InPhChMa (RoJNa08)
InPhChMa (RoJNa08)
InPhChMa (RoJNa08)
PHYSICS and
CHEMISTRY
of MATERIALS
Robert J. Naumann
INTRODUCTION TO THE
PHYSICS and
CHEMISTRY
of MATERIALS
INTRODUCTION TO THE
PHYSICS and
CHEMISTRY
of MATERIALS
Robert J. Naumann
This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been
made to publish reliable data and information, but the author and publisher cannot assume responsibility for the valid-
ity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright
holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this
form has not been obtained. If any copyright material has not been acknowledged please write and let us know so we may
rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or uti-
lized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopy-
ing, microfilming, and recording, or in any information storage or retrieval system, without written permission from the
publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://
www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923,
978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For
organizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for
identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
Contents
Preface........................................................................................................................................... xxi
Author......................................................................................................................................... xxv
v
vi Contents
Bibliography................................................................................................................................ 223
Problems ...................................................................................................................................... 223
Index............................................................................................................................. 535
Preface
The purpose of this book is to prepare students from diverse engineering and science
backgrounds for advanced study in materials science. Since the Huntsville campus did not
offer an undergraduate program in materials science when the Tricampus Materials
Science PhD program was instituted between the Huntsville, Birmingham, and Tuscaloosa
campuses of the University of Alabama system, it was customary to require students with
physics and chemistry degrees to audit an undergraduate engineering course in structures
and properties of materials in order to acquaint them with the basic concepts of materials
science and engineering. Similarly, engineering students, with little or no background in
quantum mechanics found taking a graduate course in solid-state physics from texts such
as Kittel or Ashcroft and Mermin somewhat daunting.
I felt the students would be better served by offering a course at the first year graduate
level that would build on their undergraduate backgrounds in mathematics and whatever
science and engineering courses they may have had, and give them the background they
need for more advanced work in materials science at a level commensurate with their
education. The result was a two-semester core course in our program: the first semester
emphasized chemical bonding, crystal structure, mechanical properties, phase transform-
ations, and a brief introduction to the processing of materials, while the second semester
dealt with the thermal, electronic, photonic, optical, and magnetic properties of materials.
The topics discussed in this book are arranged accordingly.
For students with little or no background in modern physics or quantum mechanics, a
chapter has been added at the beginning to introduce quantum concepts and wave
mechanics through a simple derivation of the Schroedinger equation, the electron-in-a-
box problem, and the wavefunctions of the hydrogen atom. Clearly, the electron-in-a-box
problem is essential to understand the Fermi energy in a metal or a semiconductor.
Knowledge of hydrogen-like wave functions is needed to understand chemical bonding
and the periodic table. I especially wanted students to appreciate the mathematical origin
of the quantum numbers that specify the various quantum states so they do not seem to be
simply pulled out of the air. I also tried to give an historical prospective to the development
of this knowledge so the student will have at least heard the names of some of the giants on
whose shoulders we stand.
In treating the chemical bond, hydrogen-like wavefunctions are also used to derive
the sp2 and sp3 hybrid bonds essential to understanding the structure and properties of
graphite, nanotubes, and semiconductors. Techniques for evaluating the Madelung sum
are discussed and the Coulomb potential is used to compute the lattice energies of ionically
bonded systems. At this level, the metallic bond has to be treated somewhat empirically,
but it is shown that the cohesive energy per volume for simple metals is roughly propor-
tional to the electron density and, like the ionic bond, is inversely proportional to approxi-
mately the fourth power of the atomic radius. The energy per volume of the transition
metals is shown to peak when the d-band is about half filled, and the covalent bonds
become saturated as would be expected.
The treatment of structure emphasizes the role of size and bond type in determining the
structure of materials. Point and space groups are introduced as well as Pearson and
Strukturbericht notations so the student would at least recognize these symbols in the
xxi
xxii Preface
literature. Drawings to help visualize stacking in diamond and wurtzite structures are
included. Unusual structures found in electron-deficient metals are briefly mentioned
along with those of selected polymers.
Concepts of the reciprocal lattice and Brillouin zones are introduced in the chapter on
X-rays along with a more formal treatment of the scattering of X-rays or wave-like
particles. This formal treatment for obtaining the Laue conditions for constructive inter-
ference leads to a deeper understanding of the scattering process and gives students the
ability to calculate the structure factors for various materials. The applications of X-ray
diffraction to material identification, structure determination, and crystal characterization
are briefly discussed.
Elastic behavior is introduced with the stress and strain tensors and the elastic coefficient
matrix and it is shown that for a material with cubic symmetry the coefficient matrix
reduces to only three elastic coefficients. Relationships between the various elastic moduli
are derived. The concept of a simple central force potential such as the Coulomb potential
is taken to the limits of its validity and is used to obtain the bulk modulus and to estimate
the thermal expansion of ionically bonded materials. The Coulomb potential is then
modified to represent a unidirectional potential and used to estimate the elastic coefficients
of these materials. The Morse potential is used to relate the elastic constants of metals to
their theoretical strengths.
The role of defects in solids is introduced with a derivation of vacancy defects followed
by a discussion of other point defects. Dislocations are introduced as an explanation of why
materials fall short of their theoretical mechanical properties, setting the stage for various
strengthening mechanisms introduced in later chapters that are designed to inhibit the
motions of dislocations. The concept of surface energy is introduced by using the Made-
lung sum for a (100) plane of an NaCl structure with the Madelung sum for the solid to
obtain the surface energy excess.
Methods for mechanical testing of materials are briefly introduced along with various
strengthening mechanisms. The number and surface area of the slip systems in metals and
in ceramics are shown to be responsible for the ductility (or the lack of it) and for ductile-to-
brittle transitions. Griffith’s theory of brittle fracture is used to introduce fracture mechan-
ics and to develop the concept of fracture toughness. The viscoelastic behavior of polymers
is briefly discussed.
The concept of free energy is used to describe single and multiple component phase
transitions. The classical theory of nucleation is presented along with various under-
cooling methods and their role in the development of bulk glassy metals. The role of
excess heat of mixing is demonstrated by calculating simple binary phase diagrams for
solid solution systems, intermetallic systems, and eutectic systems from their free ener-
gies. Peritectic and monotectic reactions along with their solid-state counterparts are also
described in terms of their free energies. Actual phase diagrams of similar systems are
presented as examples.
The problem of macrosegregation during alloy solidification from the melt is presented
along with methods for dealing with it. The constitutional supercooling criterion for
maintaining a planar solidification interface is developed along with a description of the
microsegregation that results when the interface breaks down into dendritic growth.
A chapter introducing the Bose–Einstein, Maxwell–Boltzman, Planck, and Fermi–Dirac
distribution functions follows before discussing the thermal, electronic, magnetic, and
optical properties for the benefit of students who have not been exposed to quantum
statistical mechanics. This chapter is a logical beginning for the second half of this book
since these concepts are essential to an understanding of these properties. Similarly, the
Maxwell equations are used to derive the equations for absorption and normal reflection of
electromagnetic waves in the chapter on optical properties. The band structure of metals
Preface xxiii
and semiconductors is examined and the relation between absorption spectra and inter-
band transitions is shown.
Recent developments in organic electronics and in polymer light-emitting diodes are
discussed along with the use of superlattices to construct quantum confinement devices
such as resonance tunneling diodes and quantum well cascade lasers. Recent develop-
ments in memory storage devices such as flash memories and magnetic storage based on
giant magnetoresistance phenomena are also taken into account.
Examples are provided to evaluate how well-simplified theories, such as assuming only
nearest neighbor interactions in lattice vibrations, predict the actual performance of various
materials. Problems are assigned wherever possible to help students apply what they have
learned.
Although the text is organized as a two semester introductory course in materials
science, it could equally serve as a one semester advanced undergraduate course in
structures and properties of materials by choosing the first 14 chapters. It could also
serve as a one semester advanced undergraduate course or first-year graduate course in
solid-state physics or chemistry by omitting chapters 9 through 14.
I would like to express my appreciation to my colleagues, Professor Michael Banish,
Professor James Baird, and Professor William Kaukler at the University of Alabama in
Huntsville and to Dr. Martin Volz, Dr. Frank Szofran, and Dr. Donald Gillies at the NASA
Marshall Space Flight Center for their suggestions and help in preparing this book. I also
thank Chantell Marsh for proofreading the manuscript. Most of all, I appreciate the
patience and understanding of my wonderful wife, Sally, for putting up with me during
the time it took to complete this book.
Author
Robert J. ‘‘Bob’’ Naumann received his BS, MS, and PhD in physics and mathematics from
the University of Alabama. After a long career as a scientist and a division chief in the
Space Sciences Laboratory at the NASA Marshall Space Flight Center, he joined the
University of Alabama in Huntsville as a professor of materials science. In addition to
his teaching duties, he also served as an interim director of the Center for Materials
Research and as the director of the Tricampus Materials Science PhD program between
the University of Alabama campuses at Tuscaloosa, Birmingham, and Huntsville.
xxv
1
Introduction to Materials Science
1
2 Introduction to the Physics and Chemistry of Materials
During the Neolithic Period (the New Stone Age) humans gradually undertook domes-
tication of animals, cultivation of crops, production of pottery, and building of towns, such
as Jericho, by around 7000 BC. Copper tools and carved ivory found in Palestine are
believed to date between 5000–4000 BC. Bronze was used by the Canaanites as early as
3000 BC and walls of their towns were fortified with a plaster-like material.
The ancient Romans made extensive use of concrete. The oldest surviving concrete
structure is the Temple of Vesta, built in Tivoli during the first century BC. Other examples
are the brick-faced concrete walls of the Camp of the Praetorian Guard, built by Sejanus in
AD 21–23, the octagonal domed fountain hall of Nero’s Golden House (AD 64–68), and
Hadrian’s Pantheon of about AD 118–128, whose solid concrete dome is 43.2 m in diameter
and 1.5 m thick and is supported by 6 m thick walls of brick-faced concrete.
Battles have been won and lost because of subtle differences in arms and armor. The
Greeks defeated the Persians at the battle of Marathon in 490 BC because their bronze
armor had been annealed to a soft condition, which maximized its work of fracture. As a
result, the Persian arrows only dented the Greek armor and did not penetrate far enough to
cause a fatal wound. Similarly the English won the battle of Crècy in 1346 and the battle of
Agincourt in 1415 partly because they used flax for their longbow strings. Flax has a higher
Young’s modulus than the animal sinew used by the French, which gave their arrows
greater kinetic energy. Also the expensive and prestigious iron armor of the French knights
contained a high carbon content, making it brittle and easily penetrated by the hard-hitting
arrows from the English longbows.
Although the French naval architecture and shipbuilding was usually superior to the
British during the eighteenth century, their cast iron naval cannons were generally less
reliable than those used by the British and the bursting of the French cannons was
influential in the 1805 battle of Trafalgar.
Copper was the first metal used by man because it can be found naturally in its
uncombined metallic state. Robert Peary, while exploring Greenland in 1894, discovered
that the local Eskimos had been breaking off chunks of a copper meteorite they had found
and were using the metal for tools and weapons. Although rare, other copper meteorites
have been found throughout the world and had apparently provided ancient civilizations
with this metal. Its use was known in eastern Anatolia (the Asian portion of Turkey,
a crossroad of civilization in its day) as early as 6500 BC. Copper was being extracted
from a quarry near Varna, Bulgaria around 4400 BC. The Egyptians were casting copper in
molds as early as 4000 BC. Much of the early Roman supply of copper came from
Cyprus (Cyprium is Latin for copper meaning metal from Cyprus) where, somewhere
around 3500 BC, the Cypriots had learned to extract copper from copper pyrites
(CuSFe2) by reducing the ore in charcoal fires. Since the melting point of pure copper is
fairly high (10838C), they also discovered, probably by accident, that the addition of small
amounts of tin would lower the melting temperature, make the material easier to cast,
and improve the strength of the material. Thus the Chalcolithic (Copper-Stone) Age gave
way to the Bronze Age.
Although the use of iron as a precious metal was known as far back as 3000 BC, its
superiority over bronze was not recognized until around 1200 BC in Europe and the
Middle East. Since charcoal or coke was heated with iron ore to reduce it to iron, the
resulting cast iron had excess carbon, which caused it to be brittle. Even so, the widespread
production of iron for tools and weapons changed the face of Europe as Asia for the next
2000 years.
Early ironworkers removed this excess carbon by heating and hammering. Japanese and
Syrian sword makers repeatedly folded layers of iron over each other and continued to
hammer to break up the carbon so it could combine with oxygen to form carbon monoxide,
although they probably did not know what they were actually doing. The result was a
Introduction to Materials Science 3
fairly soft but very tough sword that would bend before it would break. Then to harden the
edge, they packed the sword in charcoal and reheated it to diffuse the carbon back
into the cutting region to harden it. The ‘‘heat and beat’’ procedure was effective,
but very expensive and such handcrafted, high-quality steel implements were limited
to the affluent.
Despite the long history of the use of metals in ancient civilizations, the art of metallurgy
was passed on from country to country along trade routes. It was not until 1530 when
Georgius Agricola wrote de re Metallica (published posthumously in 1556), which described
the practice of mining and smelting of ores, that a comprehensive survey of the art of
metallurgy existed.
In 1784, Henry Cort discovered that it was possible to stir small ‘‘pigs’’ or puddles of
iron to mix the oxide forming on the surface so that it could combine with the excess
carbon, a process called puddling. This process eliminated the expensive process of
hammering the iron to remove the excess carbon, which made good iron more affordable
and, with the invention of the blast furnace by Abraham Darby, became the forerunner of
modern steel making.
The Bessemer converter, invented by Sir Henry Bessemer in 1856, removed the excess
carbon from the iron by blowing air through the molten pig iron and made the economical
production of steel in tonnage quantities possible. The availability of large quantities of
relatively inexpensive iron and steel made our present industrial society possible.
In the beginning of the twentieth century, the structures and machines generally were
massive enough that relatively little thought was given to analyzing the actual strength of
the structures. Even so, catastrophic failures did occur. Perhaps the most noteworthy was
the sinking of the Titanic in which it is now believed that the steel used for her plates may
have undergone a ductile-to-brittle transition at the temperature she was operating in.
Such a transition could cause a relatively small crack in her hull to propagate and cause a
catastrophic failure.
With the advent of aircraft and rockets, a great deal more attention had to be paid to
carefully choosing materials and designing structures to provide just the needed strength
with a minimum of weight. The famous Japanese Zero fighter plane that proved so
effective in the early days of World War I owed its agility to its very light weight which
was achieved using a very high strength-to-weight precipitation-hardened aluminum
alloy. This alloy loses it strength over time due to overaging or ripening of the precipitates
so that they are no longer a strengthening mechanism. As a result, such aircraft (of the few
that survived) are no longer considered flight-worthy. Of course long service life was not a
major design consideration for a combat aircraft.
Even in more recent times, catastrophic materials failures still occur. Three DeHavilland
Comets (the first jet-powered commercial airliner) disintegrated in mid-air because
their designers had forgotten (or failed to learn) the lessons their colleague, A.A. Griffin,
at the Royal Aircraft Establishment at Farnborough, had taught concerning fracture
mechanics in 1920.
strong chemical bonds make them refractory, hard, and brittle. Semiconductors are a
special class of ceramics whose valence electrons can be readily promoted to conduction
electrons which makes them the basis for modern electronic and photonic devices. Poly-
mers are perhaps the most versatile class of materials because of the virtually infinite
combinations with which polymer chemists can assemble the molecules that make up
polymers. Like ceramics, their valence electrons tend to be tied up in forming chemical
bonds and some of the best electrical insulators are polymers. However, the recent discov-
ery of conductive polymers has opened a new field of molecular electronics with the
promise of large flexible TV screens or display panels.
Glasses and composites are important forms of materials but are not considered as
specific classes of materials. Composites are simply a mixture of two or more of the
other types of materials combined in such a way as to reinforce the favorable qualities
and reduce the unfavorable qualities of the components for a specific application. Most
metals and ceramics form a polycrystalline structure when they solidify. However, it is
possible to cause both metals and ceramics to solidify in an amorphous state without the
order associated with the crystalline structure. Such an amorphous solid is called a glass.
1.4.2 Ceramics
Ceramics are generally compounds formed from two or more elements that are held
together by ionic or covalent bonds, or a mixture of both. Ionic bonds generally form
between elements on the opposite sides of the periodic table where there is a large
difference in electronegativity. Electronegativity difference is the tendency of an atom,
near the right-hand side of the periodic table, to grab one or more electrons from a metal
atom with loose electrons to become a negatively charged anion. The resulting positively
charged metal ion, or cation, is electrically attracted to the anion. Again, the resulting
ionically bonded molecules form crystalline structure comprised of repeating unit cells
whose structure depends on the ratio of the cation to anion diameter.
Introduction to Materials Science 5
Since all of the electrons are tied up in forming chemical bonds, there are no free
electrons available to carry electricity and heat. Consequently, ceramics are generally
poor conductors of heat and are good electrical insulators (an important exception is the
new class of high-temperature superconductors). They are generally transparent to light as
single crystals or glasses but will be translucent or opaque if the structure is polycrystalline
(many small crystalline grains with random orientation) because of scattering of light by
the grain boundaries.
The ionic bond is very strong which explains why ceramics generally have high melting
temperatures. Since the anions are generally larger than the cations and are oppositely
charged, it is difficult for the ions to slide over one another, which explains why ceramics
are hard and brittle. Trying to find ways to make ceramics less brittle, so that that their high
strength at elevated temperatures can be utilized, is one of the goals of modern ceramists.
1.4.3 Semiconductors
Semiconductors are a special case of ceramic materials so designated because the energy
required to free a valence electron and make it a conduction electron is on the order of the
energy of a photon of visible light. This small energy gap between the valence and
conduction bands allows some electrons to be thermally excited into the conduction
band at room temperature, making this class of materials poor conductors or semicon-
ductors. However, it is possible to dope intrinsic (pure) semiconductors with impurity
atoms to form donor states in order to create a large number of conduction electrons to
form n-type material, or to form acceptor states in order to create holes in the valence band
to form p-type material. Junctions between n- and p-type materials can be formed which
are the basis for diodes, transistors, and other devices essential to the modern electronics
revolution. Since the bandgap is on the order of infrared or visible light, photoelectric
detectors can be made from these materials. Conversely, light emitted when electrons fall
from the conduction back to the valence band is the basis for light-emitting diodes (LEDs)
and solid-state lasers.
Elemental semiconductors are the lower atomic number members of the group IV
family, i.e., carbon, silicon, and germanium that form sp3 hybrid covalent bonds.
A covalent bond forms when two atoms share their electrons to fill an electron shell.
Group IV elements have two s-electrons that sit at a lower energy level than their two
p-electrons (s- and p-designate the subshells that the valence electrons occupy). In the
heavier group IV elements, e.g., tin and lead, these outer two p-electrons detach themselves
to form a metallic bond. But in the lighter elements, one of the s-electrons is promoted to
the p-level where it can interact with other s- and p-electrons to form the very strong
directional sp3 bond. The extra energy required to promote the s-electron to p-level is
regained in the overall energy of the sp3 bond.
Covalent bonds are directional and in the case of the sp3 hybrid bond, they form an open,
diamond-like crystal structure. The diamond form of carbon is also formed from the sp3
bond that is extremely strong, which is why diamond is so hard. (Diamond is not generally
considered to be a semiconductor because its bandgap is much larger than visible light.)
The sp3 bond gets progressively weaker as the atoms in the group IV family get larger
and almost disappears with tin, the next member above germanium. Compound semicon-
ductors are formed by combining elements from group III (aluminum, gallium, indium)
with those from group V (phosphorous, arsenic, antimony) to form gallium arsenide,
indium phosphide, etc., or from group II (zinc, cadmium, mercury) with group VI (sulfur,
selenium, tellurium) to form mercury telluride, cadmium selenide, etc. These compounds
also form open structures with sp3-directed covalent bonds, although these may be mixed
with some ionic bonding.
6 Introduction to the Physics and Chemistry of Materials
Modern technology has made it possible to combine these materials into structures that
can be controlled at the atomic scale to form exotic new photonic and electronic devices
such as the new highly efficient white LEDs, highly efficient double heterojunction lasers,
single electron transistors, quantum wires, quantum dots, etc.
1.4.4 Polymers
Polymers can be thought of as a bowl of spaghetti, the noodles representing long chains of
repeating groups of molecules called ‘‘mers,’’ hence the name ‘‘polymer.’’ These mers in
the chains are covalently bonded together so that the chains themselves are quite strong.
However, generally the chains are only loosely bonded to each other through secondary
van der Waals-type bonds, which allows some degree slipping past each other, especially
at elevated temperatures, hence the plastic nature of polymers. The mechanical properties
of polymers can be controlled by the length of the chains (average molecular weight) and
by the side groups of the mer units that can make it more difficult for them to slide past one
another. Polymers are generally amorphous, meaning they have no regular structure,
although in some circumstances the chains can fold back and forth on themselves to
form regions of crystallinity in the form of ordered repeating units in a three-dimensional
array. Side chains can be attached to lower the interaction between chains by keeping the
farther apart if a noncrystalline, low density, fairly soft material is required. For applica-
tions requiring the material to resist deformation, the chains can be cross-linked by opening
bond along the length of the chains (this occurs naturally in some plastics upon exposure to
ultraviolet light) or by adding sulfur atoms to form disulfide bridges between the chains
(vulcanization of rubber). The wide variety of molecules available to the polymer chemist
and nearly infinite possible ways in which they can be arranged has produced a remark-
able array of produces from soft fabrics to bulletproof vests, from plastic wrap to high-
strength polycarbonate windows, and from rubber bands to truck tires.
Since all of the valence electrons are involved in forming chemical bonds, for the most
part, polymers are extremely good insulators. However, things can be arranged so that
charge transfer can take place along the chains so that it is possible to make conductive
polymers. Recently, materials scientists have been able to form molecular diodes and
transistors, opening up an exciting new field of molecular electronics. In the future, you
may unroll a large flat screen TV and roll it back up when you are no longer watching it.
1.4.5 Glasses
Glass is not a type of material in the sense of metals, ceramics, or polymers, but refers to the
amorphous state of a material. Generally, metals and ceramics in the solid phase prefer to
arrange themselves in a crystalline structure, meaning their atoms are arranged in regular
unit cells that repeat over and over in each direction. This repetition of the unit cells
produces what is called long-range order, meaning the atoms in one unit cell will be
arranged just as they are in any other cell many cell spacings away. Such an arrangement
allows all of the bonds to be satisfied and the solid material is said to be its lowest
energy state.
At elevated temperatures, the thermal energy is sufficient to break some of these bonds
allowing the atoms (or molecules) to move about in a random fashion. Long-range order is
destroyed and the system is said to be amorphous. In one sense, a glass can be thought of
as a liquid that has been frozen in place.
The lack of long-range order in a glass has its advantages. There are no grain boundaries
to scatter light so glasses tend to be transparent, making them an inexpensive substitute for
single crystals for applications such as windows. Since grain boundaries are attacked
Introduction to Materials Science 7
preferentially by corrosive chemicals, their absence makes glasses less vulnerable to chem-
ical attack. The random network makes it more difficult for dislocations to form and move,
making glasses strong but brittle.
In order to form a crystalline solid, one or more nucleation events must take place in
which an embryonic crystal is formed. This embryo forms a sort of template on which the
other atoms can arrange themselves to produce the long-range crystalline order. As a melt
is cooled, its viscosity increases rapidly so that the atoms or molecules cannot move about
as readily. To form a glass from the melt, it is necessary to delay nucleation until the
temperature falls to the point where the viscosity becomes high enough to prevent the
atoms or molecules from arranging themselves in an orderly manner.
Ceramic melts tend to be more viscous than metals, hence they are generally good glass
formers, especially eutectic mixtures of ceramics, which have a lower melting point than
their individual components. The glasses we are most familiar with are ceramics, usually
oxides of silicon with various additives to tailor the properties of the glass for a specific
application. Such glasses are transparent because the grain boundaries that would scatter
the light in a polycrystalline solid are absent in an amorphous solid. However, it is also
possible to form metallic glasses which, because of their free electrons, are not transparent.
Wood is another of nature’s finest products and is certainly the oldest construction
material used by man. It is a remarkable composite composed of soft cellulose and hard
lignin. It has a tensile strength per unit mass equal to that of mild steel and is a tougher per
unit mass than the toughest steels. Its unusually high work of fracture is due to the soft
cellulose which causes it to split easily along the grain. These soft layers tend to blunt crack
propagating across the grain.
Muscles provide nature’s motive power by converting chemical energy directly to
mechanical energy. Since they are not heat engines, they are not limited by the Carnot
efficiency. (If they were, we could not get out of bed.) However, they are only 30%
efficient, making them about as efficient as most internal combustion engines. Since their
efficiency is not limited by something as fundamental as the Carnot efficiency, there could
be substantial payoff if higher efficiency artificial muscles could be produced.
1.5.2 Polymers
Perhaps the greatest advances in materials made during the twentieth century have been in
polymer science. The invention of nylon in 1938 by Wallace Carothers at Dupont laid the
foundation for the synthetic fiber industry and the miracle fabrics we enjoy today. Leo
Baekeland developed a method for producing bakelite (phenol formaldehyde), the first
thermosetting plastic, in 1909. This material was the forerunner of the enormous plastics
industry that developed during and after World War II. Other polymer products from
synthetic rubber to adhesives and coatings have found their way into virtually every aspect
of our present day lifestyle.
In 1977, Alan Heeger, Alan MacDiarmid, and Hideki Shirakawa were awarded the
Nobel Prize for their work on the production of high electrical conductivity in poylace-
tylene, paving the way for the exciting new field of molecular electronics. Shortly
thereafter, molecular diodes were fabricated, followed by molecular transistors. In 1987,
Ching Tang and Steven Van Slyke at Eastman Kodak developed small-molecule organic
LEDs (OLEDs). In 1990, Jeremy Burroughes at the University of Cambridge developed
large molecule or polymer LEDs (PLEDs) that can emit brighter colors over the entire
spectrum, making flat screen video and other displays possible. Some of the OLED
displays are already on the market in cell phone and other small displays. Large size
flexible displays are still several years from market. One difficulty with such devices is
that they are relatively inefficient, as were the early semiconductor LEDs. In 1999, Steven
Forrest at Princeton University and Mark Thompson at University of Southern California
found a way to control the spin of the electrons so that all of them could emit light
instead of heat.
More recently, electroactive polymers have been developed that change shape in
response to electrical stimulus. A spin-off company, Artificial Muscles, Inc. (Sunnyvale,
California) has been created to commercialize the pioneering work by the Stanford
Research Institute in this area.
1.5.3 Superconductors
When the first superconductor with a transition temperature of 95 K was discovered
by Jim Ashburn and M.K. Wu at the University of Alabama in Huntsville in 1986, there
was great excitement in the field since the highest previous transition temperature was
only 23 K. The implications of this discovery was that now superconductors could operate
with liquid nitrogen that boils at 77 K instead of the much more expensive and difficult to
store liquid helium that boils at 4 K. Also, the fact that this was accomplished with a
ceramic material (YBa2Cu3O7 ) rather than the traditional metallic systems (Nb3Ge and
Introduction to Materials Science 9
Nb46.5 wt% Ti) suggested a possible new mechanism for superconductivity that was
stronger than the electron–phonon coupling described by the Bardeen–Schrieffer–Cooper
theory and held out hope for the possibility of a room temperature superconductor.
Despite extensive work in the field of high-temperature superconductivity (HTSC), the
mechanism involved is still not completely understood and the transition temperature has
been stalled at 134 K. Unfortunately, even with their high transition temperatures, the new
HTSC materials have found only limited use because their brittle nature makes it difficult
to fabricate them into wires for making large superconducting magnets for high-field
magnetic resonance imaging (MRI) and other applications that are presently fulfilled by
conventional Nb46.5 wt% Ti superconducting magnets cooled by liquid helium.
1.5.7 Fullerenes
In 1985, Richard Smalley at Rice University discovered the first fullerene, a hollow sphere
that comprised of 60 carbon atoms, so named because its structure resembles the geodesic
structure designed by Buckminster Fuller. Since then nanotubes or buckey tubes have
been discovered that look like rolled-up chicken wire. Single-wall nanotubes (SWNTs) are
extremely strong and have, close to the theoretical strength of the sp2 carbon bond, one of
the strongest chemical bonds. At present, these nanotubes are only a few nanometers long,
but attempts are being made to extend their lengths so they can be used as reinforcing
fibers or perhaps woven into cables.
SWNTs have other interesting properties. They can be semiconductors or insulators
depending on their chirality (whether their structure forms left- or right-handed helices).
They are superconductors with modest transition temperatures. Properly doped, they have
possible application for hydrogen storage.
1.5.8 Semiconductors
The invention of the transistor, in 1947, by John Bardeen, Walter Brattain, and William
Shockley of the Bell Laboratories, as a replacement for the bulky energy consuming vacuum
tube, promised to revolutionize the world of electronics. But the early transistors were
individually made and were unsuitable for high-frequency applications. The real break-
throughs in technology that enabled the mass production of the large variety of inexpensive
electronic components that range from watches to computer motherboards had to come
about in the production of large diameter, dislocation-free single crystal silicon wafers with
purities of less than 1 part in 10 trillion. Also needed was the development of lithographic
technology that could produce patterns with submicron details. Advances in this later
technology, brought about by collaborations between polymer chemists and electronic
and optical engineers, has made possible the ever-increasing scale of device integration
predicted by Moore’s law (the number of transistors on a chip will double every 18 months).
1.5.10 Photonics
Similarly, the development of extremely high purity silica fibers and new photonic mater-
ials that provide laser transmitters and receivers have revolutionized landline communi-
cation systems. Other developments in photonic materials have made possible highly
efficient LEDs in all colors that are being used in traffic lights and other displays. Very
small efficient short wavelength solid-state lasers have also made possible massive optical
storage which enabled the development of CDs and DVDs.
Bibliography
Ashley, S., Artificial muscles, Sci. Am., Oct. 2003, 53–59.
Encyclopedia Britannica, Deluxe CD-ROM, 2004.
Gordon, J.E., The Science of Structures and Materials, Scientific American Library, New York, 1988.
Howard, W.E., Better displays with organic films, Sci. Am., Feb. 2004, 76–81.
2
Fundamental Principles
Most students will have had the material in this chapter somewhere in their undergraduate
curriculum; it will not hurt for them to see it again and perhaps see it in a different light.
For those readers who may be unfamiliar with this material, we will not make heavy use of
it in the rest of the book, but it is the basis of much of what we will be doing, so it will help
to have some understanding of the basic principles that govern the behavior of matter. The
material is presented with a historical prospective so that the reader might appreciate how
this knowledge evolved.
where
R is the empirical Rydberg constant ¼ 109,677.58 cm1 for hydrogen
n1 and n2 can take on only integer values (see Figure 2.1)
13
14 Introduction to the Physics and Chemistry of Materials
n2 = 2
n2 = 1
n1 = 1 n1 = 2 n1 = 3
Final state
FIGURE 2.1
Schematic of the transitions between the energy levels in the hydrogen atom. The transition from n2 ¼ 2 to n1 ¼ 1
emits a photon with a wavelength of 121.5 nm which is called the Lyman-a for short because it is the principal line
in the series. The transition from n2 ¼ 3 to n1 ¼ 2 emits a photon with a wavelength of 656.3 nm and is usually
referred to as the H-a line instead of the Balmer-a line.
For
n1 ¼ 1; n2 ¼ 2, 3, 4, Lyman series (ultraviolet, 121.5 nm)
n1 ¼ 2; n2 ¼ 3, 4, 5, Balmer series (visible, 656.3 nm)
n1 ¼ 3; n2 ¼ 4, 5, 6, Paschen series (infrared 1.875 mm)
n1 ¼ 4; n2 ¼ 5, 6, 7, Bracket series (infrared 4.050 mm)
n1 ¼ 5; n2 ¼ 6, 7, 8, Pfund series (infrared 7.400 mm)
Before the twentieth century, there was no theoretical explanation for these observed
spectral lines or why other atoms should produce their own characteristic spectra. Finally,
in 1913, Niels Bohr developed the first theoretical model that offered a basis for under-
standing these observations.
Ze2 m e v2
Coulomb force ¼ 2
with the centrifugal force ¼
4p«0 r r
2
Ze Ze2 me
to obtain the velocity v2 ¼ and the momentum p2 ¼ :
4p«0 me r 4p«0 r
Ze2
KE þ PE ¼ : (2:3)
8p«0 r
Fundamental Principles 15
But classical theory allows any energy whereas the observed spectra suggested transitions
between discrete energy states. Also a charge moving around an orbit is constantly being
accelerated. Electromagnetic theory requires an accelerated charge to radiate energy. This
would cause the electron to quickly spiral into the nucleus! What is wrong here?
Here is where Bohr made the bold assumption that angular momentum was somehow
quantized in units of h=2p, where h was Planck’s constant that had been postulated in 1900
to explain black body radiation and was invoked by Einstein in 1905 to explain the
photoelectric effect. Orbits in which the angular momentum was an integral number of
h=2p were assumed to be stable against radiating energy and spectral emission (or
absorption) involved a photon with energy equal to the difference between such orbits.
Therefore, the angular momentum could be written as
nh
me vr ¼ pr ¼ ¼ n
h: (2:4)
2p
Ze2 me 2
p2 r 2 ¼ r ¼ n2 h2 , (2:5)
4p«0 r
1 Ze2 me
¼ , (2:6)
r 4p«0 n2 h2
Also, if you put the numbers into Equation 2.8 with Z ¼ 1, the
1 2p2 e4 me 1 1 5 1 1 1 1
¼ ¼ 1:097 10 ¼ R , (2:9)
l (4p«0 )2 h3 c n21 n22 n21 n22 n21 n22
where R ¼ 1.097 105 is the empirical Rydberg constant. What a major triumph! (Or not,
since the theory only works for hydrogen and is not in agreement with our present
understanding of the electron.)
The reader may be wondering why the observed spectra from the Sun were in the form
of dark lines instead of emission lines described in Figure 2.1. The answer is that the
observers were viewing the bright surface of the photosphere of the Sun through the cooler
solar atmosphere. The surface of the Sun emits a continuum of wavelengths referred to as
black body radiation, similar to an incandescent lamp. Photons with wavelengths corre-
sponding to the transitions shown in Figure 2.1 are absorbed in the cooler solar atmosphere
16 Introduction to the Physics and Chemistry of Materials
Solar atmosphere
Unabsorbed
Sun
wavelengths
FIGURE 2.2
Schematic showing how light that is
spectroscopically absorbed and rera-
diated shows up as a dark band Absorbed and reradiated
against a bright background. wavelengths
by kicking electrons from lower states n1 ¼ 1, n2 ¼ 2, to various excited states and the solar
atmosphere reradiates this energy by emitting photons corresponding to the transitions
shown in Figure 2.1. The re-emitted radiation is omnidirectional, so the radiation at these
frequencies seen by the observer is small compared to the radiation at other wavelengths
that does not get absorbed. Thus the solar atmosphere appears to absorb those wave-
lengths that correspond to allowed electronic transitions. Figure 2.2 attempts to show this
effect. Since the most plentiful elements on the Sun are H and He, their emission spectra in
the solar atmosphere exceed the black body radiation from the photosphere, although this
was not realized until 1925. (In fact, He was discovered on the Sun in this manner before it
was known on Earth.)
Actually, Bohr’s theory was an inspired (lucky) guess and, since it is valid only for the
hydrogen atom, it is not particularly useful as a general theory. Attempts were made by
Sommerfeld to extend the theory to two-electron systems, but without much success. To
move on, we must understand more about the electron.
assumption me vr ¼ pr ¼ hr=l ¼ nh=2p implies that 2pr ¼ nl, or that the allowed orbits
correspond to integral number of matter waves.
In 1927, Davisson and Germer demonstrated that electrons did indeed have wave-like
properties by showing that a beam of electrons directed toward a metallic single crystal
exhibited diffraction patterns, much like x-rays.
Amplitude
−10 −5 0 5 10 −10 −5 0 5 10
(a) Position/average wavelength (b) Position/average wavelength
FIGURE 2.3
(a) Particle is localized to 10l when Dl=l ¼ 0.1. (b) Particle is localized to 5l when Dl=l ¼ 0.2.
18 Introduction to the Physics and Chemistry of Materials
A corollary to the uncertainty principle involves energy and time. Since E ¼ p2=2m,
DE ¼ pDp=m and DxDp ¼ DxDEm=p ¼ DEDx=v ¼ DEDt h. This relationship is useful for
estimating the rest energies of particles with short lifetimes.
A formal method of dealing with the wave nature of particles was developed by Irwin
Schrödinger in 1926. A greatly simplified derivation of the time-independent Schrödinger
wave equation is presented in Section 2.3.
1 @2C
r2 C ¼ , (2:10)
c2 @t2
where
c is the velocity of propagation
C is the amplitude
where
v is the angular frequency
c(x, y, z) is the time-independent wavefunction
When Equation 2.10 is substituted back into the partial differential equation (PDE),
v2
r2 c þ c ¼ 0: (2:12)
c2
Since v ¼ 2pn ¼ 2pc=l, this can be written as
4p2
r2 c þ c ¼ 0: (2:13)
l2
m2 v2
r2 c þ c¼0
2
h
where
h
h¼
: (2:14)
2p
Now substitute mv2 ¼ 2(EV) where E is the total energy and V is the potential energy and
multiply by h=2m to get the time-independent Schrödinger wave equation that describes
the behavior of particle waves in terms of their wavefunction c,
Fundamental Principles 19
2 2
h
r c þ Vc ¼ Ec: (2:15)
2m
Hc ¼ Ec, (2:16)
where
H is the Hamiltonian operator defined as H ¼ (h2 =2m)r2 þ V
E represents the energy eigenvalues, i.e., those values of E for which a solution to
Equation 2.15 exists
h2 2
r c ¼ Ec, 0 x, y, z L: (2:17)
2m
where
k ¼ ~ikx þ~jky þ ~
kkz is the propagation or waveÐÐÐvector
A is a normalization constant chosen so that cc*dV ¼ 1
(We will make frequent use of the propagation vector k to represent the momentum of a
wave-like particle in the rest of this chapter. The scalar momentum p ¼ h=l ¼ 2ph=l ¼ hk
where k ¼ 2p=l.)
Putting this solution back into the differential equation,
2 2
h h2 2 h2 2
r c¼ kx þ ky2 þ kz2 c ¼ k ¼ Ec, (2:19)
2m 2m 2m
2 k 2
h p2
E¼ ¼
2m 2m
p ¼ hk: (2:20)
This is the energy of a free electron (not confined to a crystal). Note that this energy is
continuous (not quantized) and that it goes as k 2.
Now we apply the boundary conditions that require the wavefunction to vanish at the
walls of the box. From Euler’s formula, eix ¼ cos x þ i sin x, the solution may be written as
20 Introduction to the Physics and Chemistry of Materials
a product of sines or cosines depending on the choice of the leading coefficient. Since
cos(u) ¼ 1, this choice does not match the prescribed boundary conditions. Therefore, the
appropriate solution is
n px n py n pz
x y z
c ¼ A sin sin sin , (2:21)
L L L
where nx, ny, nz are integers. When this is put back into the Schrödinger equation, the
energy has discrete values given by
h2 p2 2 h2 p2
En ¼ 2
nx þ n2y þ n2z ¼ n2 , (2:22)
2mL 2mL2
2
where n2 ¼ n . n ¼ ~inx þ ~jny þ ~ knz .
The uncertainty principle tell us that if we confine an electron to dimension L, the
momentum uncertainty will be Dp ¼ h=L, which could be satisfied if p ¼ h=2L or if
p2 ¼ h2=4L2. This would correspond to an energy E ¼ p2 =2m ¼ h2 =8mL2 ¼ h2 p2 =2mL2 ,
which is the same as the first energy level given by Equation 2.18. Thus, in some cases,
the uncertainty principle can be used to estimate the lowest energy state of a system.
Now the problem becomes, what value of n2 corresponds to the lowest energy states of
N valence electrons we put into the crystal? We could add up the individual states. Let us
start with the lowest state with nx ¼ 1, ny ¼ 1, nz ¼ 1 or E111 which hold two electrons, E112,
E121, and E211 each hold two electrons: that accounts for the first 8 electrons. It should be
apparent we will never get to 1023 electrons this way, so we must find a better way.
Let us represent nx as points on the x-axis, ny as points along the y-axis, and nz as points
along the z-axis. The lowest energy states then become the values of nx, ny, nz that are
included inside the sector of a sphere with radius n ¼ (n2x þ ny2 þ nx2)1=2. The number of states
is just the volume of this segment of a sphere with radius n or
1 4p 3
Number of states ¼ n, (2:23)
8 3
and since two electrons can occupy each state, (see Section 2.5.1)
2 4p 3 p 3
Number of electrons (Ne ) ¼ n ¼ n : (2:24)
8 3 3
Putting this back into the expression for energy, we find that the energy of the last electron
added to the crystal is
2 p2 3Ne 2=3 h2 p2 3Ne 2=3 h2 p2 3ne 2=3
h
EF ¼ ¼ ¼ , (2:25)
2mL2 p 2m pV 2m p
where ne is the number of valence electrons per unit volume. This EF is the ground state
energy of the system and is called the Fermi energy of the electrons. Since the electron
density in metals is on the order of 1029=m3, the Fermi energy is 5 eV, which is the same
order as the metallic bonding energy. Remember that 1 eV is the energy an electron has
when accelerated by a potential of 1 V or 1.6 1019 J. The thermal energy of an atom is
given by kT where k is Boltzmann’s constant ¼ 1.38 1023 J=atom-deg. Equating the two
energies and solving for T, we find that 1 eV is equivalent to 11,000 K, so the ground state
energy of the electrons in a metal corresponds to 55,000 K.
Fundamental Principles 21
Notice that the energy En increases as n2. The density of states N(E) (the number of states
with energies E and E þ dE) can be obtained by differentiating Equation 2.24 to obtain
dN ¼ pn2 dn. Differentiating Equation 2.22, to obtain dE ¼ (h2p2=mL2)n dn,
dN p 2mL2 V 2m 3=2 1=2
N(E) ¼ ¼ n ¼ E : (2:26)
dE 2 h2 p2 2p2 h2
ÐF
E ÐF
E
EN(E)dE E3=2 dE
0 0 3
hE i ¼ ¼ ¼ EF : (2:27)
ÐF
E ÐF
E 5
N ðEÞdE E1=2 dE
0 0
2 1 @
h 2 @ 1 @2 1 @ @
r þ 2 2 þ sin u c ¼ ðE V Þc, (2:28)
2m r2 @r @r r sin u @f2 r2 sin u @u @u
where the reduced mass m ¼ M=M þ m has been substituted for m. (This allows us to work
in the center-of-mass coordinate system.)
One method for attempting to find an analytical solution to a PDE is to see if the
variables can be separated. Let c(r, f, u) ¼ R(r)Q(u)F(f) where R(r) is a function of r only,
etc. Put this back into the PDE, divide by RQF, and multiply by 2mr2 sin2 u=h2 to obtain
1 @ 2 F sin2 u @ 2 @R sin u @ @Q 2m
þ r þ sin u þ 2 r2 sin2 uðE V ðrÞÞ ¼ 0: (2:29)
F @f2 R @r @r Q @u @u h
Note that the first term is a function of F only; therefore, it must equal be to some constant,
say m2. We now have an ordinary differential equation (ODE) for F, i.e.,
d2 F
¼ m2 F: (2:30)
df2
F ¼ Aeimf , (2:31)
or
Therefore,
eim2p ¼ 1: (2:34)
Since eim2p ¼ cos 2pm þ i sin 2pm ¼ 1, we see that m is restricted to values of 0, 1, 2, . . . .
The m is the azimuthal quantum number which, for reasons that become clear later, is also
called pthe
ffiffiffiffiffiffimagnetic or projection quantum number. The normalization coefficient is
A ¼ 1= 2p.
Now if we put the m2F back into Equation 2.26 @ 2F=@f2 and divide by sin2 u,
we get
m2 1 @ @Q 1 @ 2 @R 2m
sin u ¼ r þ 2 r2 (E V(r)) ¼ 0: (2:35)
sin u
2 Q sin u @u @u R @r @r h
The left side is now a function of u only. We set it equal to a separation constant ‘(‘ þ 1).
Now the Q term can be obtained by solving the ODE
1 @ @Q m2 Q
sin u ¼ ‘(‘ þ 1)Q: (2:36)
sin u @u @u sin2 u
j mj
Q ¼ B sinjmj uP‘ (cos u), (2:37)
jmj
where P‘ (cos u) are associated Legendre polynomials whose series converge only for
‘ ¼ 0, 1, 2, 3, . . . and m ¼ 0, 1, 2, . . . , ‘. The normalization coefficient is given by
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(2‘ þ 1)(‘ jmj)!
B¼ :
2(‘ þ jmj)!
Values for the associated Legendre polynomials and the normalized polar wavefunction
Qm‘ are listed in Table 2.1.
TABLE 2.1
Associated Legendre Polynomials and Normalized Polar Wavefunction
‘ m Pm
‘ (cos u) Qm‘
pffiffiffiffiffiffiffiffi
0 0 1 1=2ffi
pffiffiffiffiffiffiffi
1 1 1 ffiffiffiffiffiffiffiffi sin u
p3=4
1 0 cos u 3=2 cos u
pffiffiffiffiffiffiffiffiffiffiffiffi
2 2 3 ffiffiffiffiffiffiffiffiffiffi sin u
p15=16
2
TABLE 2.2
Derivatives of the First Several Leguerre Polynomials
n ‘ Lnþ‘
2‘þ1
(x)
1 0 1!
2 0 2x 4
3 0 3x2 þ 18x 18
Source: From White, H.E., Introduction to Atomic Spectra,
McGraw-Hill, New York, 1934.
Putting ‘(‘ þ 1) back into Equation 2.36, the PDE for the Q term, we have the radial
equation,
1 @ 2 @R ‘(‘ þ 1) 2m
r R þ 2 (E V(r))R ¼ 0 (2:38)
r2 @r @r r2 h
to be solved for R. For a potential function V(r) Ze2 =4p«0 r, the solution becomes
4p«0 h2
a1 ¼ ¼ 0:531 108 cm (2:40)
me2
nþl (x) are derivatives of the Laguerre polynomials. The first several values of Lnþ‘ (x)
and L2‘þ1 2‘þ1
For E > 0 (unbound states), solutions exist for all values of E. For E < 0 (bound states),
solutions exist only for
mZ2 e4
1
E¼ 2 2 n2
: (2:42)
2(4p«0 ) h
3=2
1 1
c2s ¼ p ffiffiffiffiffiffi (2 r)er=2 , (2:45)
4 2p 1 a
3=2
1 1
c2p0 ¼ pffiffiffiffiffiffi rer=2 cos u, (2:46)
4 2p a1
3=2
1 1
c2p1 ¼ pffiffiffiffiffiffi rer=2 sin ueif , (2:47)
4 2p 1 a
where r ¼ r=a1.
The product cc* may be interpreted as the probability density function for the electron.
The normalization coefficients A, B, C were chosen so that
ððð ðp 1
ð
cc*dV ¼ 2p cc*r2 sin ududr ¼ 1, (2:48)
all space 0 0
which says that the probability of finding the electron somewhere is unity.
Note that the wavefunction c for the 1s electron peaks at r ¼ 0 and has no angular
dependence as shown in Figure 2.4a. But when the product 2pr2cc* is taken, the prob-
ability of finding the electron is maximized at r ¼ a, the first Bohr radius, as shown in
Figure 2.4b.
The radial probability distribution functions for 2s and 2p electrons are shown in Figure
2.5a and b, respectively. Note that there is a small probability of finding the 2s electron at
the first Bohr radius, but the most probable location is around 5 Bohr radii (recall the Bohr
model predicts the electron shells to be spaced according to n2). On the other hand the
2p electron has the highest probability density at 4 Bohr radii.
Another way to visualize the probability distributions is by the use of density plots in
Figures 2.6 and 2.7. The highest probability density for both the 1s and the 2s states is
spherically distributed as shown in Figure 2.5a and b. From these plots, one is tempted to
think of the electrons in circular orbits around the nuclei. But remember, the electron in an
0.3
0.3
Probability density
0.2
yy ∗
0.2
0.1
0.1
0 0
0 1 2 3 4 5 0 1 2 3 4 5
(a) r (Bohr radii) (b) r (Bohr radii)
FIGURE 2.4
(a) Wavefunction product cc* for 1s electrons. (b) Radial probability distribution for 1s electrons.
Fundamental Principles 25
0.2 0.6
Probability density
Probability density
0.4
0.1
0.2
0 0
0 5 10 15 0 5 10 15
(a) r (Bohr radii) (b) r (Bohr radii)
FIGURE 2.5
(a) Radial probability density for 2s electrons. (b) Radial probability density for 2p electrons.
s-state has no angular momentum. Thus, one must think of the electron as simply smeared
out over region indicated by the figure.
Because of the cos u term in the c2p0 wavefunction (Equation 2.45), the probability
density has the figure eight-like lobe about the z-axis as shown in Figure 2.7a, but the
c2p1 wavefunction is complex and it loses its f-dependence when cc* is taken, giving the
probability density in the form of a toroidal distribution about the z-axis as shown in
Figure 2.7b. What is special about the z-axis? Why should the electron be distributed in a
lobe about this axis and not the other two axes?
2 5
−2 2 −5 5
−2 −5
(a) (b)
FIGURE 2.6
(a) Probability density of an electron in a 1s state. The most likely location is at 1 Bohr radius. (b) Radial probability
density of an electron in a 2s state. There is a small probability of the electron at 1 Bohr radius, but the most likely
location is at 5 Bohr radii.
26 Introduction to the Physics and Chemistry of Materials
5 5
−5 5 −10 −5 5
−5 −5
−10 −10
(a) (b)
FIGURE 2.7
(a) Probability density for an electron in the 2p0 or 2pz state. (b) Probability density for an electron in the 2p1 state
in the u ¼ p=2 plane.
3=2
c2pþ1 þ c2p1 1 1
px ¼ ¼ pffiffiffiffiffiffi rer=2 sin u cos f
2 4 2p 1 a
3=2 : (2:49)
c2pþ1 c2p1 1 1 r=2
py ¼ ¼ pffiffiffiffiffiffi re sin u sin f
2i 4 2p a1
Notice that when using these two equations to compute the probability density,
3
1 1
px p*x þ py p*y ¼ r2 er sin2 u ¼ c2p1 c2p1
* , (2:50)
32p a1
we get the same result as we would from Equation 2.49. In fact, if one takes px px* þ py py* þ pz p*,
z
the angular dependence disappears completely and the electrons are distributed in a uniform
shell about the nucleus.
until the Schrödinger equation is satisfied. There are several procedures for doing
this known as the Hartree method and the Hartree–Fock method. However, it is possible
to gain some understanding of many electron systems by making some simplifying
assumptions.
Even though the electrons react strongly with one another, it is possible to obtain a
satisfactory description of the electronic structure by assuming that each electron moves
more or less independently in an effective field. This effective field must take into account
the screening field of the other electrons which reduces attraction of the electron in
question to the nucleus. If one time-averages the motion of the other electrons, it is a
reasonable approximation to consider their distribution as spherically symmetrical. If the
screening field is taken to be spherically symmetric, the potential function is a function of r
only, just as in the case of the hydrogen atom. This will affect the radial wavefunction and
give different energy states, but the angular portions of the wavefunction will be preserved
along with the three quantum numbers n, ‘, and m that describe the electronic states.
TABLE 2.3
Possible Quantum States through n ¼ 4
n ‘ m s Maximum Occupancy
1 0 0 1=2 1s2
2 0 0 1=2 2s2
2 1 1, 0 1=2 2p6
3 0 0 1=2 3s2
3 1 1, 0 1=2 3p6
3 2 2, 1, 0 1=2 3d10
4 0 0 1=2 4s2
4 1 1, 0 1=2 4p6
4 2 2, 1, 0 1=2 4d10
4 3 3, 2, 1, 0 1=2 4f14
Notes: Principal quantum number, n ¼ 1, 2, 3, . . . ; angular momentum quantum
number, ‘ ¼ 0, 1, 2, . . . , (n 1); magnetic or projection quantum number,
m ¼ þ ‘, ‘ 1, . . . ,‘.
