You are on page 1of 10

Preface

The main objective of this book is to provide a general physical chemistry background
in a clear manner to students and scientists working with solid–liquid interfaces. The
number of industries related to surface chemistry and physics have increased dramatically
during the past decades through the rapid development of surface coatings, adhesives, tex-
tiles, oil recovery, cosmetics, pharmaceuticals, biomedical, biotechnology, agriculture,
mineral flotation, lubrication industries and the introduction of newly emerged nano-
technology. A great number of publications on these subjects over the last decade have
resulted in a better understanding of all of these scientific fields, which are mostly inter-
disciplinary in nature, where high-vacuum physicists, colloid and polymer chemists,
biologists, and chemical, material, biomedical, environmental, and electronic engineers are
all contributing.
Unfortunately, most standard textbooks on physical chemistry pay only limited atten-
tion to surface science at present. Consequently, many graduates in the fields of chemistry,
physics, biology, engineering and materials science enter their careers ignorant of even the
basic concepts of important surface interactions and are now under pressure to solve an
increasing number of challenging new problems. Thus, knowledge of the fundamental
principles of interfacial chemistry and physics is going to become a requisite for all the
research and technical personnel in these industries. Although a number of excellent books
are available in both surface and colloid sciences, and in polymer, solid physics and chem-
istry, these can seem daunting due to their written for experts nature. This book is intended
to meet the needs of newcomers to the surface field from academia or industry who want
to acquire a basic knowledge of the subject.
The text largely contains fundamental material and focuses on understanding the basic
principles rather than learning factual information. Since it is impossible to include all
branches of surface science in such an introductory book because of its wide and multi-
disciplinary scope, a specific and narrow topic, the interfacial interactions between solids
and liquids, has been chosen for this book. For this reason, the ionic interactions, charged
polymers, electrochemistry, electrokinetics and the colloid and particulate sciences cannot
be included. Some fundamental physical chemistry subjects such as basic thermodynam-
ics are covered, and many equations are derived from these basic concepts throughout the
book in order to show the links between applied surface equations and the fundamental
concepts. This is lacking in most textbooks and applied books in surface chemistry, and
for this reason, this book can be used as a textbook for a course of 14–15 weeks.
xii Preface

The book is divided into three parts: (1) Principles, (2) Liquids and (3) Solids. In Prin-
ciples, an introduction to surfaces and interfaces, molecular interactions and the thermo-
dynamics of interfaces are presented in three chapters in their most basic form. In Liquids,
the properties of pure liquid surfaces, liquid solution surfaces, the experimental determi-
nation of liquid surface tension and the potential energy of interaction between particles
and surfaces through a liquid medium are covered in four chapters. In Solids, the proper-
ties of solid surfaces, contact angles of liquid drops on solids and some applications involv-
ing solid/liquid interfaces are presented in three chapters. The aim is that both final year
undergraduate or graduate university students and also industrial researchers will find the
necessary theory they seek in this book, and will be able to use and develop these concepts
effectively. Considerable attention is devoted to experimental aspects throughout the book.
At the end of each chapter are bibliographies to which the reader may turn for further
details. Since much of the material in this book is drawn from existing treatments of the
subject, our debt to earlier writers is considerable and, we hope, fully acknowledged. This
introductory contribution cannot claim to be comprehensive; however it does hopefully
provide a clear understanding of the field by assessing the importance of related factors.
Some errors are unavoidable throughout the book and if found can be communicated by
e-mail to (yerbil@gyte.edu.tr).
I am greatly indebted to my colleagues, who encouraged me to prepare this book, and
my wife Ayse and my children, Ayberk, Billur, Beril and Onur for their patience and under-
standing throughout its preparation.

