Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Surfaces and Interfaces

Professor Harjinder Singh


Institute of Food, Nutrition and Human Health

Suggested Books:

• D. J. Shaw (1992) “Introduction to Colloid and


Surface Chemistry”, Oxford, Butterworth-Heinemann,
U.K.
• P. Walstra (2003), “Physical Chemistry of Foods”.
Ed. Marcel Dekker, Inc. New York.
• D. Myers (1999), “Surfaces, Interfaces and
Colloids”. VCH Publishers, Germany.

1
SURFACE AND INTERFACES: GENERAL CONCEPTS

Introduction

The boundary region between two adjacent bulk phases is known as an


interface, when one of the phases is a gas or vapour, the term surface is
commonly used. Matter at an interface usually has different physical
properties and energy characteristics from that in the bulk.

Surface Tension and Surface Free Energy

Manifestations of surface tension are all around us in everyday life: the


sphericity of raindrops, the rise of liquid in a capillary tube, the bulging of the
liquid surface above the rim of an overflowing vessel, and so on.

Attractive forces exist between molecules in the liquid state. In the bulk of the
liquid there is a balance of these attractive forces. Molecules are, on average,
subjected to equal forces of attraction in all directions. However, at the surface
there are forces from below and from sides but virtually no equivalent forces
in the gas above, thus there is a resultant inward force on the surface
molecules. As many molecules as possible will leave the liquid surface for the

2
interior of the liquid, the surface will therefore tend to contract spontaneously.
For this reason, droplets of liquid and bubbles of gas tend to attain a spherical
shape. To expend the liquid, more molecules must be brought to the surface,
must do the work against the molecular forces. This work is the surface free
energy or surface tension. The surface free energy or surface tension is defined
as the work (joules) required to increase a surface area by unit amount
(metres2).

Figure 1. Attractive forces between molecules at the surface and in the


interior of a liquid

Reversible work and free energy are equivalent in thermodynamics, and so the
surface tension (y) of a pure fluid system is defined by

where G is the Gibbs free energy of the system, A is the surface area, p is the
pressure, T is the temperature, and n is the total amount of material in the
system. The surface tension (  ) of very pure water is 72.8 mN m -1 at 20 °C. The
value of  for water is much larger than that for most other liquids; ethanol, for
instance, has a value of  = 22 mM m-1. It is important to stress that these values
refer to the pure liquids. Even a trace of impurity changes the surface tension of
water.

3
The interfacial tension of a binary system (e.g. hydrocarbon + water) usually
lies between the surface tensions of the two pure liquids at the same temperature.
The interfacial tension at the hydrocarbon-water interface is of the order of 50 mN
m-1. The oil phase in a food emulsion is a triglyceride, or, more correctly, a mixture
of triglycerides plus small amounts of other lipids, many of which are surface
active. The tension at the pure triglyceride-water interface is of the order of 25 mN
m-1. At the interface between a typical food oil and water, the tension may be as low
as 10 mN m-1.

Adsorption from solution

Adsorption
Adsorption processes play a central role in determining the behaviour of food
systems. 'Adsorption' is the phenomenon whereby dissolved molecules (or
dispersed particles) accumulate at a surface in greater concentrations than one
would expect simply on the basis of the molecules (or particles) being randomly
distributed throughout the system.

Small-molecule are normally considered to adsorb reversibly, since their


transport to and from the interface occurs on a time-scale much shorter than the
typical experimental observation time-scale. The adsorption of small solute
molecules may therefore be described in terms of the equations of equilibrium
thermodynamics. However, with macromolecules or particles the adsorption
process is often regarded as being irreversible, since the time-scale for
desorption or rearrangement at the surface may be long compared with the time-
scale of observation.

The Gibbs adsorption equation


The most important equation in the surface thermodynamics of solutions is the
Gibbs adsorption equation which links the change in surface tension to the
concentration of adsorbing species at the interface.