H He
K
1s 1s2
Li Be B C N O F Ne
L 2s 2s2 2s2 2s2 2s2 2p4 2p5 2p6
2p 2p2 2p3 2s2 2s2 3s2
Na Mg Al Si P S Cl Ar
M 3s2 3s2 3s2 3p4 3p5 3p6
3s 3s2 3p 3p2 3p3 3s2 3s2 3s2
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
3d 3d2 3d3 3d5 3d5 3d6 3d7 3d8 3d10 3d10 3d10 3d10 3d10 3d10 3d10 3d10
N
4s 4s2 4s2 4s2 4s2 4s 4s2 4s2 4s2 4s2 4s 4s2 4p 4p2 4p3 4p4 4p5 4p6
4s2 4s2 4s2 4s2 4s2 4s2
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
4d 4d2 4d4 4d4 4d6 4d7 4d8 4d10 4d10 4d10 4d10 4d10 4d10 4d10 4d10 4d10
O
5s 5s2 5s2 5s2 5s 5s 5s 5s 5s 5s 5s2 5p 5p2 5p3 5p4 5p5 5p6
5s2 5s2 5s2 5s2 5s2 5s2
Cs Ba La Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
4f14 4f14 4f14 4f14 4f14 4f14 4f14 4f14 4f14 4f14 4f14 4f14 4f14 4f14 4f14
P 5d 5d2 5d3 5d4 5d5 5d6 5d9 5d9 5d10 5d10 5d10 5d10 5d10 5d10 5d10 5d10
6s 6s2 6s2 6s2 6s2 6s2 6s2 6s2 6s 6s 6s2 6p 6p2 6p3 6p4 6p5 6p6
6s2 6s2 6s2 6s2 6s2 6s2
29
30 Introduction to the Physics and Chemistry of Materials
TABLE 2.5
Selected Ionization Energies
of the Elements in Electron Volts
H 13.60
Li 5.39
Na 5.14
K 4.34
Rb 4.18
Cs 3.89
He 24.59
Ne 21.56
Ar 15.76
Kr 14.00
Xe 12.13
Ra 10.74
as Xe þ 4f 145d26s2, has the same outer electron configuration as zirconium and titanium in
the lower series. A similar extraction of the actinide series is made in the n ¼ 7 series.
Thus all of the elements at the end of the period (He, Ne, Ar, Kr, Xe, and Rn) have closed
shells, s2 for He and ns2 np6 for the others, which causes these electrons to be tightly bound
and therefore do not readily interact with other elements. Compare the ionization energy
of these so-called noble gases with the alkali metals where the single outer electron is
shielded from the nuclei by the inner core electrons in Table 2.5.
2.6 Summary
A review of the foundations of modern physics that underlies all of our understanding of
matter is given with a historical perspective. Beginning with the early attempts of the Bohr
model of the atom to explain observed spectra, a review is given of the particle and wave
nature of material and the resulting uncertainty principle. A simplified derivation of the
time-independent Schrödinger equation is presented along with its application to electrons
confined to a crystal and to the hydrogen atom. The quantum mechanical behavior of
confined electrons (electrons in a box) is key to understanding the physics and chemistry of
metals and will be frequently drawn on in later chapters. Particular attention is given to the
hydrogen orbital wavefunctions because their angular dependence are similar to the
wavefunctions that combine to form the molecular orbitals responsible for the chemical
bonding of all materials. Also the quantum numbers associated with these wavefunctions
are shown to come about naturally from the mathematical solutions to the Schrödinger
equation. These quantum numbers, together with the Pauli exclusion principle, form the
basis of the periodic table and the electronic configurations of the elements.
Bibliography
Encyclopedia Britannica, Deluxe CD-ROM, 2004.
Hoddeson, L., Braun, E., Teichmann, J., and Weart, S., Out of the Crystal Maze: Chapters from the History
of Solid State Physics, Oxford University Press, New York, 1992.
Krane, K., Modern Physics, 2nd edn., John Wiley & Sons, New York, 1996.
Fundamental Principles 31
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley & Sons, New York, 1966.
Schiff, L.I., Quantum Mechanics, McGraw-Hill, New York, 1955.
White, H.E., Introduction to Atomic Spectra, McGraw-Hill, New York, 1934.
Problems
1. Fe–heme electronic structure in hemoglobin can be modeled as a 1 by 1 nm square in a
plane with 26 delocalized electrons. Why is blood red? (Hint: Treat the molecule as
electrons in a two-dimensional box. Find the ground state energy for the 26th electron.
This is the highest occupied molecular orbital (HOMO). Now find the energy of the
next available state. This would be the lowest unoccupied molecular orbital (LUMO).
As a result of a HOMO–LUMO transition, what wavelength of light does this energy
difference correspond to? Also note that E12, corresponding to nx ¼ 1, ny ¼ 2, has the
same energy as E21, but is a different state and each state can hold two electrons.)
2. Before the discovery of the neutron by James Chadwick in 1932, it was thought that the
nucleus contained electrons and neutrons. Use the uncertainty principle to show that an
electron cannot exist inside the nucleus. (Compare the attractive Coulomb potential
to the energy uncertainty resulting from the confinement within the nucleus. Take the
radius of the nucleus to be 1015 m.)
3. Al has 18 1028 electrons=m3. (a) Find its Fermi energy in electron volts. (b) Find the
next available energy level in a 1 cm3 crystal of Al. (c) Do the same for a 10 nm cube
quantum dot.
3
Chemical Bonding
The ability of atoms and molecules to form chemical bonds is the defining feature of the
structure and properties of solids. The types of bonds that are formed determine if
the material will be a metal, a ceramic, or a polymer, and whether the material will conduct
electricity, transmit light, or be magnetic.
33
34 Introduction to the Physics and Chemistry of Materials
qi qj
UC ¼ , (3:1)
4p«0 rij
where
rij is the distance between the ith and jth ion
qi is the charge on the ith ion
If the ith and jth ions have the same charge, the potential is positive and the force repulsive;
if the charges are opposite, the force is attractive. Note the 4p«0 in the denominator. We
will use rationalized meter, kilogram, second (MKS) units throughout.
If the ions get close enough that their core electronic wavefunctions begin to overlap, the
Pauli principle—which prevents electrons from occupying the same quantum state—forces
these electrons into higher energy states, causing a strong short-range repulsive force to act.
This force is represented empirically by a positive potential that increases rapidly with
decreasing r. Born suggested this repulsive force be represented by an inverse power law
B
UR ¼ , (3:2)
rnij
is similar using either the inverse power law or the Born–Mayer equation to represent the
repulsive potential as shown in Section 3.7.1.
where the 1=2 is to avoid double counting when summing over j for each i and vice versa.
It is convenient to write rij as rpij, where r is the nearest neighbor distance and pij is a
dimensionless parameter relating the distance between the ith and jth charges in units of r.
Now instead of summing over both i and j, we arbitrarily chose some i as the 0th charge,
sum over j and multiply by N, the number of ion pairs.
0 1 0 1
e 2 X 1 B X 1 e 2
B X 1
U(r) ¼ N @ þ A ¼ N @ Aþ n A, (3:4)
4p«0 r j>0 p0j rn j>0 pn0j 4p«0 r r j>0 pn0j
P
where A ¼ j>0 1=p0j is called the Madelung sum for the lattice. The Madelung sum can
be obtained by directly summing over the lattice taking the plus sign for unlike charge
pairs and minus sign for like charges, but convergence is very slow because of the 1=r
dependence of the Coulomb potential. This sign convention is chosen to make A > 0 and
the total potential energy negative for a stable lattice. Several methods (the Evjen solution
and the method of Ewald) have been developed for computing the Madelung constant for
different lattices (see Appendix for discussions on computing the Madelung sum).
The value of A depends on the structure, which is determined by the relative sizes of
the cations and the anions, which are generally larger than the cations. Since each ion
must be surrounded by counterions, a close-packed structure is not possible. It is necessary
for the smaller cations to be able to hold off the surrounding anions so that they do not
come in contact with each other, otherwise the strong repulsive forces would make the
lattice unstable. If the cations are nearly the size of the anions, it is possible to arrange
eight counterions around each ion without them contacting each other. If the cations are
smaller than the anions, it is possible to arrange only six counterions without them
touching, and if the cation is very much smaller than the anion, only a coordination
number (number of nearest neighbors) of 4 is possible. This issue will be discussed in
more detail in Chapter 5.
A structure with a coordination number of 8 is known as the CsCl structure. It can be
described as two interpenetrating simple cubic lattices with the anions on the corners of
one lattice and the cations located on the corners of the second interpenetrating lattice
arranged such that they sit in the centers of the cubes of the anion lattice so that each ion is
surrounded by 8 counterions. The Madelung constant for this structure is equal to 1.763.
The NaCl or rock salt structure has a coordination number of 6. It consists of two
interpenetrating face-centered cubic (fcc) lattices with the anions on the lattice sites of
one lattice and the cations on the lattice sites of the interpenetrating lattice arranged such
that the cations sit on the octahedral interstitial sites (points on the edges of the face-
centered cubes half-way between the corners. Despite the terminology, the coordination
36 Introduction to the Physics and Chemistry of Materials
e2 A nB X 1
¼ n : (3:5)
4p«0 r0 r0 j>0 pn0j
P
Using this result to eliminate B j>0 1=pn0j in Equation 3.4, the potential function becomes
NAe2 r0 1 r0 n
U(r) ¼ : (3:6)
4p«0 r0 r n r
For the NaCl structure, there are 4 ion pairs per unit cell and the unit cell has a volume of
(2r0)3, so N ¼ 1=2r03. Therefore, Equation 3.7 can be written as
U e2 n1
¼ A : (3:8)
Vol 8p«0 r40 n
Thus in Equation 3.6 we have a simple relationship between the Madelung constant A, the
equilibrium nearest neighbor distance r0 (which can be measured using x-ray diffraction),
the lattice energy ULat, and an arbitrary constant n. If the lattice energy is known, n can be
computed from Equation 3.7. The bond-energy function U(r) determines mechanical
properties such as the compressibility, bulk modulus, Young’s modulus, etc. Also, a
detailed knowledge of the bond energy as a function of r can be used to determine
properties such as thermal expansion and theoretical strength of the material. These details
will be discussed in Chapter 7 and subsequent chapters. Here we are able to see how we
can begin to make predictions of macroscopic material properties from atomistic consid-
erations! The bond-energy function from Equation 3.6 is shown in Figure 3.1.
0
U/U0
−0.5
FIGURE 3.1
Potential function for the ionic bond with
n ¼ 9. Note the long range of the attractive
Coulomb potential, which is the reason why
−1.0
0 2 4 6 8 10 so many terms must be considered in com-
Relative distance between atoms puting the lattice energy.
although many texts do not make this distinction and use cohesive energy in place of lattice
energy. (Actually, this distinction is important only in ionically bonded compounds.) The
problem arises because the lattice energy is difficult to measure directly. What is generally
measured is the heat of formation, which is roughly equivalent to the cohesive energy as
defined here. To convert cohesive energy to lattice energy, the energies required to
vaporize and disassociate the individual elemental components to form monatomic gases
must first be computed and then the ionization energies are added. This process is
sometimes called the Born–Haber cycle and is illustrated in Figure 3.2.
The measured lattice energies=volume for several systems with the NaCl structure are
shown in Table 3.1. Also shown are the values of the parameter n obtained from measured
lattice energies using Equation 3.8.
Notice that the energy=volume drops rapidly with increasing distance between nearest
neighbors, as would be expected from Equation 3.8. Also note that the derived value of n
increases with increasing nearest neighbor distance. Figure 3.3 shows a log–log plot of the
lattice energy=volume as a function of r0. The measured slope is 3.87 instead of 4
expected from Equation 3.8. The difference is due to the change of n with nearest neighbor
distance.
Na+(g) + Cl−(g)
Na(g) + Cl(g)
INa ED + EV
TABLE 3.1
Derived Values for n
U0=V0 (GJ=m3) Measured r0 (nm) Measured n Derived
200
100
Energy/volume (GJ/m )
3
FIGURE 3.3
Lattice energy vs. nearest neighbor distance
for ionic crystals with the NaCl structure.
(Note the approximate 4th power dependence 10
on nearest neighbor distance as predicted by 0.1 1
Equation 3.6.) Nearest neighbor distance (nm)
where
c ¼ C1c1 þ C2c2 and c1 and c2 are the ground state hydrogen radial wavefunctions
centered about their relative positions
^ is the Hamiltonian, which is an operator that includes the potential energy of
H
attraction between the unlike charges as well as the repulsive energy between
particles with like charges.
Now we find the coefficients that minimize the energy in Equation 3.10 by taking the
partial derivatives of E with respect to C1 and C2 and setting them to zero,
@E
¼ C1 ðH11 ES11 Þ þ C2 ðH12 ES12 Þ ¼ 0
@C1
, (3:11)
@E
¼ C1 ðH12 ES12 Þ þ C2 ðH22 ES22 Þ ¼ 0
@C2
where
Ð
^
Ð dtci Hcj ¼ Hij
dtci cj ¼ Sij
It may be seen that the only nontrivial solution for C1 and C2 exists when the determinant
H11 ES11 H12 ES12
H12 ES12 H22 ES22 ¼ 0: (3:12)
Normalizing the wavefunctions, S11 ¼ S22 ¼ 1. By symmetry, H11 ¼ H22. Solving the deter-
minant for E, we get
where
s is the symmetric solution
a is the antisymmetric solution
H12 represents the lowering of the energy between hydrogen ions, hence is negative
S12, which represents the overlap of the two electron wavefunctions, is positive, but
less than 1 since they are not in the same position.
Hence, it may be seen that Es < Ea and that Ea > 0. Putting these values back into Equation
3.10, we find that C1 ¼ C2 for Es and that C1 ¼ C2 for Ea. We can now construct the
symmetric and antisymmetric wavefunctions,
Normalizing the two wavefunctions and constructing the Hamiltonian from the electro-
static attraction and repulsion between the electrons and the nuclei, the calculated value for
Es ¼ 1.77 eV, which represents the bonding state and Ea > 0 represents the antibonding
40 Introduction to the Physics and Chemistry of Materials
state. The actual disassociation energy for the hydrogen molecule is 2.78 eV. So the theory
is not exact, which is not surprising considering the approximations that were made. The
important feature of this exercise is the fact that a bonding state resulted from the
symmetric combination of wavefunctions; whereas an antibonding state resulted from
the antisymmetric combination.
A B
FIGURE 3.4
(a) Schematic of 1s overlapping wavefunctions of two atoms as they are brought together (solid lines). The long
dashed line is the sum of the two wavefunctions and the short dashed line is their difference. (b) The probability
density of the two wavefunctions in Figure 3.4b. The solid line represents the cþ2 and the dashed line represent the
c2. The negative charge between the two atoms in the case of cþ2 produces the bonding state.
Chemical Bonding 41
E−
E− y−
yA
yA EA yB
EA y− EB
yB
EB E+
E+
y+ y+
(a) (b)
FIGURE 3.5
Energies of the bonding and antibonding molecular orbitals in a diatomic molecule. Atoms A and B have energies
EA and EB, respectively, when separated. As they are brought together and their wavefunctions begin to overlap,
they form a bond and an antibond energy level as shown in (a). As they are brought closer, the more the
separation between the two levels as shown in (b).
atoms; whereas, in the s bond, the overlap of the p-wavefunction is along the line between
the two atoms as shown in Figure 3.6.
The þ and signs used in this type of diagram indicate the phases of the wavefunctions,
not the electrical charge. Overlap of two þ or two regions indicates that the two
wavefunctions add constructively; whereas, overlap of a þ and region indicates destruc-
tive interference. The s bond is generally stronger because of the greater overlap of the
wavefunction. There is no restriction to the rotation about the axis of the s-bond, whereas
rotation is not allowed about the p-bond.
To illustrate how this works with a multielectron atom, consider the N2 molecule. Here
we must accommodate 14 electrons, 7 from each atom. As the two N atoms are brought
together, each of the atomic levels split into a bonding and an antibonding state as shown
in Figure 3.7.
Note that the splitting between the 2px bonding and antibonding levels is greater than
between the splitting between the 2py, 2pz levels. This is because it is assumed that a s bond
exists along the x-axis. p bonds exist between the y and z components. Both the bonding
and antibonding orbitals are occupied for the inner electrons, but only the bonding orbitals
are occupied in the 2p states. The six 2p electrons form a very strong triple bond between
the two atoms, which is the reason why the N2 molecule is so stable.
The O2 molecule introduces two additional electrons. Since there are no more bonding
states available for them to occupy, they must go to the next available 2py or 2pz antibond-
ing states which causes that lobe to become saturated and no longer able to form a bond.
Thus the disassociation energy of the double bond in the O2 molecule is considerably less
that the N2 molecule. This accounts for the fact that the primary atmospheric constituents
at low-earth orbit altitudes are atomic oxygen and molecular nitrogen.
+ +
+ +
A B − −
FIGURE 3.6
A B Schematic of p and s bonds. The s bond is
stronger because it has more overlap, but
− −
allows rotation about the bond line. The
p bond s bond weaker p bond does not allow rotation.
42 Introduction to the Physics and Chemistry of Materials
2px
2py 2pz
2p 2p
2py 2pz
2px
2s 2s
FIGURE 3.7 1s 1s
Schematic of the bonding and antibonding
levels in the N2 molecule.
3.3.5 Hybridization
The carbon atom, with its 1s22s22p2 configuration has a valence of 4. However, in many of
the systems involving C, these valence electrons behave as if they were identical and form
directed bonds toward the vertices of a tetrahedron with a bond angle of 109.58 surround-
ing the atom (e.g., diamond, CH4, etc.). This can be explained by assuming that the 2s and
2p orbitals can form a linear combination such that the four orbitals are equivalent and
point to the opposite corners of a cube with the C atom at the center. The n ¼ 2 hydrogen-
like wavefunctions from Chapter 2 are given in Cartesian space by
z z y
y
px
s + + −
FIGURE 3.8
Schematic of the overlapping wavefunctions in the s–p bond.
Chemical Bonding 43
s + + − FIGURE 3.9
s–p bonds formed between the px and py lobes of oxy-
− gen with hydrogen to form the H2O molecule. The
mutual repulsion between the two naked Hþ ions dis-
torts the 908 bond angle to 104.58.
3=2
1 1
c2s ¼ pffiffiffiffiffiffi ð2 rÞer=2
4 2p a0
3=2
1 1
c2px ¼ pffiffiffiffiffiffi rer=2 sin u cos f
4 2p a0
3=2 (3:15)
1 1
c2py ¼ pffiffiffiffiffiffi rer=2 sin u sin f
4 2p a0
3=2
1 1
c2pz ¼ pffiffiffiffiffiffi rer=2 cos u
4 2p a0
where r ¼ r=a0 .
It is left as a problem to show that the combination
1
cI ¼ c þ c2px þ c2py þ c2pz
2 2s
1
cII ¼ c2s þ c2px c2py c2pz
2
(3:16)
1
cIII ¼ c2s c2px þ c2py c2pz
2
1
cIV ¼ c2s c2px c2py þ c2pz
2
has this property. Essentially one of the 2s electrons is promoted to the 2p state which can
then hybridize with the remaining 2s electron to form four overlapping bonds with
neighboring atoms to provide an energy reduction that more than compensates for the
additional energy needed to promote the 2s electron. This is known as the sp3 hybrid bond.
Similarly, other group IV atoms can form similar hybrids (Si, Ge, and a-Sn) as well as III–V
compounds such as GaAs, GaN, InP, BN, etc. The tendency to form this hybrid bond
diminishes as the interatomic distance increases. This accounts for the reason that Sn can
exist either as a metal or a semimetal and reason why Pb is always a metal as will
be discussed further in Chapter 20. Correspondingly, the bond strength diminishes from
7.30 eV for diamond, 4.64 eV for Si, and 3.87 eV for Ge.
44 Introduction to the Physics and Chemistry of Materials
Carbon can also hybridize in a threefold structure by taking the following linear com-
binations:
rffiffiffi
1 2
cI ¼ pffiffiffi c2s þ c
3 3 2px
1 1 1
cII ¼ pffiffiffi c2s pffiffiffi c2px þ pffiffiffi c2py : (3:17)
3 6 2
1 1 1
cIIi ¼ pffiffiffi c2s pffiffiffi c2px pffiffiffi c2py
3 6 2
This is known as the sp2 hybrid and it is left to the student to show that his forms a planar
structure with bond angles 1208 apart. Graphite is structured with sp2 bonds which gives it
a very strong planar hexagonal network. The remaining electron delocalizes and is free to
meander through the structure as a conduction electron. The layers are bonded only by van
der Waals forces, which accounts for the slippery nature of the material. This sp2 bond is
also responsible for the formation of buckyballs and carbon nanotubes, which are in the
forefront of research in nanotechnology (see Chapter 5).
10
5d
8
4d
U0 (3d)
U0 (4d)
6 U0 (5d)
U0 (eV)
3d
FIGURE 3.10
Cohesive energies of the 3d, 4d, and 5d metallic elem- 0
ents as a function of their outer shell electronic config- 0 5 10 15 20
uration. Number of electrons in outer shell
Chemical Bonding 45
FIGURE 3.11
Schematic showing the formation of the band of
n=1 n=2 n=4 n=6 n = 1023 energy states in the metallic bond.
2p
2s
Energy
1s
FIGURE 3.12
r0 Showing overlap of 2s and 2p bands that makes electrical
Distance between atoms conductivity in divalent metals possible.
46 Introduction to the Physics and Chemistry of Materials
1000
100
U/V0 (GJ/m3)
Group IA
Group IIA
Group IIIB
Group IVA
10
FIGURE 3.13
Energy per volume vs. radius for the simple 1
metals and group IVA elements. The slopes 0.01 0.1 1
for each of the groups are approximately 4. R0 (nm)
loses the directionality of the covalent bond and the ion cores tend to form close-packed
structures.
Qualitatively, we can see that bond energies increase in going from column 1 (IA) to 2
(IIA) and from column 13 (IIIA) to 14 (IVA) in Figure 3.10. This can be understood simply
by the fact that more electrons are involved in the forming the bond. Also we see that the
bonds tend to be weaker with increasing atomic numbers in the same column. As the
atoms get larger, they tend to be farther apart and the bond energy falls off with distance.
Unfortunately there are no simple theories to predict the cohesive energies of the metals
like the coulomb attraction in ionic crystals. More sophisticated quantum mechanical
theories using pseudopotential or other modeling techniques are generally required.
There are some interesting correlations, however.
Figure 3.13 plots U=V0 as a function of the radius for group I (Li, Na, K, Rb, and Cs),
group II (Be, Mg, Ca, Sr, and Ba), group III (B, Al, Ga, In, and Tl), and the diamond
structured group IV (C, Si, and Ge) elements. The group IV elements are not metals, but
were included because they fit the pattern of the simple metals. The curves almost fall on
top of each other and the slope of each of the groups is close to 4 (the average slope is
4.11). The R04 dependence of the energy=volume is quite similar to the energy depend-
ence of the ionic-bonded systems.
The cohesive energy=volume is plotted against valence electron density in Figure 3.14.
Again the curves for the different groups fall close to one another, but there are some
differences in the slopes. The slopes of groups I and II average 1.34, which is consistent
with U=V0 Nval=R04. The slopes of groups III and IV were 0.75 and 0.50, respectively.
1000
Group IA
Energy/volume (GJ/m3)
10
FIGURE 3.14
1 Energy per volume vs. valence electron dens-
1027 1028 1029 1030
ity for simple metals and group IVA elements.
Electron density (m−3) The average slope is 0.85.
so the 1s bonding orbital is filled by forming H2, and the bond is saturated. In the case of
the alkali metals such as Li, there are the 2p, 3s, 3p, etc., states that are only 1–2 eV away
from the 2s state. The presence of these available states allows the Li atom to share its 2s
electron with other atoms in order to form a metallic bond.
It is believed that metallic hydrogen can exist at pressures on the order of 100 GPa and
that the core of large planets, such as Jupiter, may consist of metallic hydrogen. Such
pressures can be created for an instant of time in the laboratory using shock compression
techniques and have produced evidence of metallic hydrogen, but the hydrogen does not
remain in the metallic form after the pressure is relieved (see Weir et al., 1996).
100
5d
4d
80
Energy/volume (GJ/m3)
60
3d
40
20
FIGURE 3.15
Cohesive energy per volume vs. column number
for the 3d, 4d, and 5d transition metals. In this
plot, the curves tend to be more symmetrical 0
about column 7 where the number of bonding 2 4 6 8 10 12 14
states should be maximized. Number of electrons in the outer shell
between the atoms with the highest atomic numbers when their d-orbitals are approxi-
mately half full. Since the atomic radii of the 5d metals remain approximately the same as
their 4d counterparts (see Section 3.5), there will be greater overlap of the larger d-shells of
the 5d elements, hence stronger bonds. This trend is reversed as the d-orbitals are just
starting to fill or are nearly saturated and the strongest bonds are formed among the lower
atomic weight elements as in the case of the simple metals.
300
200
150
FIGURE 3.16
Atomic radii of the row 4, 5, and 6 metal elem-
100 ents in the periodic table. (Atomic radii taken
0 5 10 15 20 from LANL periodic table of elements, http:==
Column number periodic.lanl.gov)
assign an atomic radius of 0.077 nm to carbon. Similarly, 1=2 the nearest neighbor distance
in sp3-bonded silicon is 0.117 nm. Silicon carbide has the same crystal structure except
that each Si atom is surrounded by 4 C atoms and vice versa. Can we add 0.077 and 0.117
to get 0.194 nm for the nearest neighbor distance in SiC? Well almost. The observed nearest
neighbor distance in SiC is 0.189 nm. Though not absolutely precise, taking the atomic
radius from nearest neighbor distances of like atoms is good enough for the most part
to predict nearest neighbor distances and crystal structures for metallic and covalent
bonding systems.
Things are decidedly different for ionically bonded systems. Generally when an electron
leaves a metallic atom to form a cation, the cation will be effectively smaller than the
atom in a metallic bond even though the valence electron in the metal is delocalized.
Similarly, the size of the anion that has gained the electron is generally larger than the
neutral atom. A self-consistent set of ionic radii have been worked out for systems with a
coordination number of 6 (rock salt structure). A correction of þ0.008 nm must be added to
the sum of the standard ionic radii for coordination number of 8 and a correction of 0.011
must be subtracted for tetrahedrally coordinated structures.
Because the electrons in a metal become delocalized, it may seem strange that the hard
sphere size of the ions cores consistent with the measured density of the solid is larger than
the ionic diameter used to calculate the size of ionic compounds. The explanation is that a
delocalized electron is not completely lost from the ion core as it is when transferred to an
electronegative atom. The delocalized electrons are shared among all of the ion cores so the
effective size of the ion core is less diminished when its valence electrons delocalized.
If d r, the bracket term can be expanded to give 1 (1 þ d=r)2 so that the field becomes
2dq 2p
E¼ ¼ , (3:19)
4p«0 r3 4p«0 r3
where p is the dipole moment which is taken to be positive in the direction of the positive
charge.
The potential energy of a dipole in an electric field is given by U ¼ p E. Therefore, an
attractive potential exists between two electric dipoles that are aligned nose to tail that goes
as p1p2=r3 and the force between them goes as p1p2=r4. Thus a dipole–dipole bond is
directional and short ranged.
The electric field from one dipole can induce a dipole moment in an adjacent atom that
does not have a permanent dipole moment according to
2ap1
p2 ¼ aE ¼ , (3:20)
4p«0 r3
where a is the polarizability of the atom. Therefore, the interaction energy between a
dipole and an atom or molecule with an induced dipole goes as ap1p2=r6 and the force
between them is maximized along the axis of the permanent dipole.
(a) (b)
FIGURE 3.17
Schematic of hydrogen bonding in polar substances (a) H2O and (b) HF. The two H atoms in H2O allow water ice
to form open tetrahedral-like structures whereas as a single functional molecule such as HF can only form chains.
Chemical Bonding 51
Because of the small size of the hydrogen ion, only two counterions can get close to
it without touching each other. Therefore the hydrogen atom is always doubly coord-
inated in a hydrogen bond. The strength of the hydrogen bond is typically 0.1–0.5 eV,
which is less than the primary bonding energies, but large enough to hold substances
together at ambient temperatures in the absence of any of the primary bonding
mechanisms. Thus the hydrogen bond is very important in polymers and other organic
solids, in biological systems, and in polar molecular systems such as water. The
tendency for such bonds to form tetrahedral open structures in the case of water is
responsible for many of the unique properties of water such as ice having a lower
density than liquid water.
But how does this process get started between two neutral nonpolar particles? Fluctuations
of the charge distribution are continuously occurring because of the zero-point energy in the
system in the absence of any other mechanism. A fluctuation in the charge distribution in
the first particle produces a temporary dipole moment which induces a dipole moment in an
adjacent particle. This induced dipole in the second particle induces a dipole moment
in the first particle and so the process continues.
This resulting force between the two particles is known as the van der Waals force (also
called the London dispersion force). Like the metallic and ionic bonds, the force acts along
the line between any two particles, so it is best described as a central force. Because there is
no charge on the particles, close-packed structures are favored in van der Waals solids.
Note that the attractive (negative) part of the potential goes as r6 (or V2, which is the
correction for the molecular interaction in the equation of state for a van der Waals gas).
As one might imagine, the van der Waals bond is extremely weak (on the order of
ambient thermal energy), but it is always there and works when nothing else does. It
provides the mechanism for condensation and solidification of noble and molecular gases.
Its strength increases with increasing number of electrons, which explains why F2 and Cl2
are gases at ambient temperature while Br2 is liquid and I2 is solid. The van der Waals
bonding becomes even more important in macromolecular systems such as polymers and
biomolecules.
1.0
0.5
Relative energy
0.0
FIGURE 3.18
Lennard–Jones potential function. The
quantity, s, is the closest distance one −0.5
particle falling from infinity can
approach another. Note the short
range of the Lennard–Jones potential −1.0
by comparing this figure with 0.8 σ 1.0 1.2 1.4 1.6 1.8 2.0
Coulomb potential in Figure 3.2. Relative distance between atoms
where A and s are constants to be determined for a particular system. The 6th power
attractive term is chosen to model the van der Waals interaction while the 12th power
repulsive term is purely arbitrary. As shown in Figure 3.18, the potential is zero when rij
equals s, which is the closest distance one particle falling from infinity can come to
the other particle. Thus s can be interpreted as the diameter of the particles in a hard
sphere model.
Differentiating to find r0 that minimizes the potential, we find r0 ¼ 21=6 s. Putting this
back into Equation 3.2, we can express the coefficient A in terms of the binding energy U0
(taken as a positive number) and write the potential as a function of r=r0.
" 6 # " 6 #
s 12 s r0 12 r0
U(rij ) ¼ 4U0 ¼ U0 2 : (3:23)
rij rij rij rij
This form is useful in modeling molecular vibrations and for relating the pressure of a van
der Waals gas to the attractive potential between the molecules.
We can apply the Lennard–Jones potential to a crystalline solid such as a condensed
noble gas in which van der Waals forces are the only bonding mechanism. We replace rij
with pijr as we did for the case of ionic bonding and write the binding energy for a
particular configuration,
" #
1X X s 12 s 6
s 12 s6
U ðrÞ ¼ U(rij ) ¼ NA ¼ NA S12 S6 , (3:24)
2 i6¼j j>0
p0j r p0j r r r
P 12
P 6
where S12 and S6 are short for the lattice sums j¼1 p0j and j¼1 p0j , respectively.
Differentiating Equation 3.24 with respect to r to find the r0 that minimizes the Ulat,
6
s S6
¼ : (3:25)
r0 2S12
Putting Equations 3.26 and 3.25 back into Equation 3.24, the Lennard–Jones potential may
be written as a function of r in the form,
r 6
r0 12 0
U(r) ¼ U0 2 : (3:27)
r r
The van der Waals bond is universal in that it applies to all atoms. But since the polariz-
ability a 1040, the force is very weak (Ulat 0.010.1 eV) and is usually overshadowed
by other stronger bonds. However, in the absence of other bonds (e.g., for the noble gases),
it becomes the primary bonding mechanism. It is short ranged and nondirectional, hence
when acting as a primary bond, it favors a high coordination number found in either fcc of
hcp structures.
X X S26
p6
0j ¼ S6 ¼ 14:45392; p12
0j ¼ S12 ¼ 12:13188; ¼ 17:220398: (3:28)
j>0 j>0
S12
X X S26
p6
0j ¼ S6 ¼ 14:45489; p12
0j ¼ S12 ¼ 12:13229; ¼ 17:222127: (3:29)
j>0 j>0
S12
It looks as if the hcp lattice has slightly lower energy and would be favored. Interestingly,
most van der Waals-bonded substances, primarily the noble gases actually form fcc
structures. The only explanation is that it is not a perfect theory. Remember, the assumed
form for the potential repulsive potential was arbitrary and has no theoretical basis.
NAe2
U(r) ¼ þ B exp(br): (3:30)
4p«0 r
Taking derivatives to obtain the equilibrium energy and configuration as before, we find that
NAe2 r0 eb(r0 r)
U(r) ¼ (3:31)
4p«0 r0 r r0 b
54 Introduction to the Physics and Chemistry of Materials
−0.6
Relative energy
−0.8
FIGURE 3.19
Comparison of Born–Mayer potential (dashed line)
with the Born potential (1=rn) for the repulsive
term (Equation 3.6) (solid line) with r0 b ¼ n. Only −1.0
0.8 1.0 1.2 1.4
the region around r ¼ r0 is shown because this is
where the difference between the two is the largest. Relative distance between atoms
Comparing this result with Equation 3.7, we see that r0b plays the same role as n in the
power-law formulation. A comparison of the two potentials is shown in Figure 3.19.
m nm
B¼A r : (3:34)
n 0
Using this result to relate the equilibrium spacing r0 to the lattice energy at equilibrium,
A n m
U(r0 ) ¼ : (3:35)
rm
0 n
Putting this back into the general equation (Equation 3.33), the potential becomes
r n
U ðr0 Þ r0 m 0
U(r) ¼ n m : (3:36)
nm r r
Chemical Bonding 55
One can see that by setting m ¼ 1, Equation 3.36 reduces to Equation 3.6. Also, the
Lennard–Jones potential is obtained by setting m ¼ 6 and n ¼ 12 which is the same as
Equation 3.27.
U0 mn 1 r0 m 1 nð1r=r0 Þ
U ð rÞ ¼ e : (3:37)
mn m r n
U ðr0 Þbr0 r0 1 bðr0 rÞ
U ðr Þ ¼ e , (3:38)
1 br0 r br0
which is the Born–Mayer potential given by substituting Equation 3.32 into Equation 3.31.
r n U ðr Þ
U ðr0 Þ r0 m
½nð1 þ xÞm mð1 þ xÞn :
0 0
U(r) ¼ n m ¼ (3:39)
nm r r nm
U ðr0 Þ
U(x) ¼ ½nemx menx : (3:40)
nm
We now set n ¼ 2m
U(x) ¼ U ðr0 Þ 2emx e2mx , (3:41)
which is the Morse potential. The Morse potential is often used for molecular modeling as
well as for metals. Changing the sign in the Morse potential gives an antibonding potential
with 4 times the energy of the bonding potential at the equilibrium point.
The Morse potential is sometimes written in the form DU(x) ¼ U(r0) U(x) or
DU(x) ¼ U ðr0 Þ Uðr0 Þ 2emx e2mx ¼ U ðr0 Þ 1 2emx þ e2mx
¼ U ðr0 Þ½1 emx 2 : (3:42)
56 Introduction to the Physics and Chemistry of Materials
3.8 Summary
The three primary chemical bonds are the ionic bond, the covalent bond, and the metallic
bond. The ionic bond requires an electron transfer one atom to another atom with a higher
electron affinity and forms between elements that have a large difference in electronega-
tivity. The attractive forces between the resulting anions and cations are purely electrostatic
and can be described by Coulomb’s law. The ionic bond is the strongest of the chemical
bonds, has a long range since the energy falls off as 1=r, and is nondirectional in nature.
Since it is necessary for each ion to be surrounded by counterions, the coordination number
(number of nearest neighbors), hence the crystal structures that are possible, is determined
by the relative sizes of the cations and anions.
The covalent bond forms by atoms sharing electrons with one another to form filled
electron shells which lowers the energy of the system. As two atoms are brought together,
their wavefunctions began to overlap, forming a bonding state and an antibonding state.
The strength of the covalent bond can be quite large and depends along with other things
on how many electrons are allowed go into the bonding state. Since the p, d, and f electrons
have angular dependence, bonds formed by these overlapping wavefunctions are direc-
tional which leads to open diamond-type structures. Covalent bonds may form between
like atoms as well as between dissimilar atoms, although the bonds between dissimilar
atoms may be partially ionic, depending on the difference in their electronegativity.
Metallic bonds form by valence electrons becoming delocalized from atoms on the left
hand side of the periodic table that have only a few loosely attached outer shell electrons.
These electrons form a sea of negative charge that surrounds and attracts the positively
charged ion cores with the electron sea filling the role of the anions in the ionic bond. The
metallic bond is generally weaker than either the ionic or the covalent bond and is
nondirectional. Since there is no restriction on the number of nearest neighbors other
than geometry, metals tend to form close-packed structures with a coordination number
of 12 although many metals prefer to form body-centered cubic structure with a coordin-
ation number of 8. The covalent bond is also formed among the d electrons in the transition
metals which accounts for their strength, especially when the d shells are nearly half full so
that none of the electrons have to fill antibonding states.
Secondary bonds are dipolar in nature. A polar molecule has a neutral charge but an
inhomogeneous charge distribution such that one end of the molecule is positive and the
other end negative. Such a molecule forms an electric dipole and is surrounded by an
electric field that attracts other dipoles. An H atom that is covalently bonded to a halogen
or to an O atom effectively loses its electron to the host atom which forms a strong electric
dipole. Dipole–dipole bonds mediated by an attached H atom are said to be hydrogen
bonded. The hydrogen bond, though considerably weaker than the primary bonds, is very
important in the bonding of organic solids and as a secondary bond in polymer systems.
The bond is short-ranged since the dipole field falls as 1=r3 and is directional along the
dipole axis. Single functional molecules such as HF and HCl tend to form chains with nose
to tail structures. Bifunctional molecules such as H2O form open structures that cause ice to
expand when it freezes.
An electric field can induce a dipole moment in a nonpolar molecule by pushing the
positive and negative charges in opposite directions. A fluctuation in charge distribution in
one nonpolar molecule can induce a dipole moment in an adjacent nonpolar molecule
causing an attractive force to occur. London dispersion forces (also called van der Waals
forces) are extremely weak and short ranged, falling as 1=r6, but they work when no other
bonding forces are present. Consequently, they provide the mechanism by which noble
and molecular gasses condense at very low temperatures. They also hold neutral layered
Chemical Bonding 57
solids such as graphite together and provide interactions between the covalently bonded
chains in polymers.
A number of semitheoretical models have been proposed to describe the bond-energy
curve in ionic-bonded and van der Waals-bonded systems in which the attractive force can
be described accurately by an inverse power law. The repulsive potential is described by an
exponential or a high order inverse power law. Such potentials are useful for relating
physical properties such as the bulk modulus and elastic modulus to the lattice energy.
Unfortunately, there are no simple lattice models that can be applied to metallic or
covalently bonded systems.
4ð1Þn Xn1
8ð1Þnþi 4
Rn ¼ þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (A:3:1)
n i¼1 n þi
2 2 n þ n2
2
+ − + − + − + − +
− + − + − + − + −
+ − + − + − + − +
− + − + − + − + −
+ − + − + − + − +
− + − + − + − + −
+ − + − + − + − +
− + − + − + − + −
+ − + − + − + − +
FIGURE A.3.1
Charges on the (100) face of an NaCl crystal.
58 Introduction to the Physics and Chemistry of Materials
TABLE A.3.1
Slow Convergence of the Lattice
Using the Direct Method
N Rn An
1 1.172 1.172
2 0.163 1.335
3 0.080 1.415
4 0.044 1.459
5 0.028 1.487
6 0.020 1.507
7 0.014 1.521
8 0.011 1.532
9 0.087 1.541
10 0.070 1.548
100 1.786 105 1.609
1000 1.786 107 1.6151
where the sign has been reversed to make the Madelung constant positive in accordance
with the convention.
The problem with this straightforward approach is that the convergence is extremely
slow. For example, let us examine the first 10 terms and their sum shown in Table A.3.1.
Notice how slowly the sum is converging. Even after 1000 squares, the sum has only
converged to the third decimal.
Convergence may be hastened by summing the contributions of the charges contained
within each of the squares instead of the ions on the squares. For example, since only 1=2 of
the 4 face atoms and 1=4 of the 4 corner atoms lie within the first square, the contribution
from the charges within the first square is S1 ¼ 2=1 þ 1. The contributions from the
charges between the first and second square include the remainder of the ions outside
the first square, 2=1 þ 3=(12 þ 12)1=2, as well the portion of the charges that lie inside the
second square, i.e., þ2=2 4=(22 þ 12)1=2 þ 1=(22 þ 22)1=2. The contributions when summed
in this manner are shown in Table A.3.2.
TABLE A.3.2
Faster Convergence by Grouping
Contributions from Squares
N Sn An
1 1.292893 1.292893
2 0.313981 1.606874
3 0.003648 1.610522
4 0.002987 1.613510
5 0.000981 1.614491
6 0.000441 1.614933
7 0.000225 1.615158
8 0.000127 1.615285
9 0.0000767 1.615361
10 0.0000491 1.615410
100 4.058 109 1.615542498
1000 3.985 1013 1.615542627
Chemical Bonding 59
Taking the contributions from the ions on each square provides the same number of
positive and negative charges, but if only the contributions from the portions of the ions
within each square are considered, there are more unlike charges than like charges and
the attractive contributions are maximized, which helps the sum to converge more
rapidly. With this latter scheme, convergence to 3 decimal places is achieved after only
7 squares and to 6 decimal places after 100 squares. If the sum of total ion contributions is
taken as in Equation A.3.2, convergence to the first decimal place is not reached until the
44th square and convergence to the third decimal place is not achieved after the 1000th
square.
Since summing over the ions between each square gives back the contribution from the
outer portion of the previous square, one can simply compute the lattice sum by using
Equation A.3.2 to sum over all of the ions up to the nth square and then simply add the
contribution from the ions with the next square, which is given by
"
XN
ð1Þi X n X n
ð1Þi ð1Þnþ1
AðnÞ ¼ 4 þ pffiffiffiffiffiffiffiffiffiffiffiffiffi þ
i¼1
i i¼1 j¼1 i2 þ j2 2ðn þ 1Þ
#
X
n
ð1Þnþ1þi 1
þ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi þ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : (A:3:3)
i¼1 ðn þ 1Þ þi2 4 2ðn þ 1Þ2
2
Bibliography
Borg, R.J. and Dienes, G.J., The Physical Chemistry of Solids, Academic Press and Harcourt Brace
Jovanovich Publishers, Boston, MA, 1992.
Gilman, J.J., The Electronic Basis of the Strength of Materials, Cambridge University Press, UK, 2003.
Ibach, H. and Lüth, H., Solid State Physics, 3rd edn., Springer-Verlag, New York, 1990.
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley & Sons, New York, 1966.
Ladbury, R., Livermore’s big guns produce liquid metallic hydrogen. Phys. Today, 49, May
1996, 17–18.
Srivasava, C.M. and Srinivasan, C., Science of Engineering Materials, Wiley Eastern, New Delhi,
1987.
Weir, S.T., Mitchell, A.C., and Nellis, W.J., Phys. Rev. Lett., 76, 1996, 1860.
Problems
1. Show that the coordination number for a close-packed structure is 12. Now show why
ionically bonded systems cannot have a close-packed structure.
2. Use the angular dependence of the wavefunctions in Equation 3.16 to construct the sp3
wavefunctions, which should produce four lobes of charge distribution. Find
the directions (u and f) of these four lobes. (Ignore the radial part of the wavefunctions.)
3. Show that the sp2 hybrid wavefunctions given by Equation 3.17 produce a three-lobed
charge distribution 1208 apart.
60 Introduction to the Physics and Chemistry of Materials
− − −
ro
a + + +
21/2 ro
−
− −
21/2 a
FIGURE 3.20
(110) face of the NaCl structure.
4. Arrange two electric dipoles with dipole moments qd nose to tail at distance r (r d)
along a line including their axis. Use Coulomb’s law to show that the force between
them varies as r4.
5. The (110) surface of the NaCl structure is shown in the Figure 3.20. Find the 2D
Madelung constant for this surface.
4
Crystals and Crystallography
61
62 Introduction to the Physics and Chemistry of Materials
of individual devices simultaneously on them was the enabling technology that permitted
the boom in consumer electronic devices.
a
b
b
g
FIGURE 4.1
Unit cell vectors and the angles between them. a
Crystals and Crystallography 63
FIGURE 4.2
Oblique unit cell in two dimensions with a basis of (0,0),
a (1=4,3=4), (3=4,1=4).
coherently diffract x-rays and makes it possible to make precise measurements of the lattice
parameters as well as to determine the position and types of atoms making up the unit cell.
The use of the word ‘‘group’’ has a special mathematical meaning. A group is defined as
a set of operations that obey the following set of rules: (1) the result of any two operations
must also be an element of the set; i.e., A B ¼ C; (2) the operations must be associative;
i.e., A (B C) ¼ (A B) C; (3) there must exist an identity operation; I A ¼ A (achieved
by setting n1, n2, n3 ¼ 0); and (4) there must be an inverse operation that when applied to a
previous operation element returns the system to its original state; i.e., A1 (A B) ¼ B,
which implies A1 A ¼ I. In general, group elements are not required to be commutative;
i.e., A B ¼ B A is not required. If the elements do happen to commute, the group is
called an abelian group. (We will not make further use of the mathematical concept of a
group in this book. Readers interested in the application of group theory to crystallography
are referred to Harrison’s Solid State Theory.)
[111]
b
[100] [110]
FIGURE 4.3
a Basic crystallographic direction vectors.
64 Introduction to the Physics and Chemistry of Materials
1/3
B
2/3
b
A 1/3
FIGURE 4.4
1/2, 1/2
Generalized vector directions. Using vector subtraction,
vector A would be designated [136] and vector B would
be [2
32]. a
However, the projections of a vector onto the axes need not be whole numbers. The
location of the central atom in a body-centered cubic (bcc) unit cell has coordinates relative
to the unit cell (1=2,1=2,1=2), so the direction of a vector from the origin to that central atom
is written as [111]. Similarly, the location of the middle atom of a face-centered unit cell in
the a–b plane is (1=2,1=2,0), so the direction from the origin to it would be written as [110].
Nor do the vectors have to pass through the origin. Consider a more general case shown in
Figure 4.4. Vector A goes from (1=3,0,1) to (1=2,1=2,0). Vector subtraction yields
(1=6,1=2,1). Multiplying by 6 to clear the fractions, vector A will have direction given
by [13 6]. Similarly, vector B goes from (1=3,1,0) to (1,0,2=3). Vector subtraction yields
(2=3,1,2=3). Multiplying by 3, we obtain the direction [232].
Notice that the lattice directions are written within square brackets with no commas.
Negative directions are indicated with a bar over the number as [110]. Families of similar
directions are denoted by an angle bracket h100i, which consists of the six vectors, [100],
[010], [001], [
100], [0
10], and [001]. Note that [100] is antiparallel to [100], etc.
The above rules apply for any geometry of the unit cell. For the case of cubic unit cells, it
is easy to obtain the angles the various direction vectors make with each other by using the
definition of the dot product A B ¼ jAkBj cos(AB). For example, the direction from the
central atom in a bcc unit cell to one of the corner atoms can be written as [111], and the
direction from the central atom to the atom in the opposite corner on the same face is [111].
Taking the dot product, the angle between these two corner atoms may be found:
Taking the arc cosine of 1=3 gives the angle between A and B, which is found to be
109.478.
2/3
(463) 1/3
b
1/2
FIGURE 4.5
a Illustration of the (463) plane.
specific plane, but to any plane that has the same orientation, so a plane that intersected at
1,2=3,4=3 would also be the (463) plane.
If a plane does not intersect an axis as in Figure 4.6a, the intercept is taken to be 1 and its
Miller index would be zero. The plane intersects the a-axis at 1=2 and the b-axis at 1. We
have to add another unit cell below the a–b plane in Figure 4.6b to see that it intersects the
c-axis at 1. So the Miller indices are (20 1). Negative indices are expressed with a bar over
the index, just as in the case of lattice directions.
What if the plane contains an axis or goes through the origin and does not intersect
any axis? For example, the plane containing the b and c lattice vectors in Figure 4.7 can
be moved out one unit along the a-direction and would be referred to as (100). Similarly, a
plane that contains the c-axis and bisects the angle between the a and þb-axes could
be moved forward one unit along the a-axis (or b-axis) so that it intercepts a at 1, b at 1, and
c at 1 and would be written as (110). A plane that contains the origin and intersects no
axes can be brought forward until it intercepts the three axes as in the case of the (111)
plane in Figure 4.7.
c (210)
b
1/2
1/2
a
b
1/2
1/2
a
(a) (b)
FIGURE 4.6
(a) Plane to be specified does not intersect one of the axes and (b) unit cell added below to allow plane (201) to
intersect the c-axis at 1.
66 Introduction to the Physics and Chemistry of Materials
c c c
(110) (111)
(100)
b b b
a a a
FIGURE 4.7
Planes that contain one or more axes or go through the origin can be shifted by unit cells so that their intercepts can
be located.
Just as in the case of lattice directions, families of planes, such as the six face planes (100),
(010), (001), (100), (0
10), (001) are designated with a special symbol, in this case, {100}.
However, note that (100) and ( 100), (010) and (010), (001) and (001) are equivalent.
As stated previously, the Miller indices refer to a family of parallel planes. However,
sometimes it is necessary to refer to a specific plane. In this case, the plane is specified
without reducing it to the lowest set of whole numbers. For example, a plane intercepting a
at 1=2, b at 1, and c at 1 would be written as (200) instead of (100). The reason is that the
atoms in that particular plane are of interest and the (200) plane may very well have a
different set of atoms than the (100) plane. This notation is sometimes referred to as Laue
notation and much use is made of this notation in x-ray crystallography as seen in Chapter 6.
So far, no assumptions have been made concerning the properties of the lattice vectors
a, b, and c, so all of the notations we have developed thus far apply to any set of lattice
vectors. For cubic systems only, the Miller indices of a plane are the same as the direction
indices of a vector that is perpendicular to it; i.e., the (321) plane is perpendicular to the
[321] vector as demonstrated in Section 4.1.5.
x y z
þ þ ¼ 1, (4:2)
x 0 y0 z 0
where x0 , y0 , z0 are the intercepts of the plane on the respective coordinate axes. For the
orthorhombic lattice, a 6¼ b 6¼ c, a ¼ b ¼ g ¼ 908, the plane hk‘ intercepts the coordinate axes
at a=h, b=k, and c=‘. We wish to construct a vector d from the origin that is perpendicular to
the plane. To enforce this condition, we require d A ¼ 0 and d B ¼ 0 where A and B lie
in the plane hk‘. We construct vectors A and B in the hkl plane by setting A ¼ îa=h þ ^jb=k
and B ¼ îa=h þ kc=‘.^ The vector from the origin that is perpendicular
to this plane
is found by taking A B to give d ¼ const ^ih=a þ ^jk=b þ k‘=c^ . Inserting this in Equation
4.2, to find the constant, we obtain
0 1 !
h=a
@ A 1
dhk‘ ¼ k=b : (4:3)
‘=c ðh=aÞ2 þðk=bÞ2 þðl=cÞ2
Crystals and Crystallography 67
The magnitude of d is
1
dhk‘ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : (4:4)
h =a þ k2=b2 þ ‘2=c2
2 2
Notice that the direction of the vector perpendicular to the (hk‘) plane is [hk‘] for the case of
cubic symmetry.
If we multiply the intercepts a=h, b=k, and c=‘ by an integer n, Equation 4.2 becomes
xh yk z‘
þ þ ¼ n, (4:7)
a b c
and the distance from the origin to the nth plane is given by
n
dnhk‘ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi , (4:8)
h2 =a2 þ k2 =b2 þ ‘2 =c2
so that Equation 4.7 represents a family of planes whose spacing between them is given by
Equation 4.4 (or Equation 4.5 in the case of cubic symmetry). Since the (hk‘) plane cuts
through the unit cell, it may not contain any lattice points or atoms. To construct a plane
containing the lattice points, chose n to be the common factor for h, k, and ‘, i.e., set n ¼ 6 for
the (321) plane.
Equation 4.7 may be used as a test to see if a particular atom lies in the plane. For
example, consider the atom in the center of the bcc system (a ¼ b ¼ c) whose x, y, z
coordinates are a=2, a=2, a=2. Setting h ¼ 1, k ¼ 1, ‘ ¼ 0, and n ¼ 1, we see that Equation
4.7 is satisfied and the central atom does indeed lie in the (110) plane.
A common misconception among students is that the central atom in the bcc system lies
in the (111) plane. Equation 4.8 with h ¼ 1, k ¼ 1, ‘ ¼ 1 is not satisfied by x ¼ y ¼ z ¼ a=2. This
misconception is caused by the assumption that a cube is divided in half by the (111) plane.