H. Yıldırım Erbil
2006
PART I

Principles
Chapter 1
Introduction to Surfaces and Interfaces

1.1 Definition of a Surface and an Interface


A phase of a substance is a form of matter that is uniform throughout in chemical com-
position and physical state. There are mainly three phases of matter namely solid, liquid
and gas. The word fluid is used to describe both gas and liquid phases. We usually classify
the phase of a material according to its state at the normal ambient temperature (20–25°C),
which is well above the melting point of most fluids. We mostly deal with two or more
phases, which coexist, in equilibrium or non-equilibrium conditions. Phase diagrams are
used as a convenient method of representing the regions of stability of solid, liquid and
gas phases under various conditions of temperature and pressure. An interface is the phys-
ical boundary between two adjacent bulk phases. The interface must be at least one molec-
ular diameter in thickness for the purpose of constructing a molecular model. In some
cases it may extend over several molecular thicknesses. We use the word surface in order
to define the physical boundary of only one of these phases, such as solid surface and liquid
surface etc. In reality, we deal with an interface in all cases other than absolute vacuum con-
ditions for solids, since every single phase is in contact with another phase such as solid–air,
liquid–air contacts. In many standard physical chemistry, surface chemistry and surface
physics textbooks, the words surface and interface are used interchangeably because the
authors neglect the small differences between the air phase and absolute vacuum condi-
tions. On the other hand, some authors define and use the word surface for only the
solid–gas and liquid–gas phase contacts, and the word interface for all of other phase con-
tacts, but this has no scientific basis and should be abandoned.
The molecules that are located at the phase boundaries (that is between solid–gas,
solid–liquid, liquid–gas and liquid1–liquid2) behave differently to those in the bulk phase.
This rule generally does not apply for solid–solid and gas–gas interfaces where atomic and
molecular bonding in the solid structure restrict the reorientation of interfacial molecules
for the former, and the ease of miscibility of different gas molecules in free space does not
allow any interface formation for the latter.
There is an orientation effect for molecules at fluid surfaces: the molecules at or near
the fluid surface will be orientated differently with respect to each other than the mole-
cules in the bulk phase. Any molecule at the fluid surface would be under an asymmetri-
cal force field, resulting in surface or interfacial tension. The nearer the molecule is to the
interfacial boundary, the greater the magnitude of the force due to asymmetry. As the dis-
4 Surface Chemistry of Solid and Liquid Interfaces

tance from the interfacial boundary increases, the effect of the force field decreases and
vanishes after a certain length. Thus, in reality there is no very sharp interfacial boundary
between phases, rather there is a molecular gradient giving a change in the magnitudes of
both density and orientation of interfacial molecules. We must remember that this layer is
also very thin, usually between one and five monomolecular layers for liquid–vapour inter-
faces. The location of the so-called mathematically dividing surface (see Section 3.2.5) is
only a theoretical concept to enable scientists to apply thermodynamics and statistical
mechanics.
For solid surfaces, we should define the depth of the molecular layers, which are regarded
as surface. We may say that only the top monolayer of surface atoms and molecules, which
is the immediate interface with the other phases impinging on it, can be called the surface.
However, in practice, it depends on the spectroscopic technique that we apply: some so-
phisticated spectroscopic instruments can determine the chemical structure of the top 2–10
atomic layers (0.5–3.0 nm), thus taking this depth as the solid surface layer. In many older
books on surface chemistry or metallurgy, the surface is regarded as the top 100 nm or so
of the solid, because some older technologies can only determine the chemical structure
in the range of 10–100 nm and the surface was therefore assumed to be in this depth range.
However, at present, this range may be called an intermediate layer between the bulk and
surface structures, and only after 100 nm is it appropriate to regard such a layer in terms
of its bulk solid-state properties. Thus, in the recent scientific literature, it has become
necessary to report the solid surface layer with its thickness, such as the “top surface
monolayer”, the “surface film of 20 nm”, the “first ten layers”, or the “surface film less than
100 nm” etc.

1.2 Liquids and Liquid Surfaces


A liquid is defined as a medium which takes the shape of its container without necessar-
ily filling it. Water is the most abundant liquid on earth and is essential for living systems,
and non-aqueous liquids are widely used as solvents in synthetic, analytical, electrochem-
istry, polymer chemistry, spectroscopy, chromatography and crystallography. Molten me-
tals are also considered to be liquid solvents. We may also consider another property that
distinguishes liquids: the response of a liquid to an applied force is mainly inelastic; that
is a liquid does not return to its original shape following the application, then removal, of
a force.
The densities of liquids under normal pressures are not too dissimilar to those of their
solids between the melting and boiling points. Generally a liquid is less dense than its solid
at the melting point, but there are a few exceptions of which water is one; ice floats on
water. These less dense solids have rather open crystal structures. Silicon, germanium and
tin are other examples. Another similarity between normal liquids and solids is that they
have a low compressibility due to there not being a great deal of space between the mole-
cules in a liquid.
If we were able to observe the molecules in a liquid, we could see an extremely violent
agitation in the liquid surface. The extent of this agitation may be calculated by consider-
ing the number of molecules that must evaporate each second from the surface, in order
to maintain the measurable equilibrium vapor pressure of the liquid. When a liquid is in
Introduction to Surfaces and Interfaces 5

