enzymology 184 (volume devoted to "Avidin-Biotin Technology")
(1990) 5-45;
N.M. Green, Adv. Protein Chem. 29 (1975) 85.
[51] C. Helm, W. Knoll and J.N. Israelachvili, Proc. Natl. Acad. Sci. USA 88 (1991) 8169.
[52] F.P. Bowden and D. Tabor, Friction and Lubrication, Methuen: 1967.
[53] J.A. Greenwood and K.L. Johnson, Phil. Mag. A 43 (1981) 697.
[54] F. Michel and M.E.R. Shanahan, C.R. Acad. Sci. Paris 310 II (1990) 17;
D. Maugis, J. Mater. Sci. 20 (1985) 3041.
[55] C.A. Miller and P. Neogi, Interfacial Phenomena (Decker, Basel, 1985).
[56] A.M. Schwartz, J. Colloid Interf. Sci. 75 (1980) 404.
[57] A.L. Weisenhorn, P.K. Hansma, T.R. Albrecht and C.F. Quate, Appl. Phys. Lett. 54 (1989) 26;
P.K. Hansma, V.B. Elings, O. Marti and C.E. Bracker, Science 243 (1988) 1586.
[58] B.R. Lawn and T.R. Wilshaw, Fracture of Brittle Solids (Cambridge Univ. Press, London, 1975).
[59] M. Sahimi and J.G. Goddard, Phys. Rev. B 33 (1986) 7848.
[60] M.D. Ellul and A.N. Gent, J. Polym. Sci. Polym. Phys. 22 (1984) 1953; 23 (1985) 1823.
[61] M.E.R. Shanahan, P. Schreck and J. Schultz, C.R. Acad. Sci. Paris 306 II (1988) 1325.
[62] F.J. Holly and M.J. Refojo, J. Biomed. Mater. Res. 9 (1975) 315.
[63] J.D. Andrade, L.M. Smith and D.E. Gregonis, in: Surface and Interfacial Aspects of Biomedical Polymers, Vol.
1, Ed. J.D. Andrade (Plenum, New York, 1985) p. 249.
[64] K.L. Johnson, K. Kendall and A.D. Roberts, Proc. R. Soc. London A 324 (197l) 301.
[65] H.M. Pollock, M. Barquins and D. Maugis, Appl. Phys. Lett. 33 (1978) 798;
M. Barquins and D. Maugis, J. M~c. Thdor. Appl. 1 (1982) 331.
[66] W.A. Zisman, Ind. Eng. Chem. 55(10) (1963) 19;
F.M. Fowkes, Ind. Eng. Chem. 56 (1964) 40;
W.A. Zisman and J. Fox, Colloid Sci. 7 (1952) 428.
[67] J.N. Israelachvili, A.M. Homola and P.M. McGuiggan, Science 240 (1988) 189.
Adhesion forces between surfaces in liquids and condensable vapours 159

[68] A.M. Homola, J.N. Israelachvili, M.L. Gee and P.M. McGuiggan, J. Tribology 111 (1989) 675.
[69] J. Van Alsten and S. Granick, Phys. Rev. Lett. 61 (1988) 2570.
[70] M. Gee, P. McGuiggan, J. Israelachvili and A. Homola, J. Chem. Phys. 93 (1990) 1895.
[71] A.M. Homola, J.N. Israelachvili, P.M. McGuiggan and M.L. Gee, Wear 136 (1990) 65.
[72] G.M. McLelland, in: Adhesion and Friction, Springer Series in Surface Sciences, Vol. 17, Eds. M. Grunze and
H.J. Kreuzer (Springer, Berlin, 1989) pp. 1-16.
[73] J.N. lsraelachvili and P.M. McGuiggan, J. Mater. Res. 5 (1990) 2223.
[74] Y.L. Chert, M.L. Gee, C.A. Helm, J.N. Israelachvili and P.M. McGuiggan, J. Phys. Chem, 93 (1989) 7057.

Vous aimerez peut-être aussi