Actually two different (111) planes can pass through a cube; one intercepting at (1,0,0).
(0,1,0), (0,0,1) and the other at (1,0,1), (0,1,1), and (1,1,0). These two planes are parallel but
distinct as may be seen by viewing the system along the [110] direction in Figure 4.8. The
central atom lies in the region between the two planes.
(0,0,1) (0,1,1)
(0,1,1)
(1,0,1) (0,0,1) (1,0,1)
(0,1,0)
FIGURE 4.8
A bcc structure showing two distinct (111) planes with their intercepts (left) and as seen from along the [110]
direction (right). The center atom lies between the two (111) planes indicated by heavy dashed lines.
(2h0 k0 ); j ¼ n=3(2k0 h0 ); k ¼ (h þ k); ‘ ¼ n‘0 ; where the primes represent the indices in the
three number system. Vector A in Figure 4.9 would be [100] in the three coordinate system.
Using the equations above, it transforms into [2110] when n is set to 3 in order to clear the
fractions. If vector A had intercepted the vertical axis at 1, making it [101], it would
transform into [ 2113]. If vector B is displaced to the right where it intersects a and b at 1,
it goes from 0, 1, 0 to 1, 0, 0. Vector subtraction yields [110]. Transforming to the four-index
system gives [1 100]. Notice that in this case, n is set to 1 so if this vector also had a vertical
component of 1, [1 11], it would transform into [1101].
The designation of planes is more straightforward and proceeds as before. Plane
A intersects the a-axis at 1=2 and axes b and c at 1. If it is perpendicular to the basal
plane, it would be designated as ( 2110). If it also intersected the z-axis at one unit, it would
be (2111). Plane B intercepts the b-axis at 1, the a-axis at 1, and the c-axis not at all. It
would thus be denoted (1 100). Again note that the sum of the first three Miller–Bravais
indices is always zero in this notation.
Vector A
[2110]
Plane B
[1100]
c b
Plane A
[2110]
c
c
b 90°
2
b
b
90°
a
a
(a) (b)
c
2
c
b
2
b
2
a a
(c) (d)
FIGURE 4.10
(a) Triclinic system, a 6¼ b 6¼ c and a 6¼ b 6¼ g 6¼ 908 or 1208; (b) monoclinic system, a 6¼ b 6¼ c and a ¼ g ¼ 908, b 6¼ 908.
‘‘2’’ indicated the twofold rotation axis and the dashed lines indicate mirror planes; (c) trigonal system, a ¼ b ¼ c and
a ¼ b ¼ g 6¼ 908 or 1208; (d) orthorhombic system, a 6¼ b 6¼ c and a ¼ b ¼ g ¼ 908.
Crystals and Crystallography 71
c 6
c
2
b
2
2 a
2
a b
(e) (f)
2 b
a
(g)
subscript d stands for diagonal and indicates the mirror planes that contain the threefold
axis and bisect the angles between the twofold axes. The complete international symbol is
32m, usually abbreviated as 3m, the bar above the 3 indicating a rotation-inversion axis
(a rotation followed by inverting the coordinates of each point along a line through a point
called the center of inversion).
the axes. The resulting crystal must then be described by C4v in the Schoenflies notation or
by 4mm in the international notation indicating a tetragonal lattice with fourfold symmetry
about the c-axis and two sets of vertical mirror planes.
Each of the other crystal systems has similar restricted symmetries and it can be
shown that there is a total of 32 unique sets of point symmetry operations or point groups.
The symmetry of every crystalline structure may be described by one of these 32 point
groups. Such classification of point symmetries is useful in the search for materials with
certain properties. For example, if one is looking for materials with permanent dipole
moments, one would look only at systems that are noncentrosymmetric, i.e., systems that
do not possess a center of inversion symmetry. The 10 noncentrosymmetric point groups
are 1, 2, 3, 4, 6, m, 2mm, 3m, 4mm, and 6mm.
Notice that rotation symmetry only exists for n ¼ 1, 2, 3, 4, and 6; five- and sevenfold
rotations are not allowed because bodies with these symmetries cannot fill all space for the
same reason that you cannot tile a floor with pentagons or with septagons. Icosahedral
quasicrystals with fivefold symmetry can form, but cannot grow into crystalline solids
in the strict sense of the word. Even so, such quasicrystals are extremely interesting
both theoretically as well as from applications that utilize their unusual properties, as
discussed later.
a
a⬘
b b⬘
FIGURE 4.11
Face-centered rectangular lattice in 2D. All the lattice points could be generated by translating along the oblique
primitive lattice vectors a and b, or by the nonprimitive vectors a0 and b0 with a basis of (0,0) and (1=2,1=2).
74 Introduction to the Physics and Chemistry of Materials
special symmetry associated with the cubic lattice, it is called the body-centered cubic or
bcc lattice and this nomenclature is generally used instead of the primitive trigonal lattice
as seen in Figure 4.12a. A similar situation arises when lattice points fall on the centers of
the faces of the cube. This lattice is also given the special name of face-centered cubic or fcc
lattice shown in Figure 4.12b.
Other members of the seven basic crystal systems can be constructed from simpler
primitive vectors if lattice points are placed in their centers or on their faces. The reader
may wonder why some of these seven basic crystal systems have separate body-centered,
face-centered, or base-centered lattices and others do not. The answer is, if the additional
lattice points produce a system of higher symmetry than the primitive lattice, the non-
primitive lattice is given a special name. A body-centered tetragonal (bct) shown in
Figure 4.12c has the symmetry of the tetragonal system, which is higher than the trigonal
primitive lattice. But then, why not a face-centered tetragonal lattice? Because it
c c
b b
a a
(a) (b)
(c) (d)
FIGURE 4.12
(a) bcc lattice, (b) fcc lattice. Both of these lattices could be represented by a primitive trigonal lattice by taking the
lattice vectors from the origin to the three nearest body-centered points in the case of the bcc or the three nearest
face-centered points in the case of the fcc. (c) bct lattice, (d) body-centered orthorhombic lattice. (e) base-centered
orthorhombic lattice, (f) face-centered orthorhombic lattice (not all lattice points shown), and (g) face-centered
monoclinic lattice (not all lattice points shown).
Crystals and Crystallography 75
would simply result in another bct. The same argument would apply to a base-centered
tetragonal lattice.
On the other hand, the orthorhombic lattice can have body-centered, base-centered, and
face-centered lattices, shown in Figure 4.12d through f, which are not redundant because
the three nonprimitive lattice vectors all have different lengths.
It is difficult to get more primitive than the tetragonal or triclinic lattice, and the base-
centered monoclinic lattice (Figure 4.12g) rounds out the 14 Bravais lattices.
M.L. Frankheim in 1842 was the first to classify the possible crystal lattices including
the special body-centered and face-centered nonprimitive lattices. However, he had mis-
takenly added a 15th structure that turned out to be redundant. Auguste Bravais was the
first in 1845 to correctly characterize the 14 unique lattices that now bear his name.
a2 FIGURE 4.13
hcp lattice. The unit cell capable of generating this structure
is outlined in heavier lines and the lattice points are black
a1
instead of gray.
76 Introduction to the Physics and Chemistry of Materials
nm
8
44
0.
1.001 nm
(110) Plane (001) Plane
FIGURE 4.14
Hg2Cl2 molecule with space group I4mmm. The smaller Hg atoms are dark gray and the Cl atoms are light gray.
Note that no atoms sit on any of the lattice points, but the environment (in this case back-to-back Cl) at each lattice
point is the same and the space group uniquely describes the configuration of the unit cell.
Space group notation begins with a capital letter, P for primitive, C for base
centered, F for face centered, I for body-centered (from the German Interzentrum), or
R for rhombohedral. The point group is designated next with subscripts indicating the
various screw axis and glide plane operations that leave the structure invariant. Care is
taken to eliminate any duplications where two different sets of operations produce the
same result such as the equivalence of 2 ¼ m and 6 ¼ 3=m. For example, an fcc crystal would
have a space group designated by Fm3m, a bcc crystal’s space group is Im3m, a bct crystal
is I4mmm, etc.
For a further example of how the point and space group describes a crystal
system, consider the mercurous chloride (Hg2Cl2) system. The crystalline structure consists
of parallel chains of covalently bonded Cl–Hg–Hg–Cl molecules aligned along the
crystallographic c-axis as shown in Figure 4.14. (The Cl anions are shown in light gray
and the Hgþ cations in darker gray. Their sizes are to scale.)
The point group can easily be seen to be 4=mmm from the symmetry. The space group is
I4=mmm. This additional piece of information tells us that the unit cell is body centered.
Note there are no ions at the center of the unit cell or at any of the other lattice points, but
the environment at the center, which is a point between two adjacent Cl ions, is the same
as at every other lattice point. (Note: We could have just as easily chosen the lattice points
to be between the two adjacent Hgþ ions.)
a FIGURE 4.15
The (100) face of an fcc structure showing the relationship between the
atomic radii and the length of the unit cell.
In the diagram of the (100) face of an fcc structure in Figure 4.15, note that the diagonal is
four times atomic radii R. Thus the relation between the unit cell length a and the atomic
radius for the fcc structure is
pffiffiffiffiffiffiffi pffiffiffi
4R ¼ 2a2 or a ¼ 2 2R: (4:9)
This same diagonal would be along the [111] direction in the bcc structure. Therefore, for
the bcc structure,
pffiffiffiffiffiffiffi 4
4R ¼ 3a2 or a ¼ pffiffiffi R: (4:10)
3
The basis for the bcc system is (0,0,0) (1=2,1=2,1=2); so there are eight corner atoms and one
interior atom for a total of two atoms per unit cell.
16p
APF ¼ ¼ 0:740: (4:15)
3 83=2
2 1
Planar density ¼ pffiffiffi ¼ pffiffiffi , (4:17)
4 3R 2 2 3R 2
3
Planar density ¼ pffiffiffi : (4:18)
8 2R 2
4R
FIGURE 4.16
The (111) high planar density face of the fcc structure showing how the
atomic planar density is calculated.
Crystals and Crystallography 79
FIGURE 4.17
The (110) face of a bcc structure has the highest planar
21/2 a density.
The atoms in the (211) and (321) bcc planes are shown in Figure 4.18. The area of a triangle,
whose sides are A, B, C, can be found from
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðA þ B þ CÞðA þ B CÞðA B þ CÞðA þ B þ CÞ
Area ¼ : (4:19)
4
p
Planar fraction ¼ pffiffiffi ¼ 0:907, (4:20)
2 3
[111]
[111]
(211) (321)
FIGURE 4.18
Atoms sitting in the (211) and (321) planes of the bcc structure. Atoms in the (211) plane have the coordinates
(1,0,0), (0,2,0), (0,0,2), (0,1,1) and (1=2,1=2,1=2) and satisfy the equation 2x þ y þ z ¼ 2. Atoms in the (321) plane
satisfy 3x þ2y þ z ¼ 6.
80 Introduction to the Physics and Chemistry of Materials
and from Equation 4.18, the planar fraction for the bcc system is
3p
Planar fraction ¼ pffiffiffi ¼ 0:833: (4:21)
8 2
4.4 Interstices
The positions between the atoms in a unit cell are called interstices or interstitial sites and
are important because they provide paths for diffusion in solids and are sites for small
impurity atoms that are often added to harden the material. They also affect the solubility
of small atoms such as carbon in iron, which plays a major role in steel making.
FIGURE 4.19
Interstitial site for a simple cubic lattice is located at (1=2,1=2,1=2).
Crystals and Crystallography 81
(a) (b)
FIGURE 4.20
(a) Edge octahedral interstitial sites in the fcc lattice shown in gray (not all points shown) and (b) central octahedral
site showing the six nearest neighbors forming the octahedron about the site.
pffiffiffi
r ¼ 1 2 R ¼ 0:414R: (4:22)
The interstitial sites located at (1=4,1=4,1=4), etc., have four nearest neighbors that form a
tetrahedron and are called tetrahedral sites. There are eight such sites per unit cell as may
be seen in Figure 4.21a. The tetrahedral coordination is best seen for the (3=4,3=4,3=4) in
Figure 4.21b.
The size of the tetrahedral interstitial site in the fcc structure can be shown to be
pffiffiffiffiffiffiffiffi
r¼ 3=2 1 R ¼ 0:225R: (4:23)
(a) (b)
FIGURE 4.21
(a) Eight tetrahedral sites in the fcc lattice and (b) the site at (3=4,3=4,3=4) is easiest to visualize because it is
coordinated with an outside corner atom and the three exposed face atoms.
(a) (b)
FIGURE 4.22
(a) Octahedral interstitial sites (gray) in the bcc lattice lie on the faces and are coordinated with the four corner
atoms plus two center atoms (not all shown) and (b) tetrahedral interstitial sites (gray) lie on the edges of the bcc
lattice and are coordinated with two corner atoms and the two middle atoms (not all shown).
Crystals and Crystallography 83
TABLE 4.1
Tetrahedral Sites
Atoms per Number of Tetrahedral Tetrahedral Number of Octahedral Octahedral
Structure Unit Cell Sites Sites r=R Sites Sites r=R
c
a2
1/4
a1 1/8
a1
(a) (b)
FIGURE 4.23
(a) Octahedral interstitial (open circle) and tetrahedral interstitial (black dot) in the (0001) planes of the hcp unit
cell and (b) octahedral interstitial (larger open circle) and tetrahedral interstitial (black dot) sites in the (1100) face
of an hcp unit cell showing their positions relative to the c-axis.
than the fcc or hcp lattice, notice that their octahedral sites are actually smaller than those
for the fcc and hcp because they are constrained by the short distance between the central
atoms in the bcc lattice system (Figure 4.23). This difference between the size of the
interstitial sites in the fcc and bcc plays a significant role in the solidification of the iron–
carbon system as discussed in Chapter 14.
4.5 Quasicrystals
Having gone to great lengths to stress the importance of long-range order in crystals
brought about by repeated translations of unit cells that fill all space, we now have to
inform the reader that nature makes exceptions to these rules of crystal growth. There is
another close-packed structure with a coordination number of 12 besides the fcc and the
hcp configurations and that is the icosahedron. The icosahedron, shown in Figure 4.24, is a
polyhedron with 20 triangular sides. It is possible for crystals to nucleate and grow in this
configuration, but since an icosahedron has six axes with fivefold symmetry, it cannot fill
all space, hence form a crystal with a periodic structure. It can, however, form an aggregate
of small icosahedral grains now called quasicrystals. The structure of the grains can be
quite complex, but there are several ways in which they can arrange themselves in a long-
range repeating structure so that x-ray diffraction (XRD) can reveal their fivefold (and
sometimes higher fold) symmetry.
84 Introduction to the Physics and Chemistry of Materials
(a) (b)
FIGURE 4.24
Vertices of an icosahedron also represent a close-packed structure with a coordination number of 12 (see Figure
5.19). Crystals can nucleate in this configuration, but since icosahedra cannot fill all space; they themselves do not
form periodic structures.
While working with rapidly solidified Al–Fe–Mn alloys at the NBS (now NIST) in 1981,
Dan Shechtman observed an unmistakable XRD pattern with tenfold symmetry. (Crystals
with n-fold symmetry, where n is odd, can diffract with 2n symmetry for reasons that will be
discussed in Chapter 6.) It was several years before the concept of quasicrystals became
generally accepted, but then intensive research was focused on searching for new quasicrys-
tals and understanding their properties. Most of the early work focused on systems pro-
duced by melt spinning (a technique for rapid solidification) because it was thought that
quasicrystals were metastable since they were not in a configuration in which all the bonds
were satisfied. Most of these systems were Al-transition metals or Al–Mg-transition metals.
However, it was later found that some quasicrystals such as Al65Cu20Fe15, Al65Cu20Ru15,
Al65Cu20Os15, and Al70Pd20V5Co3 were in fact stable.
How can icosahedra, which we know cannot fill all space, form solids with enough long-
range order to diffract x-rays? One of the simplest ways would be to place each icosahe-
dron on a Bravais lattice site and fill in the space between them with other atoms. Actually,
most quasicrystal structures are more complicated than this and will not be dealt with here.
Quasicrystals have some very unusual properties that are discussed in subsequent
chapters along with the properties of ordinary crystals. Suffice it to say at this point that,
although they consist of metals, their properties are very unmetal-like. One of their most
interesting properties is their extremely low surface energy, which means nothing wants
to stick to them. Taking advantage of this property, Lynntech, Inc. (College Station, Texas)
has developed an electrodeposition process for coating Al-3004 alloy with Al65Cu23Fe12
quasicrystals to form nonstick wear-resistant cookware.
4.6 Summary
A crystal is an array of atoms (or molecules) that repeat in three dimensions over distances
that are long compared to the interatomic spacing. Since this configuration represents
the lowest possible internal energy, most solids are in the form of crystals, although
these crystals may be microscopic in size. A polycrystalline solid consists of many
randomly oriented crystallites bonded together through their grain boundaries. Some
Crystals and Crystallography 85
applications require macroscopic single crystals in which the long-range order is extended
over millimeters or even centimeters. Noncrystalline or amorphous solids can be formed in
which the atoms themselves are randomly oriented resembling liquids that have been
frozen in place. Such solids are called glasses and have no grain boundaries. It is also
possible to form nanocrystalline materials in which the order extends only over a few
atomic spacings. Such materials possess some of the properties of ordered solids as well as
amorphous solids.
A unit cell of a crystal consists of a volume which contains the array of atoms that are
repeated. The smallest possible unit cell that can form the structure is a primitive unit cell
although is it often convenient to describe to solid using nonprimitive unit cells that have
higher symmetries. A unit cell is described by three lattice vectors that define the directions
the cell can be propagated to form the crystal lattice or framework of the structure. The
basis of the crystal describes the positions of the atoms within the unit cell. The basis plus
the lattice defines the structure. The set of all possible translations along the lattice vectors,
forms the translation group.
Directions relative to the lattice vectors are defined by the projections onto the lattice
vectors on these axes converted to whole numbers with no common denominator and set
in square brackets with no commas. Negative numbers are denoted by bars above the
number. Equivalent families of vectors, such as directions along the lattice vectors, are
denoted by pointed brackets.
Planes are denoted by their Miller indices h, k, and ‘ generated by taking the reciprocals of
the intersections of the plane with the lattice vectors and converting them to whole numbers
with no common denominator. The indices are set in parentheses with no commas and
negative numbers are denoted by bars above the number. The indices represent not only a
single plane, but all parallel planes that have the same spacing. For rectangular lattices, the
distance from the origin to the (hk‘) plane (as well as between planes with the same indices) is
given by
1
dhk‘ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi :
(h =a ) þ (k2 =b2 ) þ (‘2 =c2 )
2 2
Families of equivalent planes, such as planes perpendicular to the lattice vectors, are
denoted by curly brackets. For cubic lattices only, the vector perpendicular to a plane
has the same indices as the Miller indices of the plane.
Sometimes, particularly when describing higher order reflecting planes in XRD, it
is necessary to define a specific plane. In this case, the Miller indices are not reduced to
the nearest common denominator.
There are seven different polyhedra with different symmetries that can fill all 3-D space.
These polyhedra form the primitive cells of seven basic crystal systems. It is also possible to
form an additional seven nonprimitive systems with higher symmetry by adding lattice
points in the center or on the faces of the basic systems, thus forming the 14 Bravais lattices
that describe all crystals.
Many of the physical properties of a crystal are determined by its symmetry, which
can be described by its point group and space group. There are 32 possible point group
operations (operations than involve rotations or reflections about a point of a plane of
symmetry). By including certain other spatial operations, the 32 point groups can be
subdivided into 230 possible space groups that completely describe the symmetry of all
possible crystal systems.
Relations between the atomic radius R and the lattice dimension a in the cubic systems
are important for determining properties such as mass density, APF, planar densities, and
86 Introduction to the Physics and Chemistry of Materials
line densities. These latter two parameters determine the slip systems that are important in
the deformation of metals. For the fcc lattice
pffiffiffi pffiffiffiffiffiffiffi
a ¼ 2 2R or 4R ¼ 2a2
4 pffiffiffiffiffiffiffi
a ¼ pffiffiffi R or 4R ¼ 3a2 :
3
The interstitial spaces between the atoms in the lattice offer pathways for diffusion to occur
and places for impurity atoms to reside. There are two types of interstitial sites character-
ized by their coordination numbers: tetrahedral interstitial sites whose four nearest neigh-
bors form a tetrahedron and octahedral sites whose six nearest neighbors form an
octahedron. The small octahedral sites in bcc material result from the short distance
between the central atoms in the bcc structure and are responsible for the low solid
solubility of interstitial impurities in bcc material.
Just as we think we know everything about how matter is constructed, nature always has a
way of surprising us. Quasicrystals, crystals with fivefold symmetry that were not supposed
to exist, actually do. Their discovery has opened exciting new areas of research into their
structure and properties, especially when it was found that, even though they are metallic
systems, in many respects they behave more like ceramics and semiconductors than metals.
Bibliography
Allen, S.M. and Thomas, E.L., The Structure of Materials, Wiley MIT Series in Materials Science
and Engineering, John Wiley & Sons, New York, 1999.
Callister, W.D., Materials Science and Engineering: An Introduction, 7th edn., John Wiley & Sons,
New York, 2007.
Dunois, J.-M., Useful Quasicrystals, World Scientific Publishing, Hackensack, NJ, 2005.
Hahn, T., Ed., Space-group symmetry in International Tables for Crystallography: Volume A, 5th edn.,
Institut für Kristallographie, Technische Hochschule Aachen, Germany, 2005.
Harrison, W., Solid State Theory, McGraw Hill, New York, 1970.
Liboff, R.L., Primer for Point and Space Groups, Springer, New York, 2004.
International Tables for Crystallography, Vol. A, Th. Hahn, Ed., Springer, 2006, available online through
www.springerlink.com
For an online source of point and space groups posted by the University of Oklahoma Crystallo-
graphic Laboratory, see http:==xrayweb.chem.ou.edu=notes=symmetry.html.
Problems
1. Al forms an fcc lattice and has a density of 2.7 g=cm3 and a molecular weight of 26.98.
Calculate the lattice parameter, the atomic radius, and the atomic density.
2. Rank the planar fractions in order of the (100), (101), (111), (211), and (321) planes in the
bcc system.
Crystals and Crystallography 87
3. What are the directions of highest linear density in the planes mentioned in Problem 2?
4. Verify the relative sizes of the interstitial sites presented in Table 4.1.
5. Find the APF for the hcp structure.
6. Find the planar density fraction in the (1100) plane of an hcp lattice? What is the
direction of highest planar density?
7. What happens in a trigonal lattice when the angle between axes becomes 1208?
5
The Structure of Matter
The properties of matter are intimately tied to its structure, as we shall see in subsequent
chapters, and the structure of matter is primarily determined by the sizes of the atoms and
ions of its constituents and by the types of bonding between them. In this chapter, we will
explore some of the myriads of ways nature has found to assemble materials, starting with
simple one-component systems and then moving to compounds and finally to polymeric
molecular structures.
89
90 Introduction to the Physics and Chemistry of Materials
(a) (b)
FIGURE 5.1
(a) ABA stacking to form the hcp structure. Heavy black circles indicate where the next ball would be placed in
rows 2, 3, and 4. Note that every other vertex is uncovered. (b) ABC stacking to form the fcc structure. Heavy black
circles indicate where the next ball would be placed in rows 2, 3, and 4.
[111]
A B C
(110)
(a) (b)
(c) (d)
FIGURE 5.2
(a) (111) planes in the fcc lattice. (b) Projection of the (111) ABC planes on the (110) face. (c) High planar density
atomic configuration of the (111) A plane looking at the fcc cube along the [111] direction. (d) Addition of atoms in
the B and C planes forms the [111] corner of the fcc cube. The outline of the full cube is shown for reference.
The Structure of Matter 91
Na Mg Al
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd
Cs Ba La Hf Ta W Re Os Ir Pt Au Tl Pb
FIGURE 5.3
Periodic table of the metals that form fcc, hcp, and bcc structures. The transition metals tend to go from hcp ! bcc !
hcp ! fcc except for Mn, Fe, and Co whose magnetic properties apparently influence their structures. Mn has a
complex cubic structure with a space group I34m.
density as may be seen in Figure 5.2c. Adding the B- and C-layer atoms, we see the [111]
corner of the cube in Figure 5.2d.
Both fcc and hcp structures have the same coordination number (number of nearest
neighbors) of 12 and the same atomic packing fraction of 0.74 (ratio of the volume of
oranges to the container volume), but metals that form the fcc structure are generally
more ductile than hcp metals because there are more planes with high area fraction
which provide slip systems in the fcc structure. This will be discussed in more detail in
Chapter 9.
Metals solidifying with the fcc structure (space group Fm3m) include Al, Ag, Au, Ca,
Cu, Ir, Ni, Pb, Pd, and Yb. Metals solidifying with the hcp structure (space group
P63=mmc) include Be, Co, Gd, Ho, Mg, Ti, Tl, and Zn (see Figure 5.3). None of the hcp
structures have the ideal c=a ratio of 1.633. Mg and Co come the closest with c=a values
equal to 1.624 and 1.623, respectively. Be has the smallest with 1.568 and Zn has the
largest with 1.856.
Note that the space group for the hcp structure is P63=mmc. The point group is 6=mmm,
but the subscript 3 next to the 6 signifies additional symmetry in the space group in the
form of a threefold screw axis parallel to the sixfold rotational axis. Similarly, the c in place
of the last m signifies a glide plane parallel to the c-axis.
which structure a metal will chose to solidify in? And why do some metals solidify in bcc
structures instead of close-packed structures such as fcc or hcp? The number of nearest
neighbors for the bcc structure is 8 compared to 12 for the close-packed structures,
but the number of second nearest neighbors in the bcc structure is 6, and these are
closer than the second nearest neighbors in the close-packed structures. Therefore, with
14 nearest and not-to-far-away neighbors, one would expect the difference in binding
energy between bcc and the close-packed systems may not be very large. We know in
fact that changes in temperature or pressure can cause a metal to change its structure. Fe,
for example, changes from bcc to fcc at 9128C and back to bcc at 13948C before it melts at
15388C. These different forms of the same material are called allotropes.
So then how does a metal decide what structure it should solidify into? Examining the
periodic table of the metals that form fcc, hcp, and bcc structures in Figure 5.3 reveals some
semblance of a pattern. All of the monovalent alkali metals are bcc. Except for the magnetic
elements (Mn, Fe, and Co), the transition metals go from hcp ! bcc ! hcp ! fcc.
The structures of the other elements seem to have no particular order.
There is no simple explanation other than to say that the free energies of each of the
elements are determined by the temperature, pressure, and the distance between the ion
cores and their electronic environment. Subjecting elements to high pressure alters their
internuclear distance and makes a P–V contribution to the free energy which can produce
solid-state phase changes. The only simple explanation seems to be that those phases that
happen to have the lowest free energies under ambient conditions are the ones that appear
in Figure 5.3.
The ability to compute the free energies of various possible structures has evolved
over the years beginning with early attempts to obtain self-consistent solutions to the
Schrödinger equation for a many electron system. This computational capability was
greatly enhanced in the 1970s by the development of the density functional theory (DFT)
that earned Walter Kohn the 1998 Nobel Prize in chemistry. The essence of the DFT is the
replacement of the many-bodied wavefunction, which has 3N variables, with a density
function containing only three variables.
many unique mechanical, chemical, electronic, optical, and magnetic properties. Their uses
range from strengthening or even replacing superalloys (Ni–Al), to catalysts (Raney Ni),
protective surface coatings (Al–Cu–Fe icosohedral phase), semiconductors (GaAs and
AlN), thermoelectric materials (Bi2Te3), super-hard magnets (SmCo5 and Fe14Nd2B), and
even dental amalgams (Ag2Hg3 and Sn8Hg) that solidify at 378C and have the crush
strength of cast iron. The following few examples are intended to introduce the reader to
the enormous scope of these fascinating materials and to emphasize that there is still much
to be learned in the field of metallurgy.
FIGURE 5.4
A-15 structure of A3B compounds. The A-atoms
form aligned chains on the faces of the bcc lattice.
TABLE 5.1
Coordination Numbers
Ratio of Cation=Anion Diameter Possible Coordination Number
1.00 12
0.732–1.0 8
0.414–0.732 6
0.225–0414 4
0.155–0.225 3
<0.155 2
the surrounding anions from direct contact with each other. Otherwise, the repulsive force
between the counterions will be greater than the attractive force from the cation.
The next largest coordination number that would allow a repeating structure is 8,
which would be in the form of a simple cubic lattice with the anions at the eight corners
and the cation at the interstitial position at the center of the cube (or the equivalent in
which the cations sit on the corners of the cube and the anion is in the middle). Simple
geometry shows that the cation=anion diameter must be greater than 0.732 to prevent the
surrounding ions from touching each other. Following this line of reasoning, the coord-
ination number for various ranges of cation to anion diameter ratios is summarized in
Table 5.1.
The following are examples of the simplest and most common of the many possible
structures for ionic compounds.
FIGURE 5.5
CsCl structure with coordination number 8. This is a
simple cubic lattice, not a bcc lattice, with the counterion
in the middle.
FIGURE 5.6
Rock salt structure is fcc with the counterions occupying
all of the octahedral sites. The ionic bonds are shown as
heavy lines.
Again the charge of the cations (1 in the middle plus 12 on the edges, each contributing
1=4 to the unit cell for a total of þ4) is balanced by the charge of the anions (eight on the
corners times 1=8 plus the six on the faces, each contributing 1=2 for a total of 4).
Compounds with this structure include AgCl, BaS, CaO, CeSe, DyAs, GdN, KBr, Lap,
LiCl, LiF, MgO, NaBr, NaF, NiO, PrBi, PuC, RbF, ScN, SrO, TbTe, UC, YN, YbO, and ZrO.
Now that we have examined a few of the simpler AX ionic structures, we will look at a
few of the more complicated ionically bonded structures.
FIGURE 5.7
Fluorite structure, typical for AX2 compounds,
places the divalent A cations on the fcc sites and
the monovalent X anions on all of the tetrahe-
dral sites in order to balance the charge. Each
anion is coordinated with one corner and three
face cations. Each cation is coordinated by eight
anions.
group Fm3m) in Figure 5.7. Other compounds with the fluorite structure include AmO2,
AuAl2, AuIn2, BaF2, Be2B, CO2, CdF2, CeO2, CoSi2, EuF2, HgF2, Ir2P, Li2O, Na2O, NiSi2,
PtAl2, Rb2O, SrCl2, SrF2, ThO2, and ZrO2.
FIGURE 5.8
Perovskite structure for ABX3 compounds. The
quadravalent A cation is located in the center
(darker sphere), octahedrally coordinated by
the six divalent anions on the faces. The diva-
lent cations sit on the corners, octahedrally
coordinated by the six anions on the faces.
sites, and the B3þ ions sit on two of the four octahedral sites. Other compounds with this
structure include ZnAl2Se4, CrAl2S4, CaIn2S4, CdAl2S4, ZrCr2Se4, and ZnMn2Te4.
The inverse spinel structure has one of the B3þ cations on the tetrahedral site and the A2þ
cation and the other B3þ cation on the octahedral sites. The tetrahedrally coordinated cation
shares an O2 ligand with one of the two octahedrally coordinated cations. This feature is
responsible for the interesting magnetic properties of the ferrites, a class of ceramic
magnets that will be discussed in more detail in Chapter 25.
For rc=ra < 0.15, the maximum coordination number is 2. Because of the small size of
the Hþ ion, substances that are hydrogen bonded have a coordination number of
2 and can only form linear chains unless the molecules involved are multifunctional
such as in the case of H2O, which can form a variety of crystalline ices by hydrogen
bonding.
The reader should be cautioned that the radius rules above are by no means
absolute and in fact are only obeyed 50% of the time. More ionic compounds
have the rock salt (NaCl) structure and fewer have the CsCl structure than would be
predicted from the cation=anion radius. Still, the use of the radius ratio is a good place
to start in trying to understand the crystalline structure of binary ionically bonded
compounds.
[001]
[111]
A B
BA⬘
AC⬘
AC⬘ BA⬘ A
B CB⬘
BA⬘ BA⬘
A
B
[110]
C C⬘ A A⬘ B B⬘ C
(a) (b)
AA⬘ B B⬘ C C⬘
(c)
FIGURE 5.9
(a) Diamond structure is fcc (black) with every other tetrahedral interstitial sites occupied (gray). Some bonds are
omitted for clarity. A (dark gray) and C (light gray) planes are also shown. (b) Views of the diamond structure
are in three projections. The AA0 BB0 CC0 stacking sequence can be seen in the [110] direction. The planes
containing atoms in the [111] projection are indicated by the letters in the [110] projection. Notice that the
interstitial atoms are hidden by the fcc atoms in [111] projection. (c) View of four of the unit cells shown in
Figure 5.9b. Note the distorted hexagonal channels that run through the crystal.
assigned to one type of atom and gray to the other. Crystals with this structure include
AlAs, AlP, BN, BP, CdS, CdTe, GaAs, GaP, GaSb, HgTe, HgSe, InP, InSb, b-SiC, ZnS, ZnSe,
and ZnTe.
B
A⬘
A
B⬘
B
A⬘
A
B⬘
(a) (b)
FIGURE 5.10
(a) Structure of the (0001) planes of wurtzite. Light circles represent the B0 planes and the gray circles represent the
A planes. Heavy lines show the offset hexagonal rings in the two levels. Bonds between the light and gray atoms
have been omitted for clarity. (b) Stacking sequence of a wurtzite structure showing the AA0 BB0 stacking. Light
circles represent Zn and gray circles S. Heavy lines represent bonds (some omitted for clarity).
A
FIGURE 5.11
Graphite showing the ABAB layered structure.
Each lattice point in the A planes lies above the
center of a hexagon in the B plane and vice versa.
sp2 bonds. These planes can be rolled up like rolls of chicken wire to form graphite fibers
that are used to strengthen various composites.
It is now possible to make sheets of highly ordered pyrolytic graphite that have a
thermal conductivity of seven times that of aluminum. This material is finding its way
into thermal management systems ranging from heat sinks in laptop computers, to indus-
trial heat exchangers, to avionic components for high performance aircraft.
FIGURE 5.12
A bucky ball is basically a soccer ball with C atoms on
each of its 60 vertices forming a series of pentagons and
hexagons.
of C70) known as fullerites. These fullerites are harder and have a higher bulk modulus and
sound velocity than diamond.
It is possible to add interstitial atoms to fcc C60 to form A3C60 compounds known as
fullerides. Since there are 12 A-atoms in the unit cell, all tetrahedral and all octahedral sites
will be occupied. If A is potassium, the materials will be superconducting with a critical
temperature of 18 K. Going up the group 1 monovalent metals increases the critical
temperature (Tc) until A3 is Rb2Cs, which has a Tc of 28 K. Interestingly, Rb3C60 is not
superconducting.
Because the Cn molecule is like a hollow shell with a diameter ranging from 0.4 to 1 nm
as n goes from 60 to 240, other atoms or molecules can be placed inside of this protective
cage to form a new class of materials called endohedral fullerenes. Dipolar salt molecules
such as LiF, LiCl, and NaCl encapsulated in C60 exhibit paramagnetic ordering. Molecular
hydrogen can be encapsulated in C60. Metal atoms from group II and III encaged by C82
form endohedral mettalofullerenes. Electron transfer takes place between the metal atom
and the C cage to form a super atom such as Y3þ @C823 (the @ indicates the Y is inside the
C82 molecule).
Ch Ch Ch
b a
b a b a
FIGURE 5.13
(a) m ¼ n, armchair structure. (b) m ¼ 0, n ¼ 1, zigzag structure. (c) m ¼ 1, n ¼ 2, chiral structure.
as Ch ¼ am þ bn, where a and b are unit vectors defined as shown in Figure 5.14 and m and
n are integers. The chiral vector is taken to be parallel to the circumference of the nanotube.
If m ¼ n, the ‘‘armchair’’ structure in Figure 5.13a results. If m or n ¼ 0, the zigzag structure
in Figure 5.13b results. Structures with other values of m or n as seen in Figure 5.13c are
referred to as ‘‘chiral.’’
The electric and thermal conductivity of metallic SWNTs can be very high; electron
trajectories are almost ballistic-like with little scattering from the lattice atoms. Current
densities can be a thousand times greater than in metals. The electrical properties of
SWNTs are determined by the quantum confinement of the electrons around the circum-
ference of the SWNT and therefore are dependent on the diameter and chirality of the
structure. The SWNTs are metallic or conductive if (2m þ n)=3 ¼ an integer. Otherwise
the SWNT is a semiconductor whose bandgap depends on the diameter as well as the
m and n values. Inserting various endohedral metallofullerenes to form ‘‘pea-pod’’
structures can alter bandgaps and other electronic properties thus allowing SWNTs to
function as diodes and transistors. Clearly SWNTs will play important roles in the future
of nanoelectronics.
The development of ways to make SWNTs and control their structure as well as ways
to manipulate them and incorporate them into devices is at the forefront of research in
nanotechnology. Recent developments in CVD growth of nanotubes have resulted in the
ability to grow bundles of CNTs that can be harvested and spun into fibers to make a
super thread. The individual CNTs in a super thread are bonded by van der Waals forces,
but it is possible to alter their mechanical and electrical properties by heat treating and
irradiation. The sides of CNTs have been functionalized to bind with epoxies and
polymers to form composites and their tips have been functionalized to serve as chemical
sensors, atomic force microscope (AFM) tips, ion and electron emitters, and for other
novel applications.
FIGURE 5.14
Bk ring structure of a-boron nitride. Unlike
graphite, the layers sit on top of each other are
bonded together by a combination of ionic and
covalent bonds to make a very strong and inert
structure.
FIGURE 5.15
Layered C7 structure of Mo2S. The layers consist of S atoms (light
spheres) on either side of the Mo atoms (grey spheres) in an hcp
arrangement. Since all bonds are satisfied, the layers slip over each
other, attracted only by van der Walls forces.
The Structure of Matter 107
5.4.9 Quartz
The silicates, which are thought to have a mixture of ionic and covalent bonds, can
form a vast variety of structures. The basic building block is the orthosilicate ion, SiO4 4 ,
formed by a Si atom in the center of a tetrahedron of four O atoms as seen in Figure 5.16.
The tetrahedra can join with other tetraherda by sharing the bridging O atoms at
the corners of the tetrahedra to form silica or SiO2. Because there is little barrier to rotation,
these bonds allow a random network structure (fused quartz), which is the basis for the
silicate glasses, as well as ordered crystalline structures to be formed.
Crystalline quartz is polymorphic, meaning it can have more than one crystal structure.
In fact nine different structures of quartz are classified in mineralogy tables, several
of which are metastable forms, meaning they are only stable at high temperatures
and pressures. The stable form under ambient conditions is a-quartz or low quartz,
which is stable up to 5738C. It has a trigonal structure with space group P3221. Above
5738C, b-quartz, or high quartz (C8), is stable under normal pressure. It is hexagonal with
space group P6222. Above 8678C, hexagonal b-tridymite (C10) forms with space group
P63=mmc. Finally, b-cristobalite (C9) forms above 14708C before melting at 17138C. The Si
atoms in b-cristobalite form an Fd3m diamond structure of Si atoms with the O atoms
inserted between them. The structure of these different polymorphs gets progressively
more open as the temperatures of their stable phase increases. At high pressures, a
monoclinic form called coesite can exist and at even higher pressures, tetragonal stisho-
vite (P4=mmm) forms. These latter two metastable forms are found in impact craters.
If crystalline quartz is compressed, say by forcing the top ion toward the bottom three
ions, the Si4þ ion will be displaced from the center of the charge distribution resulting in a
displacement current. Thus quartz is a piezoelectric material. Conversely, an applied
electric field will cause the structure to distort. More details about this highly useful
property will be discussed in Chapter 23.
FIGURE 5.16
The SiO44 tetrahedron, the basic building block
of the silicates. The small Si4þ ion in the center
can barely keep the large O2 anions from
touching one another.
108 Introduction to the Physics and Chemistry of Materials
5.4.11 Zeolites
Another important class of the aluminosilate family is the zeolites. Zeolites have porous
open structures that can accommodate various alkali metal ions that can easily be
exchanged, making them useful as molecular sieves and as powerful catalysts.
There are 48 natural occurring zeolites and many more have been synthesized for
specific purposes. They are widely used in gas and water purification and as sorption
pumps in vacuum systems. They are used as desiccants because of their ability to adsorb
(a) (c)
(b) (d)
FIGURE 5.17
Structures formed by combining the basic SiO44 orthosilicate unit (a). Combining two of the units forms the
Si2O76 backbone for the pyrosilicates (b). Longer chains (c) and rings (d) are the backbone for a large number
of minerals.
The Structure of Matter 109
FIGURE 5.18
Bottom view (upper) and edge view (lower) of a
(Si2O5)2 sheet structure. Hydroxides of Al, K, and
Mg can bond with the O ions on the top of this
layer to form electrically neutral sheets that make
up layered structures such as clays, talc, and mica.
large amounts of water. Added to soils, they can store water, making plants more drought
resistant. As solid-state acids, they are used in the petrochemical industries for cracking.
The nuclear industry uses them for trapping and immobilizing radioactive waste.
FIGURE 5.19
Icosahedral cluster of boron atoms (T-50). These clusters occupy
five symmetry sites in the primitive cell of the P42=nnm space
group.
FIGURE 5.20
The 8-atom molecular rings of sulfur (A-16). The space
group is Fddd.
increases as the temperature is lowered until the glass transition temperature is reached.
Above the glass transition temperature, the material is in the state of an undercooled melt.
Below the glass transition temperature, the molecules are locked into place just as a
crystallized solid and the viscosity of the glass is virtually infinite. However, since the
molecules in a glass form a random network instead of an ordered solid, the volume of a
glass is slightly larger than its crystalline counterpart.
Since the glass consists of a random network of molecules, it lacks the long-range order
of a crystalline solid, hence a glass does not show sharp x-ray diffraction peaks. Instead
low, broad peaks in a continuum can be seen that reflect only the short range order of the
molecule. Since glass is essentially transparent to x-rays, glass tubes are used to contain
powder x-ray samples. We are accustomed to thinking of glass as being transparent to
visible light because most glasses are formed form ceramics whose bandgap is higher than
The Structure of Matter 111
the top of the visible spectrum. Being transparent to visible light is by no means a property
of a glass; metallic glasses are as reflective as crystalline metals.
In 1932, Zachariasen proposed a continuous random network model that predicted that
oxide glasses of the form AnOm could exist with configurational energies that were only
slightly higher than the crystalline state if the following conditions were met:
Examples of these rules are illustrated in Figure 5.21 (see Chapter 14 for more details on
glass formation).
(a)
(b)
FIGURE 5.21
Two different ways to form a random net structure according to Zachariasen’s rules. (a) represents SiO2 in which
the shared bridging O atoms sit at the vertices of the tretrahedra and (b) represents B2O3 in which the repeating
units are triangles.
The Structure of Matter 113
H H H H
R +C C R C C
H H H H
FIGURE 5.22
H H H H H H H H Illustration of addition or chain polymerization.
A free radical R. opens a double bond of an
R C C + C C R C C C C ethylene molecule, transferring the unpaired elec-
tron to its opposite end where it can open the
double bond of another polyethylene molecule
H H H H H H H H and so on.
atom. Thus the chains actually form random coils and their average length is N1=2 times the
length of the repeating unit, where N is the number of repeating units.
The chains are held together by secondary bonds and can slip by each other in response
to stress, which accounts for the plastic response of polymers. This response can be greatly
varied depending on the structure of the polymer chains. Increasing the molecular weight
of the polymer (by adding more mer units) increases the stiffness and yield strength of the
polymer and raises its glass transition temperature (the temperature at which chain motion
is no longer possible and the material exhibits brittle rather than plastic behavior). Under
favorable circumstances, the chains can fold back and forth on themselves to form a
crystalline structure. Crystallinity combines strength and stiffness with ductility.
A variety of different properties can be obtained by replacing a hydrogen in polyethylene
with a large anion such as Cl to form polyvinyl chloride, or with blocky side groups such
as CH3 to form polypropylene, or with an aromatic ring to form polystyrene. These
structures are illustrated in Figure 5.23.
Repeating unit
Repeating unit
Carbon Hydrogen
Polyethylene Polypropylene
Polyvinylchloride Polytetrafluoroethylene
(PVC) (PTFE)
FIGURE 5.23
Examples of linear chain polymers shown as a zigzag chain to emphasize that the bond angles on the carbon
atoms in the chain are 1098.
114 Introduction to the Physics and Chemistry of Materials
O O H H CH3 O
C C O C C O O C O C
H H CH3
O O
H O H O
C C
N C N C C C N N
C C
H H H H
6 6 O O
Nylon 6,6 Polyimide
FIGURE 5.24
Repeating units of selected polymers.
A variety of different repeating units may be used instead of ethylene, which gives the
polymer scientist much freedom and versatility in the development of new materials.
A few examples are shown in Figure 5.24.
OH
CH3 CH3
R R R R
Isotactic
R R
R R FIGURE 5.26
Example of isotactic and syndiotactic
stereoregular polymers. Polymers with
random locations of their side groups
Syndiotactic are called atactic.
the side groups to alternate from one side to the other in regular fashion. This configuration
is call syndiotactic. If the distribution of the side groups is random, the configuration is
atactic. The isotatic and syndiotactic configurations are shown in Figure 5.26. On a one-
dimensional (1D) diagram such as Figure 5.24, it would appear that one could change the
position of a side group by a simple bond rotation. However, as can be seen in Figure 5.26,
such a rotation is not possible without breaking bonds.
Isotactic and syndiotactic polymers are referred to as stereoregular. Being able to control
the positions of the side groups so that they repeat in a regular fashion is important to
make high-strength materials because the uniform structure leads to close packing of the
polymer chains and a high degree of crystallinity. The catalyst systems employed to make
stereoregular polymers are now referred to as Ziegler–Natta catalysts.
H H H H H
H
C C C C C
C
C C C C C C
H H H H H
H
CH3 H CH3 H CH3 H
cis-Polyisoprene
H H H
H CH3 H
C C C
H H H
C C C C C C
H H H
C C C
CH3 H CH3
FIGURE 5.27 H H H
Comparison of the cis- and trans-iso-
mers of the isoprene repeating unit. trans-Polyisoprene
to which side of the double bond the side groups are located on. The cis-configuration
places both side groups on the same side while the trans-configuration places them on
opposite sides. Since atoms forming a double bond cannot be rotated, the effect is to place
the CH3 groups on the same side in cis-polyisoprene and on alternating sides in trans-
polyisoprene.
Both cis- and trans-polyisoprene are natural products that come from different trees. The
cis-polyisoprene is natural rubber while trans-polyisoprene is something called gutta
percha, a tough leathery substance that was originally used for golf balls and for under-
water insulation. It has now been largely replaced by synthetics although it is still used for
packing teeth after a root canal.
5.6.7 Vulcanization
Natural rubber was too soft to be of much use. It was called rubber by Joseph Priestly,
who found it useful for rubbing out pencil marks. The vulcanization process, developed
in 1839 by Charles Goodyear, added sulfur to open the double bonds in the isoprene
molecule and form covalent sulfur bonds that cross-linked the chains together as illus-
trated in Figure 5.28. The tougher product immediately found use in bicycle tires and
later in automobile tires. The shortage of natural rubber during World War II prompted
the development of synthetic rubbers with properties far superior to the best natural
rubbers.
5.7 Summary
Because there is no directionality in the metallic bond and because the charge on the ion
cores is screened by the intervening sea of electrons, metals tend to form close-packed
structures. The two close-packed structures that can fill all space are the fcc and the hcp.
The two structures differ in the packing sequence of hexagonal layers, hcp packs as ABAB,
The Structure of Matter 117
H H H H H H
S S
H H H H H
H
C C C C C
C
C C C C C C
H H H H H
H
H CH3 H CH3 H
CH3
FIGURE 5.28
Vulcanization of cis-polyisoprene. The addition of sulfur opens the carbon double bond and forms a covalent
sulfur bridge between two adjacent chains which greatly strengthens the final product.
whereas fcc packs as ABCABC. The coordination number (number of nearest neighbors)
for both systems is 12 and they both have the same atomic packing factor, 0.74. The fcc has
a higher symmetry and more high atomic fraction planes which can act as slip planes,
making fcc metals more ductile (less brittle) than hcp metals.
Other metals form bcc structures with eight nearest neighbors and six second nearest
neighbors, which are closer than the second nearest neighbors in fcc or hcp. These 14 close
neighbors apparently produce nearly the same bond energy as the two close-packed systems
because nearly as many metals form this structure as those forming either the fcc or hcp
structures. All the monovalent alkali metals form fcc structures and the nonmagnetic
transition metals tend to follow a pattern pairwise that goes hcp ! bcc ! hcp ! fcc.
Calculation of the free energies of these metallic systems requires solving the Schrödinger
equation for a many electron system. Recent advances in this field have been facilitated
by the development of the density functional theory (DFT) by Walter Kohn (Nobel
Prize, 1998).
Ionically bonded systems must obey Pauling’s rules for ionic crystal formation which,
among other things, requires each ion to be surrounded by a polyhedron of counterions
and also that the cation must be large enough to at least touch each of the counterions.
These requirements prevent ionically bonded systems from forming close-packed struc-
tures and the ratio of the cation to anion size will play a role in determining the
coordination number. In addition, the unit cell containing the anions and cations must
be electrically neutral and must be able to fill all space by translations along its lattice
vectors.
Covalently bonded systems, because of the directionality of the covalent bond, tend to
form open systems with low coordination numbers. Like the metals, they can form cubic
and hexagonal systems by altering their stacking sequence.
The various covalent and ionic structures are summarized in Table 5.2.
118 Introduction to the Physics and Chemistry of Materials
TABLE 5.2
Summary of Common Ionic and Covalent Structures
Structure Type Lattice Unit Cell Description Examples
Diamond A fcc Fd3m A-atoms on fcc sites and on every C, Si, Ge, and a-Sn
other tetrahedral site
Sphalerite (zinc AB fcc F43m A-atoms on fcc sites and B atoms ZnS, SiC, and most
blende) on every other tetrahedral site III–V
Wurtzite A(Zn) B(s) hex P63mc Zn atoms on hexagonal sites to ZnS, some III–V and
form A layer. S atoms over Zn II–VI
atoms to form A0 layer. Zn atoms
on every other tetrahedral site to
form B layer. S atoms over Zn
atoms to form B0 layer resulting
in alternating stacks of AA0 BB0
a-BN AB hex P63mc Planar hexagonal rings alternating a-BN
A and B covalently bonded
together
Cesium chloride AX sc Pm3m A-atoms on corners of cube and B Cs–Cl, AgZn CsBr,
atom at center of cube and BeCu
Rock salt (NaCl) AX fcc Fm3m A-atoms on fcc sites and B atoms NaCl, MgO, and FeO
on every octahedral site
Fluorite AX2 fcc Fm3m A-atoms on fcc sites and B atoms CaF2, UO2, and ThO2
on every tetrahedral site
Perovskite ABX3 sc Pm3m A in the center of the cube, B on BaTiO3, SrZrO3,
the corners, and C on the faces SrSnO3,
Spinel AB2X4 fcc Fd3m X on fcc sites, A on a tetrahedral Al2O3 MgO
site and B on two adjacent Al2O3 FeO
octahedral sites
Inverse spinel AB2X4 fcc Fd3m X on fcc sites, one B on a Fe2O3 FeO
tetrahedral site. Other B and
A on octahedral sites
The stable form of carbon is graphite in which the sp2 bonds form hexagonal sheets
bonded together by weak van der Waals forces. However, carbon atoms can also form
a new class of materials called fullerenes which consists of molecules in the form Cn
(n ¼ 60–140) as well as cylindrical structures called CNTs. The C60 molecules, known as
bucky balls, can assemble in an fcc close-packed array to form fullerites, a solid form of
carbon that is harder than diamond and has many other remarkable properties, especially
when metal atoms are intercalated into the structure to form fullerides. The CNTs also have
remarkable mechanical and electronic properties ranging from super-strong composites
to nanoelectronics.
Most polymers consist of covalently bonded long chain molecules that make secondary
bonds with each other. The physical properties are dependent on the length of these chains
and on the type and arrangement of the side groups that are placed on the chains. Linear
chains with regularly oriented side groups promote crystallinity and such polymers are
stronger, more dense, and have higher glass transition temperatures than polymers with
randomly placed side groups or side branches.
Most chain polymers will soften and become more plastic when heated above their glass
transition temperature and their chain molecules are able to move past one another. Such
polymers are called thermoplastics. Chain polymers can be hardened by cross-linking
while other polymer systems form covalently bonded networks. Such systems do not
soften and primary bonds must be broken in order to melt. Such systems are called
thermosetting plastics or thermosets.