contact with its saturated vapor, the rate of evaporation of molecules is equal to the rate
of condensation in the equilibrium conditions. For example, it is possible to calculate from
kinetic theory that 0.25 g sec−1 water evaporates from each cm2, which corresponds to
8.5 × 1021 molecules sec−1 at 20°C having a saturated water vapor pressure of 17.5 mmHg.
However, the size of the water molecules permits only about ≈1015 molecules to be present,
closely packed in each cm2 of the surface layer at any particular moment, and we can con-
clude that the average lifetime of each water molecule is very short, ≈10−7 sec. This tremen-
dous interchange between liquid and vapor is no doubt accompanied by a similar
interchange between the interior of the liquid and its surface. However, although the
thermal agitation of water molecules is so violent that they jump in and out of the surface
very rapidly, the attractive forces between them are able to maintain the surface to within
one to three molecules thickness. The evidence for this can be derived from the nature of
light reflection from the liquid surfaces which results in completely plane polarized light
from very clean liquid surfaces, indicating an abrupt transition between air and the liquid
surface. On the other hand, the surface molecules of a liquid are mobile, that is they are
also in constant motion parallel to the surface, diffusing long distances. Other properties
of liquid surfaces will be described in Chapter 4.
Liquid surfaces tend to contract to the smallest possible surface area for a given volume
resulting in a spherical drop in equilibrium with its vapor. It is clear that work must
be done on the liquid drop to increase its surface area. This means that the surface mole-
cules are in a state of higher free energy than those in the bulk liquid. This is due to the
fact that a molecule in the bulk liquid is surrounded by others on all sides and is
thus attracted in all directions and in a physical (vectorial) balance. However for a surface
molecule there is no outward attraction to balance the inward pull because there are very
few molecules on the vapor side. Consequently, every surface molecule is subject to a strong
inward attraction perpendicular to the surface resulting in a physical force per length,
which is known as surface tension (see Section 3.2.4). In addition, the tendency to mini-
mize the surface area of a fluid surface can also be explained in terms of a thermodynamic
free energy concept (see Chapter 3), since the surface molecules have fewer nearest neigh-
bors and, as a consequence, fewer intermolecular interactions than the molecules in the
bulk liquid. There is therefore a free energy change associated with the formation of a liquid
surface reversibly and isothermally, and this is called the surface free energy or, more cor-
rectly, the excess surface free energy. The surface free energy term is preferred in the litera-
ture and it is equivalent to the work done to create a surface of unit area (Joules m−2 in the
SI system). It must be remembered that this is not the total free energy of the surface mol-
ecules but the excess free energy that the molecules possess by virtue of their being located
on the surface.

1.3 Surface Area to Volume Ratio


All liquid droplets spontaneously assume the form of a sphere in order to minimize their
surface free energy by minimizing their surface area to volume ratio, as we will see in the
thermodynamic treatment in Chapter 3. The surface area to volume ratio can easily be cal-
culated for materials having exact geometric shapes by using well-known geometric for-
mulas. For example, for a sphere and a cube this ratio is
6 Surface Chemistry of Solid and Liquid Interfaces

 A  = (4prsph ) = 3 ,  A 
2
6a 2 6
= 3c =
 V  sph (4 3 prsph
3
) rsph V cube ac ac
  (1)

where rsph is the radius of the sphere and ac is the edge length of the cube. A comparison
between the surface area to volume ratio of any geometrically shaped objects can be made
by equating their volumes and then calculating their respective surface areas. For example,
the edge length of a cube in terms of the radius of a sphere having the same volume is, ac
= rsph(3/4p)1/3 thus giving an area ratio of
13
Asph  p 
= ≅ 0.806 (2)
Acube  6 
The same approach may be applied to any geometric shape; the sphere has the minimum
surface area for a given volume due to the fact that it has a completely curved profile
without sharp edges (see Section 4.3 for the definition of curvature and its applications in
surface science). For materials having irregular geometric shapes, well-known integration
techniques from calculus are applied to calculate the surface area to volume ratio.

1.4 Solids and Solid Surface Roughness


An ideal solid surface is atomically flat and chemically homogenous. However, in reality
there is no such ideal surface, all real solid surfaces have a surface roughness over varying
length scales and they are also chemically heterogeneous to a degree due to the presence
of impurities or polycrystallinity differences. Surface roughness is defined as the ratio of
real surface area to the plan area and can be determined by Atomic Force Microscopy
(AFM), Scanning Electron Microscopy (SEM) and other spectroscopic methods, and also
by measuring the advancing and receding contact angles of liquid drops on substrates.
Solid surfaces are very complex: they have numerous small cracks in their surfaces and
are not single crystals, but often aggregates of small crystals and broken pieces in all pos-
sible orientations, with some amorphous materials in the interstices. Even if the solid is
wholly crystalline, and consists of a single crystal, there may be many different types of
surface on it; the faces, edges and corners will all be different. A metallic, covalent and ionic
crystal is regarded as a single giant molecule. The atoms of a solid crystal (or molecules in
a molecular crystal) are practically fixed in position. They stay where they are placed when
the surface is formed, and this may result in no two adjacent atoms (or molecules) having
the same properties on a solid surface. Crystal defects affect the crystal’s density and heat
capacity slightly, but they profoundly alter the mechanical strength, electrical conductivity
and catalytic activity. In a simple metal surface, there are many possible types of position
and linkage of the surface atoms, which confer different chemical properties in different
regions. For materials other than metals, the possibilities of variety of surface structure
from atom to atom must be far greater through the formation of elevations and cavities.
The presence of cracks in a surface decreases the strength of the crystals from that deduced
from theoretical considerations. The loss of mechanical strength may be due to surface
imperfections or internal cracks.
Wettability and repellency are important properties of solid surfaces from both funda-
mental and practical aspects. Wettability is the ability of a substrate to be covered with
Introduction to surfaces and Interfaces 7