4
Figure 2 Gibbs representation of the interface between two bulk fluid phases α and
β. The Gibbs dividing surface XY lies within the surface phase σ.

Consider a system of two bulk phase α and β (gas + liquid or liquid + liquid)
separated by a notional surface phase σ as illustrated in Figure 2. The surface
phase contains a hypothetical mathematical plane XY called the Gibbs dividing
surface. This plane XY\s not an actual physical boundary, but is just a reference
level for quantifying the extent of adsorption of any chemical component i in the
system. We define a quantity nσi, called the surface excess amount of component i,
which is taken to be the amount of component i in the surface phase in excess of
that which would have been present had the bulk phase α and β extended to the
Gibbs dividing surface with unchanging composition. A surface excess
concentration of component i is defined as

where A is the surface area. At constant temperature and pressure, the surface
excess concentrations of all the chemical components in the system are related by
the Gibbs-Duhem equation,

5
where µi is the chemical potential of component i; and dγ and dµi are total
derivatives of γ and µi, respectively. For a binary mixture of solvent (1) + solute
(2), equation reduces to

The actual values taken by Γ1 and Γ2 are dependent on the position of the Gibbs
dividing surface. If the location of the plane XY is chosen in such a way that the
solvent surface excess concentration vanishes (i.e. Γ1 = 0), then we
have

which is the commonly encountered form of the Gibbs adsorption equation. x2 is the
solute mole fraction, and R is the gas constant.

The main use of the Gibbs adsorption equation is to calculate the amount
adsorbed from measurements of the variation of surface tension (or interfacial
tension) with concentration. For a real (non-ideal) solution, we have

where f2 is the solute activity coefficient (unity for an ideal solution).

Surface Activity

Introduction
Materials that adsorb strongly at the interface, and therefore cause a
substantial lowering of the surface tension at low concentrations, are called
surface-active agents or surfactants.

6
Surface-active materials possess a characteristic chemical structure that
consists of: (1) molecular components that will have little attraction for the solvent,
normally called the hydrophobia group, and (2) chemical units that have a strong
attraction for the solvent, called the hydrophilic group (Figure 3).

Figure 3 The basic molecular structure of surface-active materials.

Materials that possess chemical groups leading to surface activity are generally
referred to as being amphiphilic ("liking both"), indicating that they have some
affinity for two essentially immiscible phases. The amphiphilic structure of
surfactant molecules not only results in: (1) the adsorption of surfactant molecules
at interfaces and the consequent alteration of the corresponding interfacial
energies, and (2) the orientation of the adsorbed molecules such that the
hydrophobic groups are directed away from the bulk aqueous phase (Figure 4).
Surface activity is a dynamic phenomenon; there is a balance between tendency for
adsorption and tendency for complete mixing due to thermal motion of molecules.

Figure 4 Schematic illustration of the preferential orientation of surfactant


molecules at interfaces.

7
Surfactant Structures
Synthetic surfactants and the natural fatty acid soaps are amphiphilic materials
that tend to exhibit some solubility in water as welt as some affinity for nonaqueous
solvents. The hydrocarbon part of the molecule is responsible for its solubility in oil,
while the polar - COOH or - OH group has sufficient affinity to water. They are
able to locate their hydrophilic head groups in the aqueous phase and allow the
hydrophilic hydrocarbon chains to escape into the air or oil phase. This situation
is energetically more favourable than complete solution in either phase.

The role of molecular structure in determining surfactant adsorption directly


related to the size of the hydrophobic and hydrophilic portions of the
adsorbing molecules. When one considers the adsorption of molecules at an
interface, it can be seen that the maximum number of molecules that can be fitted
into a given area depends upon the area occupied by each molecule. That area
will, to a good approximation, be determined by either the cross-sectional area of
the hydrophobic chain or the area required for the arrangement for closest
packing of the head groups (Figure 5). For straight-chain 1:1 ionic surfactants, it is
usually found that the head group requirement predominates, so that for a given
homologous series, the surface tension minimum obtained varies only slightly
with the length of the hydrocarbon chain.