The Structure of Matter 119
Bibliography
Papers Dealing with Methods for Computing Lattice Energies
Hohenberg, P. and Kohn, W., Inhomogeneous Electron Gas, Phys. Rev. B864, 1964, 136.
Koch, W. and Holthausen, M.C., A Chemist’s Guide to Density Functional Theory, 2nd edn., Wiley-VCH,
Weinheim, 2002.
Kohn, W. and Sham, L.J., Phys. Rev. A1133, 1965, 140.
Kohn, W., Electronic Structure of Matter—Wave Functions and Density Functionals, Nobel Lecture, Jan.
28, 1999.
Problems
1. Show that the ideal c=a ratio for an hcp structure is 1.633.
2. Verify the cation=anion radius for the various coordination numbers shown in Table 5.1.
3. Write the basis for the following structures: (a) hcp metal, (b) sphalerite, and (c) wurtzite.
4. Pd can absorb 900 times its volume of H under standard conditions. Write a formula for
this compound.
6
Reciprocal Lattice and X-Ray Diffraction
Most readers are probably familiar with the Bragg formula for x-ray diffraction (XRD),
2d sin u ¼ nl, and the selection rules that tell us that the sum of the Miller indices for body-
centered cubic (bcc) crystals must be even and the indices for face-centered cubic (fcc)
crystals must be all even or all odd in order to produce a diffraction peak. The intent of this
chapter is to give a more general derivation of the Laue conditions leading to Bragg
reflections of electrons and phonons that take place in three-dimensional (3D) crystals.
These reflections from lattice planes within the crystal are fundamental to the understand-
ing of the thermal and electronic properties of materials. In the process, the reciprocal
lattice will be introduced which will be widely used in subsequent chapters to describe
how materials behave thermally, electronically, magnetically, and photonically.
h 2ph
p¼ ¼ ¼ h k, (6:1)
l l
where k ¼ 2p=l is sometimes called the propagation vector. (For mass particles, l is the
deBroglie wavelength.)
121
122 Introduction to the Physics and Chemistry of Materials
The electron density should be invariant under any translation operation T such that
n(x þ T) ¼ n(x), where T ¼ ma. In other words, the electron density in one cell should look
the same as in any other cell. We can see that this is the case from
X X
n(x þ T) ¼ e h e2pihx=jaj e2pihm ¼
A e h e2pihx=jaj ¼ nðxÞ,
A (6:3)
h h
such that n(r þ T) ¼ n(r). To accomplish this, we must find a vector G that expresses
the periodicity of the lattice in reciprocal space since G must have the dimension of 1=r.
It may be seen that in one dimension, G ¼ 2ph=a, where h is an integer and r ¼ a.
In a 3D space, T ¼ ma þ nb þ pc, where a, b, c are the lattice vectors and m, n, and p are
integers. To enforce n(r þ T) ¼ n(r), Equation 6.3 must be written as
X
n(r þ T) ¼ e G eiðGrÞ eiðGTÞ ¼ nðrÞ:
A (6:5)
G
This requires expðiG TÞ ¼ 1, which means that G T ¼ 2p integer.
To facilitate the construction of the vector G, we define G as a translation vector in
reciprocal space:
where
A, B, and C are unit cell vectors in reciprocal space
h, k, and ‘ are integers.
(You may note we use the same notation for these integers as we did for the Miller indices.
This is no accident—you will find later that they are the Miller indices.)
Using the definition of G as the translation vector in reciprocal space,
G T ¼ ðhA þ kB þ ‘CÞ ðma þ nb þ pcÞ ¼ 2p integer:
a A ¼ 2p
a B¼0
a C¼0
b A¼0
b B ¼ 2p
b C¼0 :
c A¼0
c B¼0
c C ¼ 2p
To evaluate the const, we dot a with the first of the above equations and set it equal to 2p.
a A ¼ const a b c ¼ 2p, (6:8)
from which const ¼ 2p=a bc. Therefore, the inverse lattice vectors are given by
bc ca ab
A ¼ 2p , B ¼ 2p C ¼ 2p
a bc a bc
,
a bc
, (6:9)
and the reciprocal lattice translation vector Ghk‘ ¼ hA þ kB þ ‘C plays the same role
in reciprocal space as the translation vector T in direct space and the lattice points in
reciprocal space represent planes in direct space.
Thus we have constructed Ghk‘ so that it is perpendicular to the plane with Miller indices
h, k, ‘ and it can be shown (proof left to the student) that the distance from the origin to
lattice plane (hk‘) is given by
2p
dhk‘ ¼ : (6:10)
jGhk‘ j
2pðb cÞ 2p 2p 2p
A¼ ¼ b
x, B ¼ b C¼ bz:
a bc a a
y, and
a
(6:11)
The reciprocal lattice vectors form a simple cubic lattice in reciprocal space with a lattice
constant given by 2p=a.
We use the primitive vectors for the bcc lattice to distinguish its special symmetry from the
nonprimitive bcc lattice and use the prime symbol to distinguish these primitive lattice
vectors from the lattice vectors a, b, and c. The choice of the primitive vectors is arbitrary.
Any set will produce similar results. Using the set of relations in Equation 6.9 to transform
these vectors to reciprocal space,
0 1 0 1 0 1
1 0 1
2p @ A 2p @ A 2p @ A
A0 ¼ 1 , B0 ¼ 1 , C0 ¼ 0 : (6:13)
a a a
0 1 1
These are the Cartesian coordinates of a set of primitive translation vectors for an fcc lattice.
124 Introduction to the Physics and Chemistry of Materials
(022)
(002) (020)⬘
(010)⬘
(202)
(002)⬘ B⬘
(001)⬘ C
C⬘
FIGURE 6.1 B
Primitive and nonprimitive vectors in an fcc recip- (020)
rocal lattice in which the primitive vectors and A A⬘
their translation sites are denoted by prime sym-
bols. Translations along these primitive vectors (100)⬘
produce a nonprimitive fcc reciprocal lattice with
a basis of (000), (110), (101), (011). Notice that
primitive translations (100)0 are equivalent to (200) (200)⬘
(110), (200)0 to (220), etc. (220)
where h0 , k0 , ‘0 specify the translations along the primitive vectors. These translations along
the primitive vectors will map all of the lattice points of the fcc lattice, but the coordinates
of the lattice points will be different from the nonprimitive coordinates as may be seen
in Figure 6.1.
The nonprimitive reciprocal lattice vectors A, B, and C are given by Equation 6.9 using
the nonprimitive lattice translation vectors a, b, c. These nonprimitive reciprocal lattice
vectors are along the x, y, z axes and form a cubic unit cell with length 2p=a and a basis of
(000), (110), (101), (011)
pffiffiffi as may be seen in Figure 6.1. Note that the primitive reciprocal
lattice vector jA0 j ¼ 2jAj.
Putting the Fourier expansion for n(r) from Equation 6.4 into Equation 6.16,
ð X
E k0 dV e G exp½iðGhk‘ DkÞ r:
A hk‘ (6:17)
hk‘
crystal
The wave scattered in direction k0 can only have significant amplitude when the waves
scattered from the same point in each cell arrive in phase, which requires
This is the Laue condition for Bragg reflection. The Laue diffraction condition stated in the
Equation 6.18 is a necessary, but not sufficient condition, for a diffraction peak. The
intensity of the peak, or whether the peak may be cancelled out, depends on the contents
of the unit cell as discussed later.
The Laue condition can also be written as
k þ Ghk‘ ¼ k0 : (6:19)
Squaring both sides and making use of the fact that the amplitude of the scattered wave is
unchanged for the incident wave (elastic scattering), we can write
k⬘ P
P−r
R+r
P
r
k⬘ Δk
R
k FIGURE 6.2
0 Schematic of x-rays being diffracted.
126 Introduction to the Physics and Chemistry of Materials
2k G þ G2 ¼ 0: (6:20)
If we dot the Laue condition Ghk‘ ¼ Dk with the lattice vectors a, b, and c, we obtain the
Laue equations:
a Dk ¼ 2ph, b Dk ¼ 2pk
c Dk ¼ 2p‘: (6:21)
Each equation describes a cone of possible Dks about that particular lattice vector that
could satisfy the Laue condition. All three equations must be satisfied simultaneously in
order to meet the Laue condition. This requires that each cone must have a common line of
intersection with the other cones. Note that if G satisfies the Laue equations (Equation
6.21), G will also (h, k, ‘ denote the same plane as h, k, ‘). Therefore, the Laue
condition Equation 6.20 can also be written as 2k G ¼ G2 .
6.2.2 Ewald Construction
Ewald devised a construction in reciprocal space that provides a good way to visualize
how the Laue condition may be met. An Ewald sphere with radius k ¼ 2p=l is drawn
around the crystal. A vector k is drawn from the crystal in the direction of the incoming
x-ray beam to a reciprocal lattice point on the Ewald sphere which is taken as the origin of
the reciprocal lattice (000) as shown in Figure 6.3. If another reciprocal lattice point falls on
the Ewald sphere, as is the case for the (100) point shown in Figure 6.3 (remember,
reciprocal lattice points denote planes in the direct lattice), the Laue condition G ¼ k0 k
is met by a vector k0 drawn from the crystal to that point. The diffracted beam k0 makes
angle 2f with the incident beam and f with the G100 vector which is normal to the (100)
plane. Recall from Equation 6.10 that jG100j equals 2p=d100 where d100 is the spacing
between the (100) planes in the direct lattice. From Figure 6.4, it may be seen that
jG100 j ¼ 2jkj sin u and since jkj ¼ 2p=l,
2p 2p
jG100 j ¼ 2 sin u ¼ (6:22)
l d100
from which the more familiar form of Bragg’s law, 2d100 sin u ¼ l is obtained. Figure 6.4
illustrates this relationship in the direct lattice.
(100)
G
k=
k⬘ − k⬘
q f 2q
(000)
k Crystal
FIGURE 6.3
Ewald construction for an x-ray with wave vector k inci-
dent at angle f relative to the (100) plane. The reflection
spot will be in the direction of k0 . A simple geometrical
argument shows this construction satisfies Equation 6.20.
Reciprocal Lattice and X-Ray Diffraction 127
f f
q
(000)
What if the reciprocal lattice point corresponding to the (200) plane was sitting on the
Ewald sphere? The reciprocal lattice vector would now be jG200j ¼ 2p=d200. Bragg’s law
would be written as 2d200 sin u ¼ l. The d200 is the spacing between the (200) planes, which
falls halfway between the (100) planes. But what if there were no atoms on the (200) plane
to diffract the x-rays? Reflections from the (200) plane can be thought of as second-order
reflections from the (100) planes in which the path difference between the two rays,
2d ¼ 2d100 sin u ¼ 2l or 2d100 sin u ¼ nl, where n ¼ 2.
A useful feature of the Ewald construction is the ability to see which planes will be
able to meet the Laue condition. Consider the reciprocal lattice superimposed on the
Ewald sphere shown in Figure 6.5a. In this case, the k is 2.4 times larger than the reciprocal
lattice vector, which means the direct lattice spacing a is 2.4 times larger than the wave-
length l. None of the reciprocal lattice points lie on the Ewald sphere (at least in the
(030)
k⬘
2q
k
(000)
k
(000) (500)
(500)
FIGURE 6.5
(a) Ewald sphere lattice projected on a simple cubic reciprocal lattice. The k is 2.4 times larger than the reciprocal
lattice vector. None of the lattice points satisfy the Laue condition. (b) The crystal is rotated by 138 to satisfy the
Laue condition for the (010) reflection. The angle 2u between k and k0 is 268
128 Introduction to the Physics and Chemistry of Materials
two-dimensional [2D] projection) when the radiation is normal to the (100) face of
the crystal. However rotating the crystal also rotates the reciprocal lattice about the (000)
origin, and all of the points for which jGhk‘j < 2jkj will eventually intersect the Ewald sphere.
If polychromatic radiation is being used, as in the Laue method, the Ewald circle represents
the short wavelength limit (highest energy photon). Now all planes whose representative
points lie inside the circle represent possible reflections and the symmetry of the reflections is
the symmetry of the reciprocal lattice relative to the incident radiation. Thus the Laue method
is useful for determining the orientation of the axes in a single crystal.
DEFINITION
The smallest polyhedron centered at the origin, (000) in k-space, that is enclosed by the perpen-
dicular bisectors of G to each nearest reciprocal lattice point is called the first Brillouin zone.
Similarly, the smallest polyhedron enclosed by the perpendicular bisectors of the next set of
reciprocal lattice vectors is the second Brillouin zone, etc.
This definition can be illustrated by constructing the first two Brillouin zones for a 2D
oblique lattice as shown in Figure 6.6.
A similar construction called a Wigner–Seitz cell can be formed in direct space by taking
perpendicular bisectors between each lattice point. The polyhedra formed will fill the 3D
space and contain the volume associated with that particular lattice point. From this
analogy, the Brillouin zones can be considered Wigner–Seitz cells in reciprocal space.
The 3D Brillouin zones for the bcc, fcc, and hexagonal close-packed (hcp) reciprocal lattices
are shown in Figure 6.7.
The G denotes the origin and the other letters indicate directions of high symmetry. We
encounter these designations later in mapping electron energy bands (although the choice
of letters may vary).
FIGURE 6.6
Construction of the first two Brillouin zones
in a 2D oblique reciprocal lattice by taking
perpendicular bisectors of the reciprocal lattice (20) (20)
vectors G drawn from the origin to each
lattice point. Since the Laue condition is met
everywhere along a perpendicular bisector,
the Brillouin zone surfaces can be thought of
as mirrors that reflect radiation traversing the
(22) (02) (22)
crystal for which 2(k . G) þ G2 ¼ 0.
Reciprocal Lattice and X-Ray Diffraction 129
kz
kz
P
L
Γ H
Γ ky
ky
X N
W K kx
kx
X at 2π/a (1,0,0)
H at 2π/a (010)
K at 2π/a (1,1,0) N at π/a (110)
L at π/a (1,1,1) P at π/a (111)
(b)
W at π/a (2,1,0)
(a)
kz
Γ M
K ky
kx
K at 4π/3a (1,0,0)
M at 2π/31/2 a (0,1,0)
(c) A at π/c (0,0,1)
FIGURE 6.7
Brillouin zones for fcc direct lattice (bcc reciprocal) (a), bcc direct lattice (fcc reciprocal) (b), and hexagonal direct
lattice (hexagonal reciprocal) (c). The letter G designates the origin and the other letters designate directions of high
symmetry. This notation will be seen later in the band structure of various materials.
r
FIGURE 6.8
rj
Geometry for analyzing scattering of x-rays from electron r
density located at r relative to the jth atom in the unit cell Tmnp
translated by Tmnp from the origin.
Ð
The electron density in the vicinity if the jth atom is given by nj (r) ¼ dVcj (r)cj *(r).
The electron concentration n(r) in the crystal may be written as
X
MNP s ð
X
n(r) ¼ dVcj (r)cj*(r): (6:23)
mnp j¼1
The scattering amplitude is proportional to the electron concentration times the appropri-
ate phase factor:
ð X
MNP s ð
X
dVh(r) expirDk ¼ dVcj (r)cj*(r) expirDk : (6:24)
mnp j¼1
Ð
where we define fj ¼ dVc(r)c*(r)e rG as the atomic form factor for atom j.
We can make a simple estimate of the atomic form factor by assuming a spherical
distribution of electrons with radius Rj surrounding the jth atom and considering the
phase shift from the scattering across the region.
ð
Rj
ðp ð
Rj
irG cos a sinðrGÞ
fj ¼ 2pnj r dr e
2
sin a da ¼ 4pnj r2 dr , (6:26)
rG
0 0 0
where nj is the electron density in the cloud surrounding the jth atom. We assume G
satisfies the Bragg equation (Equation 6.22) for a beam with wavelength l being scattered
at u and write Gr as
4pr
Gr ¼ sinðuÞ ¼ x sinðuÞ (6:27)
l
1.0
x=1
Atomic form factor F(q, x)
x=2
0.5
x=3
FIGURE 6.9
x=5 x=4 Angular part of the atomic form factor for vari-
x = 10 ous values of x ¼ 4p particle size=wavelength.
X-rays with wavelengths much shorter than the
0 particle size will be scattered in the forward
0 90 180 direction. As the wavelength increases, the scat-
2q tering becomes more omnidirectional.
where x0j ¼ 4pRj=l. Thus we see that the atomic form factor is just the total charge on the
jth atom times an angular scattering factor that depends on the ratio of atomic radius to
wavelength. Plots of this scattering function for different values of the radius to wave-
length ratio are shown in Figure 6.9. If the wavelength of the radiation is short compared
to the size of the atoms, the scattering will be more intense in the forward direction. As
the wavelength becomes comparable to the size of the atoms, the scattering becomes
more omnidirectional and backscattering can be observed.
Actual values for the atomic form factors for various atoms as a function of scattering
angle can be found in the International Tables for X-Ray Crystallography or from the online
NIST database http:==physics.nist.gov=PhysRefData=FFast=Text=cover.html.
If the Laue condition is met, G ¼ Dk in Equation 6.24 and the scattering amplitude will be
proportional to the sum of the contributions from each of the j atoms in all of the M3 unit
cells, or
X
MNP X
s X
MNP
E fj ei(rj þTmnp ) G ¼ SG ei Tmnp G ¼ SG MNP ¼ SG Nunit cells : (6:30)
mnp j¼1 mnp
Thus the scattered E-wave is directly proportional to the structure factor SG, which
contains details of the unit cell, and to the sum over all of the lattice phase
factors, which is just the number of unit cells. The scattered intensity Isc will be then be
proportional to
E20 Nunit
2 *
cells SG SG
Isc jEj2 ¼ EE* ¼ , (6:31)
R2
where R is the distance from the crystal to the observation point. Since the atomic form
factors in the structure factors are proportional to the Z of their atoms, the intensity of the
diffraction peak will be proportional to Z2. This fact makes if more difficult to use XRD
techniques with very low Z material.
132 Introduction to the Physics and Chemistry of Materials
Expanding rj G ¼ (xja þ yjb þ zjc) (hA þ kB þ ‘C) ¼ 2p(xjh þ yjk þ zj‘). The structure
factor becomes
X
s
SG ¼ fj e2pi(xj hþyj kþzj ‘) , (6:32)
j¼1
where xj þ yj þ zj are the positions of the jth atom in the unit cell. For a simple cubic
structure with a single atom at 0, 0, 0; Shk‘ ¼ f1 for all h, k, ‘ and all reflections permitted
by the Laue criterion will occur.
For a CsCl structure, which can be represented by a simple cubic lattice with a basis of
(0,0,0) and (1=2,1=2,1=2), the structure factor is
It can be seen that now Shk‘ ¼ 4 f1 if h, k, ‘ are either all odd or all even and Shk‘ ¼ 0 if they
are mixed. Thus the (100), (101), (210), (211), etc., lines will be absent.
The disappearance of the (100) line in the case of both the bcc and fcc structure can be
easily understood by the fact that in both cases, there are atoms on the (200) planes halfway
between the (100) planes. When the path difference between (100) planes is l, meeting the
Laue condition for (100) reflections, the path difference between the (200) planes is l=2 and
destructive interference results. Thus the reflections from the (200) planes cancel the
reflections from the (100) planes. When the path difference between (200) planes is l,
meeting the Laue condition for (200) reflections, the reflections from the (100) planes
(which technically are also (200) planes) reinforce the reflections from the (200)
planes and the intensities are enhanced.
Note that we could have represented the bcc and fcc structures by their primitive lattices
and since these primitive unit cells contain single atoms, there is no interference and all
h0 k0 ‘0 reflections would occur. Now if we take the primitive h0 k0 ‘0 indices for the fcc
reciprocal lattice shown in Figure 6.1 and map these primitive indices into the nonprimitive
fcc reciprocal lattice as shown in Figure 6.10a, we generate points at (200), (110), (220), etc.
in the first layer (‘ ¼ 0) and (101), (211), (301), etc. in the second layer (‘ ¼ 1) layer. These
points correspond to the allowed reflections of a bcc direct lattice in accordance with the
structure factor calculations (Equation 6.34), which requires hk‘ to be net even. Similarly,
one can see in Figure 6.10b that the primitive vectors in the bcc reciprocal lattice will map
out (200), (220), (222), (400), etc. in the even-numbered ‘ planes and (111), (311), (331), etc.
Reciprocal Lattice and X-Ray Diffraction 133
FIGURE 6.10
(a) First two layers of the reciprocal lattice points for an fcc reciprocal lattice generated by translating the primitive
reciprocal lattice vectors shown in Figure 6.1. These are the Laue indices of the allowed reflections for a bcc direct
lattice. Note that the net sum is even. (b) First two layers of reciprocal lattice points of a bcc reciprocal lattice
generated by translating the primitive bcc reciprocal lattice vectors. These are the Laue indices of the allowed
reflections for an fcc direct lattice. Note that the indices are all even or all odd.
in the odd-numbered ‘ planes, which are the allowed reflections in the fcc direct lattice in
accordance with the structure factor calculations (Equation 6.34), which requires that hk‘ be
all even or all odd.
I = 78.4%
I = 68.2%
Intensity
I = 24.1%
I = 22.0% I = 21.3% I = 20.9%
I = 19.3%
10 20 30 40 50 60 70
2q
FIGURE 6.11
XRD pattern for MgF taken with CuK-a radiation. (Photograph courtesy of M. Banish, University of Alabama,
Huntsville.)
When equipped with a hot stage, the powder method is also useful for observing solid-
state phase changes or peritectic reactions in which a solid reacts with a liquid to form a
new solid.
FIGURE 6.12
Back reflection Laue photograph along the [111] axis
of GaAs. The spots showing sixfold symmetry are
from 6 of the 24 {642} planes that lie in the Ewald
sphere. The spots showing threefold symmetry are
from 4 of the 12 {422} planes that lie in the Ewald
sphere. (Picture courtesy of D. Gillies, private com-
munication.)
Strain in the crystal would broaden the width of the rocking curve and twist, tilt, or small
angle grain boundaries would show up as multiple peaks as illustrated in Figure 6.13b.
Reflection amplitude
FIGURE 6.13
(a) Simulated rocking curves of an ideal crystal and (b) crystal with a small angle grain boundary.
136 Introduction to the Physics and Chemistry of Materials
will cause every spot within 2k of the origin of the Ewald diagram to cross the Ewald
sphere, thus producing rows of spots that can be indexed and measured.
when the Laue condition G ¼ Dk is satisfied. Also, recall that the complex Fourier trans-
form that transforms a time varying signal f(t) to the frequency domain is given by
ð
1
1
F(v) ¼ pffiffiffiffiffiffi dtf (t)eivt : (6:37)
2p
1
It is easy to see by analogy that the integral over the electron density function times the
phase shift is just a complex 3D Fourier transform that transforms a periodic function in
direct lattice space to reciprocal lattice space. Therefore, one may think of the diffraction
pattern as a Fourier transform of the direct lattice. In principal, one should then be able to
measure the intensities and locations of the reflections in the diffraction pattern, take the
inverse Fourier transform, and recover the electron density function. The inverse transform
is obtained by
1 XXX
n(x,y,z) ¼ SG(h,k,‘) e2pi(hxþkyþ‘z) : (6:38)
V h k l
! !
X X
I(h,k,‘) SG SG* ¼ fj e2pi(xj hþyj kþzj ‘) fj e2pi(xj hþyj kþzj ‘) , (6:39)
j j
assuming the number of measured I(h,k,‘) 3n. (This is why a large number of spots
must be measured.) There are algorithms for solving a set of nonlinear overdetermined
equations, but they are time-consuming and the above approach is feasible only for small
molecule crystals where n < 100 atoms per unit cell.
6.5 Summary
Much of the wave nature of solids is carried out in reciprocal space because the momentum
vector of photons, phonons, and particles such as electrons and neutrons can be expressed
directly in terms of their vector k as jkj ¼ 2p=l. The reciprocal lattice vectors A, B, and C are
constructed such that they are perpendicular to the direct lattice vectors a, b, and c and
their lengths are proportional to the inverse length of the direct lattice vectors. The lattice
points in reciprocal space correspond to planes in direct space and a reciprocal transla-
tional vector is defined as Ghk‘ ¼ hA þ kB þ ‘C, where hk‘ are the Miller indices of these
planes. It is shown that Ghk‘ is perpendicular to the (hk‘) plane and the distance of this
plane from the origin is given by dhk‘ ¼ 2p=jGhk‘j.
138 Introduction to the Physics and Chemistry of Materials
TABLE 6.1
Applications of XRD
Method Radiation Sample Information Obtained
The Laue condition for diffraction (including x-ray, electron, and neutron diffraction) can
simply be stated as Ghk‘ ¼ Dk, where Dk is the vector difference between the incident
and scattered wave vector. For elastic scattering, the Laue condition can also be written as
2k G ¼ G2 which is shown to be equivalent to the more familiar Bragg equation,
2dhk‘ sin u ¼ l.
The Laue condition is satisfied everywhere on a perpendicular bisector to Ghk‘. Thus
planes that are perpendicular bisectors of the various Ghk‘ vectors act as diffraction mirrors
for waves or particles that meet the Laue condition. The Brillouin zones, polyhedra
bounded by these perpendicular bisecting planes, will play an important role in the
electron theory of metals as will be shown in later chapters.
Satisfying the Laue (or Bragg) condition is a necessary but not sufficient condition for a
diffraction spot. The presence or absence of the hk‘ diffraction peak, as well as its intensity, is
P
determined by the structure factor Shk‘ ¼ sj¼1 fi e2pi(xj hþyj kþzj ‘) where xj, yj, zj are the coord-
inates of the jth atom in the unit cell, fj is the atomic form factor for the jth atom and is
proportional to the number of electrons surrounding the atom, and the sum is taken over
all of the atoms in the unit cell. Therefore, the reflected intensity is proportional to the
square of the number of unit cells in the beam and to the square of the atomic numbers of
the atoms in the cells.
For primitive systems with only one atom per unit cell, all peaks are visible. But for
systems with multiple atoms per unit cell, many of the peaks cancel out. For a bcc
structure, the sum of hk‘ must be even to produce a diffraction peak. For an fcc structure,
hk‘ must either be all odd or all even to produce a diffraction peak.
There are a number of different ways in which XRD is used to study materials.
A summary of these methods is given in Table 6.1.
Bibliography
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia, 1976.
Christman, J.R., Fundamentals of Solid State Physics, John Wiley and Sons, Inc., New York, 1988.
Ibach, H. and Lüth, H., Solid State Physics, 3rd edn., Springer-Verlag, New York, 1990.
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley and Sons, Inc., N.Y., 1966.
McPherson, A., Crystallization of Biomolecular Molecules, Cold Spring Harbor Laboratory Press,
MA, 1999.
Reciprocal Lattice and X-Ray Diffraction 139
Problems
1. Show that the h, k, ‘ components of the G vector are the Miller indices.
2. Show that the spacing between lattice planes is given by dhk‘ ¼ 2p=jGhk‘j for any lattice.
3. Show that the Bragg condition is met everywhere on a perpendicular bisector of any
reciprocal lattice vector G.
4. Show that the Bragg condition is met by any reciprocal lattice point that falls on the
Ewald sphere.
5. Diamond structure is fcc with every other tetrahedral interstitial site occupied. Write the
structure factor. Write the Laue indices for the first five diffraction peaks.
6. GaAs has the zinc blende structure, which is the diamond structure with one type of
atom on the fcc sites and the other on the interstitial sites. Write the structure factor and
the Laue indices for the first five diffraction peaks.
7. Find the reciprocal lattice vectors for a hexagonal lattice and verify the size of the
hexagonal Brillouin zone shown in Figure 6.7c.
pffiffiffi
8. Basis for the hcp system is (0,0,0) and a=2, a=2 3, c=2 . Find the structure factor and
identify a criterion for determining whether or not a peak would be visible.
9. Assume the nearest neighbor distances for the fcc and hcp form of a given material are
the same. Find the distinguishing features of their XRD patterns that would identify
which form you have.
7
Theory of Elasticity
141
142 Introduction to the Physics and Chemistry of Materials
Zz
Yz
Zy
Xz
Zx
Yy
Yx
Δz Xy y
Xx Δx
FIGURE 7.1 Δy
Stresses acting on the faces on a stationary elastic
cube anchored at the origin. x
where each subscript runs from 1 to 3, giving the coefficient matrix 81 components.
However, since Sij and ek‘ each have only six unique components, it is more convenient
to represent them as vectors with six components and write
(∂x/∂y)Dx
qx
Dy
0 1 0 10 1
Tx C11 : : : : C16 exx
B Ty C B : : CB eyy C
B C B CB C
B Tz C B : : CB ezz C
B C¼B CB C , (7:4)
B Sx C B : : CB eyz C
B C B CB C
@ Sy A @ : : A@ ezx A
Sz C61 : : : : C66 exy
where the Cij are the modulus of elasticity or stiffness coefficients. It is easy to show that the
Cij elements are symmetric by the following argument. Consider an increment of work dU
where
Since @U=@exx ¼ Tx and @U=@eyy ¼ Ty , @Tx =@eyy ¼ @ 2 U=@exx @eyy ¼ @Ty =@exx . Since Tx ¼
C11exx þ C12eyy þ and Ty ¼ C21 exx þ C22 eyy þ , (@Tx =@eyy ) ¼ C12 and (@Ty =@exx ) ¼ C21 ;
hence, C12 ¼ C21. Similarly it can be shown that Cij ¼ Cji for all i and j. Since the coefficient
matrix is symmetric, the number of unique components is reduced to 21.
The above results are general and apply to any coordinate system—not necessarily an
orthogonal system. However, if we have cubic symmetry, C11 ¼ C22 ¼ C33, C44 ¼ C55 ¼ C66,
C12 ¼ C13 ¼ C23 ¼ C21 ¼ C31 ¼ C32. Since there are mirror planes perpendicular to each of the
axes, shears in opposite directions must have opposite signs. But we have already shown
that all of the matrix elements are symmetric; hence, the off-diagonal terms must vanish.
Now the coefficient matrix can be written as
0 1 0 10 1
Tx C11 C12 C12 0 0 0 exx
B Ty C B C12 C11 C12 0 0 0 C B eyy C
B C B CB C
B Tz C B C12 0 C B C
B C¼B C12 C11 0 0 CB ezz C (7:5)
B Sx C B 0 0 C B C
B C B 0 0 C44 0 CB eyz C
@ Sy A @ 0 0 0 0 C44 0 A@ ezx A
Sz 0 0 0 0 0 C44 exy
and only three elastic moduli or stiffness coefficients are needed to characterize the
response of the material.
Using the last two equations to eliminate eyy and ezz in the first equation, we can write
2C212
E ¼ C11 : (7:7)
C11 þ C12
C12
n¼ : (7:8)
C11 þ C12
The velocity of a longitudinal wave moving in the [110] direction can be shown to be
given by
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
vL [110] ¼ (C11 þ C12 þ 2C44 )=2r: (7:11)
If the material is isotropic, the velocity in the [100] direction will be the same as in the [110]
direction; hence
This is not true for single crystals because they are generally anisotropic, i.e., the velocity of
propagation is not the same in the [100] as in the [110] directions. However, polycrystalline
materials and glasses are isotropic because of the random nature (or absence) of their grain
structure. Therefore, the above relationship holds for bulk polycrystalline solids. This
allows a number of relationships between B, E, G, and n to be derived. Using the relation
between the Poisson ratio and C11 and C12 to eliminate C12, the bulk modulus may be
related to Young’s modulus by
Using Equation 7.12 to relate C44 to C11 and C12 for isotropic materials, we can write
E ¼ 2G(1 þ n) (7:14)
and
3B 1 2n
G¼ : (7:15)
2 1þn
where
x0 ¼ r 0
A ¼ br20
From Equation 7.19 we see that C11 ¼ cB=b and from Equation 7.6
1 C11 b
C12 ¼ ð3B C11 Þ ¼ 3 1 (7:20)
2 2 c
C12 3b c
n¼ ¼ : (7:21)
C11 þ C12 3b þ c
For isotropic media, the shear modulus C44 may be estimated from
C11 C12 3C11 b
C44 ¼ ¼ 1 : (7:22)
2 4 c
0
U/U0
FIGURE 7.3
Equilibrium position of an atom in the bond-
energy diagram at 3kT=U0 (dashed line). The
−0.5 asymmetry of the bond energy causes this equi-
3kT/U 0 librium position to shift to larger values as the
temperature increases, which is responsible for
thermal expansion dr=r0. Note that this shift is
−1.0 highly exaggerated in this drawing. Recall that
0 2 4 6 8 10
the melting point occurs at only 4% of the
dr/r0 Relative distance between atoms bond energy for most materials.
where a is the coefficient of linear thermal expansion (CTE) which is O (106 K1).
A similar expression can be written for volume expansion and it is easy to show that since
a is small, the coefficient of volumetric expansion is 3a. Thermal expansion occurs because of
the asymmetry in the bond-energy function. This asymmetry produces anharmonicity in the
oscillation of the atoms in the lattice; i.e., instead of oscillating with equal amplitude on both
sides of their equilibrium position, they can swing farther to the soft side of the bond-energy
function than toward the steep side. As a result, the average positions of the atoms are
displaced from the equilibrium position at the 0 K location on the bottom of the bond-energy
curve as seen in Figure 7.3. The average displacement of the lattice atoms can be found from
Ð
1
drebUðdrÞ ddr
0
hdri ¼ 1
Ð , (7:24)
ebUðdrÞ ddr
0
where U(dr) is the bond energy as a function of the displacement dr from the equilibrium
position and b ¼ 1=kT.
The bond-energy curve can be approximated by U(dr) ¼ Cdr2 Ddr3 þ in which C
represents the harmonic part of the potential and D is the anharmonic part. It is assumed
that bDx3 1 so that exp(bDx3) can be expanded as 1 þ bDx3 þ . Keeping only terms
through second order, the integrals can then be written as
Ð 2
1
ebCdr dr þ bDdr4 þ ddr
1 3DkT
hdri ¼ 1
Ð ¼ : (7:25)
4C2
ebCdr2 ð1 þ bDdr3 þ Þddr
1
U ðr0 Þ h r0 m r n i
0 U 0 h r0 m r n i
0
U(r) ¼ n m ¼ n m , (7:28)
nm r r nm r r
where U0 is a positive quantity. Differentiating Equation 7.28
U0 mn hr0 m r0 n i
U0 ðrÞ ¼ , (7:29)
ðn mÞr0 r r
U0 mn h r m r n i
U 00 ðrÞ ¼
0 0
(m 1) þ(n 1) , (7:30)
ðn mÞr0
2 r r
from which
U0 mn U0 mn
U000 ¼ and U0000 ¼ ðn þ m þ 3Þ: (7:31)
r20 r30
U0 h r0 r0 n i
U(r) ¼ n : (7:32)
n1 r r
For the NaCl structure, c ¼ 8, and b ¼ 4. From Equations 7.17 and 7.31 we have
c
C11 ¼ B ¼ 2B, (7:34)
b
3b c 1
n¼ ¼ , (7:36)
3b þ c 5
TABLE 7.1
Elastic Properties of Covalent Compounds
B Calculated B Measured C11 Calculated C11 Measured
U0=V (GJ=m3) n Values (GPa) Values (GPa) Values (GPa) Values (GPa)
Gilman points out that the shear modulus cannot be estimated from a simple nearest
neighbor central force potential. He represented the cross section of a rod using nine atoms
connected by springs of one stiffness for nearest neighbors and springs of a different
stiffness for second nearest neighbor. His conclusion was that the shear modulus for
ionic compounds could be approximated as C44 ¼ G ¼ 3=5B. Using Equations 7.34 and
7.37, we estimated G ¼ 3=4B for ionic compounds. However, we did not just use a nearest
neighbor central force potential.
We now check our predicted elastic constants against their measured values in Table 7.1.
There is reasonably fair agreement between the predicted and measured values of B and
C11. We will now examine the other predictions in Table 7.2.
The values for B calculated from Equation 7.6 may be slightly different from the
measured values quoted in Table 7.1, but were used to be consistent with other values
that could be computed from the elastic constants in this table.
We see from Table 7.2 that the predicted values for C12=C11 ¼ 0.25, B=C11 ¼ 0.50, and
n ¼ 0.20 are somewhat lower than the mean observed values, but are well within the
standard deviations of the observed values.
TABLE 7.2
Predicted Elastic Properties of Covalent Compounds
C11 Measured C12 Measured B Equation n Equation
Values (GPa) Values (GPa) 7.6 (GPa) C12=C11 B=C11 7.8
TABLE 7.3
Estimated Thermal Expansion fo Covalent Compounds
Material U (kJ mol1) N aCal (K1) aMeas (K1)
Recall in Figure 3.3 we saw that the specific energy (U=V) exhibited approximately a
fourth power relationship with the nearest neighbor distance. Since the bulk modulus and
the elastic coefficients are directly related to (U=V), one would expect a similar fourth
power dependence of these quantities on the nearest neighbor distance.
We now investigate the ability to estimate thermal expansion from Equations 7.23
and 7.28.
The results are shown in Table 7.3 for a few selected compounds. The agreement is only fair.
1000
Bulk modulus (GPa)
100
10
1 FIGURE 7.4
1027 1028 1029 1030
Bulk modulus versus electron density for simple
Electron density (m−3) metals. The slope of the line is 1.22.
as was the case for the simple metals. Instead, as seen in Figure 7.5, they more closely
follow the same behavior as their specific energies shown in Figure 3.15.
The elastic properties of selected transition metals are tabulated in Table 7.4.
Note that the ratio of B=U0 tends to increase as one moves down and to the right of the
periodic table.
The M-shell transition metals have magnetic interactions that weaken their bonds
because some of the d-electrons are locked in spin states, which prevents them from
forming covalent bonds, (discussed in Chapter 3). The bulk moduli of the N- and O-shell
elements show a very strong (inverse eighth power dependence) on their atomic radii
(Figure 7.6). Recall from Figure 3.16 that there is a very small difference between the atomic
radii of the N- and O-shell transition metals, which would account for the strong depend-
ence on their radii.
5d
4
Bulk modulus (1011 Pa)
4d
3
2 3d
0 FIGURE 7.5
0 5 10 15 20
bulk modulus of the 3d, 4d, and 5d transition
Number of electrons in outer shell metals.
152 Introduction to the Physics and Chemistry of Materials
TABLE 7.4
Elastic Properties of Transitions Metals
Element Atomic Number Group U0 (GJ=m3) B (Gpa) C11 (Gpa) B=C11 B=U0
1000
Bulk modulus (GPa)
100
FIGURE 7.6
Bulk moduli of the N- and O-shell transition metals 10
as a function of their atomic radii. Note the approxi- 0.1 1
R0 (nm)
mate eighth power dependence on their radii.
The second derivative has a root at xmax ¼ x0 þ j ln 2, so the maximum stress will be
given by
U 0 ðxmax Þ U0 cU0 =V
smax ¼ ¼ ¼ : (7:40)
bx20 2bx30 ðj=x0 Þ 2bðj=x0 Þ
If there are no defects in the crystal, the elastic bonds theoretically will stretch until x
reaches xmax. Beyond this point the system energy is lowered by increasing the separation.
In other words, if the stress is released at x < xmax, the system will recover elastically, but if
strained beyond xmax the bonds will be broken and the system will separate into two parts.
Since from Equation 7.19
the maximum stress can be obtained in terms of C11 by substituting Equation 7.41 into
Equation 7.40 to give
9C11 ðj=r0 Þ
smax ¼ : (7:42)
4
The j=x0 must be found from Equation 7.41 using the measured values of C11 and U0=V.
Putting this into Equation 7.42 yields
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
9cC11 U0 =V
smax ¼ : (7:43)
8b
pffiffiffi
For Al, U0=V ¼ 32 GJ=m3, C11 ¼ 114.3 MPa, and c=b is 2= 2 for fcc. Using these values, the
maximum cohesive strength for Al is predicted to be 76.3 GPa. The tensile strength of pure
annealed 1100-0 Al is only 90 MPa, almost three orders of magnitude lower than the
theoretical cohesive strength.
Often the Young’s modulus is used instead for C11 and the theoretical strength for brittle
fracture is generally estimated as E=10 (see Section 9.5). For Al, the Young’s modulus is 69
GPa, which gives an estimated maximum brittle strength of 6.9 GPa. The tensile strength of
7075 T-6 Al alloy is 572 MPa which is more than an order of magnitude lower than the
theoretical maximum.
Sz
R0
qxy
x
(a) (b)
FIGURE 7.7
(a) Geometry of a crystal of thickness d sheared by distance x. (b) Illustration of the barrier in moving planes of
crystal over one another.
will occur at x ¼ R0=2 as the planes of atoms are at the top of each other. Therefore, the
theoretical maximum critically resolved shear force is given by
R0 G R0
scrss ¼ NG ¼ : (7:44)
2d 2 d
pffiffiffi pffiffiffi
For fcc, d ¼ 2R0 and for bcc, d ¼ 2= 3R0 . As a general rule the theoretical shear strength
is estimated to be G=4.
The shear modulus for Al is 40 GPa which translates into an estimated theoretical shear
strength of 10 GPa. The actual yield strength of 7075 T-6 Al is 500 MPa, about 1=20 of its
estimated theoretical value. For pure Al (1100-0 annealed) the yield strength is only
34 MPa.
It should be understood that the elastic properties we have been discussing are a
function only of the bonds and are not affected by any of the strengthening mechanisms
(except for alloying) that will be discussed in the later chapters. The elastic properties of
alloys will generally follow the rule of mixtures (see Section 10.3.1).
7.6 Summary
The stress and strain tensors can be written as six-component vector (three tensile and three
shear) that are related by an elastic coefficient matrix, which can be reduced from 36 to 21
components by symmetry arguments. For systems with cubic symmetry, the elastic coef-
ficient matrix can be further reduced to only three elastic constants, C11, C12, and C44. These
constants can be determined by measuring the velocity of sound in the medium in different
directions in a single crystal. Once these elastic constants are known, other properties such
as the bulk modulus B, the Young’s modulus E, Poisson’s ratio n, and the shear modulus G
may be derived, i.e.,
1 2C212 C12
B ¼ ðC11 þ 2C12 Þ, E ¼ C11 , n¼ , and G ¼ C44
3 C11 þ C12 C11 þ C12
Using similar relations, we then estimated the theoretical or maximum tensile and yield
strength of a material from its bond strength and elastic coefficients. For Al, the estimated
theoretical yield and tensile strengths proved to be more than three orders of magnitude
greater that the strengths of pure Al in the annealed state. Strengthening mechanisms (to be
discussed in subsequent chapters) increase the actual strengths by about an order of
magnitude, but they are still an order of magnitude short of the theoretical values for
brittle fracture. So we have much room for improvement in developing stronger materials.
Bibliography
Ashcroft, N. and Mermin, N.D., Solid State Physics, Brooks Cole, Belmont, MA, 1976.
Barsoum, M.W., Fundamentals of Ceramics, Institute of Physic Publishing, Bristol and Philadelphia,
2003.
Gersten, J.I. and Smith, F.W., The Physics and Chemistry of Materials, John Wiley & Sons, New York,
2001.
Gilman, J.J., The Electronic Basis of the Strength of Materials, Cambridge University Press, UK, 2003.
Hosford, W.F., Physical Metallurgy, Taylor & Francis, CRC Press, Boca Raton, FL, 2005.
Kingery, W.D., Bowen, H.K., and Uhlmann, D.R., Introduction to Ceramics, 2nd ed. John Wiley & Sons,
New York, 1976.
Kittel, C., Introduction to Solid State Physics, 7th ed., John Wiley & Sons, New York, 1966.
McMillan, N.H., Review: The theoretical strength of solids, J. Mat. Sci. 7, 1972, 239–254.
Russell, A.M. and Lee, K.L., Structure–Property Relations in Nonferrous Metals, John Wiley & Sons,
New York, 1966.
Theoretical Strength of Materials, Prepared by the Materials Advisory Board, Publication MAB-221-M,
National Academy of Sciences–National Research Council, Aug. 1966 (available online).
Problems
1. Show that for an isotropic solid the Young’s modulus E is related to the shear modulus G
and the Poisson ratio n by E ¼ 2G(1 þ n). Also show that E ¼ 9 BG=(G þ 3B).
2. If the density of a material remained constant while it was stretched elastically, what
would the Poison ratio have to be? What would the bulk modulus have to be? Is it
realistic to assume that the density remains constant while a material is deformed
elastically?
3. What would be the theoretical maximum elongation of Al? What would be the stored
energy?
4. The Materials Advisory Board (Aug. 1966) introduced a simple bond-energy function,
U(r) ¼ U0(1 þ r=D) exp(r=D) where r is the displacement from equilibrium and D
is an adjustable parameter. Use this function to compute the theoretical maximum
cohesive strength at rupture, and the strain at rupture. How does this result compare
with the predictions using the Morse potential.
8
Defects in Crystals
As we shall see, nature does not allow a crystal to be perfect, but even if it did, the
performance of a perfect crystal would not necessarily be improved. In fact the defect
structure very much determines the mechanical, electrical, thermal, optical, and magnetic
properties of a material and our ability to understand the role of defects and to be able to
control their formation is key to the development of useful materials.
157
158 Introduction to the Physics and Chemistry of Materials
the Gibbs energy (or free energy) G is minimized—not the internal energy U. The Gibbs
energy is G ¼ U þ pV TS. (Note: For solids at atmospheric pressure, the pV contribution is
negligible and the Helmholtz energy F ¼ U TS is often used in place of G.)
Minimizing the free energy requires dF ¼ dU TdS ¼ 0. If there are Nv vacancies, the
change in internal energy is just Nv Q, where Q is the activation energy per atom to form a
vacancy. Now we have to find a way to determine dS.
Despite what you may have learned about entropy in your undergraduate thermo-
dynamics, there is a simple definition of entropy that you should always remember:
S ¼ k ln W, (8:1)
where
k is Boltzmann’s constant
W is the number of accessible states the system can be in for a given temperature,
pressure, and volume.
Clearly, if all of the sites were occupied with atoms, there would be only one accessible
state. However, if there were Nv vacancies and if the energy U did not depend upon which
sites were vacant and which were occupied, there would be many more accessible states.
The S would increase until the T dS term is balanced by the increase in U.
Now let us quantify the number of accessible states. If we have N sites and Nv vacancies,
just as there are 52! ways of arranging a deck of cards, there are N! ways of arranging
N sites with Nv vacancies. However, since atoms are indistinguishable, rearranging the
N Nv atoms among the sites they occupy does not constitute different states. Thus we
must divide the N! ways of arranging the N sites by (N Nv)! to eliminate the redundancies
from the fact that atoms are indistinguishable. Similarly, since interchanging one vacant
site for another does not constitute a different state, we must also divide N! by Nv! to
remove this redundancy. Thus the number of accessible states is given by
N!
W¼ : (8:2)
(N Nv )!Nv !
As a sanity check, one can see that if all the sites were occupied, W ¼ 1 (remember that
0! ¼ 1). If there were only 1 vacancy, there would be N places that vacancy could occur.
If there are 2 vacancies, W ¼ N(N 1)=2 etc.
However, we must take the natural log of W. Since for macroscale systems both the
number of sites and the number of vacancies are very large, we can employ Stirling’s
approximation ln x! ¼ x ln x x, which holds for large x. Thus
ln W ¼ N ln N (N Nv ) ln(N Nv ) Nv ln Nv : (8:3)
or
Nv Nv
¼ ¼ eQ=kT : (8:6)
N Nv Natoms
This is just Boltzmann’s law, which says that the ratio of atoms (in this case, vacancies)
occupying a state with energy Q to those in the ground state at temperature T is given by
the Boltzmann factor, exp(Q=kT). This type of behavior is also said to follow the Arrhe-
nius law. If one plots the log(N=Nv) vs. 1=T, a straight line should result and the activation
energy can be determined from its slope. Conversely, any property whose log plots a
straight line against 1=T is said to have Arrhenius-like behavior, meaning that there is an
activation energy involved. Diffusion and viscosity coefficients are two other materials
properties that exhibit Arrhenius-like behavior.
Atoms or ions leaving a lattice site to form a vacancy can either jump into an interstitial
site or move to the surface of the solid. Other atoms or ions may move into the site they left
behind; thus vacancies may meander or diffuse through the solid, a process known as
vacancy diffusion. In metals, single-atom vacancies are possible, but in ionically bonded
compounds, vacancies must occur in pairs in order to maintain charge balance.
Schottky
defect
Frenkel
defect
FIGURE 8.1
Illustration of Frenkel and Schottky
defects in an ionic crystal. Larger light
gray circles represent anions and smal-
ler darker gray circles represent cations.
160 Introduction to the Physics and Chemistry of Materials
Applied
tensile stress Applied
tensile stress
f
l
Slip
plane Slip
direction
When the resolved shear stress exceeds the critical resolved shear stress tcrss (a material
property), slip will occur as shown in Figure 8.2. The yield stress can then be written as
t crss
t yield ¼ , (8:8)
cos f cos l
which is known as Schmid’s law. The cos f cos l is called the Schmid factor. The maximum
Schmidt factor is 0.5 and occurs when f ¼ l ¼ 458. The slip system having the largest Schmid
factor will be the first to yield. It should be noted that if the Schmidt factor is zero, the yield
strength is infinite. In a polycrystalline system, there will always be grains with a nonzero
Schmidt factor; they will slip, and the materials will yield through a combination of this slip
combined with grain boundary slip. However, it is possible to orient a single crystal so that
there is no resolved stress in the direction of the applied tensile stress. This is the reason for
making single crystal turbine blades for high performance jet engines.
b b
A B A B A B
C D C D C D
FIGURE 8.3
Mechanism for dislocation movement. Under a shear stress, bonds between B and C are broken and reform
between C and A. This moves the dislocation line to between C and D. The process continues with bonds
reforming between D and B until the dislocation moves through the edge of the solid causing a displacement of
one lattice vector.
kink at one end and moving the kink across the carpet rather than trying to move the entire
carpet at once.
The line along the core where the extra half plane of atoms terminates is called a
dislocation. If this line emerges perpendicularly to the face of the crystal, it is called an
edge dislocation. The symbol ? denotes an edge dislocation with the vertical part in the
direction of the extra half plane. The mechanism by which crystals might deform by the
motion of edge dislocations is shown in Figure 8.3.
t̂
FIGURE 8.4
Schematic showing a screw dislocation on the face of a
F crystal.
t̂ b
b
FIGURE 8.5
An example of a mixed dislocation. The disloca-
tion line enters the crystal as an edge dislocation
on one face, curves through the crystal, and
emerges as a screw dislocation. The magnitude
and direction of the Burgers vector is conserved
F throughout.
164 Introduction to the Physics and Chemistry of Materials
FIGURE 8.6
Etch pits in Ge single crystals grown with and without wall contact. Notice the 100-fold decrease in dislocation
density that results from eliminating wall contact during the growth. (From Schweizer, M., et al. J. Crystal Growth,
235, 161, 2002. With permission.)
Defects in Crystals 165
B
FIGURE 8.7
A Frank–Reed dislocation generating mechanism. As
stress is applied, a dislocation line (A), becomes
pinned between two points, moves progressively
upward (B, C, D) until it folds back on itself and
pinches off (E) to form a dislocation loop. This loop
continues to expand while the process starts over with
a new line (A).
A
B
C
A
C
B
(a) (b)
FIGURE 8.8
Coalescence of vacancies into dislocation loop. Vacancies coalesce in the plane C to form the empty region in the
diagram to the right (b). This produces the half plane vacancy in the region bounded by the dislocation loop
shown on the left (a).
growing crystal since they can satisfy more bonds. As they attach, the spiral keeps
advancing, always providing additional low energy attachment sites for new atoms or
molecules to be incorporated. Superstrength single crystal whiskers may be grown from
the vapor in this manner. Since the screw axis runs along the length of the crystal, the
Schmid factor is zero along this axis and such whiskers can achieve near theoretical
strength.
interfacial energy and express it as J=m2. In other words, this is the energy it costs to make
new surface, which is the same as the surface excess energy. Surface tension and interfacial
energy are numerically as well as dimensionally the same. Interfacial energies for metals
are on the order of 1–10 J=m2.
Some crystals can be cleaved along planes that are parallel to faces of their growth habit.
These are generally planes of minimum interfacial energy since the growing crystal likes to
surround itself with surfaces that cost the least amount of energy in order to lower the total
free energy. Another way of saying this is that the growth atoms tend to attach to higher
energy surfaces where the force of attraction is the greatest. It is often said that you never
see the fastest growing surface of a crystal because these surfaces grow themselves out,
leaving the slower growing surfaces exposed, which are generally the planes with the
highest atomic density.