water or other liquids. Wettability can be seen when a liquid is brought into contact with
a solid surface initially in contact with a gas or another liquid (see Chapters 9, 10). Many
processes may happen: this liquid may spread on the substrate without any limit, displac-
ing the entire original fluid surface from the entire solid surface area available, corres-
ponding to a contact angle of 0° (see Section 9.1 on the definition of contact angles).
Alternatively, the liquid may move out over the solid displacing the original fluid; however
it halts when the contact angle between the liquid–fluid and solid–liquid interfaces reaches
a certain value. This shows the magnitude of the wettability of the substrate with this liquid.
Lastly, there is no wettability such that the substrate repels the liquid forcing it to form a
spherical drop corresponding to a contact angle of 180°. Thus, wettability can be deter-
mined by measuring the contact angle of a water drop resting on a solid substrate. When
the effects of surface stains or adsorption of other materials are ignored, the wettability of
the solid surface is a characteristic material property and strongly depends on both the
surface energy arising from the surface chemical structure and the surface roughness.
Wenzel and later Cassie and Baxter provided different expressions showing the relation-
ship between the contact angle and the roughness of a solid surface. The determination of
wettability is important in adhesion, detergency, lubrication, friction, coating operations,
flotation, catalysis and many other processes in chemical, mechanical, mineral, metallurgy,
microelectronics, biomedical and biological industries.

1.5 Chemical Heterogeneity of Solid Surfaces


The chemical structure of the top surface layers of a solid determines its surface free energy.
If these top surface layers consist of the same chemical groups, it is called chemically homo-
geneous, and if they consist of different chemical groups, it is called a chemically hetero-
geneous surface. The presence of two or more chemically different solid substances in a
surface layer enormously multiplies the possibility of variety in the type of surface. Copoly-
mer surfaces and catalysts having many different atoms at the surface are good examples
for this. If desired, it is possible to impart chemical heterogeneity locally, for example con-
centric cylinders, stripes, patches etc. for various special applications. Measuring the
contact angle hysteresis of liquid drops on a surface or measuring the heat of adsorption
can check the presence of chemical heterogeneity (see also Chapter 9). The surface free
energy is different for each region (or patch) for chemically heterogeneous surfaces and
during an adsorption process, the higher free-energy spots (or patches) are filled first. Some
dynamic experiments, for example adsorption chromatography, give information on sur-
face heterogeneity.
Chemical heterogeneity of a surface is an important property affecting adhesion, adsorp-
tion, wettability, biocompatibility, printability and lubrication behavior of a surface. It seri-
ously affects gas and liquid adsorption capacity of a substrate and also the extent of a
catalysis reaction. As an example, the partial oxidation of carbon black surfaces has an
important, influence on their adsorptive behavior. In a chemically heterogeneous catalyst,
the composition and the chemical (valence) state of the surface atoms or molecules are
very important, and such a catalyst may only have the power to catalyze a specific chemi-
cal reaction if the heterogeneity of its surface structure can be controlled and reproduced
during the synthesis. Thus in many instances, it is necessary to determine the chemical
8 Surface Chemistry of Solid and Liquid Interfaces

composition, structural and electronic states and bonding properties of molecules at the
solid surface. Such a complete surface investigation can be done by applying several ad-
vanced spectroscopic surface analysis techniques.

References
1. Adam, N.K. (1968). The Physics and Chemistry of Surfaces. Dover, New York.
2. Adamson, A.W. and Gast, A.P. (1997). Physical Chemistry of Surfaces (6th edn). Wiley, New York.
3. Lyklema, L. (1991). Fundamentals of Interface and Colloid Science (vol I). Academic Press, London.
4. Atkins, P.W. (1998). Physical Chemistry (6th edn). Oxford University Press, Oxford.
5. Murrell, J.N. and Jenkins, A.D. (1994). Properties of Liquids and Solutions (2nd edn). Wiley,
Chichester.
6. Aveyard, R. and Haydon, D.A. (1973). An Introduction to the Principles of Surface Chemistry.
Cambridge University Press, Cambridge.

You might also like