The introduction of polar groups, such as double bond: ether, ester, or amide
linkages; or hydroxyls located well away from the head group, usually results in a
significant lowering of both the efficiency and effectiveness of the surfactant as
compared to a similar material with no polar units. Such a result has generally been
attributed to changes in orientation of the adsorbed molecule with respect to the
surface due to interactions between the polar group and the water (Figure 5c).

8
Figure 5 Schematic illustration of the role of surfactant tail and head group in
determining packing efficiency at an interface: (a) straight chains favour close,
efficient packing, limited by the size of the head group; (b) branched or bulky
chains hinder efficient packing, (c) a second polar group in the surfactant on its
orientation at the interface.

Classification of Surface Active Agents


The most useful chemical classification of surface-active agents is based on the
nature of the hydrophilic group, with subgroups being defined by the nature of the
hydrophobic group. The four general groups of surfactants are defined as follows:

1. Anionic, with the hydrophilic group carrying a negative charge such as


carboxyl (RCOO-), sulfonate (RSO3-)( or sulfate (ROSO3-).

2. Cationic, with the hydrophilic group bearing a positive charge, as for


example, the quaternay ammonium halides (R 4N+Cl-).

3. Nonioic, where the hydrophilic group has no charge but derives its water
solubility from highly polar groups, such as polyoxyethylene (--OCH2CH2O--).

4. Amphoteric (and zwitterionic). in which the molecule contains, or can


potentially contain, both a negative and a positive charge, such as the

9
sulfobetaines, RN+(CH 3)2CH2CH2SO3-.

Table 2 gives some examples of commonly encountered surfactants falling


into the four categories of nonionic, anionic, cationic, and zwitterionic. Food
surfactants are mostly nonionic or zwitterionic.

Micelle Formation by Surfactants

Introduction
On dissolving in water, surfactant molecules can achieve a segregation of
tail-groups from the solvent through association to form micelles. Micelle
formation is the property that surface-active solute have of forming organised
aggregates in solution. Micelle formation or micellization affect the interfacial
phenomenon, such as interfacial tension reduction.

Critical Micelle Concentration (CMC)


In dilute solutions the surfactant acts as a normal solute and exists as a
monomer. However, at fairly well defined concentration, abrupt chan ges in
several physical properties, such as osmotic pressure, turbidity, electrical
conductance and surface tension, take place (Figure 6). The rate at which

10
osmotic pressure increases with concentration becomes abnormally low and
the rate of increase of turbidity with concentration is much enhanced, which
suggests that considerable association is taking place. Molar conductance
curve show a sharp break in it, indicating a sharp increase in the mass per
unit charge of the material in solution. This behaviour could be explained in
terms of formation of micelles, of the surfactant monomers in which the
hydrophobic hydrocarbon chains are oriented towards the interior of the
micelle, leaving the hydrophilic groups in contact with the aqueous medium.
The concentration above which micelle formation becomes appreciable is
termed the critical micelle concentration (CMC).

Osmotic pressure

Figure 6 Physical properties of sodium dodecyl sulphate solutions at 25°C.

Micelle formation is an alternative mechanism to adsorption - reduces the


free energy of the system. In considering the energetics of micellisation, two
factors are important.

1. intermolecular attractions between hydrocarbon chains in the interior of the


micelles - energetically favourable.
2. Strong water-water interactions in micellisation.

11
Structure of Micelles
The structure of micelle in aqueous medium, at concentrations not too far
above the CMC, can be considered roughly spherical. Hydrophobic groups of
the surface molecules are situated in the interior region and these are
surrounded by an outer region which contains hydrophilic groups and bound
water. However, at concentrations much higher than the CMC (10 times),
various crystalline phases tend to form, for which the laminar micelle is the
basic model (Figure 7).