The basal planes of mica or annealed pyrolytic graphite can easily be separated with a
razor blade. Ordinary table salt has a cubic habit and can be cleaved easily along the {100}
faces by tapping a suitably aligned razor blade. However, attempts to cleave along a {110}
will cause the crystal to shatter. The {100} faces have the highest area density which means
the ions are closer together, hence their bonds are stronger and the excess surface energy
will be minimized. The {110} unit cell faces have the same number of atoms as the {100}
faces but have approximately 40% more area, so they have a much higher surface energy
(see Problem 3.5). The {111} faces also have a high area density, but they are not electrically
neutral. All the ions on these faces have the same charge; therefore, the face would have an
extremely high surface excess energy.
Since Aplane is smaller than A, ion pairs at the interface will have a higher energy (not as
deep in the energy well) as those in the interior. This excess energy is the interfacial energy.
Using Equation 3.7, the surface excess energy is given by
Aabove
A Aplane
Abelow
FIGURE 8.9
Schematic for estimating the surface energy of an ionic solid.
Defects in Crystals 167
Ns e2 n 1 A Aplane e2 n1
g ¼ (Ainterface A) ¼ , (8:11)
4p«0 r0 n 2 8p«0 r30 n
where the surface density Ns ¼ 2 ion pairs per 4r02. The nearest neighbor distance, for NaCl,
r0 ¼ 0.282 nm and n was found to be 8.98 (from Table 7.1). Using these values, the
interfacial energy of the [100] face is found to be 0.306 J=m2. This is close to the observed
value of 0.32 cited in Barsoum (2003, p. 103).
FIGURE 8.10
Grain boundaries of a polished and etched steel nail at
60. Note the elongated grains that result from the
drawing process. (Photograph courtesy of Dr. Michael
Banish, University of Alabama in Huntsville. Personal
communication.)
168 Introduction to the Physics and Chemistry of Materials
FIGURE 8.11
Tilt boundaries. Formation of a low-angle tilt boundary
through a series of dislocations.
8.4.6 Twinning
A twin is formed when a crystal grows in a mirror image of itself about a plane which is
not a normal mirror plane. In the simple cubic and body-centered cubic systems, the {111}
planes are possible twin planes (they are not mirror planes). Similarly, in the face-centered
cubic system, ABCABCBACBA, the middle C (also a {111} plane) is a twin plane. In the
hexagonal close-packed system, ABABCBABA, C would be a twin plane. Twining is
caused by stress during cooling and complex vacancy and dislocation mechanisms. The
energy involved in forming twins is about the same as in producing stacking faults. In
space crystal growth experiments in which crystals were grown in the absence of hydro-
static pressure and wall contact, both the dislocation density and number of twins were
dramatically reduced.
8.6 Diffusion
Atoms can move through solids by diffusion just as they are able to do in liquids and gases,
except the process is much slower. Mechanisms by which atoms are able to move through
solids include vacancy exchange diffusion, in which atoms fill vacancies creating new
vacancies; interstitial diffusion, in which atoms can move through the interstitial sites;
grain boundary diffusion, in which atoms can move along grain boundaries; and pipe
diffusion, in which atoms move along dislocations.
Diffusion can be considered to be a random process in which an impurity atom is just as
likely to jump one way as the other. However, think of two herds of sheep, a small herd
and a larger herd separated by a fence. The sheep start jumping over the fence. The sheep
in either herd are just as likely to jump the fence, but more sheep from the larger herd will
actually jump the fence, simply because there are more of them. So the diffusion of sheep
will be driven by the concentration gradient of the herd. Eventually, there will be the same
number of sheep on each side of the fence and they will be jumping back and forth at the
same rate. We would say an equilibrium has been reached.
So it is in a solid. If there are more atoms of one variety in one part of the solid, they will
tend to spread out until a uniform concentration is reached throughout the solid by the
mechanism of solid-state diffusion; however, this process can be extremely slow, especially
at ambient temperatures.
There are no general rules governing diffusion coefficients; therefore, they must be deter-
mined experimentally. Each of the diffusion processes has its own activation energy, which
depends not only on the process involved but also on the host and the diffusing material.
At higher temperatures, vacancy diffusion is generally prevalent since it has the highest
activation energy. Since the activation energy for vacancy diffusion requires the breaking
of bonds, one would expect metals with higher melting points to have higher activation
energies, thus lower diffusion coefficients. Grain boundary and pipe diffusion have lower
activation energies and become more important at lower temperatures. For very small
molecules, interstitial diffusion will generally dominate. Naturally one would expect a
small atom such as carbon to diffuse through metals faster than other atoms with sizes
comparable to their host. Gases diffuse through metals much faster than atoms of other
metals, which is why vacuum chambers are constructed from thick walls of stainless steel.
He diffuses faster than H2 because of its smaller molecular size. Values of the D0 and the
activation energy for different diffusion couples are found in various handbooks
(see bibliography).
create vacancies and even voids in the region being depleted by the faster diffusing
component. This effect has been implicated in the failure of solder joints and Al–Au
interconnects as they age, which is sometimes referred to as the purple plague. However,
researchers at the Max Planck Institute of Microstructure Physics have exploited the
Kirkendall effect to make hollow spinel (ZnAl3O4) nanotubes by coating ZnO nanowires
with Al2O3 and letting the Zn diffuse into the Al2O3 coating (Nature Materials, Aug. 2006).
Because the diffusion rates of two materials may be different, it is necessary to define an
e BA ¼ xDAB þ (1 x)DBA, where x is the mole fraction of component B
effective diffusivity D
in A, DAB is the diffusion coefficient of A into B, and DBA is the diffusion of B into A. This
distinction is only important for nondilute alloy systems. For dilute systems, x is small and
e BA ! DBA.
D
Fick’s first law is useful for solving simple problems in which the concentration is known
and the instantaneous flux is required, but most problems in diffusion require solving the
time-dependent partial differential equation known as Fick’s second law (as modified by
Darken), i.e.,
@CB @ e @CB @ e @CB @ e @CB e BA rCB :
¼ DBA þ DBA þ DBA ¼r D (8:14)
@t @x @x @y @y @z @z
The De BA is included inside the second derivative to take into account a possible spatial
variation in the diffusion coefficient. If there is no spatial variation, Fick’s second law
simplifies to
2
@CB e @ CB @ 2 CB @ 2 CB ~ BA r2 CB :
¼ DBA þ þ ¼D (8:15)
@t @x2 @y2 @z2
Note the similarity between this equation and Fourier’s equations for heat flow:
2
@T @ T @2T @2T
¼k þ þ ¼ kr2 T:
@t @x2 @y2 @z2
Heat flow is also by thermal diffusion in solids and k in the above equation is the thermal
diffusivity. Both coefficients have unit of m2=s.
0.8
0.6
C(x,t)/c
0.4
100 h
0.2 10 h
1h
0 FIGURE 8.12
0 20 40 60 80 100 Diffusion profile at different times for a diffusion coef-
x (μm) ficient D ¼ 2.5 1015 m2=s.
C ¼ C0 for 0 x 1 and t ¼ 0
C ¼ Cs for x ¼ 0 and t 0.
The solution is
x
C(x,t) C0 ¼ ðCs C0 Þ erfc pffiffiffiffiffiffi , (8:16)
2 Dt
where erfc(z) is the complimentary error function equal to 1 – erf(z). The error function is
defined as
ðz
2
ey dy
2
erf(z) ¼ pffiffiffiffi (8:17)
p
0
estimating the diffusion depth. If, for example, you wanted to diffuse As into Si in order to
make an n-type layer d mm thick by heating the Si in the presence of As vapor with
concentration Cs. You know the diffusion coefficient of As in Si at the temperature you are
operating is 2.5 1015 m2=s and you want to know how long it would take for the
diffusion front (C 0.5Cs) to advance 10 mm into the Si. Using the approximation,
t2 x=D ¼ 3 106=2.5 1015 ¼ 1.2 109 s2 or 9.6 h. This estimate agrees quite well
with the actual solution shown in Figure 8.12.
This same approximation is useful for estimating the temperature at depth d after time t
in a material whose surface temperature is Ts. Substituting the thermal diffusivity k for the
chemical diffusion coefficient D, the temperature T at depth d will be 0.5(Ts T0) when
t ¼ d2=k.
2 x3
erf(x) ¼ pffiffiffiffi x þ , x 1 (8:18)
p 3
and
2
ex 1
erf(x) ¼ 1 pffiffiffiffi 1 2 þ , x
1: (8:19)
x p 2x
8.7 Summary
Defects play a major role in the performance of materials, some wanted and some unwanted.
Therefore, it is necessary to understand what they do, how they form, and how to control
them. Defect are categorized by their dimensionality: point defects (zero dimensional), line
defects (1-D), planar defects (2-D), and volume defects (three dimensional [3-D]).
Point defects include vacancies, impurity or substitutional, and interstitial. Vacancy
defects are the result of a thermodynamic equilibrium and are unavoidable. Substitutional
and interstitial defects may contain unwanted impurity atoms, or may be engineered to
improve the material’s properties. If impurity atoms are added that have different valences
from the host atoms, the material is nonstoichiometric and vacancies must form to main-
tain charge neutrality.
Dislocations are line defects that form when a half plane of atoms is inserted (or removed)
causing a mismatch in the number of atoms above and below a certain plane. Such defects
can be edge, screw, or mixed, depending on how the material is distorted. They are charac-
terized by the Burgers vector, which remains invariant along the dislocation line. The Burgers
vector is defined as the missing step in a square loop drawn around the dislocation line.
Dislocations allow the material to slip along slip planes and in slip directions that have high
atomic density by only breaking bonds along a line rather than along an entire surface and are
thus responsible for materials being far weaker than their theoretical strength. Strengthening
mechanisms generally focus on preventing the motion of dislocations.
Any surface, whether it be the face of a crystal or a grain boundary in a polycrystalline
solid, is considered a defect because the atoms on the surface have unsatisfied bonds,
Defects in Crystals 173
which give them a higher energy that those atoms deep inside the crystal. This interfacial
energy plays a major role in the surface properties of solids. Stacking faults and twin
boundaries are other forms of planar defects.
Volume defects in the form of a second phase that is purposely added or precipitated from a
supersaturated solid solution are often used to improve the mechanical properties. Voids due
to the clustering of vacancies or from trapped gases are examples of unwanted 3-D defects.
Diffusion is a process by which different components can move through their
host material. Diffusion mechanisms include vacancy exchange diffusion, diffusion
along grain boundaries, through interstitials, and along dislocation lines. Diffusion in solids
is a very slow process, especially at low temperatures. Being a thermally activated process,
the diffusion coefficients follow an Arrhenius law and are generally expressed in the form
D (T) ¼ D0 eQ=RT, where the D0 and the activation energy Q are determined experimentally.
The dimensions of the diffusion coefficient are expressed in m2=s. The diffusion flux is given
by Fick’s first law that says the rate of atoms crossing unit area is the product of the diffusion
coefficient and the concentration gradient of the diffusing component. Time-dependent
diffusion is governed by Fick’s second law that is a second-order partial differential
equation, similar to Fourier’s heat flow equation, which usually requires a numerical
solution. However,
pffiffiffiffiffiffi it is often possible to obtain the information needed from the diffusion
length, d ¼ Dt where d is the distance the diffusion front has moved in time t.
Bibliography
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia, 1976.
Barsoum, M.W., Fundamentals of Ceramics, Institute of Physics Publishing, Bristol=Philadelphia, PA,
2003.
Callister, W.D., Materials Science and Engineering: An Introduction, 7th edn., John Wiley & Sons,
New York, 2007.
Carslaw, H.S. and Jaeger, J.C., Conduction of Heat on Solids, 2nd edn., Oxford University Press, Oxford,
1959.
Crank, J., The Mathematics of Diffusion, 2nd edn., Oxford University Press, Oxford, Reprinted in 1976.
Glicksman, M.E., Diffusion in Solids: Field Theory, Solid-State Principles, and Applications, John Wiley &
Sons, New York, 2000.
Hosford, W.F., Physical Metallurgy, Taylor & Francis, CRC Press, Boca Raton, FL, 2005.
Kingery, W.D., Bowen, H.K., and Uhlmann, D.R., Introduction to Ceramics, 2nd edn., John Wiley &
Sons, 1976.
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley & Sons, New York, 1966.
Schaffer, J.P., et al., The Science and Design of Engineering Materials, Richard D. Irwin, 1995.
Schweizer, M., et al., Defect density characterization of detached-grown germanium crystals,
J. Crystal Growth, 235, 161–166, 2002.
Problems
1. Activation energy for vacancies in Cu is 1.28 eV. Find the fractional number of vacancies
at the melting temperature. Estimate the contribution of vacancies to the thermal
expansion at the melting temperature.
2. In close-packed metals, we generally can ignore Frenkel defects. Given this assumption,
devise a method for determining the concentration of vacancies as a function of tem-
174 Introduction to the Physics and Chemistry of Materials
perature. Is it necessary to be able to ignore Frenkel defects for your scheme to work?
Why?
3. Activation energy for diffusion of C into Ni is 146 kJ=mol. By what factor does the
steady-state flux of C into Ni increase if the temperature increases from 258C to 508C?
4. I want to diffusion-bond a Si chip to a Au-coated substrate and I want the bond to
penetrate to roughly 0.1 mm. Au has a diffusion coefficient in Si of 7 107 m2=s at
11008C and 109 m2=s at 8008C. Create a time–temperature plot for this process.
5. Diffusion coefficient of As in Si is 1015 m2=s at 14008C and is 1018 m2=s at 10008C.
I have a wafer of intrinsic Si that I need to make n-type with a carrier concentration
of 1018 electrons=m3 at a depth of 1 mm (which means a concentration of 1018 As at
1 mm). I propose to achieve this by exposing the wafer to As vapor. Devise a practical
method for achieving this doping level in terms of the temperature and pressure of
the As gas and the exposure time required. (Assume the As can be considered as an
ideal gas.)
6. Find the interfacial energy of the (110) face of NaCl. (see Problem 3.5).
9
Mechanical Properties of Materials
In previous chapters, we saw how materials are structured, chemically bonded, and how
their elastic properties are related to their structure and their bonds. Now we must deal
with how materials actually perform and how we can improve their performance.
175
176 Introduction to the Physics and Chemistry of Materials
Gage length
FIGURE 9.1
Sample geometries indicating the gage
length. Gage length
amount of elastic energy the material can store and is termed the resiliency. The resiliency
can be expressed as
where E is the Young’s modulus. Materials with high resiliencies are used for springs and
other applications where elastic energy storage is required.
The departure from the linear relationship between stress and strain indicates an irre-
versible change in the structure. Bonds are now being broken as individual grains undergo
slip by motion of dislocations as their resolved stress exceeds their critically resolved shear
stress. Individual grains begin to move past each other by grain boundary slip. Once a
material is stressed beyond its yield point, it will suffer a permanent deformation when
stress is relieved, as is indicated by the dashed unloading line.
True stress
s tensile
s yield
Stress
ading
Unlo
TABLE 9.1
Strength and Ductility of Selected Alloys
Alloy E (GPa) syield (MPa) stensile (MPa) eyield (%) efracture (%)
1100-0 Al 69 34 90 0.04 40
7075-T6 Al 71 504 572 0.7 11
Ti (pure annealed) 103 170 240 0.16 30
Ti6Al4V 114 760 900 0.65 14
Ti6Al4V (age hardened) 114 1103 1172 0.9 10
1020 low carbon steel 207 295 395 0.14 36.5
17-7PH SS 204 1310 1450 0.64 3.5
440A SS (tempered) 200 1650 1790 0.8 5
VIT-001 93 1863 1863 2.0 2.0
Source: From Callister, W.D. Materials Science and Engineering: An Introduction, 7th edn., John
Wiley & Sons, New York, 2007.
178 Introduction to the Physics and Chemistry of Materials
increase in both yield strength and tensile strength by the various hardening techniques is
quite apparent, but note that this increase comes at the expense of ductility (the strain at
fracture). The elastic limit is obviously increased along with the yield strength. The VIT-001
is a bulk metallic glass and will be discussed in Section 9.3.6.
ðf
L
dL Lf DL
et ¼ ¼ ln ¼ ln 1 þ ¼ lnð1 þ eÞ e: (9:2)
L L0 L0
L0
After the material begins to yield, the true stress can be empirically related to true strain by
where
K is a constant
n is the strain-hardening exponent (n < 1)
However, for most applications, it is a common practice to simply use the engineering
stress in evaluating materials.
3Ff
sf ¼ : (9:4)
2bd2
Arbitrary hardness scales have been established that allow some correlation between
hardness and yield strength. There are a variety of different methods including Brinell,
Rockwell, Vickers, and Knoop, each with its own hardness scale. The reader is referred to
ASTM Standards E10, E18, E92, and E384 for details. The primary value of hardness testing
is that it provides a nondestructive method for assessing the strength of a material and is
useful for certifying that the actual materials used for specific components in a system meet
the required material specifications.
TABLE 9.2
Slip Systems
Crystal System Slip Planes Slip Directions Number of Slip Systems
[111]
[111]
(211) (321)
FIGURE 9.3
Higher index bcc slip systems.
b
FIGURE 9.4
Edge dislocation in the (100) face of a rock salt structure. The Burger’s
circuit is shown by dashed line and the Burger’s vector b has a
magnitude of (r1 þ r2) in the [010] direction, which makes (001) the
slip plane. Dislocations cannot propagate to the right because they are
blocked by the strong coulomb repulsive force from the large anions
that must move past each other.
ionically bonded system because a dislocation cannot form with the insertion of a single
extra half plane of atoms. This would result in like atoms touching like atoms. Dislocations
can form in ionic crystals by inserting two complimentary half planes of atoms as shown in
Figure 9.4. Of course this will be more costly in energy than forming dislocations in metals
because of the increased lattice strain. In addition, it is more difficult for dislocations to
move in an ionic system. Consider the (100) plane of the rock salt structure as shown in
Figure 9.4. The large anion is blocked from moving to the left and filling the vacancy left by
the dislocation as it comes into contact with another anion by the Coulomb repulsion. In
the [110] plane, it is possible for an ion to move without direct contact with other ions of the
same sign, but this is not a high-density plane. The extra energy required to form disloca-
tions in ionic systems, together with the inhibited motion of dislocations due to ionic
repulsion, accounts for the hardness and brittleness of ionic compounds.
which is known as the Hall–Petch relationship. The k is a constant for a given material.
Methods for controlling the grain structure will be discussed in Chapter 14. Work hard-
ening and grain refining strengthening mechanisms are effective in elemental systems as
well as in alloys.
lend themselves to precipitation hardening. It also has the advantage of remaining effective
throughout long exposures to elevated temperatures without the overaging problem
encountered in precipitation-hardened systems.
9.4 Creep
Creep is a thermally activated phenomenon in which a material slowly becomes elongated
in response to applied stress. The onset of creep usually begins at around 30% of the
absolute melting temperature and can become significant above 50% of the absolute
melting point. Naturally, this phenomenon becomes of great concern in the design of
rotating machinery such as the turbine blades in a gas turbine or in a jet engine that
must operate at high temperatures.
There are three stages of creep. In the initial or primary stage, creep is fairly rapid but
diminishes with time as the material strain hardens. During the secondary stage, creep
continues at a constant rate as the rate of strain hardening is balanced by the rate of thermal
softening or recovery (to be discussed in Chapter 11). In the final stage, the creep velocity
accelerates rapidly as gross defects begin to form just before rupture.
There are several mechanisms that give rise to creep. At high temperatures there are
more vacancies as well as enhanced diffusion which allows the vacancies to migrate. In the
Nabarro-Herring creep mechanism, atoms migrate in the direction of applied stress while
vacancies migrate to the lateral surfaces, causing a net elongation in the stress direction.
Coble creep is similar except the vacancies migrate along grain boundaries. Another creep
mechanism is dislocation climb. Atoms in the extra half plane of a dislocation migrate into
vacancies, which moves the dislocation line in the direction of the applied stress (Figure
9.5). Other creep mechanisms include grain boundary sliding in response to the resolved
shear stress.
Ceramics are much more resistant to creep than are metals because of the difficulties in
forming vacancies and dislocations as well as to the lack of mobility of vacancies and
FIGURE 9.5
Mechanism of dislocation climb. Atoms from
dislocation plane migrate to vacancies making
the specimen longer and thinner.
Mechanical Properties of Materials 183
dislocations due to the charges on the atoms. However, creep can become significant even
in ceramics at very high temperatures through the diffusional mechanisms as well as
through grain boundary slippage.
Some ceramics may have a glass phase connecting the individual crystallites. Above the
glass transition temperature, glasses deform by viscous flow that increases exponentially
with temperature. Therefore, these ceramics as well as metallic glasses and polymers can
creep by viscous flow.
Also, it should be noted that creep is not just a high temperature phenomena found in
high performance gas turbine engines, but can occur at ambient temperatures in polymers
and solders.
where
2L is the length of the hole
r is the tip radius.
184 Introduction to the Physics and Chemistry of Materials
One can appreciate the effectiveness of the stress concentration in a crack whose tip radius
is on the order of an atomic dimension. No doubt the reader is familiar with how much
easier it is to tear open a plastic packet if there is a small indentation or crack to start the
tear. It is also a common practice to drill a small hole at the end of a crack in a metal plate to
‘‘blunt the crack’’ by increasing the tip radius to relieve the stress concentration.
One can make a simple model for fracture by estimating the stress at the tip of the crack
by multiplying Equation 9.6 by the applied stress and equating this to the theoretical
tensile strength, estimated to be E=10,
rffiffiffi! rffiffiffi
E L L
¼ sapplied 1 þ 2 2sapplied (9:7)
10 r r
or
rffiffiffi
E r
sfail : (9:8)
20 L
However, the Inglis stress concentration by itself may not be sufficient to cause a crack to
self-propagate even though the applied stress is greater than sfail in Equation 9.8 because
energy is required to do the work of fracture, i.e., the energy to break bonds and create new
surfaces. Since the applied stress is doing no work on the system, this energy must come
from somewhere else if the crack is to continue to propagate.
Correcting for the decrease in the force between the planes of atoms as they are separated
and integrating over x yields
rffiffiffiffiffiffi
gE
stheoretical ¼ : (9:10)
x
Griffith wanted to test his theory against the actual strength of materials. However, the g
was not known and he did not have a way to measure it directly for metals. So Griffith
turned to glasses. The measurement of the surface tension of a liquid is straightforward.
Since a glass is essentially a frozen-in liquid, Griffith reasoned that he could extrapolate the
measured surface tension of molten glass to the room temperature value, which he found
to be 0.56 N=m. A simple tensile test determined the Young’s modulus of his glass samples
to be 62 GPa. Taking the distance of separation to be 1 nm, the theoretical tensile strength
Mechanical Properties of Materials 185
was estimated to be 13 GPa. (Note that this is the same order of magnitude as estimated in
Chapter 7 as the stress required to break bonds by pulling planes of atoms apart.)
Beginning with 1 mm diameter glass rods, Griffith found they broke at 0.17 GPa, almost
two orders of magnitude below the theoretical estimate. However, when he tested smaller
diameter rods, he found them to be progressively stronger as the rods decreased in
diameter. A rod of 0.0025 mm, the smallest he could make and measure accurately,
exhibited a tensile strength of 3.4 GPa and extrapolation suggested that the theoretical
strength would be attained at vanishingly small diameters.
Why should strength increase with decreasing diameters? Griffith postulated that the
fracture of the brittle rods was initiated by microscopic cracks in their surface. The strain
energy per unit volume, we recall, is s2=2E. The strain energy that is released in the vicinity
of a crack with length L in a material with thickness d is approximately pdL2s2=2E.
The energy required to make new surfaces is 2Ldg. The energy of the system under
stress s is
s2 pdL2 s2
U¼V þ 2Ldg, (9:11)
2E 2E
where V is the volume. Initially the energy increases with increasing L as energy is required
to make new crack surfaces. But eventually the energy released in the vicinity of the crack
more than compensates for that required to create the new surface. The maximum energy
is found by differentiating with respect to L and is
2gE
LC ¼ : (9:12)
ps2
Cracks shorter than the critical LC cost more surface energy than they release in order to
propagate and are therefore stable. Those longer than LC release more strain energy than it
costs to make their surfaces and will spontaneously propagate. The smaller the diameter of
the glass rods used by Griffith, the smaller the crack they could contain and therefore they
could withstand higher tensile stresses before whatever cracks they may have had would
become critical.
You may have noticed that glass cutters scribe the surface of a glass plate they wish to
cut with a diamond point and then apply a thin coating of kerosene or light oil that wets
the surface. The scratch serves as the stress concentrator and crack initiator and the wetting
agent reduces the interfacial energy making it easier for the crack to propagate through
the glass.
GE
LC ¼ : (9:13)
ps2
300
0.01 0.11
Impact energy (J)
200
0.22
FIGURE 9.6
Impact testing data for pearlitic steels as a
function of temperature for various wt% 0.31
carbon content. High carbon steels are 0.43
100
stronger but become brittle at ambient tem- 0.53
peratures, whereas low carbon steels are 0.63
ductile until 508C, when they suddenly
become brittle. (From Reinbolt, J.A. and 0.67
0
Harris, W.J., Trans. ASM, 43, 1951. Reprinted
−200 −100 0 100 200
with permission of ASM International. All
rights reserved.) Temperature (°C)
Mechanical Properties of Materials 187
performance of the engine, but caused also Rolls Royce to declare bankruptcy in 1971
because they could not meet their delivery schedule. This delay also seriously affected
Lockheed who was waiting for these Rolls Royce engines for their new L1011 passenger jet.
The delay in getting the L1011 into service costs Lockheed considerable market share and
was one of the factors that caused them to abandon their commercial passenger business.
(The above discussion on Griffith cracks was taken from Gordon (1988).)
The best example of a self-propagating crack is a balloon when punctured by a pin.
Instead of allowing the air to be slowly released through the pinhole, the high stress in the
skin results in a very small critical crack length. Similarly, designers of pressurized
containers, including the monocoque design of modern passenger jets that utilizes the
stressed skin for rigidity, need to be more concerned with critical crack length than just
yield strength. If one wants to design such a vessel that will leak before bursting, the
stresses in the skin should be limited to a critical stress intensity factor in which the critical
crack length is equals to or greater than the skin of the vessel.
for a smaller critical crack length, which allows the maximum applied stress s in Equation
9.14 to be larger. Subdividing structures can significantly increase their resistance to tensile
failure. A stranded cable is an excellent example. Smaller individual strands can sustain
higher stress and remain stable than a single thicker strand because of their smaller critical
crack length. In addition, the stranded cable has the advantage of redundancy: several
smaller strands can break while the rest carry the load if there is no mechanism for
transferring the release energy from one strand to another. This same principle carries
over to structures with plates that are welded together vs. plates that riveted or bolted
together. The tendency for modern aircraft to use a monocoque design in which the stresses
are carried predominantly by the skin improves their strength-to-weight ratio, but makes
them more susceptible to catastrophic failure from a crack exceeding its critical length.
Presumably this is what happened to the Comet passenger jets that disintegrated in midair.
9.5.7 Fatigue
Fatigue is a primary cause of structure failure. Perhaps the most spectacular failure was the
sudden loss of the skin over the forward passenger cabin of a Boeing 737–200 used on
Aloha Airlines Flight 243 in 1988. The only casualty was a flight attendant who was blown
overboard. Miraculously, the pilot was able to land the plane safely with the passengers
strapped in their seats. The investigation found that metal fatigue, exacerbated by stress
corrosion, was the reason for the failure. Similar incidents have occurred on other aircraft,
trains, and structures.
Fatigue can be thought of as a gradual wearing out of a material due to cumulative internal
structural damage brought about by repeated stresses and strains such that the material
eventually fails at a stress that is well below the normal tensile strength of the material.
Repeated bending of a tab on a drink can or paper clip will cause a fracture at a small fraction
of the material’s tensile strength. Why? The first bending generates dislocations that intersect
one another and work hardens the material. (Recall the copper wire bending experiment.)
Repeated bending creates more dislocations that continue to pile up making the material
more brittle because the dislocations can no longer move. Eventually, the lattice becomes so
disordered that cracks begin to form and propagate, resulting in brittle fracture.
Metals are tested for fatigue by repeated loading or flexing at a given stress level until they
fail. The test is then repeated at different stress levels until enough points are generated to
construct what is termed an S–N (stress vs. number of cycles to failure) curve illustrated
schematically in Figure 9.7. Since at low stress, as many as 107 cycles may be required for a
single datum point, it can be appreciated that fatigue tests can be time-consuming and
B
Stress
Fatigue limit
FIGURE 9.7
Fatigue strength
Typical S–N fatigue curve. Material A exhibits a fatigue
limit and can withstand an unlimited number of cycles if
the stress is held below this level. Material B does not have a
102 104 106 108
fatigue limit. Its fatigue strength is defined as the average
stress that results in failure after 107 cycles. Cycle number
Mechanical Properties of Materials 189
expensive. To make things worse, there will be significant statistical scatter in the measured
fatigue life, the average number of cycles before failure at a given stress. This means that
enough tests must be run to determine the average number of cycles to failure with some
level of confidence. Stresses that cause some plastic deformation cause failure at much fewer
cycles O(104–105) and are termed low cycle fatigue. If the deformations are purely elastic, the
fatigue life is greatly extended in what is termed high cycle fatigue.
Notice that curve A approaches a lower stress limit called the fatigue limit or endurance
limit below which fatigue failure does not occur, where in curve B the fatigue stress
continues to decrease with increasing cycles. In this case the fatigue strength is defined
as that strength for which failure occurs at 107 cycles. Steels tend to have a fatigue limit,
which favors their use for applications such as train wheels, axels, crankshafts, connecting
rods, and other components of reciprocating machines. Aluminum, copper, and other fcc
metals do not exhibit an endurance limit.
Surface treatments to prevent surface cracks forming in steels include case hardening,
which consists of carburizing or nitriding in which carbon or nitrogen is diffused into the
interstities of the bcc lattice to put the surface into compression. Another surface treatment
used on connecting rods and other components of high performance engines is shot
peening. Small diameter shot (0.1–1 mm diameter) is impinged at high velocity on the
component. The small dents from the impacts work harden the surface and seal incipient
surface cracks. Both of these techniques raise the S–N curve as well as the fatigue limit.
Ceramics are not generally subject to fatigue since they do not deform plastically because
dislocation motion is inhibited.
Necking Chains
begins align
Glass
Semicrystalline
Stress
Elastomer
FIGURE 9.8
Schematic stress vs. strain behavior of different
Strain classes of polymers.
190 Introduction to the Physics and Chemistry of Materials
stretching or bending of covalent bonds in their crystalline chains and reversible displace-
ments of the molecules in their amorphous regions. At the onset of plastic deformation, the
covalent chains in the amorphous regions start to move and align themselves along the
direction of the applied stress and the specimen begins to form a neck. Next the crystallized
regions begin to align themselves in the direction of the stress and separate into segments
of crystal linked by the covalent strands. At this point the material actually becomes
stronger as all of the stress is being borne by aligned covalent chains. This can be
demonstrated by taking the plastic (polyethylene) band from a six pack of drinks and
stretching it. Observe how the material forms a neck of aligned fibers as it stretches and
how much stronger the aligned material in the neck is compared to the original material.
9.6.2 Elastomers
The class of polymers known as elastomers do not crystallize. Instead their randomly
coiled covalent chains are loosely cross-linked. Initially they are easily deformed, but as
their chains uncoil and become aligned, their stiffness increases giving their stress–strain
behavior a sort of ‘‘lazy-J’’ appearance as shown in the sketch in Figure 9.8.
Stretching an elastomer increases the order in the system as the strands become aligned.
When the stress is relieved, the second law of thermodynamics requires the entropy to
increase, causing the covalently bonded chains to form random coils again and, guided by
the loose cross-links, the elastomer returns to its original shape.
FIGURE 9.9
Time response of different rheological
systems to applied forces. The Maxwell
model gives steady creep with some Elastic Viscous Maxwell Kelvin–
post stress recovery, representative of Voigt
a polymer with no cross-linking. The
Kelvin–Voigt model gives a retarded Force Force Force Force
viscoelastic behavior expected from a applied applied applied applied
cross-linked polymer. Time
Mechanical Properties of Materials 191
viscous behavior. Many polymer melts are non-Newtonian and have to be modeled by
various combinations of springs and dashpots.
In the Maxwell model of the viscoelastic system, initially the entire force is transmitted
through the dashpot causing the spring to stretch in proportion to the force. As the dashpot
begins to move, the force is divided between that required to move the fluid and to stretch
the spring allowing the spring to begin to relax. The partial relaxing of the spring accounts
for the bending of the response curve and limited recovery when the force is removed. In
the Kelvin–Voigt model, the spring response is damped by shock absorber as in an
automobile suspension, which results in a delayed elastic response. Adding additional
elements gives more variables to fit the behavior of any type of viscoelastic material.
s(t)
ER (t) ¼ : (9:15)
«0
A specimen is rapidly strained by an amount «0 and then the stress s(t) required to hold
that strain is measured as a function of time. For an elastic material, ER(t) is the Young’s
modulus and is constant in time. For a purely viscous material, ER(t) quickly goes to zero.
To obtain the temperature behavior of a viscoelastic material, ER(t) is evaluated at some
particular time, usually 10s, during the relaxation process. The values of ER(10) are then
plotted against temperature as shown in Figure 9.10.
All thermoplastic polymers start out with a glassy behavior below the glass transition
temperature Tg. As T is increased above Tg, they all become more ductile or leathery and
later become softer or more rubber-like. A semicrystalline isotactic polymer will retain its
strength longer than an amorphous polymer because its crystalline structure permits more
elastic deformation while the others deform viscoelactically. Eventually, semicrystalline
isotactic polymer as well as the amorphous polymer will turn into a viscous liquid as they
eventually melt. A cross-linked polymer softens somewhat at the glass temperature but
does not melt because of the strong covalent cross-link bonds. It will decompose before
these bonds are broken.
Glassy
Viscoelastic
Relaxation modulus
Rubbery
Melt
FIGURE 9.10
Tg Schematic of the relaxation modulus plotted against
Temperature temperature for a viscoelastic polymer.
192 Introduction to the Physics and Chemistry of Materials
9.7 Summary
At low stresses, metals tend to deform elastically. The strain is proportional to the stress
(Hooke’s law), no bonds are broken, and there is complete recovery when the stress is
relieved. At a certain stress level (the yield strength), ductile metals will began to deform
plastically, bonds are being broken, and a permanent deformation will result when the
stress is relieved. As more stress is applied, ductile metals will begin to form a neck and
eventually rupture. The maximum recorded stress is the tensile strength of the material.
However, the recorded stress at rupture will usually be less than the maximum recorded
stress because of the area reduction as the neck forms. The strain at rupture is a measure of
the ductility, usually expressed in percent elongation. These engineering stress and strain
measurements can be converted into true stress and strain by taking into account the area
reduction when the neck is formed. The total area under the stress–strain curve is a
measure of the toughness of the material. The area under the stress–strain diagram at the
elastic limit is the elastic work stored in the material and is called the resiliency and is given
by syield eelastic =2 ¼ Ee2elastic =2 ¼ s2yield =2E, where E is the Young’s modulus.
Brittle metals and ceramics undergo little or no plastic deformation before rupturing.
Rupture occurs at stress well below the theoretical strength because of surface cracks that
open and eventually exceed their critical length and propagate through the material
leaving a planar surface. Very ductile materials tend to neck down and simply pinch off
at stresses well below their theoretical strength. Less ductile metals tend to form voids due
to vacancy coalescence during the necking process. These voids form an elliptical crack that
propagates across the specimen, leaving a characteristic cup and cone fracture.
Metals deform by dislocations moving along slip planes, planes with high atom densities
(atoms=area), in directions that have high linear atom densities. The ductility of a metal
depends on the number of active slip systems. Face-centered lattices have the most number
of active systems and are generally the most ductile. Body-centered systems actually have
more slip systems than fcc systems, but the slip planes have lower atom densities, hence
are not as active. Some of these slip systems become inactive at low temperatures causing a
ductile-to-brittle transition to occur. Hexagonal close-packed systems have only one active
slip system. Since their slip planes are parallel, cross-slip cannot occur and hcp systems are
generally more brittle than other metals.
Ceramics tend to be very brittle because (1) it is more difficult to form dislocations
because a dislocation requires two half planes of ions to be inserted; and (2) it is difficult for
the dislocations to move because of the Coulomb repulsion as large anions must slip past
each other.
Since plastic deformation involves dislocation motion, most strengthening methods seek
to inhibit the motion of dislocations. There are essentially five ways of accomplishing this:
work hardening, solid solution hardening, grain refining, dispersion hardening, and pre-
cipitation hardening. Glasses can be strengthened by tempering, which is accomplished by
chilling the exterior to put it into compression and sealing off any insipient surface cracks.
Metals can also be hardened by diffusing small molecules such as C or N into their
interstities or by shot peening to seal any surface cracks. Steels and ceramics may be
transition toughened by incorporation of a metastable phase that will undergo a transition
under stress to seal cracks.
Creep is a thermally activated process under which materials will slowly stretch under
stress when the temperature exceeds 50% of the melting temperature. Creep mechanisms
include dislocation climb, vacancy migration, and grain boundary slip. Creep is especially
important in gas turbine engines and a class of Ni-based superalloys has been developed to
minimize its effects.
Mechanical Properties of Materials 193
Modern fracture mechanics was pioneered by A.A. Griffith who showed that brittle
materials could fail catastrophically by cracks that become self-propagating even at
stresses much lower than their tensile strength. Griffith’s theory was expanded to include
ductile materials and has led to the definition of a material property called fracture
toughness that allows the prediction of critical crack lengths.
The mechanical properties of polymers are quite different from those of metals and
ceramics and depend on the structure of the polymer and temperature. Semicrystalline
polymers initially deform elastically much like metals by the stretching of their covalent
bonds. Plastic deformation occurs when the covalent chains start to move and align
themselves with the applied stress. This alignment of the strong covalent chains causes
an increase in strength before rupture.
Elastomers have tightly coiled chains that can easily be deformed by uncoiling the
chains. Driven by entropy and guided by a light cross-linkage, the chains recoil to their
original state when stress is relieved.
Amorphous polymers deform viscously above their glass transition temperature. Poly-
mers generally exhibit non-Newtonian behavior (viscosity is not constant) and their time-
dependent behavior can be modeled mechanically by a combination of springs and
dashpots.
Bibliography
Barsoum, M.W., Fundamentals of Ceramics, Institute of Physic Publishing, Bristol=Philadelphia, PA,
2003.
Callister, W.D., Materials Science and Engineering: An Introduction, 7th edn., John Wiley & Sons,
New York, 2007.
Gordon, J.E., The Science of Structures and Materials, Scientific American Library, New York, 1988.
Hosford, W.F., Physical Metallurgy, Taylor & Francis, CRC Press, Boca Raton, FL, 2005.
Kingery, W.D., Bowen, H.K., and Uhlmann, D.R., Introduction to Ceramics, 2nd edn., John Wiley &
Sons, New York, 1976.
Lawn, B., Fracture of Brittle Solids, 2nd edn., Cambridge University Press, UK, 1993.
Reinbolt, J.A. and Harris, W.J., Trans. ASM, 1951.
Russell, A.M. and Lee, K.L., Structure–Property Relations in Nonferrous Metals, John Wiley & Sons,
New York, 1966.
Problems
1. Simply plotting the observed stress against the observed strain in the elastic region does
not give the true value for the Young’s modulus because it does not take into account the
narrowing of the sample due to the Poisson’s ratio. Derive the true Young’s modulus
taking into account true stress and true strain. How much error is incurred in using the
engineering value?
2. Explain how you could derive the modulus of elasticity from a bend test using the
relation given by Equation 9.4.
3. Find the area density fraction for the {211} and {321} families of planes for the bcc
system.
4. Find the area density fractions for the (10
10) and the (1011) planes in the hcp system.
How do these compare the (211) and (321) planes in the bcc system?
194 Introduction to the Physics and Chemistry of Materials
5. Comment on Griffith’s choice of 1 nm for the length at fracture. Compare his estimate of
the theoretical strength of his glass rods against the discussion on theoretical strength in
Chapter 7.
6. You have an automobile with 16 in. Al alloy wheels. Should you start worrying
about fatigue failure of these wheels at 100,000 mi.? Estimate the safe lifetime of these
wheels.
7. You are asked to design a pressure sphere of 0.5 m in. diameter that will hold 100
atmospheres with minimum weight. A safety factor of 2 is required below ultimate
strength. What material would you chose? How thick would the walls be? What would
the weight be?
8. The requirements in Problem 5 have changed. The customer wants the pressure sphere
to leak before it ruptures. How would this change your design?
9. Silly putty bounces when thrown against the ground but slumps when left sitting. How
would you model this response?
10
Composites
The ability to join dissimilar materials has greatly expanded the use of composite materials
in applications where great strength, light weight, and dimensional stability are required.
This chapter will review the present state of the development and the use of composites
with some insight into how they are able to increase the performance of materials.
195
196 Introduction to the Physics and Chemistry of Materials
consumer use although we are now seeing aluminum metal matrix composites (ALMMCs)
appearing in bicycle frames and in the drive shafts of drag racers and Corvettes. Carbon
fiber-reinforced silicon carbide (SiC) is now used to eliminate brake fade in the brake
rotors used by Porsche as well as in Formula 1 race cars and in some aircraft. Other
manufacturers are considering the use of these high-performance ceramics for brake rotors
and clutch plates in less exotic vehicles.
slabs from cracking. The high pH of the concrete prevents the steel from rusting. To
provide the necessary strength for concrete beams used for highway bridges and over-
passes, the beams are cast with a number of small cylindrical holes running the length of
the beam. After the beam is cured, steel cables are threaded through these holes, tensioned
hydraulically, and swaged at both ends. This prestressing keeps the beam in compression
thus taking advantage of the high compressive strength of the concrete.
Brick mason’s mortar is basically concrete without the gravel. The sand acts as a filler
and the cement binds the sand together with the bricks by penetrating into small pores in
the brick. The strength of the mortar depends greatly on how well the cement is able to
penetrate into the brick, which is determined by the porosity of the brick and the moisture
content at the brick interface.
10.2.2.2.3 Cermets
Cermets are a form of metal matrix composites (MMCs) usually composed of a large
volume percent of ceramic in a metal host phase. Cutting tools consisting of very hard
ceramics such as WC, TiC, or even industrial diamond, embedded in a metal matrix
combine the cutting ability of the ceramic with the fracture toughness as well as the heat
dissipation provided by the metal. For example, carbide drills and saw blades are often
made from a tough cobalt matrix with tungsten carbide particles inside. The metal isolates
the hard ceramic particles from each other and prevents cracks from propagating from one
particle to another.
the bond between the fiber and the matrix. This relationship as well as other mechanical
properties will be discussed in Section 10.3. Here we will briefly survey some of the
applications of fiber-reinforced composites.
rotors, bicycle frames, and sporting goods. Al MMC drive shafts are advertised to be 30%
stronger than 6061-T6 shafts.
aircraft, high-temperature molds, and heat shields. The nose and wing leading edges on
the space shuttles are C–C composites. These composites owe their high strength to the sp2
bond in graphite. The carbon filaments are made by oxidizing and pyrolyzing polyacrylo-
nitrile (PAN). By heating to 30008, the fiber turns to almost pure graphite with a tensile
strength of 5 GPa and an elastic modulus of 500 GPa. Layers of these continuous carbon
fibers are impregnated with a resin, usually a phenolic, and allowed to cure. The material is
then pyrolyzed or heated to the destruction of the binder, leaving only a carbon matrix
reinforced by carbon fibers.
One difficulty in using C–C composites in space is that they can be degraded by
exposure to atomic oxygen, which is present in near-earth orbit. For this reason a protective
layer of SiC must be added to the portions of the C–C materials that are exposed to the
space environment.
baking at 1258C. The high thermal conductivity, low density, and high stiffness make this
composite ideal for fabricating highly efficient radiators for deep space missions that use
solar=electric of nuclear=electric power. Rejecting the large amount of waste
heat from such power systems is a major design challenge calling for highly efficient
radiators. One such prototype design was fabricated by ATK Space Systems using
multiple layers of a prepreg containing K13D2U in EX1551 cyanate ester to produce a
layered tapered fin. Wrapping the prepreg around a Ti=H2O heat pipe effectively con-
ducts the heat from the heat pipe to the fin where it can radiate to space. The assembly
was laid up in a graphite mold with a grove to accept the Ti tube. Crossing the fibers at
different angles in the laminated fin provided stiffness and controls the CTE. The finished
assembly is then cured under pressure in an autoclave. Since the radiator must operate
over a range from 1008 to 550 K, it is important to match the CTE of the graphite fiber-
reinforced polymer composite with the Ti heat pipe so it does not delaminate during
thermal cycling.
With the discovery of CNTs and their remarkable properties came great interest in
trying to incorporate them into the matrix material as well as use them in the fiber to
improve the strength of fiber-reinforced composites. Originally the nanotubes were pro-
hibitively expensive to experiment with on the scale needed for such experimentation and
their lengths were shorter than the critical length needed to achieve good coupling to the
matrix. Also, ways need to be found to functionalize the nanotubes in order for them to
bond with the matrix material.
Recent developments in synthesis and assembling CNTs into continuous fibers have
opened the doors to more widespread research in the effort to develop the ultimate fiber
(see Section 5.4.7). When subjected to high pressure, some of the sp2 bonds can convert to
sp3 bonds, which may provide a means for cross-linking the CNT fibers into a super-strong
fabric—carbon threads joined by diamonds.
Substituting the net Young’s modulus times the area of the composite for kDx and the
Young’s modulus times the areas of each component for k1Dx and k2Dx, the net modulus
becomes
E1 A1 E2 A2
E¼ þ ¼ E1 V1 þ E2 V2 ¼ EMatrix ð1 VParticle Þ þ EParticle VParticle , (10:2)
AC AC
where it is assumed that the volume fractions V1 and V2 of the components are given by
their cross-section area divided by the total cross-section area.
Composites 205
s fiber 5
4
s Composite
2 FIGURE 10.1
Upper and lower bounds of the performance of a particle-
reinforced composite set by the rule of mixtures. In this
illustration the strength of the fiber is five times the matrix.
s Matrix 1 The dashed line in the middle is the estimated performance
0 1 of a randomly oriented discontinuous fibers (1=3 aligned
Volume fraction of fiber
with the applied stress, 2=3 normal to the applied stress).
The other extreme is represented by placing the springs in series. Now the FA is the same
on each spring, but the displacements will be given by Dx1 ¼ FA=k1 and Dx2 ¼ FA=k2. Now
the net displacement is given by Dx ¼ Dx1 þ Dx2 and the net spring constant becomes
FA 1 k1 k2
k¼ ¼ ¼ : (10:3)
Dx1 þ Dx2 1=k1 þ 1=k2 k1 þ k2
If we substitute the Young’s modulus times the area of the composite for kDx and identify
the strains «1 ¼ FAA1=E1 and «2 ¼ FAA2=E2
FA Dx FA FA E1 E2
EA AC ¼ kDx ¼ ¼ ¼ ¼ : (10:4)
Dx1 þ Dx2 «1 þ «2 FA A1 =E1 þ FA A2 =E2 E1 A2 þ E2 A1
E1 E2 EMatrix EParticle
E¼ ¼ : (10:5)
E1 V2 þ E2 V1 EMatrix ðVParticle Þ þ EParticle ð1 VParticle Þ
One can see in Figure 10.1 that the second model will give the lowest effective modulus
since the weakest spring in a series will stretch the most.
These derivations, while far from rigorous, do set upper and lower limits on the
combined effect of the two components. The net thermal conductivity and CTE can also
be estimated in this manner.
exert a restraining force given by dFz ¼ pD sBdz, where sB is the shear strength of the bond
between the fiber and the matrix. The stress in the fiber is
ðz ð
L=2
4 4
s ð zÞ ¼ dFz ¼ sB dz; s < sF , (10:6)
pD2 D
0 0
where sF is the tensile strength of the fiber. When the stress in the fiber reaches sF, the
maximum load has been transferred to the fiber. Thus there is a critical fiber aspect ratio
given by
L sF
¼ : (10:7)
D C 2sB
If the aspect ratio of the fibers is less than the critical value, the bond between the fibers and
the matrix will fail before the fibers break and full advantage of the fiber’s strength will not
be realized.
sF Fiber
sTC
Stress
sYC
sM
Matrix
sM(eF)
sYM
FIGURE 10.2
Schematic stress–strain behavior of aligned composites. It is
assumed that the matrix is ductile and the fibers fail by brittle
eYM eF eM
fracture. The yield strength of the composite is sYC and its
tensile strength is sTC . Strain
Composites 207
As additional stress is applied, the matrix will continue to transfer load to the fibers as it
yields plastically until the strain at fracture eF of the fiber is reached. The tensile strength of
the composite is given by
The factor of 2 in the denominator comes about since the average stress in a fiber the
critical length is half the maximum stress. For aligned fibers LC, the modulus is given by
When the fiber length L 15 LC, the fibers are considered to be fully effective as reinforce-
ments to the matrix.
Aligned discontinuous fibers offer a negligible advantage in the transverse direction. To
obtain isotropic performance, the fibers may be randomly distributed. The modulus may
be written as
where K is a fiber efficiency factor that varies from 1 for aligned fibers to 0.2 from
randomly distributed fibers.
10.4 Summary
Both nature and man have made extensive use of composite materials in which two or
more different materials are joined in such a manner that they maintain their identity but
work together to add their strengths and decrease their weaknesses. Composites can be
classified into three categories: (1) Laminates, in which sheets of different materials are
laminated together; (2) particle-reinforced composites, in which particles of one material
are imbedded in a matrix of a second material; and (3) fiber-reinforced composites, in
which fibers of one material are encapsulated in a matrix of a second material. Particle-
reinforced composites can be subdivided into small particle composites, where the par-
ticles are incorporated into the microstructure, such as dispersion-hardened alloys, and large
particle composites, where the matrix simply supports the particles. Fiber-reinforced com-
posites may have continuous versus discontinuous fibers and aligned versus randomly
oriented fibers, which can provide anisotropic versus isotropic properties. Composites
combine all combinations of metals, ceramics, and polymers into MMCs, where a metal
208 Introduction to the Physics and Chemistry of Materials
is the matrix material, ceramic matrix composites (CMCs), and PMCs. Sometimes all three
materials are combined to form hybrid composites.
There is a critical aspect ratio of the reinforcing fibers, which goes as the ratio of the fiber
strength to the strength of the bond between the fiber and the matrix. If the fibers do not
exceed this critical aspect ratio, the matrix will separate from the fiber before the fibers
break and the full strengthening effect of the fibers will not be attained.
Upper and lower bounds on the properties of a particle- or fiber-reinforced composite
can be estimated using the law of mixtures. For aligned composites the upper
bound corresponds to the longitudinal properties and the lower bound to the transverse
properties.
The development of super-strong wires and continuous fibers of boron, graphite, and
polymers such as Kevlar, Spectra, and M-5 have led to families of advanced composites. Of
these, perhaps graphite is the most interesting because of its high thermal conductivity and
negative CTE as well as its strength. Graphite fibers in MMCs, CMCs, and PMCs are being
used as highly efficient thermal conductors and radiators. Their negative CTE allows the
tailoring of composites to match the CTE of other materials as well as the construction of
light, strong, dimensionally stable structures.
Bibliography
Agarwal, B.D. and Broutman, L.J., Analysis and Performance of Fiber Composites, 2nd edn., John Wiley
& Sons, New York, 1990.
Callister, W.D., Materials Science and Engineering: An Introduction, 7th edn., John Wiley and Sons,
New York, 2007.
Clyne, T.W. and Withers, P.J., An Introduction to Metal Matrix Composites, Cambridge University
Press, Cambridge, 1993.
Petersen, R.C., Discontinuous fiber-reinforced composites above critical length, J. Dent. Res., 84(4),
365–370, 2005.
Rawal, S., Metal matrix composites for space applications, JOM 53=4, 2001, 14–17.
Stark, N.M. and Rowlands, R.E., Effects of wood fiber characteristics on mechanical properties of
wood=polypropylene composites, Wood Fiber Sci., 35(2), 167–174, 2003.
Suresh, S., Mortensen, A., and Needleman, A., Fundamentals of Metal Matrix Composites, Butterworth-
Heinemann, Boston, MA, 1993.
Problems
1. Sketch a stress–strain diagram for a composite with a smaller volume fraction of the
fibers to that shown in Figure 10.2 in which the yield strength of the composite is lower
than the fracture strength of the matrix.
2. Write expressions for the yield strength and the tensile strength of the composite in
Problem 1.
11
Phase Equilibria in Single Component Systems
209
210 Introduction to the Physics and Chemistry of Materials
Thermal
arrest
Temperature
Tm
FIGURE 11.1
Typical melting curve showing the thermal arrest. Time
only one degree of freedom. Specifying any one variable automatically fixes the other two
along the solid–vapor line, the liquid–solid line, or the liquid–vapor line on the phase
diagram. If three phases are present, there are no degrees of freedom and three phases can
exist in equilibrium only at a single fixed value of p, V, and T, namely the triple point.
In dealing with materials, we generally operate at constant pressure so there is only one
independent variable. A single component system would then have only one degree of
freedom with a single phase present, and no degrees of freedom with two phases present.
Therefore, a melt can only be in equilibrium with its solid at a fixed temperature, i.e., the
equilibrium melting temperature TM. (Note the emphasis on equilibrium.) Solids can
coexist with their melts indefinitely at the equilibrium melting temperature TM.
G U þ pV TS ¼ H TS: (11:1)
Phase Equilibria in Single Component Systems 211
0.12
0.1
0.08
ΔH/Utot
0.06
0.04
0.02
0
Li Na Mg B Al Ti Cr Mn Fe Co Ni Cu Zn Ga Nd Ru Pd Ag In W Ta Re Pt Au Hg Si Ge
FIGURE 11.2
Ratio of the enthalpy of fusion DH to the binding energy of various crystalline solids. This ratio is only 4% for
most metals.
The Gibbs free energy is constructed such that it will be a minimum when the system is at
equilibrium, or we might say that nature tends to want to minimize the free energy of a
system. Whichever phase has the lowest free energy will be the stable phase.
Let DG ¼ GL GS ¼ DH TDS be the difference in Gibbs energy between the liquid
phase and the solid phase. The change in enthalpy DH ¼ HL HS > 0 since the solid is
more tightly bound than the liquid. At low temperatures where TDS < DH, GS < GL
and the solid is the stable phase. The slope of the free energy of the liquid, @GL=@T ¼
@HL=@T SL T@SL=@T SL since the first and last terms tend to cancel each other.
Similarly, the slope of the solid free energy, @GS=@T ¼ @HS=@T SS T@SS=@T SS. The
free energies of both the solid and the melt decrease with increasing temperatures because
of the TS term, but the free energy of the melt decreases more rapidly because the entropy
of the melt is greater than the solid (SL > SS) since there are more accessible states in the
melt (the molecules are free to move about). Eventually the melt free energy curve crosses
and falls below the free energy curve of the solid at the equilibrium melting temperature
TM, as seen in Figure 11.4, and the liquid becomes the more stable phase at temperatures
>TM. A similar argument applies to the transition between liquid and vapor phase at even
higher temperatures.
212 Introduction to the Physics and Chemistry of Materials
6
Entropy of fusion (cal/mol °C)
0
Li Na Mg B Al Ti Cr Mn Fe Co Ni Cu Zn Ga Nd Ru Pd Ag In W Ta Re Pt Au Hg Si Ge
FIGURE 11.3
Entropy of fusion for various crystalline solids. For most metals, DS is only a few calories per mole.
F
liq
uid
Free energy
Fs
olid
Solid
Liquid
FIGURE 11.4
Tmelt
Schematic of the free energy whose slope becomes
discontinuous during a first-order phase transition. Temperature
Phase Equilibria in Single Component Systems 213
DHDT
DG(DT) ¼ GL GS ¼ DH TDS ¼ DH (TM DT)DS ¼ DTDS ¼ (11:2)
TM
11.3.2 Recalescence
Incipient melting begins when a solid is raised to its equilibrium melting temperature TM.
The reverse is not always true when a melt is lowered to the TM. If there is no crystalline
material in contact with the melt, a nucleation event is required to initiate solidification and
the temperature can fall well below TM before nucleation occurs. (Fahrenheit first noticed
this when he was trying to establish the zero point on his temperature scale.) When
nucleation does occur, the heat of fusion (DH) is released causing the melt to return to
TM before cooling further. This sudden increase in temperature, illustrated in Figure 11.5, is
called recalescence. Recalescence can be observed as a sudden brightening of the melt.
If the DT is so large that the DH is not sufficient to bring the recalescence temperature
back up to TM, the melt is said to be hypercooled. This large undercooling can occur
because of a significant DG barrier to initiating the solidification from a pure melt as shown
214 Introduction to the Physics and Chemistry of Materials
Temperature
TM
ΔT
FIGURE 11.5
Typical cooling curve during solidification of a metal
showing the recalescence as the release of latent heat
causes a brief increase in temperature. Time
in Figure 10.7. It should be noted that solidification of an undercooled melt after recales-
cence can be very rapid since it has already given up its latent heat.
where VL and VS are the molar volumes of the liquid and solid.
At equilibrium, dGL ¼ dGS. Subtracting the above equations yields
ðVL VS Þdp ðSL SS ÞdT ¼ 0, (11:5)
dTM ðVL VS Þ TM 0
¼ ¼ ðVL VS Þ: (11:6)
dp ðSL SS Þ DH
You should recognize this as the Clapeyron equation, which gives the slope of the melting
0
curve at some reference TM , which in this case is taken as the equilibrium melting point of a
planar surface. Since DH > 0, materials that contract upon freezing will have dTM=dp > 0.
Materials with open structures such as group IV materials with diamond structure (e.g.,
Ge, Si), group III–V and II–VI compounds with zinc blende or wurtzite structures (e.g.,
GaAs, ZnS, etc.), and, of course, H2O, expand when they freeze; hence dTM=dp < 0.
FIGURE 11.6
Hoop stress in a sphere with a pressure difference.
A small solid sphere in its melt also feels a similar pressure difference except g now
refers to the interfacial energy between the solid and melt. Just as in the case of surface
tension, this interfacial energy comes about because of unsatisfied or partially satisfied
bonds at the interface which give rise to a net radial inward force on the surface atoms.
From Equations 11.3 and 11.4, the differential Gibbs energies can be written
2gVS
dGS ¼ SS dT þ VS dp ¼ SS dT þ (11:7)
r
and
Note that the pressure term is absent in the liquid because the pressure is only felt in the
solid. Again equating the DG terms to obtain phase equilibrium, we find
VS dp 2gVS 2gTM0
VS
dTM ¼ ¼ ¼ : (11:9)
ðSS SL Þ rðSS SL Þ rDH
Since DH ¼ HL – HS > 0, we find that the local melting temperature for the solid sphere is
lowered by an amount that is inversely proportional to its radius. This lowering of the local
melting point by interfacial curvature is sometimes known as the Gibbs–Thompson effect.
Notice that it does not depend on whether the material expands or contracts when
freezing. In either case, the effect of positive curvature is to lower the melting point.
Furthermore, we see that the equilibrium melting temperature only applies to a planar
interface. Small particles will therefore melt before larger ones and larger particles will
grow at the expense of the smaller particles in a constitutionally undercooled melt (a melt
whose local freezing temperature is lower than the local temperature because of compos-
itional differences). This phenomenon, known as Ostwald ripening, is important in grain
growth and precipitation hardening and is discussed in greater detail in Chapter 13. The
lowering of the melting temperature with curvature plays an important role in sintering by
lowering the melting point at the sharp corners of the grains which allows fusion of the
material to take place below the normal melting temperature.
It is also important to understand the difference between the lowering of the melting
temperature predicted by Equation 11.9 and that predicted by the Clapeyron equation
(Equation 11.6). The Clapeyron equation applies when both the solid and the liquid phase
are under the same pressure. It predicts a raising of the freezing temperature with pressure
of a substance such as a metal that contracts when it freezes. Equation 11.9 applies when
only the solid is subjected to pressure such as occurs at the point of contact between
216 Introduction to the Physics and Chemistry of Materials
powdered metal grains as they are being compressed by a process known as hot isostatic
pressing (HIP). This local pressure, applied only to the solid particles, depresses the local
freezing point below the ambient temperature and causes localized melting. Similarly, the
lowering of the local melting point may play a role in friction or stir welding in which a
rotating tool is pressed with great force against the surfaces to be joined.
0
2gTM VS
rcritical ¼ : (11:10)
DHDT
4 DG
DG*(r) ¼ pr3 þ 4pr2 g, (11:11)
3 VS
where DG refers to the difference in Gibbs energy per unit volume between the melt and
solid for a plane interface and the second term is the added surface energy. From Equation
11.2, this can be written as
4 DHDT
DG*(r) ¼ pr3 0 þ 4pr2 g: (11:12)
3 TM VS
Figure 11.7 is a typical plot of DG*(r) as a function of r. The function initially rises because
of the positive r2 term, but the negative r3 term eventually dominates and reduces DG*(r).
1
ΔG∗(r)/ΔG∗
FIGURE 11.7
Barrier to nucleation. Embryonic nuclei with r < rcritical −1
decrease their free energy by melting; hence are not viable. 0.0 0.5 1.0 1.5
Nuclei with r > rcritical decrease their free energy by grow-
r/rcritical
ing; hence become stable.
Phase Equilibria in Single Component Systems 217
The peak DG* at rcritical is the barrier to nucleation. An embryonic nucleus with r < rcritical can
reduce its free energy by dissolving, hence it is not viable. On the other hand, if r > rcritical, the
embryonic nucleus can reduce its free energy by growing and becomes a stable nucleus.
Differentiating Equation 11.11 with respect to r to find rcritical yields
d DHDT
DG*(r) ¼ 4pr2 0 þ 8prg ¼ 0, (11:13)
dr TM VS
from which
0
2gTM VS
rcritical ¼ , (11:14)
DHDT
which is the same result as found previously in Equation 11.10 using a different but equiva-
lent argument. Putting this value for rcritical back into Equation 11.6 to find the barrier to
nucleation,
0 2 2
16 pg3 TM VS
DG* ¼ : (11:15)
3 DH 2 DT2
where I0 contains the number of atoms per unit volume, an activation energy for an atom
to move from one site to another, and some frequency term. There have been several
methods proposed in an attempt to derive I0.
David Turnbull, a pioneer in the study of nucleation in metals, estimated
I0 ¼ N ðkT=hÞ expðDF=kT Þ, where N is the number of atoms=cm3, the kT=h is an estimate
of the fluctuation frequency based on the uncertainty principle, and DF is the transport
activation energy. According to Turnbull, exp(DF=T) is the on the order of 102 at typical
solidification temperatures. Altogether Turnbull estimates I0 ¼ 10331 cm3 s1.
Turnbull conducted extensive undercooling experiments on very small (10 mm in
diameter) metallic droplets and observed that the maximum undercooling was 20% of
the Tm for a large number of metals. He further observed a correlation between the
interfacial energy and the heat of fusion if the interfacial energy is expressed as a molar
quantity, suggesting that the energy of an atom at the solid–liquid interface has a certain
fraction l of the difference in enthalpy between the two states or g m ¼ lDH.
The molar interfacial energy gm can be related to g by setting gm=A ¼ g, where A is the area
of a mole of atoms with a thickness of one atomic layer. Let d be the diameter of the atom in
question. The molar volume is Vs ¼ NAd3 and A ¼ NAd2 from which A ¼ NA1=3V2=3 where NA
is Avogadro’s number. Turnbull found l ¼ 0.45 for most metals, and ¼ 0.32 for semimetals
and water, generally systems with a more open structure.
Substituting this relationship for g into Equation 11.15,
3
1=3 2=3 0 2 2
16 p lDHN A V s TM VS 16 pl3 DHT 02
DG* ¼ ¼ M
: (11:17)
3 DH 2 DT 2 3 NA DT2
218 Introduction to the Physics and Chemistry of Materials
where Q is T=TM 0
. As discussed previously, the entropy of fusion for most metals is
approximately 2.3 cal=K mol. Thus we can write a general Boltzmann factor for the
nucleation of metals as
DG* 16 p 0:453 2:3 1:767
¼ ¼ (11:19)
kT 3 R Qð1 QÞ2 Qð1 QÞ2
0 0 1=3
2gTM VS 2lVS TM
rcritical ¼ ¼ : (11:20)
DHDT NA DT
1=3
Using Turnbull’s estimate for the pre-exponential term to be 10331 in cm3 s1, the
nucleation rate is given by
!
1:767
IV ¼ 10 331
exp 2
cm3 s1 : (11:21)
Qð1 QÞ
Flemings estimates I0 by starting with the number of critical nuclei n* in equilibrium with
N atoms n* ¼ N exp(DG*=KB T). He then takes the number of atoms adjacent to a critical
nucleus, 4p(rcritical)2=a02 and estimates the nucleation rate as the rate at which these atoms
join the critical nucleus to push it over the energy barrier. The rate at which the adjacent
atoms join a critical nucleus is given by DL=a02, where DL is the liquid diffusion coefficient
and a0 is an atomic radius. If one takes N 1022 atoms cm3, rcritical=a0 4, and
DL 104 cm2 s1; 4p(rcritical)2=a02 102, DL=a02 1011 s1 and I0 1035 cm3 s1.
Kingery et al. suggest a model similar to Flemings’ except they estimate the frequency
that atoms cross from the melt to the embryonic solid as kT=3pa03h, where h is the
absolute viscosity. For typical metallic melts, h 0.01 g cm1 s1. For T 1000 K, the
jump frequency DL=a02 1011 s1 which is the same order of magnitude as estimated
by Flemings.
As it turns out, the pre-exponential term is relatively unimportant because the nucleation
rate is dominated by the exponential term in Equation 11.21. As Q increases past 0.8, the
absolute value of the argument increases rapidly, presenting a very large barrier to
nucleation. The exponential of a very large negative number becomes so small that the
nucleation rate is nil regardless of the pre-exponential factor.
Perhaps the best way to represent the nucleation rate is through the use of Poisson
statistics. For a Poisson distribution, the probability P(n) of observing n events, when the
expected number of events is nexp, is given by
nnexp exp(nexp )
P(n) ¼ : (11:22)
n!
Phase Equilibria in Single Component Systems 219
1
Probability of no nucleation events
0.8
0.6
0.4 0.01 s
0.1 s
1.0 s
(Poisson statistics apply when nexp, is small. As nexp becomes large, the Poisson distribution
becomes a Gaussian distribution.) We can now use Equation 11.22 to compute the prob-
ability of observing no nucleation event in a volume V (cm3) after t s.
" !#
1:767
P(0) ¼ exp½IV Vt ¼ exp 10 331
Vt exp : (11:23)
Qð1 QÞ2
Turnbull used very small drops to maximize the chance of eliminating impurities that
could serve as nucleation sites. Assume a 5 mm drop. Taking the pre-exponential term as
1033, the probability of observing no nucleation events in 0.1 s as a function of the
dimensionless temperature is shown in Figure 11.8. From this plot, it can be seen that
half of the drops would nucleate at Q ¼ 0.7895 or undercool by 21.05%. Practically all of
the drops would undercool by at least 20% and virtually none of the drops would under-
cool by more than 21.5%. To illustrate the insensitivity of the degree of undercooling to
the pre-exponential factor, curves were also plotted for t ¼ 0.01 and 1.0 s. Varying the pre-
exponential factor by two orders of magnitude only shifts the undercooling for P(0)
from 0.784 to 0.795 or 0.7%. In other words, a two-order magnitude variation in the
pre-exponential factor only shifts the expected amount of undercooling from 20.5% to
21.6% of the melting temperature.
Equation 11.23 can also be used to find the average time a given droplet size can be held
at a given Q before it will be nucleate homogeneously.
!
ln(0:5) 0:693 10331 1:767
tave ¼ 331 ¼ exp : (11:24)
10 V V Qð1 QÞ2
Figure 11.9 illustrates the time before nucleation as a function of undercooling for different
size molten droplets using Turnbull’s theory. Droplets of virtually any size can be held
more or less indefinitely at DT=T ¼ 0.9% or 10% undercooling. However, the time before
nucleation decreases rapidly as the undercooling approaches 20%, especially for larger
droplets. It became generally accepted that the maximum undercooling that could be
achieved in a melt was 20% of the melting temperature and this became known as the
Turnbull limit for undercooling.
220 Introduction to the Physics and Chemistry of Materials
1.0
0.9
1 − ΔT/T 0.8
0.7
5 cm
0.6 0.5 mm
5 μm
0.5
−40 −20 0 20 40 60 80
Log time (s)
FIGURE 11.9
Time before probability of nucleation reaches 0.5 for 5 mm droplets (solid), 0.5 mm (dashed), and 5 cm (dotted)
computed from Turnbull’s model. Any size droplet can be undercooled by 10% more or less indefinitely. Only
very small droplets can be undercooled by 20%.
q Melt
Solid
FIGURE 11.10 r
Molten sessile drop making contact angle u with
a solid surface.
Phase Equilibria in Single Component Systems 221
Since a melt completely wets its own solid, the contact angle goes to zero and the barrier
to nucleation disappears once the solid has formed. Nucleation can be inhibited by
surrounding the melt with nonwetting surfaces. As the contact angle approaches 1808, the
f(u) approaches unity and the nucleation barrier approaches the DG* for homogeneous
nucleation.
undercooled state. A similar facility was later constructed to operate in the beam line of the
Advanced Photon Source at Argonne National Laboratory where it is being used to study
the structure of materials during the nucleation process.
11.7 Summary
In dealing with a single component system, the Gibbs phase rule tells us the melting point
is a function of pressure only. Since most of the processes we will be dealing with take
0
place at ambient pressure we will concern ourselves with TM , the temperature at which a
plane front solid can remain indefinitely with its melt.
The Gibbs free energy is defined as G ¼ H – TS. Nature always strives to minimize the
free energy of a system. The liquid phase has a higher (less negative) enthalpy than a solid
because bonds have been broken. It also has higher entropy than a solid (more disorder,
more available states). At low temperatures, the lower enthalpy of the solid outweighs the
larger TS term of the liquid and the solid is the stable phase. As the temperature is
increased, the TS of the melt eventually dominates and liquid becomes the stable phase.
0
The TM occurs when the Gibbs energy of the solid equals that of the melt.
There is a discontinuity in the slope of the free energy when plotted against temperature
at the melting point as the solid free energy curve goes over to the steeper liquid free
energy curve. A phase transition in which there is a discontinuity in the first derivative of
the free energy is called a first-order phase transition and is accompanied by a release (or
absorption) of the enthalpy of fusion. A second-order phase transformation has a discon-
tinuity in the second derivative of the free energy and there is a discontinuity in the heat
capacity rather than in the enthalpy.
If both the solid and melt are under the same pressure, the dependence of the melting point
with pressure is determined by the Clapeyron equation and depends on whether the solid
expands or contracts on freezing. However, if only the solid is under pressure, the melting
point is always lowered by increasing the pressure on the solid. A solid with a convex solid–
liquid interface is under pressure due to its interfacial energy; therefore, the local melting
point is lowered (Gibbs–Thompson effect). Very small (nanometer-size) particles can have
0
melting points well below the TM of a plane face solid. These effects become important in
understanding grain growth and other ripening phenomena as well as in nucleation theory.
When a heated crystal reaches its melting temperature, its temperature remains constant
until the enthalpy of fusion is absorbed and the crystal is completely melted before the
temperature will continue to rise, causing a thermal arrest. When a melt is cooled to its
melting temperature, it does not immediately solidify. In order to drive solidification, the free
energy of the solid must be lowered, which is generally done by lowering the temperature
below the melting point, which is called undercooling. Very little undercooling is required
for a melt in the presence of its solid, but in the absence of a crystalline particle or surface to act
as a nucleation site, melts can be undercooled by as much as 33% of their absolute melting
temperature before nucleation occurs. When nucleation does occur, the release of the latent
heat (enthalpy of fusion) brings the melt back to the melting temperature before additional
cooling can occur. This sudden increase in temperature is called recalescence.
Even though the release of latent heat raises the temperature of an undercooled melt
back to the melting point, this heat has already been removed so that solidification is very
rapid. Therefore, the ability to undercool a melt is very useful in developing rapid
solidification processing for producing metastable states or other novel microstructures.
In the absence of any nucleation site, homogeneous nucleation will eventually occur as
the atoms cluster together to form embryonic nuclei. There is a critical radius for these
Phase Equilibria in Single Component Systems 223
nuclei to become viable and grow into a solid, which is obtained from the Gibbs–
Thompson equation. The free energy required to create a viable nucleus forms a nucleation
barrier. Thus homogeneous nucleation is a stochastic process in which we can only
estimate the probability of a thermal fluctuation large enough to surmount the nucleation
barrier for a cluster of atoms to form a viable nucleus.
Several different models have been proposed for estimating the nucleation rate and
Poisson statistics can be used to estimate the average amount of undercooling one would
expect for a given system, or one can estimate the length of time a given system can be held
before nucleation occurs. These models contain simplifying assumptions and have large
uncertainties, but still are useful in developing processes for glass formation where it is
necessary to undercool bulk melts to the point where the viscosity of the melt becomes high
enough to prevent crystal formation.
Nucleation events are found in many aspects of materials processing besides the solidi-
fication of a melt. A nucleation event is required for precipitates to form in precipitation-
hardened alloys, for the decomposition of immiscible systems, for new grains to form in a
casting, for graphite flakes to form in cast iron, phase selection in alloy solidification and in
crystal growth, etc. Understanding of the nucleation process is key to being able to control
it in order to achieve the desired result in a process.
Bibliography
Bayuzick, R.J. et al., Review on long drop towers and long drop tubes, Collins, E.W and Koch, C.C.,
Eds., Undercooled Alloy Phases, ASM International, 1986, 207.
Chalmers, B., Principles of Solidification, Robert E. Krieger, Huntington, New York (Reprinted by
arrangement from 1964 version published by John Wiley & Sons).
Flemings, M.C., Solidification Processing, McGraw Hill, New York, 1974 (McGraw Hill Series in
Materials Science and Engineering).
Gocken, N.A. and Reddy, R.G., Thermodynamics, 2nd edn., Springer, Berlin, 1996.
Ragone, D.V., Thermodynamics of Materials, Vol. II, John Wiley & Sons, New York, 1995.
Rathz, T.J., et al., The Marshall space flight center drop tube facility, Rev. Sci. Instrum., 61=12,
1990, 3846.
Rhim, W.K., An electrostatic levitator for high-temperature containerless materials processing in 1-G,
Rev. Sci. Instrum., 64=10, 1993, 2961–2970.
Schmelzer, J.W.P., Ed., Nucleation Theory and Applications, Wiley-VCH, Weinheim, 2005.
Turnbull, D., Formation of crystal nuclei in liquid metals, J. Appl. Phys., 21, 1950, 1022–1027.
Problems
1. We generally ignore the atmospheric pressure when dealing with materials. (a) Find the
difference between the melting point of Fe at 1 atm (atmospheric pressure 1 bar ¼ 105
Pa) and in a vacuum. (b) What would be the melting point of Fe deep inside the Earth
where the pressure is 5 Mbar. Assume the solid is immersed in its melt. (c) How would
we know if the Fe was molten or not at that pressure?
2. Powdered metals are sintered together using a hot isostatic press (HIP) typically oper-
ating at 2 kbars. If 1 mm Fe particles are pressed at 2 kbars, how much would their
melting temperature be lowered by the applied pressure? (Assume the particles have
only solid–solid contact.) How much would the local melting point be lowered by their
tip radius?
224 Introduction to the Physics and Chemistry of Materials
3. Use the Poisson probability distribution to find the number of times you would expect
heads (or tails) to occur in 100 coin tosses.
4. How many times can you expect to roll a pair of dice without a 7 showing up?
5. Calculate the critical radius of an Fe nuclei in a melt that is undercooled by 0.33 of its normal
melting temperature. How many atoms would be contained in such a cluster?
6. Show volume of a spherical cap sitting on a surface with contact angle u is given
that the
by 4pr3 =3 ð1=4Þ½ð2 þ cos uÞð1 cos uÞ.
12
Phase Equilibria in Multicomponent Systems
Phase diagrams are the basic maps required for determining the melting points of alloys or
compounds and for designing processes to obtain a desired composition of a system to be
solidified from the melt. In this chapter we will develop not only a working knowledge of
how to use phase diagrams but also an understanding of how the various phase diagrams
evolve from free energy considerations.
N! N!
W¼ ¼ : (12:1)
NA !NB ! (N NB )!NB !
225
226 Introduction to the Physics and Chemistry of Materials
0.8
0.6
Smix/Nk
0.4
0.2
0.0
0.0 0.2 0.4 0.6 0.8 1.0
FIGURE 12.1
Dimensionless entropy of mixing vs. mole fraction. A Mole fraction, x B
Since the Ns are very large, the factorials can be represented by Stirling’s approximation, ln
x! x ln x x. The entropy of mixing can then be expressed as
N! N NB N NB
Smix ¼ k ln ¼ k NB ln N ln : (12:2)
(N NB )!NB ! NB N
It is more convenient to work with the mole fraction x NB=N. Introducing this in the
above expression,
1x
Smix ¼ Nk lnð1 xÞ þ x ln ¼ Nk½ð1 xÞ lnð1 xÞ þ x lnðxÞ: (12:3)
x
This function is symmetrical about x ¼ 0.5 as may be seen in Figure 12.1. Note also that the
slope of S approaches infinity at x ¼ 0 and minus infinity at x ¼ 1 (which can easily be
verified by differentiating Equation 12.3). This behavior has significant implications in the
contamination of melts by their containers (see Figure 12.18).
p
UA ¼ NA ½(1 x)uAA þ xuAB , (12:4)
2
where p is the number of nearest neighbors. The factor 1=2 is included since each bond is
shared between two atoms. Similarly, the total energy for the B atoms is
p
UB ¼ NB ½(1 x)uAB þ xuBB : (12:5)
2
Phase Equilibria in Multicomponent Systems 227
e = −uAA/2
− 0.5
e=0
U0/Np
−1.0
e = uAA/2
FIGURE 12.2
−1.5
0 0.2 0.4 0.6 0.8 1 Heat of mixing from Equation 12.8 for uAA ¼ 2U0=Np,
uBB ¼ U0=Np, and « ¼ uAA=2 (dashed line), ¼ 0.0 (solid
A Atomic fraction, x B line), and ¼ uAA=2 (dotted line).
Since the number of A atoms is N (1 x) and the number of B atoms is Nx, the total energy is
p
Utot ¼ N (1 x)2 uAA þ x2 uBB þ 2x(1 x)uAB : (12:6)
2
Note that if uAB is the average of uAA and uBB, Equation 12.6 reduces to
p
Utot ¼ N ½(1 x)uAA þ xuBB : (12:7)
2
If uAB is written as huAA þ uBBi=2 þ «, where « is the excess heat of mixing per atom,
Equation 12.6 can be written as
p
Utot ¼ N ½(1 x)uAA þ xuBB þ 2x(1 x)«: (12:8)
2
Note that Equation 12.8 plots as a straight line in Figure 12.2 for « ¼ 0 (no excess heat of
mixing), curves upward for « > 0 (positive excess heat of mixing), and curves downward
for « < 0 (negative excess heat of mixing).
Np
F(x,T) ¼ ½ð1 xÞmAA þ xmBB þ 2xð1 xÞ« þ NkT ½(1 x) lnð1 xÞ þ x lnðxÞ: (12:9)
2
The free energy per atom as a function of temperature is shown in Figure 12.3 for different
excess energies. Taking the second derivatives of Equation 12.9
@ 2 F(T,x) NkT
¼ 2Np« þ : (12:10)
@x2 x(1 x)
228 Introduction to the Physics and Chemistry of Materials
−0.6
.5
=0
T / T C .8
−0.8 =0
T / T C
=1
T/T
C
F/NkT
−1.0
=2
TC
T/
−1.2
FIGURE 12.3
Free energy plot for various values of T=TC, where the −1.4
consolute temperature TC is defined by Equation 12.12.
The mixed phase is stable for all x if T ¼ TC (solid line)
0.0 0.2 0.4 0.6 0.8 1.0
but the system will become segregated if T < TC as will
be explained in Section 12.4.1. A Atomic fraction, x B
The curvature of the free energy will continue to be positive (curve upwards) so long as
kT 2p«x(1 x) 0 (12:11)
and the mixed phase will be stable over the entire range of x. Clearly this will always be the
case if the excess energy « < 0 (the atoms prefer their counterparts to their own kind).
However, if « 0, the free energy curve may exhibit an inflection point if Equation 12.11
becomes 0. The lowest temperature where this can happen is called the consolute tem-
perature TC, where for this system
Also note that no matter how large the heat of mixing may be, the initial free energy
curves always have an initial drop at x ¼ 0 and 1. This is because the entropy of mixing
has an infinite slope at x ¼ 0 and 1 as was shown previously. This implies that no matter
how dissimilar the two components may be, there will always be a limited region of solid
phase solubility, even though it may be vanishingly small. This will have significant
consequence in the purification of materials, especially when we later discover that
certain trace impurities at concentrations far less than our ability to detect chemically
(less than parts per billion) can have a dramatic influence of the electrical properties of
semiconductors.
12.4.1 Miscibility
Systems in which the bonding between atoms A and B are different, e.g., a polar substance
such as water and a nonpolar substance such as oil, will have « > 0. Solids in which the
A atoms have substantially different sizes than B atoms will not form bonds between each
other that are as strong as bonds between themselves; hence « > 0. However in these cases,
a single phase will form at T > TC because of the lowering of the free energy due to the TS
term, but for T < TC there will be a range of compositions over which the system will
separate into two phases. This region is termed as a miscibility gap and can be present in
either the solid or the liquid phase, or both.
Phase Equilibria in Multicomponent Systems 229
−0.95
f = F/N (a.u.)
−1.00
FIGURE 12.4
Free energy curve at some arbitrary tempera-
ture for a binary system with an immiscibility
gap. The free energy of the system is minimized
−1.05 when the system separates into two phases, one
with x ¼ C1 (a phase) and the other with x ¼ C2
(b phase). The compositions are determined by
constructing a line that is mutually tangent to
−1.10 the lowest regions of the free energy curve. The
0.0 0.2 0.4 0.6 0.8 1.0 compositions of the two phases correspond to
C1 x C2
the x-values of the two tangent points.
where
Na and Nb are the number of atoms in the a and b phases, respectively
f(xa) and f(xb) are the respective values of the free energy curve at xa and xb
We shall proceed to show that the free energy of the segregated system is lower than the
free energy of the homogeneous system in the miscibility gap region. We shall also
demonstrate that the segregated system with the composition found by the method of
tangents is in equilibrium. But first we have to find the amount of the two phases present,
namely Na and Nb.
xβ − xα
x0 − xα xβ − x0
FIGURE 12.5
Illustration of the lever law. The fulcrum
represents the starting composition Nα Nβ
x0N ¼ number of B atoms in the system. Na
is the number of atoms in the a phase and
xaNa is the number of B atoms in the a
phase. Likewise, Nb is the number of
0 xα x0 xβ 1
atoms in the b phase and xbNb is the num-
ber of B atoms in the b phase. Composition, x
we first specify the average composition x0 of the system. Since Na þ Nb ¼ N, and xaNa þ
xbNb ¼ NB (total number of B atoms) ¼ x0N, we can solve these two equations simultan-
eously to obtain
N a xb x0 N b x0 xa
¼ , ¼ : (12:14)
N xb xa N xb xa
This simple but important result is known as the lever rule because it has the same form as
the equation for balancing a lever arm as shown in Figure 12.5.
Assume the left end of the lever arm is located at xa, the right end is at xb, and x0 is the
fulcrum. If x0 is closer to xa than to xb, Na must be greater than Nb to balance the system.
The lever rule also applies to phase diagrams in weight fraction or wt% if the xs in
Equation 12.14 are expressed in terms of weight fraction or wt% and mole fractions Na=N
and Nb=N are replaced by weight fractions Wa=W and Wb=W, respectively.
Putting these results into Equation 12.13, the free energy of the segregated system
becomes
_ F 1
f (x) ¼ ¼ xb x f ðxa Þ þ ðx xa Þf xb , xa x xb : (12:15)
N xb x a
This function is linear in x and intersects f(xa) at xa and intersects f(xb) at xb, which is just
the mutual tangent curve we constructed previously in Figure 12.5. Since the free energy
of the segregated system f (x) is lower than the homogeneous system f(x) between xa and
xb, the stable configuration of the system will be two distinct phases with compositions xa
and xb.
even though, strictly speaking, the functions are not continuous, hence are not differentiable.
For the system to be in equilibrium, each component must have the same chemical potential
in the coexisting phases; otherwise, the free energy could be reduced by a net migration of
components between phases. To show that the above system is indeed in equilibrium with
compositions xa and xb, we must show that mA(xa) ¼ mA(xb) and mB(xa) ¼ mB(xb).
To proceed, we introduce f ¼ F=N, where f is the average free energy per particle as used
in Figure 12.5. For a two-component system
@Nf @f @f @x @f x
mA ¼ ¼f þN ¼f þN ¼f þN ,
@NA T,p,NB @NA @x @NA @x N
which reduces to
@f
mA ¼ f x : (12:17)
@x
@f
mB ¼ f þ (1 x) : (12:18)
@x
f 0 ¼ mB mA : (12:19)
Since f 0 (xa) ¼ f 0 (xb) because of the mutual tangent line, mB(xa) mA(xa) ¼ mB(xb) mA(xb),
from which mB(xa) ¼ mB(xb) and mA(xa) ¼ mA(xb), QED.
0.6 TC
0.7 TC
0.8 T
TC
Free energy
0.9 T
TC
TC
0.9 TC
0.8 TC
L1 + L2 Spinoid
0.7 TC
0.6 TC
FIGURE 12.6
Phase diagram (left hand side) and spinodal (right hand side) for a completely immiscible system derived from
free energy considerations. Mixtures cooled into the spinodal separate spontaneously whereas mixtures between
the spinodal and the region of immiscibility are metastable and require a nucleation event to separate.
positive, a nucleation barrier exists that must be overcome in order to form the segregated
phase. However, in the region between the two inflection points where the second deriva-
tive of the free energy is negative, there is no barrier to nucleation and decomposition
occurs spontaneously. This spontaneous decomposition is known as spinodal decompos-
ition and can occur in the solid as well as in the liquid phase.
The spinodal is located within the miscibility gap and can be mapped by projecting the
points of inflection of the free energy curves (where @ 2F=@x2 ¼ 0) on to their corresponding
isotherms as shown in Figure 12.6.
Given the similarity of atomic radii between the 4d and 5d transition metals, one would
expect those pairs in the same column to be isomorphous. This is the case for Hf–Zr, Mo–W,
Ru–Os, and Au–Ag. Si–Ge is also isomorphous as are several compound systems such as
HgTe–CdTe, Al2O3–Cr2O3 that exhibit pseudobinary isomorphous phase diagrams.
Some mixed valency systems such as Cu–Ni and Ag–Pd are isomorphic. Other systems
meet all of the Hume–Rothery criteria but are immiscible in the liquid state (e.g., Nb–Ta) or
are completely miscible in the melt but decompose spinodally in the solid state (e.g., Pd–Pt).
One must conclude that there is nothing hard and fast about the Hume–Rothery rules and
they should be used only as a guide to estimating which systems may be isomorphous.
We will now set out to compute the free energies for the solid and liquid phases of an
ideal isomorphous binary system and the resulting phase diagram using a very simple
model. If we add the entropy of mixing (Equation 12.3) to the chemical potentials of pure
A and pure B solid m0A and m0B , we can write the free energy of the solid as
where x is the mole fraction of component A. The free energies of the pure liquids are
m0A þ DSðTA T Þ and m0B þ DSðTB TÞ, where TA and TB are the melting temperatures of
A and B, respectively. Taking the pure solid A and B and the standard states, we may set
their chemical potentials to 0. Assuming no excess heat of mixing, the free energies of the
solid and liquid then become
and
Taking DS ¼ 2.3 cal mol1 ¼ 9.6 J mol1, TA ¼ 1683 K, and TB ¼ 1210 K (melting points for
Si–Ge), the free energy curves were calculated for different temperatures in Figure 12.8. At the
top are the free energies at 1210 K. Note that the FS and FL meet at x ¼ 1 (pure Ge) and that the
FS curve is always lower than the FL curve, indicating that the solid phase is stable below
1210 K for all x. As the temperature is increased above the melting point of Ge, notice that the
FL curves drop because of the TDS term and move through the FS curve until at T ¼ 1683 K,
the FL curve is always lower than the FS curve (except where they meet at 1683 K), indicating
that the liquid phase is present at temperatures above 1683 K for all x. At each temperature
between the two melting points, mutual tangents can be drawn between the two free energy
curves near the point of their intersection. The x-values of these points of intersection can be
projected onto a T versus x plot, which maps out the phase diagram for the Si–Ge binary alloy
seen in Figure 12.7. This may be compared to the actual Ge–Si phase diagram in Figure 12.8.
0
T = 1210
−5
−5 T = 1400
−5
T
T ==1500
1500
−10
0
T
T ==1683
1683
−5
−10
FIGURE 12.7
1683
Temperature (K)
C3 C0 C0 C2
NS ¼ N ; NL ¼ N : (12:23)
C3 C2 C3 C2
Phase Equilibria in Multicomponent Systems 235
1400
1300
L
Temperature (°C)
1200
Solid + liquid
(GeSi)
1100
1000
900
0 10 20 30 40 50 60 70 80 90 100
Ge At% Si Si
FIGURE 12.8
Phase diagram for the Ge–Si system. The Ge is on the left in this diagram and the liquid–solid two-phase region is
somewhat fatter than in the calculated case, possibly due to small departures from ideality of the system. (Reprinted
from Massalski, T.B., Handbook of Binary Alloy Phase Diagrams, Vols 1, 2, and 3, ASM, 1990. Reprinted with
permission of ASM International. All rights reserved.)
TA
Liquidu
T1 s
T2
L+S
T3
Solidu
Temperature
s
TB
A C1 C2 C0 C3 C4 B
Composition, x
FIGURE 12.9
Equilibrium solidification process in a solid solution system. When a melt with composition C0 is lowered to T1,
the first-to-freeze component will have composition C1. The partition coefficient k ¼ C1=C0. At T2, solid with
composition C2 will be in equilibrium with liquid whose composition is C3. Finally at T3, solid with composition C0
will be in equilibrium with the vanishingly small amount of the last-to-freeze liquid whose composition is C4. But
see Figure 13.3.
236 Introduction to the Physics and Chemistry of Materials
Instead of using the C0 as the fulcrum, it is often convenient to use the end points of the tie
line as a pivot point. For example, taking the pivot point at C2 for T2, the mole fraction of
liquid in equilibrium with the solid can be found by balancing (C0 C2) times N against
(C3 C2) times NL resulting in
NL C0 C2
xL ¼ ¼ : (12:24)
N C3 C2
It is assumed here that solidification is being carried out as a series of equilibrium states in
which the solid continually adjusts through solid-state diffusion to the value correspond-
ing to the intersection of the tie-line with the solidus. Eventually as the temperature falls to
T3, the final solid will have a uniform composition equal to the starting value.
T = 2250
Free energy
T = 1900
T = 1683
Liquid
2250
a+L
1900
b+L
1683
Solid solution
Temperature (K)
1210
FIGURE 12.10
Hypothetical phase diagram for a system similar to Ge–Se
but in which the solid excess heat of mixing was 30 kJ
mol1. This high negative excess heat of mixing in the
solid pushes the solidus line toward higher temperatures,
0 which in this case allows congruent melting=solidifica-
A Mole fraction B tion to occur at the point of highest temperature.
energy term x(1 – x)« is added to the solid free energy and « was set to 30 kJ mol1. Notice
that the melting point was raised above the melting points of the pure components and
that the two-phase region collapsed at one point where the system melts congruently. Only
at the composition corresponding to a congruent melting point, can melting and freezing
take place without a change in composition.
A similar but opposite effect will occur if the solid had a higher excess energy than the
liquid. In this case the melting points will be below those of the pure components.
1800
1600 L
1400
AlNi
Temperature (⬚C)
1200 (Ni)
1000
800
Al3Ni2
600
Al3Ni
Al Al3Ni5 AlNi3
400
0 10 20 30 40 50 60 70 80 90 100
Al At% Ni Ni
FIGURE 12.11
Al–Ni phase diagram. Intermetallic phases Al3Ni, Al3Ni2, AlNi, Al3Ni5, and AlNi3 are seen. Note that the
intermetallic AlNi phase has a higher melting point than either Ni or Al. This stoichiometric Al-50 At % Ni
phase has a broad range of existence (45%–58% Ni) at ambient temperature. (Reprinted from Massalski, T.B.,
Handbook of Binary Alloy Phase Diagrams, ASM, 1990. Reprinted with permission of ASM International. All rights
reserved.)
2800
Liquid
MgAl2O4 (ss) + Al2O3
Liquid +
2400 Liquid
MgO (ss)
Temperature (°C)
+
Liquid
MgO
2000 (ss)
MgAl2O4
(ss)
MgO (ss)
1600 +
MgAl2O4 (ss) MgAl2O4
(ss)
+
1200 Al2O4
0 20 40 60 80 100
MgO Composition (wt% Al2O3) Al2O3
FIGURE 12.12
MgO–Al2O3 phase diagram. In this case the composition in wt% is along the bottom and At% is along the top. The
compound spinel (MgAl2O3) forms at 50 At% and has a wide coexistence region near the melting temperature because
of vacancy defect formation (see Problem 12.6). (From Halsedt, B., J. Am. Ceramic Soc., 75=6, 1992. With permission.)
Phase Equilibria in Multicomponent Systems 239
1200
1098°C
1100
L
1000
900
Temperature (ºC)
800
αCdTe
700
L + αCdTe
L + αCdTe
600
500
446.2°C 449.57ºC
400
321.108ºC 321°C
300 Te + αCdTe (Te)
Cd + αCdTe
(Cd)
200
0 10 20 30 40 50 60 70 80 90 100
Cd At% Te Te
FIGURE 12.13
Cd–Te phase diagram. The line compound at 50 At% is CdTe. (Reprinted from Massalski, T.B., Handbook of Binary
Alloy Phase Diagrams, ASM, 1990. Reprinted with permission of ASM International. All rights reserved.)
line compound is formed, the phase diagram is sometimes drawn with this compound as one
of the endpoints on the composition axis as shown in Figure 13.2 in which the Fe–C phase
diagram is shown for compositions ranging from pure Fe to the compound Fe3C (cementite).
The free energy of an intermediate phase becomes much narrower in the vicinity of its
stoichiometric composition, as illustrated in Figure 12.14a and b, because only then can the
stronger ionic or covalent bonds really form. The free energy configuration for such a
system must have a sharp wedge-like shape shown in Figure 12.14a so that when tangents
are drawn from it to the liquid free energy, they meet at a point. The solid free energy curve
for a system that produces a solid solution intermediate phase would have a more blunt or
rounded nose so there is some space between the intersections of the tangents as illustrated
in Figure 12.14b.
T1
T1
L
L L L
γ γ
T0 T0
T1 T1
TA A+γ γ+B
TB TA A+γ γ γ+B TB
A γ B
A B
(a)
(b)
FIGURE 12.14
(a) Free energy diagram for a line compound. (b) Free energy for a solid solution intermediate phase.
At low temperatures (T < TE), the liquid free energy (dashed line at T ¼ 700 in Figure
12.15) is above the mutual tangent drawn between the two lowest points on the solid free
energy curve and two segregated phases a and b exist in equilibrium as shown. The
relative amounts of these two phases are determined from the overall composition using
the lever rule.
As temperature is increased, the liquid free energy curve drops because of the TDS
term in the free energy, and encounters the mutual tangent line at TE. At this temperature,
three phases can coexist, a liquid plus the two solid phases. The Gibbs phase rule tells us
that, at constant pressure, we have no additional degrees of freedom; therefore, there is
only one composition and temperature for which this situation can occur. We call this
particular combination of temperature and composition the eutectic point. This particular
temperature is called the eutectic temperature and the composition is called the eutectic
composition. Since temperature and composition are fixed at this particular point, it is
called an invariant point. At this invariant point we say a eutectic reaction occurs. We will
encounter other invariant points characterized by different invariant reactions.
As temperature is further increased, the liquid free energy curve descends below the
mutual tangent line between the two solid phases so that now two tangent lines may be
drawn, one from the solid free energy curve on the A-rich side to the liquid free energy
curve, and the other from the liquid free energy curve to the solid free energy curve on the
B-rich side. Mapping these compositions onto the temperature-composition plot produces
the typical eutectic phase diagram shown at the bottom of Figure 12.15. Note that the
temperature will have to be increased above TE to push the liquid free energy curve to the
point where it intersects the solid free energy curve at x ¼ 0 or 1. Thus it may be seen that a
eutectic system will have a lower melting point than either of its pure constituents.
This lowering of the melting temperature is exploited to form low melting point alloys
such as Pb-61 wt% Sn which melts at 1838C; whereas, the melting point of pure Pb is
327.48C and pure Sn is 231.98C. A similar situation is seen in the Ag–Cu system as shown
Phase Equilibria in Multicomponent Systems 241
T = 1500
T = 1100
Free energy (kJ mol−1)
T =1000
=1000
1000
T =940
=940
940
TT==700
700
1500
Temperature (K)
α+L L
α
1100
1000 b +L
940 b FIGURE 12.15
700 If the solid has a large enough positive heat of mixing, it
α+β will tend to separate into an A-rich a phase and a B-rich
b phase when it solidifies. At the eutectic point, these
two phases are in equilibrium with the melt. The eutec-
tic temperature TE will be lower than the melting point
A Mole fraction B of either pure A or B.
in Figure 12.16. Note that Ag and Cu both crystallize into the fcc structure, but the atomic
radius for Ag is 0.175 Å and Cu is 0.157 Å, which accounts for their relative insolubility.
There is, however, some solid solubility on each side of the phase diagram denoted as the a
phase on the A-rich side the b phase on the B-rich side. The line indicating the terminal
solid solubility on either side is called the solvus. In the two-phase region, layers or
lamellas of one phase will be interspersed between layers of the other phase. The relative
amount and composition of these two phases can be determined from the lever rule.
Systems in which A and B crystallize into different structures have separate solid free
energy curves that bend up sharply as the second component is added. If the two structures
are vastly incompatible, such as fcc Au and diamond-like Si, the bend is extremely sharp,
which accounts for the deep eutectic in this system shown in Figure 12.17. Pure Au melts at
10638C and Si at 14048C, while the eutectic composition Au 31 At% Si melts at 3708C.
This means a Si chip can be bonded to a Au-coated ceramic substrate by raising the
temperature to 3708C at which point a melt will form with the eutectic composition. When
cooled back down, even though the melt solidifies back into almost pure Au and pure Si
because there is practically no solid solubility between the two phases, there will be a
fusion weld in which there is intimate contact between the two components. Thus we have
a practical, low temperature method for bonding Au interconnects to Si chips or Si chips to
Au-coated ceramic substrates.
242 Introduction to the Physics and Chemistry of Materials
1200
1000
Temperature (°C) L
L + Cu
800 Ag + L (Cu)
(Ag)
600
Ag + Cu
400
200
0 10 20 30 40 50 60 70 80 90 100
Ag At% Cu Cu
FIGURE 12.16
Ag–Cu phase diagram. (Reprinted from Massalski, T.B., Handbook of Binary Alloy Phase Diagrams, ASM, 1990.
Reprinted with permission of ASM International. All rights reserved.)
1500
1300
1100
L
Temperature (°C)
900
700 Si + L
Au + L
500
383 ± 3°C
300
(Au) Au + Si (Si)
100
0 10 20 30 40 50 60 70 80 90 100
Au At% Si Si
FIGURE 12.17
Au–Si phase diagram. Note that the Au–Si system has virtually no region of solid solution solubility between
Au and Si because of their incompatible lattice structure. The large excess heat of mixing drops the eutectic
temperature dramatically. (Reprinted from Massalski, T.B., Handbook of Binary Alloy Phase Diagrams, ASM, 1990.
Reprinted with permission of ASM International. All rights reserved.)
Phase Equilibria in Multicomponent Systems 243
−10−6
800 K
−10−4
Free energy (J mol−1)
FIGURE 12.18
−10−2 Contamination of a material by its cru-
1000 K
cible. It is assumed the binary melt-
crucible has a 100 kJ/mol excess heat
of mixing. The lowest point on the free
energy curve is the mole fraction of
−1 1500 K crucible material dissolved in the
melt at equilibrium at that particular
temperature. At 1500 K, 2 parts per
10−7 10−6 10−5 10−4 10−3 thousand of the crucible material
Mole fraction would be in equilibrium with the melt.
Recall that the derivative of the entropy of mixing becomes infinite at both 0 and 1,
which means that the segregated free energy curve initially starts downward before going
positive. This is demonstrated in Figure 12.18 in which the free energy of a segregated solid
was assumed to have an excess heat of mixing of 100 kJ mol1. Assume that A component
is the crucible in which the material B is being heated. There will be 0.5 ppm of the
crucible component in equilibrium at the interface with material B at 800 K and 10 ppm at
1000 K. The choice of excess heat of mixing in this example may or may not be realistic, but
the example serves to illustrate the difficulty in maintaining purity in materials during
processing at high temperatures.
TA
Liq
uid
us Liquid TB
Soli
dus
α α+L L+β
TE
β
s
Solvu
α+β
FIGURE 12.19
A generic eutectic phase diagram used as an example of the
A C0 C1 CE C2 B evolving microstructure when the starting compositing inter-
Composition sects the solvus line.
244 Introduction to the Physics and Chemistry of Materials
precipitate and grow within the a phase. The relative amounts and the compositions of the
two phases are determined by the lever rule.
If the starting composition C0 happens to be the eutectic composition CE, the system will
remain molten until the eutectic temperature TE is reached. At this point the melt trans-
forms into a with composition C1 and b with composition C2. The relative amount of the
two phases is determined by the lever rule. Using one end of the tie-line as the pivot point,
e.g., C1 and balancing the torque from N atoms acting on lever arm (CE C1) against Nb
acting on (C2 C1) we obtain
Nb CE C1
xb ¼ ¼ : (12:25)
N C2 C1
C2 CE
xa ¼ : (12:26)
C2 C1
R
Melt Melt
A B A B A B
FIGURE 12.20
Schematic of the mechanism by which a melt of the
l
eutectic composition solidifies into alternating a
α β α β α β
and b lamellas as described by the Jackson–Hunt
theory.
Phase Equilibria in Multicomponent Systems 245
Hunt–Jackson relation described above. In either case, slightly more kinetic undercooling is
required to drive the solidification of a eutectic because of the higher interfacial energy.
CE C0 C0 C1
x0a ¼ ; xL ¼ : (12:27)
CE C1 CE C1
When the eutectic temperature is reached, the remaining melt solidifies with lamellas of a
and b phases so that the final solid will be composed of grains of primary a (sometimes
called proeutectic a) interspersed in lamellas of eutectic a and b. The total mole fraction of
a is given by
C2 C0
xa ¼ : (12:28)
C2 C1
Solidification of a hypereutectic system takes place in a similar manner except that primary
b replaces the primary a that is interspersed throughout the eutectic matrix in the final
solid.
It should be mentioned here that eutectic-like reactions can also take place in which a
solid of composition a decomposes into two solids with compositions b and g at a specific
temperature. Such a reaction is called a eutectoid reaction and is symbolized as a $ b þ g.
The Pb–Bi system shown in Figure 12.23 contains a eutectoid reaction.
TA
Liq
uid
us
Soli
Liquid TB
dus
α+L
TE α L+β
β
s
Solvu
α+β
A C1 C0 CE C2 B FIGURE 12.21
Solidification of a hypoeutectic melt descri-
Composition bed with a generic eutectic phase diagram.
246 Introduction to the Physics and Chemistry of Materials
T = TP
Free energy
TA Liquid
Temperature
L+α
TP α
α+β L+β
TB
β
FIGURE 12.22
Schematic showing the construction Pure A Xp Pure B
of a peritectic phase diagram from
free energy curves. Composition, x
Phase Equilibria in Multicomponent Systems 247
327.502°C
300
271.442°C
Pb + L
L
200 187°C
Temperature (°C)
Bi + L
ε+L
125.5°C 99.5
100 (Pb) ε (Bi)
Bi + ε
0 Pb + ε
−46°C
Pb + Bi
−100
0 10 20 30 40 50 60 70 80 90 100
Pb At% Bi Bi
FIGURE 12.23
Pb–Bi system. There is a peritectic reaction at 1878C with composition 30 At% Bi. There is also a eutectoid reaction
at 468C with 27.5 At% Bi. (Reprinted from Massalski, T.B., Handbook of Binary Alloy Phase Diagrams, ASM, 1990.
Reprinted with permission of ASM International. All rights reserved.)
S1 S2
Free energy
T = TM
S2
S1
T = TE
TA Liquid
L1
Temperature
L1 + α L1 + L2 L2
TB
TM α
α + L2
TE β
α+ β
As a system with monotectic composition is solidified from the melt, the liquid trans-
forms to a plus B-rich melt (L2) at the monotectic temperature. Again the relative amounts
are determined by the lever rule. This second phase liquid eventually transforms to a
eutectic microconstituent at TE. Since the B-rich phase is incorporated into the a phase as a
liquid, it tries to reduce its interfacial energy by spherodizing. Therefore the final eutectic
microconstiuents are generally found in the form of spherical particles.