Changes in temperature, concentration of surfactant, additives in the liquid


phase, and structural groups in the surfactant all may cause change in the
size and shape of the micelle, with the structure varying from spherical
through rod- or disk-like to lamellar in shape.

Figure 7 Micellar structures in surfactant solution, (a) Spherical (anionic)


micelle, (b) Laminar micelles.

Monomolecular Films (Insoluble Films)

Introduction
Some materials, when placed on an interfaces, form the so-called Insoluble
monolayers - that is, monomolecular layers or films of adsorbed molecules
which have very low solubility in the supporting liquid phase, so that they are
essentially isolated on the surface. In these films, all the surfactant is at the

12
interface. The molecules in a mono molecular film, especially at high surface
concentrations, are often arranged in a simple manner, and much can be
learned about the size, shape and orientation of the individual molecules by
studying various properties of the monolayer.

The study of insoluble films at oil/water (O/W) and water/air (W/A)


interfaces provides a means of understanding the behaviour of emulsions and
foams.

Surface Pressure
The surface pressure of a monolayer film,  , is defined as the difference
between the surface tension of the pure supporting liquid,  0, and that of the
liquid with an adsorbed film,  .

 0= surface tension of clean interface.


 = surface tension in the presence of monolayer.
i.e.  is the decrease in surface tension due to monolayer.

The surface pressure, as defined by equation 1.7, represents an expanding


pressure exerted by the monolayer acting against the (contracting) surface
tension of the pure liquid substrate (Figure 9). Analogous to the pressure -
volume (P- V) curve of a three-dimensional bulk material, one can construct a
pressure—area (  --A) curve for a monolayer.

Figure 9 An illustration of the pulling action of surface tension working against


the "pushing" of adjacent adsorbed molecules. 13
Experimentally, one can either work with a fixed surface area and increase
the pressure by incrementally adding more of the adsorbed material, or a
known amount of material can be added to a surface and the pressure
increased by slowly decreasing the available area. In practice, the latter
approach, as represented by the classical Langmuir trough, is much preferred,
since it is easier to measure area reproducibly than to measure the addition of
small amounts of dilute solutions. The Langmuir-Adam surface balance (or
trough) uses a technique for containing and manoeuvring insoluble
monolayers between barriers for the direct determination of  -A curves. The
film is contained (Figure 10) between a movable barrier and a float attached to a
torsion-wire arrangement. The surface pressure of the film is measured
directly in terms of the horizontal force which it exerts on the float and the area of
the film is varied by means of the movable barrier.

With a known amount of material on the surface, the  -A curve allows one
to determine something about the physical nature of the film and some
molecular characteristics of the adsorbed material.

Figure 10 The principle of the Langmuir-Adam surface balance.

14
Surface Rheology
Because of the mobility of molecules in the surface of a pure liquid, such
surfaces have very little elasticity. For that reason, pure liquids cannot support a
foam. In the presence of an adsorbed monolayer film, however, the
Theological properties of the surface can change dramatically. By analogy with
bulk phases, the physical state of a surface film can be distinguished by its
viscosity.

Surface viscosity is the change in the viscosity of the surface layer brought
about by the monomolecular film. Monolayers in different physical states can
readily be distinguished by surface viscosity measurements.

It is the increase in surface viscosity produced by adsorbed films (insoluble


and Gibbs monolayers, adsorbed polymers,etc) that leads to the production of
persistent foams, helps stabilize emulsions, and explains the role of spread
monolayers in dampening surface waves, among other important interfacial
phenomena.

Protein film surface rheology depends on the molecular structure of the


proteins in the adsorbed layer and the solvent conditions (pH, ionic strength,
etc.). Generally speaking, all the surface Theological parameters reach
maximum values at the isoelectric point. This is interpreted as being due to an
optimization of intra- and intermolecular cohesion within the film under
conditions of minimum electrostatic repulsion and hydration.