Attempts to form alloys with compositions under the two-liquid dome will result in
an almost completely segregated system because of the density differences between the
two immiscible liquid phases. Even attempts to form such alloys under microgravity
conditions led to similar results, but for a different reason. Instead of buoyancy effects,
the phases became separated because of their relative interfacial energy differences with
the container walls (see Cahn, J.W., J. Chem. Phys.).
A number of binary systems will form an AB solid solution at higher temperatures
because of the entropy of mixing term, but will become immiscible at lower temperatures,
segregating into a1, an A-rich AB solid solution and a2, a B-rich solid solution. In some
cases a monotectoid reaction, a2 $ a1 þ b, will also occur. Examples include Al–Zn, Hf–Ta,
Nb–Zr, and Nb–U. The monotectoid reaction in the Al–Zn system is shown in Figure 12.26.
1200
L2
L
1000
L + L2
84.2°C
Temperature (°C)
800
710°C
(βCa)
600 βCa + L2
447°C
400
(αCa) αCa + L2
97.558⬚C
200
αCa + βNα (βNα)
0 10 20 30 40 50 60 70 80 90 100
Ca At% Na Na
FIGURE 12.25
Ca–Na monotectic system. There is a solid phase transition from aCa to bCa at 4478C. The Na–aCa eutectic
temperature is only 0.458C below the melting point of Na. (Reprinted from Massalski, T.B., Handbook of Binary
Alloy Phase Diagrams, ASM, 1990. Reprinted with permission of ASM International. All rights reserved.)
800
700
660.452°C
600
L
Temperature (°C)
500 Al + L
(Al) 410.58°C
400 381°C
361.5°C (Al2)
300 Al + Al2 Al2 + Zn (Zn)
277°C
200
Al + Zn
100
0
0 10 20 30 40 50 60 70 80 90 100
Al At% Zn Zn
FIGURE 12.26
Al–Zn system. The monotectoid reaction occurs at 2778C with a composition of 59 At% Zn. The Al2 is a
solid solution of Al and Zn that is richer in Zn. These two components form a solid solution above 361.58C up
to 60 At% Zn, but below that temperature, they will segregate into a Zn-rich phase and a Zn-poor phase.
(Reprinted from Massalski, T.B., Handbook of Binary Alloy Phase Diagrams, ASM, 1990. Reprinted with permission
of ASM International. All rights reserved.)
250 Introduction to the Physics and Chemistry of Materials
1100
900
β
Temperature (°C)
500
ε
Zn
fcc
300
38% 46%
100
0 10 20 30 40 50 60 70 80 90 100
At% Zn
FIGURE 12.27
Cu–Zn phase diagram. Note the order–disorder transition in the b phase at 4608C. Also note the extent of the a
phase to 38.27 At% Zn and the b phase to 48.2 At% indicated by the numbers on the diagram. (Reprinted from
Massalski, T.B., Handbook of Binary Alloy Phase Diagrams, ASM, 1990. Reprinted with permission of ASM Inter-
national. All rights reserved.)
when divalent Zn is added to monovalent Cu. Note in Figure 12.27 that a brass, which
forms up to 38 At% Zn, is fcc. For higher concentrations of Zn, the system undergoes a
phase transformation to b brass which is bcc up to 48 At%. At this point it becomes g,
a complex cubic cell containing 52 atoms. The addition of still more Zn produces the e and
h phases which are hcp with different c to a ratios. Cu–Al, Cu–Sn, Ag–Zn, Ag–Al, Ag–Cd
and other mixed valence systems show similar behavior. As mentioned previously, this
type of phase transformation can be understood in terms of the Fermi-sphere expanding
through the Brillouin zone of the initial phase requiring a new phase with a larger Brillouin
zone to accommodate the additional electrons. This subject will be treated in more detail in
Chapter 19.
12.7 Summary
In two-component systems, the Gibbs phase rule allows one degree of freedom, usually in
the form of a temperature vs. composition line. If there is little or no excess heat of mixing
and the components are highly compatible, i.e., nearly same atomic radius and electro-
negativity, same crystal structure, and same valence (Hume–Rothery rules), the system can
have complete solid and liquid solubility over the entire range of composition. Such
systems are said to be isomorphous. However, even isomorphic systems do not solidify
Phase Equilibria in Multicomponent Systems 251
congruently, i.e., do not go from melt to solid with the same composition. Instead, the
phase diagram contains a lenticular region, bounded by an upper liquidus line and a lower
solidus line, in which the liquid and solid can coexist. The segregation coefficient k is
defined as the ratio of the solid to liquid composition.
The relative amounts of the two phases can be determined using the lever rule. A tie line
is constructed across the two-phase region at the temperature of interest. Using the initial
composition as the fulcrum, the amount of one phase times its lever arm (difference
between its composition and the initial composition) must balance the amount of the
second phase times its lever arm.
The first-to-freeze will have composition C0=k, where C0 is the initial composition of the
melt. Equilibrium solidification in which the solid has the same composition as the melt is
only possible if the solidification is carried out very slowly so that solid-state diffusion can
take place. Otherwise, the grains are cored, meaning that centers of the first-to-freeze
component are surrounded by the last-to-freeze composition.
Ordered phases can form at low temperatures in solid solutions. These will become
disordered as the temperature is increased because of the TDS term in the free energy
where DS is higher entropy of the disordered phase. Order–disorder transitions are
examples of second-order phase transitions that take place over a range of temperatures
with no heat of transition.
If the solid and the liquid has a negative heat of formation, which can result if there is
a large difference in electronegativity or if covalent bonds can form, the freezing points
of all compositions become elevated and there is a tendency for intermetallic or
other intermediate compound phases to form. Sometimes these intermediate phases
form by congruent melting in which a melt goes directly into a solid with the same
composition.
If the solid has a positive heat of mixing, which often occurs because of a lattice
mismatch, the melting point is depressed for all composition and either a eutectic or
peritectic reaction will occur. In a eutectic reaction, the liquid transforms into two solids,
an A-rich a and a B-rich b phase. In a peritectic reaction, the liquid and the solid phase
transform into a second solid phase. In either case, the solid will tend to segregate into two
or more distinct phases. There is limited solid solubility in these two phases, and there will
be a range of compositions between these phases where a and b will coexist in alternating
layers or lamella. The spacing between these lamellas is governed by the Hunt–Jackson
relation, which tells us that the spacing is inversely proportional to the square root of the
cooling rate. The terminal solubility of these two phases is separated from the mixed a þ b
region by the solvus line.
Solidifying an off-eutectic composition from the melt will result in pro-eutectic grains
with the first-to-freeze composition. The remaining melt then freezes with the eutectic
composition. Attempts to solidify from the melt at the peritectic composition will result in
highly cored grains because when the b first forms, it blocks the remaining a from reacting
with the liquid to form new b phase.
Monotectic reactions, in which a liquid decomposes into a solid and a second liquid, can
occur in systems that have excess heats of mixing in both the solid and liquid phase.
Monotectic systems also have a region of liquid phase immiscibility. Unless done very
rapidly, attempts to solidify through this two-phase liquid region will result in almost
complete phase separation because of interfacial effects. Monotectoid reactions are also
possible in which an AB solid solution separates into an A-rich solid solution and a B-rich
b phase.
The various invariant points in which three phases can be present are described by the
diagrams in Table 12.1.
252 Introduction to the Physics and Chemistry of Materials
TABLE 12.1
Summary of Invariant Reactions
Eutectic α+ L L β +L
L↔α+β α β
α +β
Eutectiod α +γ γ β +γ
γ ↔α+β α β
α +β
Peritectic α +L
α L
α+L↔β α +L β β +L
Peritectoid α +β
α β
α+β↔γ α +γ γ β +γ
Monotectic α + L1 L1 L1 + L2
α L2
L 1 ↔ L2 + β α + L2
Monotectoid α1 + α2 α2 α2 + β
α1 β
α↔ β +γ α1 + β
Systems in which the components have different valences produce complex phase
diagrams with a number of different solid phases and peritectic reactions as well as
eutectoid and peritectoid reactions. These transitions can be explained in terms of inter-
actions between the Fermi sphere and the Brillouin zones.
Bibliography
Askeland, D.R., The Science and Engineering of Materials, Books=Cole Engineering Division,
Wadsworth, Bellmont, CA, 1984.
Barsoum, M.W., Fundamentals of Ceramics, Institute of Physic Publishing, Bristol and Philadelphia,
2003.
Cahn, J.W., Critical point wetting, J. Chem. Phys. 66, 3667, 1977.
Callister, W.D., Materials Science and Engineering: An Introduction, 7th edn., John Wiley & Sons,
New York, 2007.
Gersten, J.I. and Smith, F.W., The Physics and Chemistry of Materials, John Wiley & Sons, New York,
2001.
Halsedt, B., Thermodynamic assessment of the system MgO–Al2O3, J. Am. Ceramic Soc., 75=6, 1992.
Kingery, W.D., Bowen, H.K., and Uhlmann, D.R., Introduction to Ceramics, 2nd edn., John Wiley &
Sons, New York, 1976.
Levin, E.M., Robins, C.R., and McMurdie, H.F., Phase Diagrams for Ceramists, The American Ceramic
Society, Columbus, OH.
Massalski, T.B., Senior Editor, Handbook of Binary Alloy Phase Diagrams, Vols. 1, 2, and 3, ASM,
Materials Park, OH, 1990.
Moffat, W.G., Handbook of Binary Phase Diagrams, Genium Publishing, New York, 1976.
Muller, O. and Roy, R., The Major Ternary Structural Families and the Major Binary Structrual Families, in
Crystal Chemistry of Non-Metallic Materials, Vols. 3 and 4, Springer, Berlin, Heidelberg,
New York, 1974 and 1977.
Ragone, D.V., Thermodynamics of Materials, Vol. II, John Wiley & Sons, New York, 1995.
Rosenberger, F.E., Fundamentals of Crystal Growth, 2nd edn., Springer-Verlag, New York, Heidelberg,
Berlin, 1981.
Shunk, F.A., Constitution of Binary Alloys, McGraw-Hill, New York, 1969.
Phase Equilibria in Multicomponent Systems 253
Problems
1. How would you know if a solid solution was ordered or not?
2. Si–Ge forms a solid solution over the entire range of compositions. But III–V com-
pounds, which have similar diamond-like lattice structures, do not. Why do you
suppose this is?
3. What is the wt% of Te in CdTe?
4. Identify all of the invariant points in the Cu–Zn phase diagram (Figure 12.27). Specify
the reactants, the composition, the temperature, and the type of reaction that is taking
place.
5. I am solidifying a 50 At% Ag–Cu alloy. Sketch what you would expect the microstruc-
ture to look like. With the help of Figure 12.16, describe the microstructure in terms of
the amount and composition of the proeutectic and the amount and composition of the
two eutectic phases.
6. In the MgO–Al2O3 phase diagram in Figure 12.12, it may be seen that the spinel
(MgAl2O4) solid solution, which forms at 50 At% extends for nonstoichiometric com-
positions ranging from 40 At% Al2o3 to 80 At% Al2o3.
(a) What is the type and fraction of vacancy defects at 80 At% Al2O3?
(b) What is the type and fraction of vacancy defects at 40At% Al2O3?
13
Alloy Solidification
We saw in Chapter 12 that binary and multicomponent systems do not solidify congru-
ently. As a result, the final solid will not be compositionally homogeneous. Consequently, a
number of solidification methods have been devised in an attempt to control the final
microstructure on both a micro- and macroscale. We briefly explore some of the techniques
in this chapter along with some exotic methods that are now employed in order to control
the composition of the final product at the atomic scale.
255
256 Introduction to the Physics and Chemistry of Materials
Water spray
Mold wall
From
the Melt Solid
tundish
FIGURE 13.1
Schematic of a continuous casting process as an example of direction solidification.
Quartz
Melt ampoule
Hot zone
Adiabatic
zone
Solid
Cold
FIGURE 13.2
zone
Schematic of a Bridgman–Stockbarger furnace with an adiabatic zone. Heat
flows into the melt from the hot zone, through the solidification interface in the
adiabatic zone, and is extracted through the solid in the cold zone as indicated
by the arrows. The growth ampoule is slowly lowered at the desired solidifi-
cation velocity. In order to maintain a near-planar solidification front, the
solidification interface is kept in the adiabatic zone where the isotherms are
nearly perpendicular to the walls.
moving parts. However, this technique lacks the control of the thermal environment in the
vicinity of the growth interface offered by the Bridgman–Stockbarger method.
13.2.2 Macrosegregation
Macrosegregation refers to an unwanted variation in composition of a material on a macro-
or global scale. As seen in Chapter 12, one of the components in an alloy system will almost
always solidify at a higher temperature than the others. As the temperature of the melt is
lowered to the liquidus curve as in Figure 13.3, this component will be the first to freeze. As
solidification continues, the loss of this first-to-freeze component changes the composition
of the remainder of the melt and produces one form of macrosegregation.
Placing the hot zone above the cold will create a thermal gradient that will tend to
stabilize the melt against convective flows, although the radial thermal gradients will still
cause some mixing. If the rejected component of the alloy system is more dense than the
TA
Liquid
T1
Liquid +
T2 solid
FIGURE 13.3
Typical phase diagram of a binary system with complete liquid
Solid and solid solubility. The first-to-freeze component will have com-
TB
position C1. The rejected B component will be pushed ahead of
the solidification front and, if convectively mixed with the
remaining melt, the solid will become richer in B as the process
A C1 C0 C2 B
continues.
258 Introduction to the Physics and Chemistry of Materials
starting composition, the system will be both thermally and solutally stable. However, if
the rejected component is less dense than the bulk melt, the system will be solutally
unstable and will be subject to overturning convective flows. Under these circumstances,
the system can still be unstable even if the thermal gradient is sufficient to cause a net
positive density gradient in the downward direction. This phenomenon is known as
double diffusive convection and occurs because the thermal diffusivity of the melt is
always greater than chemical diffusivity. Therefore if a parcel of fluid that is rich in the
lighter component is displaced upward, it can accommodate thermally faster than com-
positionally and will continue to rise (see McFadden et al., Physics of Fluids).
Recall from Chapter 12 that the partition coefficient k was defined as the ratio of solid
component CS to the liquid component CL at the solidification interface. In the phase
diagram shown in Figure 13.3, the k ¼ C1=C0 at T1 and ¼ C0=C2 at T0. In this case, k is
seen to vary with temperature. For dilute systems (alloys in which the second component is
a small fraction of the host component), k can be taken to be a constant and can be defined
as the ratio of the slope of the liquidus to the slope of the solidus. Even for nondilute alloys,
assuming k to be constant is a reasonable approximation for most systems.
We will consider two limiting cases: The first case assumes the rejected solute is
uniformly distributed throughout the remaining liquid during the solidification process
(complete convective mixing). The second case assumes that there is no convective mixing,
so the rejected solute builds up ahead of the advancing solid to form a diffusion layer.
Solid-state diffusion is so small that it will be completely neglected, so whatever compos-
ition the solid had when it froze will remain unchanged.
which can be integrated with the initial condition, for f ¼ 0, CS ¼ kC0, to obtain the Scheil
equation
1
k = 0.7
k = 0.5
k = 0.1
0
0 0.2 0.4 0.6 0.8 1
Fraction solidified f
FIGURE 13.4
Final composition profile of a directionally solidified ingot with complete convective mixing. The composition of
the solid varies continually with distance and the solute that was rejected initially is later incorporated into the
solid so that the overall average composition is the starting composition C0.
Consider a reference system in which the solidification front remains stationary and x0 is
the distance into the melt. The melt flows in from the right-hand side at the solidification
rate R. Balancing the solute flowing into the diffusion layer against the rate the solute is
diffusing to the right,
d2 CL dCL
D 02 þ R ¼ 0: (13:3)
dx dx0
Under steady-state conditions, the composition of the solid being formed CS is the same as
the bulk melt, C0. The boundary conditions are CL(0) ¼ C0=k; CL(1) ¼ C0. The solution can
be written as
1k R
CL (x) ¼ C0 exp x0 þ 1 ; x 0, (13:4)
k D
which can be verified by direct substitution. A typical plot of this solution is shown in
Figure 13.5 after the initial transient. The ratio D=R is the diffusion length l. Diffusion
coefficients for liquid metals are typically about 5 cm2=day. Therefore, solidification rates
of 5 cm=day will produce a diffusion length of 1 cm.
C2
C0
C1 Solid Liquid
f d
FIGURE 13.5
Composition along a directionally solidified sample in which a fraction f has been solidified with no convective
mixing. The first to freeze has composition is C1 ¼ kC0. The rejected B-rich component builds up a diffusion layer
ahead of the advancing solidification front. When the composition in the melt C2 ¼ kC0, the solid will have the
composition C0.
where x is the distance into the solidified material and the characteristic length of the
transient region is D=(Rk). The composition after a certain fraction of an ingot has been
solidified looks like Figure 13.5.
After steady-state conditions have been established, the balance of solute flowing into
the solid at the same rate it flows from the liquid into the diffusion region continues until
the tail of the diffusion layer reaches the end of the melt. At this point, the concentration of
the rejected component increases rapidly to make up for the initial transient so that the
average composition equals the composition of the starting melt as seen in Figure 13.6.
Although the composition of the final solid is by no means uniform, there is at least a
region over which the composition of the solid is the same as the starting composition.
The ends of the final solid where the composition is not uniform can be cut off and
remelted. Obviously, the sample length must be longer than the initial transient and the
diffusion length for this technique to be effective.
C0
Final
C1 transient
Initial transient Diffusion controlled
FIGURE 13.6 = D/R
Final composition profile of a directionally = D/kR
solidified ingot under diffusion-controlled
conditions (no convective mixing). Distance along ampoule
Alloy Solidification 261
G1
G2
T1
T0
FIGURE 13.7
Local freezing temperature along the growing sample. The temperature ahead of the solidification interface must
increase at least at the rate shown by G1 to stabilize the plane front growth interface. If the thermal gradient is
insufficient, as illustrated by slope G2, the growth interface breaks down into a dendritic structure with a mushy
zone (mixture of solid dendrites and interdendritic fluid) that extends to the point where the temperature is above
the local freezing point.
solid has the same composition C0 as the starting melt, which solidifies at T0. The solid is in
equilibrium with the melt composition C2. The local freezing temperature of the melt
in front of the solidification interface can be related to the composition of the mixture by
T(x) ¼ TA þ mCL(x), where TA is the freezing point of pure A and m is the average slope of
the liquidus line (usually negative). From Equation 13.4, the local freezing temperature
of the melt is
0 1k R 0
T(x ) ¼ TA þ mC0 exp x þ 1 : (13:6)
k D
At the interface (x0 ¼ 0), the T(0) ¼ TA þ mC0=k ¼ T0. However, at a large distance from the
interface, T(1) ¼ TA þ mC0 ¼ T1 > T0. If the imposed temperature in front of the solidifica-
tion interface increases faster than the local freezing point of the melt ahead of it, as
illustrated by the slope G1 in Figure 13.7, the solidification front will be stable. However,
if the imposed temperature gradient is such that temperature falls below the local freezing
temperature, as illustrated with gradient G2, the region over which the local freezing point
is higher than the local temperature is said to be constitutionally undercooled.
As local perturbations in the solidification interface that move slightly ahead of the front
find themselves in a region where the local temperature is lower than the local freezing
point. The constitutionally undercooled interface will break down into columns or fingers.
These microscopic fingers then continue to race ahead of the average solidification front
until they reach a temperature above their freezing point. These projections also break
down along their sides to form arms or branches and eventually form a dendritic or tree-
like structure (see Section 13.5). The region between the dendrite tips and the main
solidification front is called the mushy zone, the extent of which is illustrated by the
interception of slope G2 with the local freezing temperature.
A simple constitutional supercooling (CS) criterion was developed by Rutter and Chalmers
(1954) that predicts the ratio of the gradient required to stabilize the interface to the growth
velocity for a given solidification system. Taking the derivative of Equation 13.4 at x0 ¼ 0,
dT R 1k
¼ mC0 ¼ G1 , (13:7)
dx 0 D k
262 Introduction to the Physics and Chemistry of Materials
where G1 is the gradient required to stabilize the growth front against breakdown due to
constitutional undercooling. Often this relation is written in terms of the G=R ratio, which
is the gradient to growth rate required to stabilize the growth front,
G mC0 1 k
¼ : (13:8)
R D k
Later, Mullins and Sekerka (1964) developed a more rigorous theory based on a stability
analysis that included the liquid–solid interfacial energy, which can provide a stabilizing
effect on the interface. However, the difference between the two theories is so small
that, for the most part, the more conservative CS criteria can be used for experiment
design.
where l is the zone length. Setting CL ¼ CS=k and integrating with the boundary condition
CS ¼ kC0 at x ¼ 0,
h i
CS ðx0 Þ ¼ C0 1 ð1 kÞekx=l : (13:10)
When distance traversed by the zone x l, the composition of the solid will equal C0.
If the starting material is macroscopically homogeneous but microscopically inhomo-
geneous (e.g., a pressed powder sample with two or more components), a single pass
Alloy Solidification 263
C2
C0
FIGURE 13.8
T1 Composition profile (upper) and local melt-
ing temperature (lower) of material being
processed by the traveling zone method. If
T0 the material in the molten zone is com-
pletely mixed, its composition will be C0=k
and it will have the same melting tempera-
ture as the newly formed solid when
Solid Liquid Solid steady-state growth is achieved. Since the
f L material in the zone is not constitutionally
undercooled, interfacial breakdown is not a
Distance along ampoule problem.
with a short zone will homogenize most of the specimen. This process is called zone
leveling. A similar process in which a seed crystal or seed selector is used to start the
growth can be used to grow single crystals of near uniform composition. This technique is
called the traveling zone method of crystal growth. The advantage over the Bridgman
method is that the zone essentially replaces the diffusion layer so that it may be possible to
obtain a shorter initial growth transient if l can be made <R=D. Since the traveling zone
relies on convective mixing to homogenize the solute in the zone, it can be used in cases
where it is not possible to prevent convective mixing. Also, since the composition of the
melt in the traveling zone is not constitutionally undercooled, it is easier to prevent
interfacial breakdown. Care must be taken, however, to keep the convective flows in the
zone from becoming turbulent, which can cause temperature and compositional fluctu-
ations resulting in compositional striations in the resulting solid.
r gl2
B0 ¼ : (13:11)
g
An analysis by Coriell, Hardy, and Cordes (1977) showed that the limiting Bond number
that can contain the molten floating zone was B0 ¼ 7.5. With the density r ¼ 5 g=cm3 and
the surface tension g ¼ 500 dyn=cm, the maximum length l ¼ 8.75 mm in normal gravity
regardless of diameter of the starting material. Since it is difficult to melt a zone whose
length L is less than the diameter of the starting material, this type of crystal growth is
264 Introduction to the Physics and Chemistry of Materials
somewhat restricted to small samples for most materials. However, in conductive samples
with nonvolatile components, it is possible to heat the sample directly electromagneti-
cally, which also provides levitation forces for additional zone confinement. Ultra-pure
dislocation-free silicon can be produced by this technique.
l
dCn ðxÞ ¼ ½Cn1 ðx þ lÞ Cn ðxÞdx: (13:12)
k
For n ¼ 1, the Cn1 ¼ C0 and Equation 13.15 reduces to Equation 13.12. A number of
analytical solutions to Equation 13.15 have been found that involve double and triple
series (see Pfann, Zone Melting). Milliken was able to simplify the solution to a single series,
Cn ðjÞ 1 k kj X1
ðs þ jÞs2
¼ e ðs nÞks esk ½jðs 1Þ þ ð1 kjÞðs þ jÞ, (13:13)
C0 k s¼nþ1
s!
where j ¼ x=l. The solution is valid until x ¼ L l‘. This solution is plotted in Figure 13.9
for k ¼ 0.2 and convergence of the series was reached for n ¼ 10 after 50 terms. Conver-
gence is much slower for larger values of k.
The initial concentration after n zone passes for different k values is shown in Table 13.1.
It may be seen that the number of passes needed to achieve high purity goes up quickly
with increasing k as would be expected. Often the material to be purified is placed in a long
1.0 n = 1
n=2
n = 10
C/C0
0.5
n = 20
n = 30
FIGURE 13.9
Zone purification after n passes for k ¼ 0.2. Not 0
0 20 40 60 80 100
shown is the portion j > L l where the rejected
material piles up at the end. Distance/zone length
Alloy Solidification 265
TABLE 13.1
Fractional Amount of Impurities after n Zone Passes
n k ¼ 0.2 k ¼ 0.4 k ¼ 0.6 k ¼ 0.8
ampoule and multiple molten zones are repeatedly moved along through the solid.
Clearly, this process is most effective if k 1.
Crystal
RF coils
Melt
FIGURE 13.10
Schematic of a crystal being pulled from the melt using
the Czochralski method for growing single crystals from
Susceptor the melt.
266 Introduction to the Physics and Chemistry of Materials
FIGURE 13.11
Thermal dendrite growing in undercooled succinonitrile.
(Photo taken from the RPI Isothermal dendrite growth
experiments. See http:==www.rpi.edu=locker=56=000756=.)
13.6 Casting
13.6.1 Continuous Casting
Most steel is now produced by a continuous casting process that takes molten steel direct
from the furnace and converts it into slabs or beams of uniform cross section in a single
operation. Molten metal from the tundish, or reservoir, is forced through a water-cooled
mold with a cross section of the final product where the surface begins to solidify. The slab
emerges from the mold in a partially solidified state, as indicated in Figure 13.1, and is
further cooled by a water spray. When completely solidified, the product is cut to length
Alloy Solidification 267
and removed. As the name implies, this is a continuous process. Ladles from the furnace
keep replenishing the tundish as the melt flows out. Aluminum and copper slabs can also
be produced by a similar process.
13.7 Sintering
13.7.1 Sintering Process
Sintering involves the forming of solids from powders without completely melting the
components. Powder metallurgy has many advantages of casting and other fabrication
techniques in that it is relatively inexpensive, produces little waste material, and is highly
versatile. But the primary advantage is the powder can be densified without raising it to
its melting temperature, which lends itself to the manufacture of refractory metals and
ceramics. Another major advantage of powder metallurgy is the ability to achieve a high
degree of chemical uniformity without the segregation problems encountered when solidi-
fying from the melt.
Recall from Chapter 9 that decreasing the radius of curvature lowers the melting point of
a material. This makes it possible to fuse small sharp-pointed particles together at temper-
atures well below their normal melting points. This is equivalent to saying the process is
driven by nature’s propensity to reduce free energy, in this case the surface energy
component. Two processes are involved in fusing small particles together. In one case,
surface diffusion causes material on the surface to migrate to the point of contact to form a
neck between the two particles. As the process continues, the total surface area is reduced
as material flows from the surface to the neck region. However, the centers of the two
particles remain in the same place, so no densification results. In the second case, material
flows by volume diffusion through the point of contact to the neck. Because material is
flowing from their interiors, the centers of the particles will move toward each other and
densification does occur.
268 Introduction to the Physics and Chemistry of Materials
FIGURE 13.12
Hg2Cl2 crystal growing by PVD in a closed ampoule. Hg2Cl2 can only be grown from the vapor because it is not
soluble and it decomposes before it melts. (Photo courtesy of Joo Soo Kim, University of Alabama in Huntsville.)
An extremely high vacuum (1011 Torr), free of any electrically active impurity (see
Chapter 20), is required for successful production of good superlattices. The inner walls
usually have a cryogenic liner filled with LN2 to trap any condensable residual gases. Since
pumping such a system down to the vacuum levels is a time-consuming process, a sample
Rotating
sample holder
FIGURE 13.13
Schematic of an MBE facility. Materials
to be deposited are stored in the effu-
Effusion cell with shutter sion cells with shutters to control the
To vacuum flow. The sample is heated and rotated
pump by a motor to assure a uniform coating.
270 Introduction to the Physics and Chemistry of Materials
exchange air lock is usually provided. Because of possible trace contamination, such a
facility is usually dedicated to a single set of materials such as GaAs and AlGaAs. The
quality and thickness of the growing films are monitored by a reflection high energy electron
diffraction (RHEED) system mounted in the analysis ports. Such a facility can grow epitaxial
layers with thicknesses of atomic precision. However, their expense and their limited
throughput, makes them primarily research facilities. Higher throughput with similar pre-
cision can be obtained without the extremely high vacuum requirements by chemical beam
epitacy (CBE) or metalorganic chemical vapor deposition (MOCVD) techniques.
13.9 Summary
Alloy solidification is a complex process in which a number of different things are
happening. Since most metals do not solidify congruently, the first-to-freeze metals
will have a different composition from the bulk melt which leads to macrosegregation
in the solid. Constitutionally undercooled melts lead to the breakdown of planar solidi-
fication fronts and the formation of dendrites. Differences in the density of the melt due
to both thermal and solutal gradients drive convective flows that break off dendrite arms
and redistribute solute as well as these dendrite tips throughout the melt. These dendrite
tips grow larger and eventually form the microstructure of the solid.
In order to control the structure and properties of the final product, it is necessary to
understand how each of these processes work. The solidification process can be simplified
and controlled by extracting heat unidirectionally through the solid at the advancing
solidification interface in a controlled manner using the Bridgman technique. A planar
growth interface can be maintained if the thermal gradient to growth rate (G=R) ratio meets
the CS criterion. The ability to control growth rate and thermal gradient allows the growth
of single crystals or aligned in situ composites and also provides a means for studying
interfacial stability, dendrite growth, particle=solidification front interaction, and other
solidification phenomena.
Alloy Solidification 271
Zone melting and traveling zones offer effective means of avoiding macrosegregation,
eliminating constitutional supercooling, and purifying materials. Floating zone crystal
growth eliminates contact with container walls, which are sources of contamination and
defect formation.
Bibliography
Chalmers, B., in Principles of Solidification, Krieger, R.E., Ed., Huntington, New York (Reprinted by
arrangement from 1964 version published by John Wiley & Sons).
Choudhary, K. and Mazumdar, D., Mathematical modeling of fluid flow, heat transfer and solidifi-
cation phenomena in continuous casting of steel, Steel Res., 66(5), 1995.
Coriell, S.F., Hardy, S.C., and Cordes, M.R., J. Colloid Interface Sci., 60, 1977, 126–136.
Flemings, M.C., Solidification Processing, McGraw Hill Series in Materials Science and Engineering,
McGraw Hill, New York, 1974.
German, R.M., Sintering, Theory and Practice, Wiley Interscience, New York, 1996.
Hosford, W.F., Physical Metallurgy, Taylor & Francis, CRC Press, Boca Raton, FL, 2005.
Jackson, K.A., Kinetic Processes: Crystal Growth, Diffusion, and Phase Transitions, Wiley-VHC,
New York, 2004.
Juretzko, F.R., et al., Particle engulfment and pushing by solidifying interfaces: Part I. Gound
experiments, Metallurgical Materials Trans., 29A, 1998, 1691.
McFadden, G.B., Coriell, S.R., and Boisvert, R.F., Double-diffusive convention with side walls, Physics
of Fluids, 28, 1985, 2716.
Mullins, W. and Sekerka, R., J. Appl. Phys., 34, 1964, 323.
Pfann, W.G., Zone Melting, 2nd edn., John Wiley & Sons, New York, 1966.
Rutter, J.W. and Chalmers, B., A prismatic substructure formed during solidification of metals. Can.
J. Phys., 31, 1954, 15.
Thomas, B.G., Continuous casting: modeling, in The Encyclopedia of Advanced Materials, Vol. 2,
Dantzig, J., Greenwell, A., Michalczyk, J., Eds., Pergamon Elsevier Science, Oxford, United
Kingdom, 1999.
Problems
1. You are designing a crystal growth experiment in which you are required to grow a
10 cm length of Si0.80Ge0.20 with no more than a 1% variation in the Ge content.
The conductivity of the material and the heat transfer in the Bridgman furnace
you want to use is such that the maximum gradient you can obtain is 3008C=cm. The
diffusion coefficient in the melt is 2.0 105 cm2=s, the slope of the liquidus line on the
Si-side of the Ge–Si phase diagram is 58C=At%, and the segregation coefficient is 0.8.
(a) What is the maximum growth rate you can use and maintain a planar solidification
interface?
(b) What will be the minimum initial growth transient before the desired Ge compos-
ition is reached?
(c) How much time will be required to complete the growth process, not counting the
final transient?
2. You decide to investigate to possibility of using the traveling zone process. Again with
the maximum gradient of 3008C=cm, what is the minimum zone length you could
reasonably expect to use? What would be the minimum initial growth transient before
the desired Ge composition is reached?
14
Transformation Kinetics
The strength of a material can vary over an order of magnitude or more depending on
its microstructure, which in turn depends on the way it is heated and cooled during the
solidification process. In order to be able to control the evolution of the microstructure, we
must understand the kinetics of the various transformation processes.
VT VU
¼1 ¼ 1 exp(ktn ): (14:1)
V0 V0
This relationship is known as the Avrami equation (also known as the Kolmogorov–
Johnson–Mehl–Avrami or KJMA equation). Half of the material will be transformed
when ktn ¼ 0.693. Only 1% of the transformation has occurred when ktn ¼ 0.01 and the
transformation is 99% complete when ktn ¼ 4.6. The transformation time is generally
specified as the time required for half of the material to be transformed, which is given by
0:693 1=n
t0:5 ¼ : (14:2)
k
A transformation described by the Avrami equation with its typical sigmoid shape is
shown in Figure 14.1 for k ¼ 0.0001=s and different values of n. We will now consider
several time-dependent, thermally activated transformations.
273
274 Introduction to the Physics and Chemistry of Materials
1.0
Fraction transformed
0.8
n=5
0.6 n=4
n=3
0.4
0.2
FIGURE 14.1 0
1 10 100
Fraction transformed versus time on a log scale as
described by the Avrami equation. Time (s)
1600
L+δ
δ
1400
L+γ
Liquid
1200 L+C
1447°C
Temperature (°C)
γ (Austenite)
Fe3C (Cementite)
800
727°C
0.76 wt%
600 α + Fe3C
α (Ferrite)
0.027 wt%
400
0 5 10 15 20 25
At% Fe3C
FIGURE 14.2
Fe-Fe3C phase diagram. The eutectoid point at 7278C and 0.76 wt% Fe2C is the focal point for steel making.
(From Massalski, T.B., Senior Editor, Handbook of Binary Alloy Phase Diagrams, Vols. 1–3, American Society for
Metals, 1990. Reprinted with permission of ASM International. All rights reserved.)
Transformation Kinetics 275
a-ferrite (bcc) to g–austenite (fcc) and to d-ferrite (bcc) before melting (there is no b phase).
Ferrite is the name given to the bcc structure and austenite is the fcc structure.
Austenite is a solid solution of g–Fe and C atoms that reside in the interstities of the
fcc structure. There is a very limited region of solid solubility of C in a-Fe (0.022 wt% C).
Recall from Chapter 3 that the interstitial sites in the fcc lattice are considerably larger
than in the bcc lattice, which account for the much greater terminal solid solubility of C in
austenite.
Of particular importance to steel making is the eutectoid point at 7278C 0.76 wt% C.
When austenite of this composition is cooled to this temperature, it undergoes euctectoid
reaction g , a þ Fe3C. Further cooling produces lamellas of alternating a-Fe and cementite
called pearlite because its appearance resembles mother-of-pearl. The spacing of the
lamella can be controlled by the cooling rate as discussed in Chapter 12.
Steels with less than 0.76 wt% C are called hypoeutectoid steels. Before they reach the
euctectoid temperature, some proeutectoid a-Fe is formed. Upon further cooling to
the eutectoid temperature, the remaining g is transformed to pearlite, composed of euctec-
toid a and cementite. The final structure then consists of proeutectoid a and pearlite. The
relative amounts of each phase are determined by the starting composition and can be
found by applying the lever rule.
Steels with greater than 0.76 wt% C are called hypereutectoid steels. The development of
their microstructure is similar to the hypoeutectoid steels except the proeutectoid cementite
forms before reaching the eutectoid temperature and the final solid consists of proeutectoid
cementite and pearlite.
The phase diagram tells us only what phases can exist in equilibrium. The final micro-
structure is controlled by the rate of transformation. A sketch of a typical isothermal
time-temperature-transformation (often called a T-T-T diagram) map for the austenitic to
ferritic transformation is shown in Figure 14.3.
The curves labeled 1%, 50%, and 99% indicate, respectively, the beginning of the
transformation, the point at which half of the material has been transformed, and
the point at which virtually all of the material has been transformed. The region to the
left of the 1% line represents untransformed austenite. The region to the right of the 99%
line and above the nose of the curve represents pearlite, ferritic iron interspersed
with cementite. The spacing of the cementite becomes finer as the transition temperature
is lowered because the diffusion length is shorter at the lower temperature and because
800
Eutectoid temperature
700
600 1%
Pearlite
Temperature (°C)
500 99%
400
300
Bainite
1%
200 FIGURE 14.3
50% Isothermal transformation map for the austenitic-ferritic-
99%
100 martensitic transformation. (From Boyer, H., Ed. Atlas
Martensite of Isothermal Transformation and Cooling Transformation
0 Diagrams, American Society for Metals, p. 28. Reprinted
−1 0 1 2 3 4 5 6 with permission of ASM International. All rights
Log time (s) reserved.)
276 Introduction to the Physics and Chemistry of Materials
the transition takes place more rapidly. Both of these factors limit the distance the carbon
atoms can diffuse before they react to form cementite. Fine pearlite is stronger but less
ductile than coarse pearlite.
To the right of the 99% line and below the nose lies bainite. Bainite can be thought as the
extreme limit of fine pearlite. The cementite is in the form of very fine needles or platelets,
so fine that they can only be seen by an electron microscope. Bainitic iron is stronger and
more ductile than pearlitic iron because of the fine scale of its microstructure.
800
Eutectoid temperature
700
1% Pearlite
600
99%
Temperature (°C)
500
400
300 Bainite
1%
FIGURE 14.4
200
Isothermal transformation map of the austenitic- 50%
ferritic system with various cooling curves. 99%
100 Martensite
(From Boyer, H., Ed. Atlas of Isothermal Trans-
formation and Cooling Transformation Diagrams, A B C D E
American Society for Metals, p. 28. Reprinted 0
with permission of ASM International. All rights −1 0 1 2 3 4 5 6
reserved.) Log time (s)
Transformation Kinetics 277
In curve D, the sample is quenched and held until 50% is converted to fine pearlite. Then
the temperature is lowered rapidly into the bainite region and held until the rest of the
sample is converted to bainite. As before, these structures are retained as the material is
quenched to room temperature.
In curve E, the temperature is lowered to just below the eutectoid temperature and held
until it is 100% converted to coarse pearlite.
It should be noted that the diagrams shown above are supposed to be isothermal
transformations. That is, the temperature is presumed to be changed in a time much
shorter than the holding time. This may not always be possible, especially when
the holding times are shorter than the times required to change the temperatures of the
melt in question. The isothermal transformation diagrams may be used for continuous
temperature changes, although the positions of the transformation curves may be shifted
somewhat.
14.2.3 Spherodizing
Spheroidite does not appear on the transformation map. Either pearlite or bainite can be
transformed into spheroidite by soaking below the eutectoid temperature for an extended
period. Driven by the tendency to reduce interfacial energy by reducing the surface area,
solid-state diffusion will cause the cementite lamella or needles to become more spherical
in shape. Spheroidite is not as strong as coarse pearlite, but is more ductile and tougher
because the sharp edges of the cementite particles have become rounded off, reducing
their effectiveness as stress risers. The large number of distributed spherical particles act as
crack stoppers.
14.2.4 Tempering
Tempered martensite is formed by first rapid quenching of austenite to form highly
strained martensite. The martensite can then be reheated to below the eutectoid tempera-
ture so that solid-state diffusion can operate effectively. The strain is relieved as the
bct structure converts to bcc with the excess carbon combining with Fe to form cementite,
just as in the process of precipitation hardening. Unlike spheroidized pearlite or bainite,
the resulting cementite forming from bct martensite is on the nanoscale, much finer that
the microscale of spherodized pearlite or bainite, and is more uniformly distributed.
Controlling the aging process, i.e., the time and temperature, provides control over the
size of the precipitates. Tempered martensite can be almost as hard as untempered
martensite, but has increased ductility and toughness. The hardness decreases with
aging but the ductility increases as the precipitated cementite particles grow in size.
14.2.5 Annealing
In many forming processes, it is desired to work with a soft ductile material. The effects of
cold working can be removed by annealing, a process whereby the material is heated
below its melting temperature and some of the lattice strain is reduced as the atoms, driven
by the strain energy, try to move back to their equilibrium configuration. At the same time,
new crystals begin nucleating and start to grow, free of dislocation strain, by diffusion. The
driving force for this recrystallization phase is the lower energy of the strain-free crystals.
These new crystals will continue to grow, driven by the lowering of the interfacial energy
(larger crystals have less surface per volume), until they completely replace the old-
strained crystals. As this process progresses, the strength diminishes while the ductility
278 Introduction to the Physics and Chemistry of Materials
FIGURE 14.5
Effects of annealing on grain structure. Elongated grains in a steel nail (a) have grown larger and more equiaxed
(b) after a 4 h anneal. (Courtesy of M. Banish, University of Alabama in Huntsville.)
increases to the soft preworked values. Controlling the temperature and time allowed for
grain growth provides a means for grain refining, i.e., achieving the desired grain struc-
ture. The recrystallization temperature is defined as the temperature at which grain
growth is complete within 1 h. It is generally 30%–50% of the absolute melting tempera-
ture. Figure 14.5 illustrates the effect of annealing on strained crystals.
14.2.6 Hardenability
Hardenability, not to be confused with hardness, is the ability of a material to be hardened
by a heat treatment. In steels, it is a measure of the ability to form martensite and is
measured by the Jominy end-quench test. A Jominy bar is fashioned to AISI specifications
Transformation Kinetics 279
(25.4 mm diameter 100 mm long). After austinizing the bar at a specified temperature for
a specified period of time, the bar is quenched at one end with a specified water spray to
provide a reproducible cooling rate that varies along the length of the bar. Microhardness
is measured along the length of the bar. A plot of hardness with distance from the quench
end gives the hardenability curve.
where
R(0) is the average initial radius
KLSW is a constant which contains the relevant material properties such as the
interfacial energy and diffusion coefficient.
However, the classical LSW analysis is based on a mean field theory that ignores the
finite volume of the dispersed phase. Extensive theoretical and experimental work has
been done recently to extend the model for ripening to finite volume fractions (Ratke and
Voorhees, 2002).
TA
Tsolidus
Liquid TB
+L Rapid quench
TE
Tsolvus
A C0 C1 CE B
(b) Time (s)
(a) Composition
FIGURE 14.6
(a) Generic phase diagram of a material than can be precipitation hardened. The starting composition must be
heated through the solvus line in order to dissolve the B-component. (b) Time-temperature profile of the
precipitation hardening process. After dissolving the B-component, a rapid quench leaves a metastable solid
solution. Additional heating or ageing will allow the metastable B-component to precipitate out in a controlled
manner.
This process is used to strengthen Al–Cu alloys. The initial precipitates are thin disc-
shaped clusters of Cu atoms called GP zones after Guinier and Preston who first identified
these structures. The GP-1 zones are 1–2 atoms thick and 25 atoms in diameter that form
platelets parallel to the (100) planes. These zones maintain the same lattice structure as the
Al matrix, but the smaller Cu atoms produce lattice strain, which provides some strength-
ening of the material.
As time progresses, these GP zones evolve into a new metastable tetragonal phase
(Al3Cu) called the u00 phase and precipitates in the form of platelets approximately 30 nm
in diameter and 2 nm thick. (Note that the a-phase in the Cu–Al system starts on the
Cu-rich side so the u phase is actually Al2Cu.) The a and b lattice parameters in the
tetragonal u00 phase match the 0.404 nm Al lattice exactly, but the c lattice parameter is
0.786 nm, which is slightly less than twice the length of the Al unit cell. Thus the u00 is
coherent on all faces, but the poor fit in the c-direction produces large lattice strains which
results in the very high strength of the alloy with this phase present.
With additional aging, the precipitate grows to 100 nm and the u00 phase gives way to
another metastable u0 (Al5Cu2) phase, which is also tetragonal with the same a and b
lattice parameters as before, but the c lattice vector is now 0.580 nm, so the coherence in
the c-direction is lost resulting in a decrease in strength. If the material is aged further, the
u0 phase goes into the stable Al2Cu u phase, which is also tetragonal, but the lattice
parameters no longer match the Al lattice and all coherence is lost. The dispersed u
particles still provide some strengthening by pinning dislocations, but the relative soft
Al2Cu is not a particularly good material for dispersion hardening. The loss of strength
that occurs after the peak strength that was attained with the formation of the u00 phase is
called overaging.
Why do the GP zones and then the u00 phase form before the u0 and the u phase?
Recall from Section 11.4 that the barrier to nucleation is directly proportional to the
interfacial energy. The small clusters of Cu atoms and coherent u00 phase have much
lower interfacial energy than the incoherent u phase. However, the u phase is the stable
phase so that eventually the metastable u00 and u0 will proceed to the stable u phase.
Transformation Kinetics 281
630, also known as 17–4 PH is the most common of the precipitation-hardened stainless
steels that are now beginning to replace Al in airframes because of their strength and
corrosion resistance.
Duplex stainless steels are a mix of austenitic and ferritic stainless steels. The idea is
to combine the strength of the ferritic steels with the corrosion resistance of the
austenitic steels.
8008C–9008C (above the eutectoid temperature) and slow cooling. This allows the cemen-
tite to convert to graphite, which takes the form of rosettes in a matrix of either an a-iron or
pearlite depending on the cooling rate.
Adding Si and slower cooling promotes graphite rather than cementite formation. Gray
irons have from 2.5–4 wt% C and the graphite exists in the form of flakes, resembling corn
flakes. Gray iron is relatively weak and brittle; the sharp edges of the graphite flakes can
act as stress risers. It is machinable and has excellent damping properties and is often used
as bases for large machinery, engine blocks, cylinder heads, and transmission housings.
Adding small traces of Mg or Ce promotes the formation of nodules of graphite rather
than flakes. This morphology eliminates the stress concentration near the graphite precipi-
tates and makes the iron more ductile. The strength of ductile cast irons approaches those
of steel. Ductile cast iron is used for pressure vessels, pump housings, gears, and rollers.
alloy contains 5% Al and 2.5% Sn for solid solution hardening and has a yield strength
of 784 MPa. Addition of V lowers the transition temperature, but not enough to change
the transition temperature to ambient. The Ti-10 V-2Fe-3Al alloy (10 wt% V, 2 wt% Fe, 3 wt%
Al) contains b stabilizers, which allow it to be heated into the b-region and rapidly cooled
to produce a metastable b-structure. It may then be age hardened to increase its strength
to 1150 MPa.
By combining various combinations of a and b alloys, it is possible to optimize the
desired properties. Ti-6Al-4 V is used for prosthetic devices because of its strength,
corrosion resistance, and biocompatibility. Ti-10 V-2Fe-3Al has the best combination of
strength and toughness. Its yield strength of 1150 MPa and ductility of 10% elongation is
tougher than any steel.
14.5.9 Superalloys
Special nickel-based superalloys with exotic names such as Astroloy, Hastelloy, and
Waspaloy, have been developed for high temperature turbine blades and research con-
tinues in this high payoff area. Even a small increase in operating temperature translates
into significant improvements in performance and huge savings in fuel. The superalloys
have very high creep resistance at temperatures as high as 70% of their melting temperat-
ures. This creep resistance is achieved by various combinations of solid solution hardening,
dispersion hardening, precipitation hardening, and grain boundary stabilization.
The host phase is a solid solution in the Ni-rich corner of the Ni-Al-Ti ternary phase
diagram (gamma phase). The fcc Ni is supersaturated with Al and Ti and also contains
Transformation Kinetics 285
a large fraction (up to 30%) of combinations of Cr, Mo, Ta, Re, and W for solid solution
hardening. These high density metals also act as diffusion inhibitors but care must be taken
to prevent their forming topologically close-packed phases, which would weaken the
structure. Cr is also effective in preventing corrosion. Co is added to reduce the stacking
fault energy in order to promote cross slip from the {111} planes to the {100} plane where
they become locked. Very small (submicron) oxides such as Y2O3 are sometimes added as
dispersion hardeners. Small quantities of B, Zr, Hf, and C may also be added to stabilize
the grain boundaries, although the problem of grain boundary slip can be eliminated by
growing single crystal turbine blades. Most of the newer jet engines use the third and
fourth generation single crystal alloys such as René N6 and the TMS-series for the turbine
blades in their high temperature sections.
The most important mechanism for improving high temperature creep resistance is the
precipitation hardening through the formation of large volume fractions (up to 70%) of
the g0 (gamma-prime) phase. The g0 can be Ni3Al, Ni3Ti, or Ni3(Al,Ti). The low interfacial
energy between the g and the g0 phase allows the g0 phase to grow as cubes that are
three-dimensionally coherent with the g host phase. Since coarsening of the precipitate is
driven by interfacial energy, this low interfacial energy prevents unwanted coarsening at
high temperatures that could destroy the coherency. As we saw in Chapter 5, the L12
structure of Ni3Al (as well as Ni3Ti) is primitive fcc with Ni on the face sites and Al on
the corners. This ordered structure together with the slight misfit between the g and g0
phases is very effective in preventing dislocation motion. The amount and sign of the
misfit can be controlled by the Al–Ti ratio. A negative misfit (the g0 is smaller than the
host phase) puts a strain of the precipitated phase, which is very effective in preventing
dislocation climb.
A second precipitating phase, known as g00 , can be formed by substituting V or Nb
(sometimes called Columbium [Cb]) for the Al to form Ni3V or Ni3Nb. These form a bct
structure with the V or Nb on the corner lattice sites. Although the coherency is lost in one
dimension (1-D), this phase is useful for strengthening alloys such as Inconel 718 where the
operating temperatures are not as extreme.
dVC
¼ 4pr2 drIV t ¼ 4pu3 t3 IV dt, (14:4)
VL
where VL is the volume of the untransformed melt. The volume of untransformed melt
VL is the original volume V0 times the probability that it has not yet been transformed in
time t or
VC VL
¼1 ¼ 1 exp kt4 , (14:5)
V0 V0
286 Introduction to the Physics and Chemistry of Materials
where
ðt
kt ¼ 4pu3 t3 IV dt ¼ pu3 t4 IV :
4
(14:6)
0
where
DG is the free energy difference between the solid and melt
a0 is an atomic dimension
n is the jump frequency given by Kingery (see Chapter 11)
kT
n¼ , (14:8)
3pa30 m
where m is the viscosity. At deep undercoolings, the DG=kT is large, so the limiting growth
rate is given by
a0 kT
u¼ : (14:9)
3pa30 m
F @vx
¼ Sxy ¼ m , (14:10)
A @y
where the viscosity m can be thought of as shear stress required to produce a velocity
gradient in the medium. (This is different to the shear modulus introduced in Chapter 8,
which related the shear force to the static strain.) This viscosity is sometimes called the
absolute viscosity or the dynamic viscosity. The SI unit for viscosity is kg=m s or Pa s. The
cgs unit is poise whose symbol is P and 10 P ¼ 1 Pa s. The viscosity of H2O is 0.01 P or
1cP ¼ 1 mPa s. To confuse matters even further, it is common to introduce a term called
kinematic viscosity which is the absolute viscosity divided by the density and symbolized
by Greek symbol v. It has units of m2=s and is call a Stoke, abbreviated St. If Equation 14.10
is written as
we see that n is the coefficient of momentum flux and Equation 14.11 takes the form of the
heat and mass transport equations.
The viscosity of a low density gas can be computed from the theory of an ideal gas and is
given by
pffiffiffiffiffiffiffiffiffi
2 mkT
m ¼ 3=2 , (14:12)
3p d2
where d is the molecular diameter. Note that the viscosity for low density gases increases
with the square root of temperature. This occurs when the mean free path is long compared
to the spacing between plates and the momentum transport is directly proportional to the
velocity of the gas molecules.
Things are quite different for condensed phases when molecules cannot move independ-
ently from each other. The Stokes–Einstein equation relates the diffusion coefficient to the
viscosity of a fluid:
kT
D¼ : (14:13)
6pmR
kT NA 1=3 1 DE
mð T Þ ¼ ¼ m0 exp : (14:14)
3p VM D0 exp½DE=RðT TG Þ Rð T T G Þ
Volume
Glass
al
Cryst
FIGURE 14.7
TG TM
Schematic illustration of the glass transition indicated by rate
of volume change with temperature. Temperature
VT
¼ 1 exp(kt4 ) ¼ 106 ¼ kt4 ¼ pu3 t4 IV : (14:15)
V
TABLE 14.1
Viscous Behavior of Glasses
Benchmark Viscosity (Pa s) Descriptive Behavior
13
Strain point 310 No plastic deformation, approximately equals the TG
Annealing point 1013 Diffusion fast enough to eliminate internal stresses
Softening point 4107 Can be deformed plastically
Working point 103 Easily deformed but does not flow
Melting point 10 Liquid enough to pour
Source: From Callister, W.D., Materials Science and Engineering: An Introduction, 7th
edn., 2007.