Amongst the milk proteins, those film formed from the globular whey
proteins are much more viscoelastic than those formed from the disordered
caseins, both at oil-water and air-water interface. Even amongst the caseins
themselves, there are significant differences: the surface shear viscosity of
αs1-casein at the oil-water interface is significantly larger than that for β-
casein, but much less than that for  -casein.

15
The Physical State of Monolayer Films
Like bulk materials, monolayer films exhibit characteristics that can
(sometimes with a bit of imagination) be equated to the solid, liquid, and
gaseous states of matter. For films, the equivalent states are roughly defined as:

a. Condensed (solid) films, which are coherent, rigid (essentially


incompressible), and densely packed, with high surface viscosity. The
molecules have little mobility and are oriented perpendicular (or almost so)
to the surface (Figure 11 a).
b. Expanded films, roughly equivalent to the liquid state, in which the
monolayer is stilt coherent and relatively densely packed but is much more
compressible than condensed films. Molecular orientation is still
approximately perpendicular to the surface, but the tails are less rigidly
packed (Figure 11 b).
c. Gaseous films, in which the molecules are relatively far apart and have
significant surface mobility. The molecules act essentially independently,
much as a bulk phase gas and molecular orientation will be random
(Figure 11c).

Each type of monolayer film exhibits its own special characteristics


analogous to the corresponding bulk phase, as well as distinct "phase
transitions" which are useful in characterizing the nature of the film in terms if its
equation of state, molecular orientation, interfacial interactions, etc.

Figure 11 Schematic illustration of the primary "states" of adsorbed


monolayer films: (a) the condensed state—tight packing of head groups and
limited mobility of tails; (b) liquid expanded—relatively tight head group
packing, but significant mobility in the tails; (c) gaseous—wide separation of
molecules with little interaction between neighbors.

16
Wetting and Spreading

The wetting of a surface by a liquid and the ultimate extent of spreading of


that liquid are very important aspects of practical surface chemistry. The
following discussion is intended to introduce the fundamental concepts
underlying modern theories of wetting and spreading processes.

The Contact Angle


One of the primary characteristics of any immiscible, two- or three-phase
system containing two condensed phases, at least one of which is a liquid, is the
contact angle of the liquid on the other condensed phase. The contact angle of
one liquid on another, is important in certain multilayer coating processes, in
which several liquid layers are coated simultaneously on a solid surface. In such
processes, it is important that the wetting properties of each layer on the one
below be such that smooth, uniform coverage is obtained. Of more practical and
widespread importance is the contact angle of a liquid directly on a solid,

When a drop of liquid is placed on a solid surface, the liquid will either
spread across the surface to form a thin, approximately uniform film (Figure
12a) or it will spread to a limited extent, or remain as a discrete drop on the
surface. The final condition of the applied liquid on the surface is taken as an
indication of the wettability of the surface by the liquid or the wetting ability of the
liquid on the surface. The quantitative measure of the wetting process is taken to
be the contact angle, 0, which the drop makes with the solid as measured
through the liquid in question (Figure 12b).

In the case of a liquid that forms a uniform film (i,e, where  = 0°), the solid is
said to be completely wetted by the liquid. If a finite contact angle is formed (  >
0°), say between 30° and 89° the system would be 'partially wetting." And 90°
and above, the system is considered to be nonwetting.

17
Figure 12 Schematic illustration of the various degrees of wetting: (a)
complete wetting; (b) partial wetting; (c) complete nonwetting.

Thermodynamics of Wetting
The basic framework for the application of contact angles and wetting
phenomena lies in the field of thermodynamics. However, in practical
applications it is often difficult to make a direct correlation between observed
phenomena and basic thermodynamic principles. Nevertheless, the
fundamental validity of the analysis of contact angle data and wetting
phenomena helps to instill confidence in its application to nonideal situations.

1. Young's equation
If one considers the three-phase system depicted in Figure 13, in which the
liquid drop is designated fluid L, the surrounding medium G, and the solid
surface S, then at equilibrium the contact angle θ will be given by Young's
equation as

where  LG ,  SL , and  SG are the interfacial tensions at the L/G, L/S and S/G
interfaces, respectively.

18
Figure 12. Schematic illustration of the mechanical equilibrium of surface
force leading to contact angle formation as given by Young's equation.

While Young's equation provides a thermodynamic definition of the contact


angle, its experimental verification is prevented by the fact that the values of
 SL and  SG cannot be directly determined experimentally. In this sense, the
contact angle of a liquid on a solid differs from that of a liquid on a second
liquid since in the latter case all three interfacial tensions can be determined
independently and the relationship can therefore be verified directly.

When the liquid of interest (1) is found to spread completely over the solid
surface, θ = 0 and equation (1.8) reduces to

2. The spreading coefficient


Young's equation is usually found to be a very useful and adequate means of
describing wetting equilibria in most circumstances. However, it is
sometimes found useful to define another term that indicates from a
thermodynamic point of view whether a given liquid-solid system will be
wetting (θ = 0°) or nonwetting (θ > 0°). Such a term is the spreading
coefficient, S.

For a spontaneous process (such as spreading) to occur, the free energy of


the process must be negative. In terms of surface free energies, then, one can
write the relationship

19
SSLG   SG   SL   LG (1.10)

where the subscript SLG refers to the initial spreading coefficient for the
spreading of liquid L over the solid S in the presence of (or displacing) gas G.

3. Work of cohesion and work of adhesion


Although the mathematical relationships encountered in wetting
phenomena are usually quite simple, they are found to be very useful in many
practical applications. Their combinations and variations have given rise to still
more relationships, which further expand their utility without expanding the
amount of information necessary for their application. Two thermodynamic
relationships that can be useful in the analysis of wetting and spreading
phenomena are the works of cohesion and adhesion.

The work of cohesion, Wc, is defined as the reversible work required to


separate two surfaces of unit area of a material with surface tension  , given
simply as:

WC  2 LG
(1.11)

The related quantity the work of adhesion, W a, is defined as the work


required to separate unit area of interface between two different materials.

WA   SG   LG   SL
(1.12)

Wetting of Milk Powders


The rate of dissolution of skim milk powder into water to form a milk colloid is
of considerable importance to the food technologist. A crucial stage of the
dissolution process is the wetting of the solid particles by the aqueous
medium. The thermodynamics of wetting is determined by the value of the

20
contact angle θ between the solid particle and the water. The surface of a
powder particle is rough, and contact angles are hard to measure with any
reliability. Values of θ =20º for skim milk powder and θ = 50° for whole milk
powder, have been reported.

As a general rule, whole milk powders wet readily in warm water (> 40°C) but
poorly in cold water. But skim milk powders (or powders made from skim milk
recombined with a low-melting-point dairy fat fraction) wet adequately in both
warm and cold water. Following forced immersion, a poorly wetting powder
floats to the liquid surface with a substantial part of the particle surface
remaining repellent to water. This observation is consistent with the view that at
least part of the 'wetting problem' arises from disruption of protective layers
around fat globules during spray-drying, with the released fat forming a water-
repellent layer around the dried solid particles. If the fat is liquid, however, the
particles are wettable because the oil film is able to retire into lenses or minute
globules. The problem can be minimized in practice by using procedures that
discourage disruption of the globule membrane.

In the absence of a solid fat barrier, contact with water induces rapid
hydration of the lactose and swelling of the proteins. This may lead to the
formation of a gelatinous layer around a group of particles, thus hindering
further penetration of solvent. With a coarse powder such aggregated lumps can
be avoided, but the rat of solubilization may be unacceptably slow (especially
in cold water). Rapid solubilization is achieved through the production of
sponge-like conglomerates (2-4 mm diameter). Capillary action draws water
through the fine pores, leading to fast wetting and dissolution.

21

You might also like