Transformation Kinetics 289
TM
Undercooled melt
Crystal
formation
A
FIGURE 14.8
Schematic T-T-T diagram for glass forma-
tion. If the cooling rate is too slow (cooling
TG curve A), a polycrystalline solid will form.
B Glass To form a glass, the cooling trajectory must
miss the nose of the T-T-T curve as does
Log time (s) the cooling curve B.
KT 4pr2crit
I0 ðT Þ ¼ N, (14:18)
3pa30 mðT Þ a20
where m(T) is given by the Vogel–Fultcher relationship (Equation 14.15). We use Kingery’s
model for I0, which is more appropriate for ceramics than Turnbull’s or Flemings, because
it is inversely proportional to viscosity and hence decreases with temperature as the
viscosity increases and goes to zero at the glass transition temperature.
The time for the crystallization to reach 1 ppm is found by solving Equation 14.15 for t,
!1=4
106
t(T) ¼ , (14:19)
puðTÞ3 IV ðTÞ
TABLE 14.2
Selected Thermophysical Properties of Glass-Orming Ceramics
Glass Soda Lime Quartz
2200
Tmelt 1050
2000
1000 Tmelt
1800
950
Temperature (K)
Temperature (K)
1600
900
1400
850
1200
800
1000 750
Tglass
800 700
Tglass
600 650
0 10 20 30 40 50 10 20 30 40 50
(a) Log time (s) (b) Log time (s)
FIGURE 14.9
T-T-T curve for (a) fused quartz calculated from Equation 14.18 and (b) T-T-T curve for soda-lime glass calculated
from Equation 14.18.
Although most glasses are based on silica, glasses can also be made from random chains
of the group VI elements, sulfur, selenium, and tellurium. The chains are twofold coord-
inated, but can be cross-linked by threefold or fourfold-coordinated elements such as
arsenic, antimony, or germanium. These chalcogenide glasses have interesting semicon-
ducting properties and photoconductive properties.
In an effort to develop optical fibers that transmit in the infrared, attention has turned to
replacing the oxygen and silicon in the oxide glasses with heavier elements such as fluorine
and various heavy metals to lower the infrared absorption frequency. The most studied of
these heavy metal fluoride glasses are the ZBLANs, which contain fluorides of zirconium,
barium, lanthanum, aluminum, and sodium. Compared to quartz, the melts of these
systems are much less viscous, making them reluctant glass formers and their glasses
have a tendency to devitrify (crystallize).
Fe80B11Si9 was later developed, which reduced the core loss to 30% of the best Si-steels
currently in use. However, the conventional transformers were already 99.7% efficient,
so cutting the core loss by a factor of 3 only made a marginal overall improvement. But
considering the amount of power going through transformers on the power grid, even a
fractional gain in efficiency translates into huge energy savings.
Splat quenching produces a cooling rate of 105 deg.=s but can only form small chips
that are 25–50 mm thick, suitable only for research purposes. Melt spinning (squirting a
jet of molten metal onto a rapidly spinning drum) enabled the production of amorphous
wires and thin ribbons. Allied Signal extended this technique to produce wide sheets,
which enabled the commercial production of metallic glasses. However, the requirement
for rapid solidification always restricts at least one dimension of the final product. In order
to produce bulk metallic glass, ways had to be found to relax the critical cooling rate.
The solution is to find alloying components that increase the entropy of fusion, increase
the viscosity of the melt, lower the melting temperature, and inhibit crystal formation.
Adding more components increases the number of states in the melt, which increases its
entropy and confuses the crystallization process. The addition of elements such as Co, Pd,
Zr, Nb to the more common metals increases the viscosity of the melt
Using containerless techniques to measure the thermophysical properties such as vis-
cosity, thermal conductivity, and entropy of fusion of metallic glass candidates in the
undercooled state, William Johnson at the California Institute of Technology was able to
formulate a system that could form a glass from a bulk melt. This effort led to the
development of a new class of bulk metallic glasses such as VIT-001 or Vitreloy, a Zr-based
bulk metallic metallic glass now being produced by LiquidMetal Technologies. Not only
are bulk metallic glasses stronger and more resilient than metals with grain structures, as
was discussed in Chapter 8, but the lack of grain boundaries makes them more resistant to
corrosion and other forms of chemical attack.
The effectiveness of such an amorphous structure in blocking dislocations can be seen in
Figure 14.10 which compares the elastic portions of the stress–strain diagrams of several
alloy systems with VIT-001 or Vitreloy. The mechanical properties of this alloy are com-
pared with other high strength alloys in Table 9.1.
Although the Young’s modulus of Vitreloy is lower than the stainless steel and the
titanium alloys, its yield strength is higher and its elastic limit is more than twice the others.
2000
pjwstk|402064|1435428860
1500
Stress (MPa)
1000
7075T6
500 Ti6AlV aged
17-7PH SS
440A SS tempered
FIGURE 14.10
VIT-001
Stress–strain diagram comparing the perform-
ance of a bulk metal glass with other high 0
strength materials. Data for these plots were 0 0.5 1 1.5 2 2.5
taken from Table 9.1. Elongation at yield (%)
Transformation Kinetics 293
Its resiliency is nearly three times larger than the stainless steel and titanium alloys. It is,
however, brittle and fractures with a shower of sparks when its elastic limit is exceeded.
Even so, taking the theoretical strength to be E=10, it achieves only 5% of its theoretical
strength for reasons that were discussed in Section 9.5.
14.7 Summary
Time and temperature-dependent transformations can dramatically alter a material’s
properties. Examples include the austenitic to ferritic transformation, followed by different
heat treatments to obtain the desired microstructure of steels, the precipitation hardening
of various alloys, annealing to improve ductility so that metals can be worked, grain
refining to obtain the optimum grain structure, and the formation of glasses. Given a
theoretical or an empirical transformation rate, the Avrami equation predicts the fraction of
the material that has been transformed in a given time. Isothermal time-temperature-
transformation (T-T-T) curves showing the onset, the halfway point, and the completion
of a transformation are developed to guide the heat treatment process.
The fcc g-phase of Fe (austenite) is stable above 7278C but there is a euctectoid reaction at
0.76 wt% C, where the austenite transforms into bcc a-ferrite and Fe3C cementite. The
lamellar structure of alternating layers of ferrite and cementite is called pearlite. Hypoeu-
tectoid steels (steels with lower carbon content than the eutectoid composition) are called
low carbon steels and are weaker but more ductile than the high carbon hypereutectoid
steels. The strength and ductility of steels also depends on the lamellar spacing of the
pearlite, which can be controlled by time and temperature profiles during cooling.
Rapid quenching of austenite results in a displacive or martensitic transformation in
which the atoms are instantly displaced or rearranged from fcc to bct, trapping the carbon
atoms in the process before they have time to diffuse away (also called a diffusionless
transition). Martensitic steel is extremely hard and brittle but can the tempered by heating
and ageing to increase the ductility.
Annealing is a process in which a material is heated to below the melting point in order
to remove the effects of work hardening in the case of metals, or to remove internal stresses
in the case of a glass.
Hardenable materials are alloy systems that are capable of being hardened by heat
treating. The alloying component for precipitating hardening must be capable of forming
a solid solution with the host above a certain temperature and then precipitate out at a
lower temperature, usually in the form of intermetallic phases. Maximum strengthening
occurs when these intermediate phases form structures that are coherent with the host
structure, but with slightly different lattice constants so as to strain the host lattice.
Glasses are formed when a material is cooled from the melt without crystallization
occurring. The material transforms from an undercooled melt to a rigid solid at the
glass transition temperature, which is accompanied by a change in the rate of thermal
expansion. Good glass formers generally have low melting points resulting from deep
eutectics and have high viscosities that increase rapidly as the temperature is reduced.
The T-T-T curve governing glass formation is a C-shaped curve representing 1 ppm
crystallization, which is the limit of detection of crystallinity in a glass. The T-T-T curve
lies between the melting temperature, where crystal formation is limited by the nucle-
ation barrier, and the glass transition temperature where the growth of crystals
that have nucleated is limited by the viscosity of the material. In order to form a
glass, the cooling curve must miss the nose of the T-T-T curve as it descends to the
glass transition temperature.
294 Introduction to the Physics and Chemistry of Materials
Metals can also form glasses if the cooling rate is fast enough, but rates in excess of
106 deg=s are required for most metals. By adding components to increase the viscosity and
working at deep eutectic points, some metallic systems can form glasses at cooling rates of
105 deg.=s, which can be attained by splat quenching or melt spinning. However, these
methods restrict the thickness of the metallic glass. Recently, new glass-forming alloys
have been developed that have reduced the cooling requirements so that bulk metallic
glasses can be formed.
Having no crystal structure gives metallic glasses unusual properties. They are strong
and have a high resiliency, but are subject to brittle fracture at their elastic limit. With no
grain boundaries subject to chemical attack, they are highly corrosion resistant. Also,
the lack of grain boundaries to pin magnetic domains makes magnetic metallic glasses
ideal for reducing hysteresis losses in transformer cores.
Bibliography
Boyer, H., Ed. Atlas of Isothermal Transformation and Cooling Transformation Diagrams, American
Society for Metals, p. 28 (1977).
Callister, W.D., Materials Science and Engineering: An Introduction, 7th edn., John Wiley & Sons,
New York, 2007.
Daniel, J.S., Introduction to advanced high strength steels, STAMPING J., Nov. 2004.
DeCristofaro, N., Amorphous metals in electric power distribution applications, Mater. Res. Soc. Bull.,
23(5) 1998, 50–56.
Güntherodt, H.-J. and Beck, H., Eds., Glassy Metals I. Topics in Applied Physics, 48, Springer-Verlag,
Berlin Heidelberg, and New York, 1981.
Hosford, W.F., Physical Metallurgy, Taylor & Francis, CRC Press, Boca Raton, FL, 2005.
Kingery, W.D., Bowen, H.K., and Uhlmann, D.R., Introduction to Ceramics, 2nd edn., John Wiley &
Sons, 1976.
Massalski, T.B., Senior Editor., Handbook of Binary Alloy Phase Diagrams, Vols. 1–3, American Society
for Metals, 1990.
Ratke, L. and Voorhees, P.W., Growth and Coarsening: Ostwald Ripening in Material Processing,
Springer, 2002.
Reed, R.C., The Superalloys: Fundamentals and Applications, Cambridge, 2006.
Reinbolt, J.A. and Harris, W.J., Affect of alloying elements on notch toughness of pearlitic steels,
Trans. Am. Soc. Metals, 43, 1951.
Russell, A.M. and Lee, K.L., Structure-Property Relations in Nonferrous Metals, John Wiley & Sons,
New York, 1966.
Sims, C.T., Stoloff, N.S., and Itagel, W.C., Superalloys II, John Wiley & Sons, 1987.
Problems
1. I want to make a high carbon steel with 0.078 wt% Fe3 C. Using the Fe-Fe3 C phase
diagram (Figure 14.2)
(a) What will be the final composition in terms of proeutectoid Fe3C and pearlite?
(b) What will be the composition of the pearlite?
2. Do a sensitivity analysis of the parameters in Table 14.2 on the ability to form a glass.
Which parameter has the largest influence on the cooling rate required?
Transformation Kinetics 295
3. Construct a T-T-T diagram for making a glass from pure Al. Pure metals do not show
significant temperature variations near the melting point, so do not use the Vogler–
Fultcher relation. Instead, take the viscosity as 0.01 Pa s at the melting point. What is the
critical cooling rate required for glass formation?
4. Estimate the cooling rate for a material at temperature T with thermal diffusivity k and
thickness d after it is splatted against a cooling block at temperature T0. (Make whatever
reasonable simplifying assumptions necessary.) What thickness would be required to
achieve a cooling rate of 106 deg=s?
15
Distribution Functions
In the subsequent chapters in which we will be investigating the thermal, electrical, optical,
and magnetic properties of materials, it will be necessary to be able to determine the energy
distribution of electrons, holes, photons, and phonons. To do this, we need to introduce
some quantum statistical mechanical concepts in order to develop the distribution func-
tions needed for this purpose. We will develop the Bose–Einstein (B–E) distribution
function that applies to all particles except electrons and holes (and other fermions) that
obey the Pauli exclusion principle and show how this function becomes the Maxwell–
Boltzmann (M–B) distribution in the classical limit. Also, we will show how the Planck
distribution results by relaxing the requirement that particles be conserved. Next we
develop the Fermi–Dirac (F–D) distribution that applies to electrons and holes and
becomes the basis for understanding semiconductors and photonic systems.
297
298 Introduction to the Physics and Chemistry of Materials
For example, 1c, 2, 3bd, 4ae, . . . would correspond to particle c in cell 1, nothing in cell 2,
particles b and d in cell 3, etc.
(gi þ Ni )!
Wi ¼ : (15:1)
gi ! N i !
Now we must determine the most probable distribution of the N particles among the Ei
energy states with the following constraints:
P
1. Number of particles constant. N ¼ Ni ¼ constant.
P
2. Total energy is constant. U ¼ Ei Ni ¼ constant.
Q
3. Most likely distribution. W ¼ Wi ¼ maximum value.
P Q P
This last constraint can be enforced by maximizing ln Wi since lnð Wi Þ P
¼ ln Wi .
This corresponds to requiring the entropy, which is defined as S k ln W ¼ k ln Wi , to
be a maximum, which is the same as requiring the system to be in thermodynamic
equilibrium.
The above conditions can be rewritten as
dN ¼ Sd Ni ¼ 0
dU ¼ S Ei dNi ¼ 0 (15:3)
d(ln W) ¼ Sd(ln Wi ) ¼ 0:
Taking the variation of Equation 15.2 with respect to Ni (same as taking the derivative with
respect to Ni),
X
gi þ N i
ln dNi ¼ 0: (15:5)
Ni
X gi þ Ni
ln l bEi dNi ¼ 0: (15:6)
Ni
Since dNi 6¼ 0 (we are free to interchange particles between cells so long as the three
constraints are satisfied),
gi þ N i
ln ¼ l þ bEi
Ni
gi
Ni ¼ : (15:7)
e bEi þl 1
X X
gi þ N i
dS ¼ k d( ln Wi ) ¼ k ln
Ni
X X X
¼k (bEi þ l) dNi ¼ kb Ei dNi þ kl dNi
or
We can expand dS as
@S @S @S
dS ¼ dU þ dV þ dN: (15:9)
@U V,N @V U,N @N U,V
1 m
dS ¼ dU þ dN: (15:11)
T T
300 Introduction to the Physics and Chemistry of Materials
1 m
b¼ and l¼ : (15:12)
kT kT
Thus the B–E distribution function becomes
gi
Ni ¼ : (15:13)
e(Ei m)=kT 1
The chemical potential m can be expressed in terms of absolute activity A ¼ em=kT and the
B–E distribution is written in the form
gi
Ni ¼ : (15:14)
A1 eEi =kT 1
which is the (M–B) distribution function. Note that this implies that A1 is very large which
means that A must be very small.
In this classical limit, we can compute A from the requirement that
X X
N¼ Ni ¼ A gi eEi =kT : (15:16)
To proceed, we must evaluate gi which is the number of states that have energies corre-
sponding to Ei. It is more convenient to convert the sum to an integral, which we can do
by replacing gi with the density of states; i.e., g(E)dE which is simply the number of states
with energies in the range of E to E þ dE. Since E ¼ p2=2m, the volume in momentum
space corresponding to energies between E and E þ dE is the volume between two con-
centric spheres with radii p and p þ dp. Thus, Dp3 ¼ 4pp2 dp, where p2 ¼ px2 þ py2 þ pz2.
Since pdp ¼ mdE,
The total volume of a cell in phase space is Dx3Dp3 ¼ h3, where Dx3 ¼ V, the total volume of
the system. Therefore, the number of cells of phase space between E and E þ dE is
Putting this in Equation 15.16 and replacing the sum with an integral,
ð
1
4p21=2 m3=2 V
N¼A E1=2 eE=kT dE
h3
0
3=2 3=2
2 (pmkT) V
¼A (15:19)
h3
Distribution Functions 301
or
nh3
A¼ : (15:20)
(2pmkT)3=2
gi !
Wi ¼ : (15:22)
Ni !(gi Ni )!
X gi Ni
ln l bEi dNi ¼ 0 (15:24)
Ni
from which
gi Ni
ln ¼ l þ bEi
Ni
or
gi
Ni ¼ : (15:25)
elþbEi þ 1
gi
Ni ¼ : (15:26)
e(Ei m)=kT þ 1
1
F(E) ¼ : (15:27)
e(Ei m)=kT þ 1
which can be interpreted as the probability of a state with energy E being occupied
at temperature T. A plot of the Fermi function at different temperatures is shown in
Figure 15.1.
The density of states is found as before, except that we have made the cells in phase
space half as small to assure no more than single occupancy. Thus Equation 15.18 becomes
2.0
1.5
F(E,T )
m/kT = 10
0.5
FIGURE 15.1 0
Fermi function as a function of E=m for different values of 0 0.5 1.0 1.5 2.0
m=kT. At T ¼ 0, F(E,T) ¼ 1 for E < m, and 0 for E > 0. E/m
Distribution Functions 303
However, if we put the numbers corresponding to electrons in a metal, say Al, the electron
density n ¼ 1.8 1023 electrons=cm3 and m ¼ 9.11 1028 g into Equation 15.20, we see that
the value for A computed in the classical manner is 1.6 104, which of course violates the
assumption A 1 that was made in taking the classical limit. Therefore, F–D statistics must
be used in treating electrons in a metal.
ð
1
g(E) dE
N¼ (15:29)
e(Em)=kT þ 1
0
and solving for m. This cannot be done in closed form. However, at T ¼ 0, the Fermi
function is 1 for E < m0 and 0 for E > m0. Therefore, the integral reduces to
ð0
m ð0
m
4p23=2 m3=2 V 4p23=2 m3=2 V 2 3=2
N¼ g(E)dE ¼ E 1=2
dE ¼ m (15:30)
h3 h3 3 0
0 0
from which
2=3
1 3h3 N=V h2 2 2=3
m0 ¼ ¼ 3p N=V : (15:31)
2m 8p 2m
The chemical potential at 0 K is the same as the Fermi energy which we obtained in
Chapter 2 by summing over the states in the ‘‘electron in a box problem.’’ The Fermi
energy is a constant for a given material.
Sommerfeld obtained a series solution for Equation 15.29 which provides a good
approximation of the chemical potential for TF T,
p2
m ¼ m0 1 ðT=TF Þ2 þ , (15:32)
12
where the Fermi temperature TF is given by kTF ¼ m0. Since m0 ¼ O(5 eV), which is equiva-
lent to TF
50,000 K, this approximation will be good for most of our applications.
The chemical potential of electrons in a Fermi distribution is also called the Fermi level.
The energy required to remove an electron from the Fermi level to infinity (the vacuum
state) is the work function. Since the difference in chemical potential determines the flow of
particles, when two materials with different Fermi levels are brought together as illustrated
in Figure 15.2, electrons will flow from the material with the higher Fermi level (smallest
work function) to the material with the lower Fermi level until equilibrium is reached. This
transfer of charge results in the contact potential between the two materials.
In a metal at 0 K, the Fermi energy is the ground state energy of the sea of electrons. It is
the energy level above which no states are occupied and below which all states are
occupied. The Fermi level can also be defined as the energy at which the probability of
higher states being occupied is the same as lower states being unoccupied. With this
304 Introduction to the Physics and Chemistry of Materials
Vacuum level
f1
f1
f2 e− f2
EF1 EF
EF2
FIGURE 15.2
When material 1 with workfunction f1 and Fermi level EF1 is brought into contact with material 2 with work-
function f2 and Fermi level EF2, the vacuum levels line up and electrons flow from the material with the higher
Fermi level (chemical potential) to the lower Fermi level (chemical potential) until equilibrium is reached and the
Fermi levels in the two materials are the same.
definition, the Fermi level of a semiconductor will lie, not at the valence band, but in the
bandgap between the valence and conduction bands (see Chapter 20).
15.5 Summary
Distribution functions are derived for various types of particles and waves. The distribu-
tion function specifies the number of particles that will that occupy a given energy state for
a given temperature. Such distribution functions can be used for determining the average
energy for an ensemble of particles (or waves) by multiplying the distribution function
times the energy and summing (or integrating) over all energies. The distribution function
contains two parts, the density of states, which is the number of states corresponding to a
given energy, and the probability that those states are filled.
All particles can be classified into two categories, bosons and fermions. Bosons include
protons, neutrons, most atoms, and molecules. They have integral units of spin and have
no restriction on the number of particles that can occupy a quantum state. They obey B–E
statistics and the B–E distribution function is given by Equation 15.13,
gi gi
Ni ¼ m)=kT
¼ , (15:33)
e (Ei 1 A e i =kT
1 E 1
where
Ni is the number in energy interval Ei
gi is the number of available states in interval Ei
A ¼ em=kT
In the classical limit, A becomes small so that the 1 in the denominator can be neglected
and the distribution reduces to the M–B distribution given by
P Ð 3=2 E=kT
1
Ei gi AeE=kT E e dE
Etotal 0 3
hEi ¼ ¼ iP ¼ Ð 1=2 E=kT
1 ¼ kT: (15:35)
N gi AeE=kT 2
i E e dE
0
In order to be within the classical limit under which M–B statistics apply, A ¼ nh3 =(2pmkT)3=2
must remain small. For an ideal gas such as N2 at STP, A
O(107). For condensed matter
in which the number density, n
104 larger than for an ideal gas, M–B statistics can still
be used provided the temperature is not too low. At cryogenic temperatures, B–E statistics
are required.
The number of wave-like particles, such as photons and phonons is not conserved since
they can be created and destroyed. To lift the requirement that the number of particles be
preserved, the chemical potential is set to zero and the B–E distribution reduces to the
Planck distribution given by Equation 15.21,
gi gi
Ni ¼ ¼ : (15:36)
eEi =kT 1 ehv=kT 1
Fermions include electrons, protons, neutrons, and 3He. They have half-integral units of
spin and the Pauli exclusion principle prevents more than one particle from having the
same quantum state. As a consequence, such particles must be treated using the F–D
distribution function, given by
gi
Ni ¼ : (15:37)
e(Ei m)=kT þ 1
Because the mass of the electron is so small, the A ¼ em=kT will generally be much too large
to ignore the þ1 in the denominator.
The Fermi function F(E) defined as
1
F(E) ¼ (15:38)
e(Ei m)=kT þ 1
can be interpreted as the probability of a state with energy E being occupied at temperature
T. As T approaches 0, the Fermi function is 1 for E < m0 and 0 for E > m0. In this limit
ð0
m ð0
m
g(E)dE
N¼ ¼ g(E)dE: (15:39)
e (Em)=kT þ1
0 0
2 2
h
m0 ¼ 3p N=V : (15:40)
2m
The chemical potential at 0 K is defined as the Fermi energy which we obtained in Chapter
2 by summing over the states in the ‘‘electron in a box problem.’’ The Fermi energy is a
constant for a given material. The chemical potential, which is a function of temperature
and the number of free electrons in the material, is frequently called the Fermi level,
especially when dealing with materials with a band gap (see Chapter 20).
306 Introduction to the Physics and Chemistry of Materials
Appendix
A.15.1 Derivation of the Identities
The derivation of the identities
@S 1 @S m
¼ and ¼ (A:15:1)
@U V,N T @N U,V T
or
@S 1
¼ Q:E:D: (A:15:5)
@U V,N T
For constant U and V, this may be combined with the expansion of dG to give
@G
dG ¼ dN ¼ mdN: (A:15:8)
@N U,V
which, for constant U and V, may be combined with the above result to give
or
@S m
¼ Q:E:D: (A:15:13)
@N U,V T
!3=2
h2
2p
m ¼ kT ln þ kT ln (N=V): (A:15:15)
mkT
!3=2
2ph2
G ¼ Nm ¼ NkT ln þ NkT ln (N=V): (A:15:16)
mkT
5 G
S ¼ NkT : (A:15:17)
2 T
Since for photons the energy E ¼ cp, the allowed energy states are
nphc qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
E¼ , n¼ n2x þ n2y þ n2z : (A:15:20)
L
The lowest values of n fall within a sector of a sphere of radius n in nx, nx, nx
space. The density of states in n-space is given by 2 1=8 4pn2dn. The extra 2 accounts
for the two possible polarization states for each wave. The density of states in energy
is given by
LE 2 dn L 3 2
gðEÞdE ¼ p dE ¼ p E dE: (A:15:21)
p
hc dE phc
Using Planck’s distribution function, the radiation density in the cavity is given by
EgðEÞdE E 3 dE
uðEÞdE ¼ ¼p : (A:15:22)
V ðexp (E=kT) 1Þ phc ðexp (E=kT) 1Þ
Setting E ¼
hv ¼ hn, we can write the distribution in terms of v,
v 3 dv
h 8phn3 1
uðvÞdv ¼ p ¼ dn: (A:15:23)
pc ðexp (
hv=kT) 1Þ c 3 ðexp (hn=kT) 1Þ
The spectral radiance is the amount of radiation emerging from a small hole in the cavity.
Recall from kinetic theory, the number of molecules striking an area is given by nv=4,
where n is the number density and v is the average velocity. Similarly, the photon flux
emerging from a hole in the cavity is nc=4 and the energy flux will be u(n)c=4 and from
Equation A.15.23
cuðnÞdn 2phn3 1
RðnÞdn ¼ ¼ dn: (A:15:24)
4 c2 ðexp (hn=kT) 1Þ
Distribution Functions 309
In terms of l
We can integrate the spectral radiance in terms of frequency given by Equation A.15.24 by
using the integral representation of a Riemann–Zeta function
ð
1
X1
x3 dx 1 p4
¼ 6 ¼ (A:15:27)
ex 1 n¼1
n4 15
0
to obtain
ð
1
2p5 k4 T 4
R¼ RðnÞdn ¼ ¼ sT 4 : (A:15:28)
15c2 h3
0
The Stefan–Boltzmann constant s ¼ 2p5 k4 =15c2 h3 was computed in 1900 by Planck from
the radiation data that was available which allowed him to determine the value for the
Planck constant. His value of 6.56 1034 J s agreed very well with the value that Milliken
determined in 1915 from the photoelectric effect and the presently accepted value of
6.626 1034 J s.
Bibliography
Constant, F.W., Theoretical Physics, Addison-Wesley, Reading, MA, 1958.
Kittel, C. and Kroemer, H., Thermal Physics, W.H. Freeman, San Francisco, CA, 1980.
Ragone, D.V., Thermodynamics of Materials, Vol. II, John Wiley & Sons, New York, 1995.
Sears, F.W., Thermodynamics, Addison-Wesley, Reading, MA, 1953.
Problems
1. Under what circumstances is it permissible to approximate the B–E distribution with the
M–B distribution when dealing with solids?
2. Under what circumstances is it permissible to approximate the Fermi distribution with
the M–B distribution when dealing with solids?
3. Under what circumstances is it permissible to approximate the Planck distribution with
the M–B distribution when dealing with solids?
16
Lattice Vibrations and Phonons
The vibrations of the atoms in the crystalline lattice are important in understanding the
thermal properties of both metallic and nonmetallic solids. The energy involved in these
vibrations represents thermal energy; hence lattice vibrations are primarily responsible for
the heat capacity of solids. Also, these vibrations are able to transport heat and are the
dominant source of thermal conductivity in nonmetals. Therefore, in order to understand
thermal properties of solids, it is necessary to start with a general understanding of the
nature of lattice dynamics.
@2u @2u
rADx ¼ C 11 ADx
@t2 @x2
or
@ 2 u C11 @ 2 u
¼ , (16:1)
@t2 r @x2
which we recognize as a wave equation in which the wave is propagating with velocity
sffiffiffiffiffiffiffi
C11
v0 ¼ : (16:2)
r
The solution can be written in the form of a traveling wave, u(x, t) ¼ j expi(vt kx).
Differentiating this twice with respect to t and to x and putting the derivatives back into
311
312 Introduction to the Physics and Chemistry of Materials
Δx u
FIGURE 16.1
Schematic for analyzing waves in a linear homogeneous medium. The strain at x is e(x) ¼ @e=@x and at x þ Dx ¼ e
(x) þ @e=@xDx.
the wave equation, we see that v2 ¼ C11=rk2 which gives us a dispersion relationship
pffiffiffiffiffiffiffiffiffiffiffiffi
v ¼ C11 =r k ¼ v0 k that plots as a straight line with slope v0 on an v vs. k plot. Recall
that free electrons have a parabolic dispersion relationship (v k2). This is because
for a particle with mass, E ¼ hv ¼ p2=2m ¼
h2k2=2 m, or v ¼ hk2=2m. However, the momen-
tum of a free massless particle such as a photon is p ¼ hk ¼ E=c ¼ hv=c, so the dispersion
will be a straight line with slope c, which for photons, represents the velocity of light.
Therefore, a wave in a continuous elastic media would also be expected to have a straight-
line dispersion relationship whose slope is the velocity of propagation.
where b is the spring constant. Note that the force is proportional to the relative displace-
ment of the atoms. (If all were displaced the same amount, there would be no force.) Now
the equation of motion becomes
in which the x in the previous solution has been replaced by na, where n is the particle
index and a is the equilibrium spacing between atoms. Substituting this back into the
equation of motion,
v2 mjei(vtkna) ¼ bjeivt eik(nþ1)a þ eik(n1)a 2eikna : (16:6)
FIGURE 16.2
Schematic for analyzing waves on chain of
un −1 un un +1
like atoms.
Lattice Vibrations and Phonons 313
The reader should recognize exp(ika) þ exp(ika) as 2 cos(ka), and from the double angle
formula in trigonometry, cos(2u) 1 ¼ 2 sin2(u). Therefore, the solution requires
or
pffiffiffiffiffiffiffiffiffiffiffiffi
v¼ 4b=m sin (ka=2): (16:7)
This is quite a different dispersion relationship than we had before. For kap 1 (wave-
ffiffiffiffiffiffiffiffiffiffiffi
ffi
length long compared with atomic spacing), the sin(ka=2) ka=2 and v 4b=m ka=2.
Going from the microscopic to the macroscopic, we can replace m with rAa. F ¼ bdu ¼ C11
Ae ¼ C11A du=a, we can identify b as C11A=a. Thus
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
4C11 A=a pffiffiffiffiffiffiffiffiffiffiffiffi
v ka=2 ¼ C11 =rk, ka 1, (16:8)
rAa
which is the same as the result from the linear homogeneous model. Thus if the wavelength
is much longer than atomic dimensions, the discreteness of the atoms makes little differ-
ence, and the material can be treated as homogeneous. pffiffiffiffiffiffiffiffiffiffiffiffi
However, note what happens as ka ! p. The sin(p=2) ¼ 1 and the v ! 4b=m and
no longer involves k. This means the group velocity vg ¼ @v=@k ¼ 0, and this is the
maximum frequency the lattice can propagate. Note also that as k ! p=a, the solution
becomes un(x, t) ¼ j exp i(vt np), which represents a standing wave. Physically, what is
happening at the lower frequencies is the atoms are more or less moving in the same
direction as the wave propagates through, but as k ! p=a, the atoms are vibrating in
opposition to one another.
FIGURE 16.3
Dispersion relationship for linear chain
of like atoms. The v(k) repeats over an
−2π/a −π/a 0 π/a 2π/a interval of 2p=a, so all of the pertinent
information is contained between p=a
Reciprocal lattice vector k and p=a.
314 Introduction to the Physics and Chemistry of Materials
of the nearest neighbor reciprocal lattice points. It can easily be shown that the Bragg
condition 2d sin u ¼ l is satisfied at k ¼ p=a by setting u ¼ p
p=2, dffi ¼ a, and k ¼ 2p=l.
ffiffiffiffiffiffiffiffiffi
At the lower frequencies, the velocity is given by n0 ¼ a b=m, which is the velocity of
sound in the medium, typically 3 103 m=s. The cutoff frequency, or the ffi maximum
pffiffiffiffiffiffiffiffiffi
frequency that can be propagated by the lattice is given by vmax ¼ 2 b=m ¼ 2v0 =a
(2 3 105 cm=s)=(3 108 cm) ¼ 2 1013 Hz, which corresponds to the far-infrared
region of the spectrum.
2p
k¼ j, j ¼ 1, 2, 3, . . . : (16:9)
L
The only values of k that produce unique frequencies lie between p=a and p=a. If the
allowed values of k are spaced 2p=L apart, the total number of states is
Thus we arrive at the important conclusion that a finite chain having N þ 1 particles has N
normal modes of longitudinal vibration. The same arguments used above can be general-
ized to three dimensions and it can be shown that two transverse modes of vibration are
also possible. Therefore, a chain of N þ 1 atoms can have 3N discrete modes of vibration.
Thus the k-vector, which was continuous for an infinite chain, is now quantized into 3N
discrete states. The energy of each of these states is assigned to be hv. Because of the close
resemblance of these quantized vibrational states to photons in a cavity, these quantized
waves are called phonons.
Now we have a coupled set of second-order differential equations that must be solved
simultaneously. Notice that coupling comes about through the odd nearest neighbors of
the even atoms and vice versa. The solutions can be written as
Lattice Vibrations and Phonons 315
which have to be solved simultaneously. Identifying (exp (ika) þ exp (ika)) ¼ 2 cos (ka) and
writing these equations in matrix form, we get
(2b mv2 ) 2b cos (ka) j 0
¼ : (16:16)
2b cos (ka) (2b Mv2 ) h 0
If the inverse of the above matrix exists, the only solution possible is for j and h ¼ 0, which
is the trivial solution. The only other way to satisfy this equation is for the inverse matrix
not to exist, which requires its determinant to vanish. Thus the solution to the simultaneous
differential equations requires that
or
4b2
mv4 2bv2 ¼ sin2 (ka), (16:18)
mþM
FIGURE 16.4
Dispersion relationship for chain of alternating atoms with different
masses. For this plot, M ¼ 1.2 m. There is an energy gap at p=2a
−π/a 0 π/a which closes as m ! M. The frequencies in this gap cannot be propagated
by the chain. If M ¼ m, the gap would close and the lower branch would
Reciprocal lattice vector k be the same as Figure 16.3.
316 Introduction to the Physics and Chemistry of Materials
2b 2b
v2 ¼ ; ðkaÞ2 : (16:20)
m mþM
The first solution represents the cutoff frequency, or the highest frequency the lattice can
support. The second solution represents the low frequency limit where the v is linear with
k and the slope is the velocity of sound in the medium given by
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2b
v0 ¼ a: (16:21)
(m þ M)
One can see that if m ¼ M, these solutions reduce to those obtained for the chain of identical
atoms, as they should.
Next consider the case where ka p=2. Now, with some algebraic manipulation, the two
solutions are
v2 ¼ 2b=m; 2b=M: (16:22)
− + − + − + − + − k
FIGURE 16.5
Electromagnetic wave propagating along a chain of oppositely charged ions can excite a transverse optical (TO)
mode. To excite a longitudinal optical (LO) mode, the k-vector must make an angle with the chain in order to
project a component of its E-vector along the chain (Berreman effect).
In ionically bonded systems, the heavy and light ions will have different charges, which
means they can be excited by an electromagnetic wave as illustrated in Figure 16.5. A wave
propagating along a chain of crystals can excite a transverse-polarized mode, but not a
longitudinal-polarized mode because its E-vector would have no component along the
chain. In order to excite a longitudinal mode in a crystal, the incident radiation must enter
obliquely to one of the (100) faces. Only the portion of the wave polarized normal to the
face can excite the longitudinal mode.
A photon–phonon interaction requires the conservation of both energy and momen-
tum. Recall the momentum of a photon is hv=c and the momentum of a phonon is hv=v,
where v is the velocity of sound. Because of the much larger momentum=energy of the
acoustical phonons, exchange between photons and acoustical phonons is limited to
inelastic collisions in the form of radiant heating. However, in the vicinity of k ¼ 0
in the optical branch, the energy=momentum of phonons can more easily match that of
photons and there will be a strong absorption peak near the top of the optical branch
in the vicinity of k ¼ 0 where the frequency is (2b=m)1=2. This phenomenon is discussed in
more detail in Chapters 23 and 24.
It should be noted that sp3-bonded materials such as Si and Ge also exhibit both an
optical and an acoustic branch even though they contain only one type of atom. The
covalent sp3 bond places much of the electron density between the atoms so that the charge
distribution is similar to that of an ionic system.
The optical and acoustic modes for NaCl were computed from the elastic coefficients in
Table 16.1 and displayed in the reduced zone scheme in Figure 16.6. As k increases beyond
p=2a, the dispersion curve is the same as for kp=a; therefore, all the pertinent information
is contained in the reduced zone from 0 kp=2a. These optical and acoustic modes can be
mapped by neutron inelastic scattering.
TABLE 16.1
Comparison of Calculated and Measured Optical Cut Off Frequencies
C11 (GPa) C44 (GPa) a (nm) m (amu) vT (cal) vT (act) vL (cal) vL (act)
w
LO
3 ⫻ 1013
2 ⫻ 1013 LA
FIGURE 16.6 TO
Dispersion relations for NaCl in the reduced zone scheme for
the various modes of oscillation: longitudinal optical (LO),
transverse optical (TO), longitudinal acoustic, and transverse 1 ⫻ 1013 TA
acoustic. The curves were computed from Equation 16.19
using the elastic coefficients for NaCl given in Table 16.1. It
should be noted that the a in this figure is the distance
between the individual ions. Usually this dispersion relation 0
−π/2a 0 π/2a
is shown plotted between p=a and p=a where a is taken as
the distance between ion pairs. Reciprocal lattice vector k
instead of p=a to p=a as before as seen in Figure 16.6. So now the extent of k-space is p=a
and the spacing between each node is 2p=L, which means the number of modes is L=2a.
Since a is the nearest neighbor distance, L=a is the number of ions and L=2a is the number of
ion pairs. The number of modes is now 3N (two transverse and one longitudinal) in the
acoustic branch and 3N (two transverse and one longitudinal) in the optical branch, where
N is the number of ion pairs. (The result obtained here would be the same if we had defined a
as the distance between repeating ions, which would be the length of an ion pair. The k-space
would extend over 2p=a and the number of modes would be L=a ¼ N ion pairs.)
C11 ðm þ MÞ C44 ðm þ MÞ
b11 ¼ and b44 ¼ , (16:24)
2ra2 2ra2
where a is the nearest neighbor distance. For the NaCl lattice, r ¼ 4(M þ m)=(2a)3, so
Equation 16.24 simplifies to
16.5 Applications
One of the more useful predictions from the model is that the absorption peak from
Equation 16.20 is proportional to the square root of the lattice stiffness divided by the
reduced mass of the ion pairs. In the design of glasses and other optical components that
must operate in the far-infrared, one wants to operate at frequencies well above the
pffiffiffiffiffiffiffiffiffiffiffi
absorption at v ¼ 2b=m. Therefore, one looks for materials with heavy ion pairs and
low sound velocities. Single crystalline NaCl is used for infrared windows out to about
15 mm while KBr is good out to about 25 mm. KRS-5, (thallium bromoiodide) is used for
infrared applications out to 35 mm. The development of heavy metal glasses is discussed
further in Chapters 14 and 24.
16.6 Summary
For continuous p media,
ffiffiffiffiffiffiffiffiffiffiffiffithe propagation velocities for longitudinal and transverse waves is
pffiffiffiffiffiffiffiffiffiffiffiffi
given by vL ¼ C11 =r and vT ¼ C44 =r and the dispersion relation is v ¼ v0k. For a
monatomic
pffiffiffiffiffiffiffiffiffiffiffiffi chain of atoms, the dispersion relation is given by Equation 16.7,
v ¼ 4b=m sin(ka=2). The ratio of b=m can be related to the elastic stiffness coefficient
C11 (or C44) by b=m ¼ C11=ra2, where a is the spacing between the atoms
ffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffi ffi in the chain. There
is a cutoff frequency at k ¼ p=a given by vL ¼ 4b=m ¼ 4C11 =ra2 ¼ 2vL =a.
The dispersion relation simply repeats itself between p=a > k > p=a, which is the first
Brillouin zone for a simple cubic lattice, so that all of the relevant information is contained
in this zone. When a finite length L of the chain of N atoms is specified, the dispersion
relation is quantized into N modes that are spaced at 2p=L apart throughout the interval
p=a > k > p=a. Each state has energy hv where the v is related to k by the dispersion
relation. Traveling waves with these discrete values of k are called phonons. Since a wave
can have three polarization states, two transverse and one longitudinal, there are 3N
normal modes.
320 Introduction to the Physics and Chemistry of Materials
for the NaCl lattice. Comparison of this result with actual measurements for several alkali
halides shows only fair agreement.
Materials for infrared applications must have their absorption peak well below the
frequency they must transmit. This requires heavy atoms with small elastic coefficients.
Bibliography
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia, 1976.
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley & Sons, New York, 1966.
Problem
1. For a diatomic chain of atoms, use Equation 16.19 to obtain Equation 16.22 and find the
energy gap.
17
Thermal Properties of Solids
Now that we know something about how the molecules in a solid vibrate, we are in a
position to connect these vibrations to the thermal properties such as heat capacity, thermal
conduction, and thermal expansion.
For MB statistics, N(E) ¼ AeE=kTg(E)dE and for a three-dimensional (3-D) system, recall
that g(E)dE ¼ CE1=2 dE where C is a constant. Putting this in the above,
ð1 ð1
3=2 E=kT
p4 ep =2mkT dp
2
E e dE
1
hEi ¼ ð01 ¼ ð0 : (17:2)
1=2 E=kT 2m 1 2 p2 =2mkT
E e dE pe dp
0 0
The kinetic energy of a molecule with a momentum component p is p2=2m and dE ¼ pdp=m.
Carrying out the integration,
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 3=8 32p(mkT)5 3
hEi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ kT: (17:3)
2m 2
1=4 8p(mkT)3
321
322 Introduction to the Physics and Chemistry of Materials
Now consider a 1-D gas confined to a tube of length, L. In this case the volume in
phase space, DxDpx ¼ h. The number of states between px and px þ Dpx is given by
LDpx=h or
m1=2 LdE
g(E)dE ¼ : (17:4)
h(2E)1=2
It is useful to remember that a distribution function such as g(E)dE is the number of states
between E and E þ dE. As such it must be dimensionless. It is also useful to remember that
if there is a relationship between two variables such as p2 ¼ 2mE so that dE and dp can be
related, then their distribution functions can be written as g(E)dE ¼ g(p)dp.
Now the average energy per molecule is
ð1 ð1
1=2 E=kT
p2x epx =2mkT dpx
2
E e dE
1
hEi ¼ ð 10 ¼ ð0 :
1=2 E=kT 2m 1 p2x =2mkT
E e dE e dpx
0 0
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3=2
1 1=4 8p(mkT) kT
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ : (17:5)
2m 2
1=2 2p(mkT)1=2
Similarly one can show that a molecule of gas with two degrees of freedom has average
energy of kT. This example illustrates the classical theorem of equipartition of energy,
i.e., thermal energy equal to kT=2 is assigned to each degree of freedom. (It can also be
shown that the same partition of energy per degree of freedom can be assigned to each
rotational and vibrational mode provided the temperature is high enough to excite that
particular mode.)
The total energy of an oscillator such as an atom in a crystal lattice is the sum of the
kinetic and potential energies. The kinetic energy is just px2=2m. The potential energy is
ðx ðx
kx2
Potential energy ¼ F dx ¼ kx dx ¼ : (17:6)
2
0 0
pffiffiffiffiffiffiffiffiffi
Since the natural frequency v ¼ k=m, the potential energy is given by mv2x2=2, and the
total energy may be written as E ¼ p2x =2m þ mv2 x2 =2.
In thermal equilibrium, the average energy can be found by
ð1 ð1
p2x epx =2mkT dpx x2 emv x =2kT dx
2 2 2
1 0 mv2 0
hEi ¼ ð þ ð1 : (17:7)
2m 1 p2x =2mkT 2 mv2 x2 =2kT
e dpx e dx
0 0
We previously showed that the first term was kT=2. By a similar process, it can be
shown that the potential term also becomes kT=2. Therefore, the average energy for the
x-component of an oscillator is kT. The y- and z-component also each contribute kT, so
the total average energy is 3kT. Therefore, the equipartition theorem assigns kT=2 to
each of the six degrees of freedom, three kinetic and three potential, of an atom in a
crystal lattice.
Thermal Properties of Solids 323
@ hEi
CV ¼ ¼ 3k per atom or 3R per mole, (17:8)
@T
where R is the gas constant, which is Avogadro’s number times the Boltzmann constant.
Since R 2 kcal=mol, one would expect the heat capacity of solids to be 6 kcal=mol.
The observed heat capacity of most solids approaches this value at moderate to high
temperatures and this value is called the Dulong–Petit or classical limit of heat capacity.
However, the observed heat capacity is much lower at low temperatures and, for non-
metals, approaches 0 with a T3 dependence.
Metals provided additional problems for the classical theory of heat capacity. Metals are
generally much better conductors of heat than nonmetals because most of the heat is
carried by the free electrons. According to the classical theory, this electron gas should
contribute an additional (3=2)R to the heat capacity. But the measured heat capacity of
metals approached nearly the same 3R Dulong–Petit limit as the nonmetals. How can the
electrons be a major contributor to the thermal conductivity and not provide significant
additional heat capacity?
Careful measurements at low temperatures indicated that metals do have a small
electronic contribution to heat capacity that approaches 0 with a first power dependence
on T, as will be shown later. In fact, at very low temperatures, the electronic contribution
can be greater than the lattice contribution. In order to explain these departures from
classical theory, we must reformulate the problem quantum mechanically.
ð0
v
hvC(v)
hEi ¼ dv, (17:9)
ehv=kT 1
0
where
C(v) is the number of modes between v and v þ dv (the density of states)
v0 is the cutoff frequency
Now the trick is to find the appropriate C(v). We start by finding the distribution
in k-space and use the relation C(v)dv ¼ W(k)dk to obtain C(v). To get W(k), first take
the number of states on the surface of a sphere in k-space, which can be found by
dividing the volume of a shell in k-space, 4pk2dk, by the volume of an individual state,
which we found from Equation 16.9 to be (2p=L)3, and multiply by three polarization
states to get
4pk2 dk 3k2 L3 dk
W(k)dk ¼ 3 ¼ : (17:10)
(2p=L)3 2p2
324 Introduction to the Physics and Chemistry of Materials
To compute dk=dv, we should use the dispersion relation developed in Chapter 16 for the
chain of atoms. From Equation 16.7
2 dk 2
k¼ sin1 (v=v0 ) ¼
a dv a v2 v2 1=2
0
2 2
k2 L3 dk 3L3 2 2 1 12N sin1 (v=v0 )
C(v) ¼ 3 2 ¼ sin (v=v0 ) ¼ 2 1=2 :
2p dv 2p2 a v2 v2 1=2 a p v20 v2
0
Debye dodges this little problem by using the dispersion relation for a homogeneous solid
which has no cutoff frequency and in which k ¼ v=v0 and dk=dv ¼ 1=v0.
dk 3k2 L3 3v2 L3
C(v) ¼ W(k) ¼ 2 ¼ : (17:11)
dv 2p v0 2p2 v30
To be more precise, 3=v03 is sometimes written as 1=v3‘ þ 2=v3t to distinguish between the
different velocities associated with the longitudinal and transverse modes. Debye then
proceeds to determine a cutoff frequency by integrating the density of states (Equation
17.11) to a cutoff frequency vD that produces the required number of modes, i.e.,
vðD vðD
3v2 L3 v3D V
3N ¼ C(v)dv ¼ dv ¼ : (17:12)
2p2 v30 2p2 v30
0 0
It is convenient to set x ¼ -hv=kT. The integral in Equation 17.14 can then be written as
xðD
3k4 T4 V x3
hE i ¼ dx, (17:15)
2p2h3 v30 ex 1
0
In the low temperature limit, xD can be allowed to ! infinity and the integral may be
recognized by the more mathematically astute as the integral representation of a Riemann–
Zeta function (the rest of us look it up in tables) and has the value
ð
1
X1
x3 dx 1 p4
¼ 6 ¼ : (17:17)
ex 1 n¼1
n4 15
0
Therefore, the energy of the ensemble of oscillators in the low temperature limit becomes
@ hEi 12p4 Nk T 3
Cv ¼ ¼ , (17:19)
@T V 5 QD
2.5
2 Debye cutoff
1.5
w/w0
m
iu
ed
m
1 Chain of atoms
us
eo
en
og
om
FIGURE 17.1
H
3.5
10
2.5
1
2
CV/R
CV/R
1.5
0.1
1 Debye Debye
Diamond Diamond
Pb Pb
0.5 Al Al
Cu Cu
0 0.01
0 1 2 3 4 5 6 7 8 0.01 0.1 1 10
(a) T/ΘDebye (b) T/ΘDebye
FIGURE 17.2
(a) Heat capacity predicted from Debye theory compared with measured values. The Debye model appears to
slightly under-predict the actual data in the region where it departs most from the theoretical dispersion relation.
(b) Log–log plot of the heat capacity predicted from Debye theory. The third power dependence at low temper-
atures is more evident in this plot.
ð
1
g(E)EdE
hU i ¼ : (17:20)
e(Em)=kT þ 1
0
3 p2 nk2 T 2
hU i ¼ n«F þ þ , (17:21)
5 4 «F
@U p2 nk2 T p2 nkT p2 RT
Ce ¼ ¼ ¼ ¼ : (17:22)
@T 2 «F 2 TF 2 TF
Note that the electronic contribution to the heat capacity is linear in T, but for T << TF, this
contribution is much less than the 3R contribution from the ion cores. This explains why at
normal temperatures the electronic contribution is negligible. Only near absolute zero,
where the ionic contribution approaches zero as T3, does the electronic contribution
dominate as shown in Figure 17.3.
Thermal Properties of Solids 327
0.01
CV/R
Lattice
0.005
FIGURE 17.3
Electronic
Lattice heat capacity and electronic heat cap-
acity at very low temperatures. The electronic
heat capacity is drawn for a material whose
0
0 0.02 0.04 0.06 0.08 QD=TF ¼ 0.006. At very low temperature, the
electronic heat capacity can exceed the lattice
T/ΘD
heat capacity.
_ ¼ KDT,
Q (17:23)
where
_ is the heat flux (W=m2)
Q
K is the thermal conductivity (W=m deg)
dV
dw
FIGURE 17.4
dA Geometry for obtaining the thermal conductivity of an ideal gas.
328 Introduction to the Physics and Chemistry of Materials
then the number of paths in solid angle dv is dv=4p. Since the element of area dA subtends a
solid angle dv given by dA cos u=r2, the number crossing dA per increment of time is
dA cos u
nndVdt: (17:24)
4pr2
Now we ask how many molecules emerge from dV and encounter dA without making
another collision? From elementary Poisson statistics, the probability of an event not
occurring in time Dt when t is the average time between events is given by exp(Dt=t).
It follows therefore that the probability of a particle traveling a distance r without making a
collision is given by exp(r=l) where l is the mean free path. Now we replace dV with r2dr
sin ududf and integrate over the hemisphere above dA,
ð
2p ð
p=2 ð
1
dAdtvn 1
dp du sin u cos u drer=l ¼ nnldAdt: (17:25)
4p 4
0 0 0
Note that the mean free path l is just the average velocity v times the average time between
collisions, which is just the reciprocal of the collision frequency n. Thus the above result is
identical to the well-known equation for the number of particles incident on a surface per
unit time, which is given by 1=4n v.
However, the reason for going through all this trouble is that we need the average height
above the surface dA at which the molecules make their last collision. To get this, we put
h ¼ rcos u into the above integral and normalize by dividing by the integral without the rcos u.
ð
2p ð
p=2 ð
1
dAdtvn
dp du sin u cos u
2
rdrer=l
4p
0 0 0 2
h hi ¼ ¼ l: (17:26)
1 3
nnldAdt
4
Now assume there is a uniform vertical temperature gradient, dT=dy in the system.
Molecules crossing the surface from above will have a temperature given by T þ hhidT=dy,
while molecules crossing the surface from below will have temperature given by
T hhidT=dy. Therefore, the net heat transferred to surface dA in time dt will be given by
dQ 1 dT 1 dT
¼ n
vcV 2hhi ¼ nvcV l , (17:27)
dAdt 4 dy 3 dy
where cv is the molecular heat capacity. (The negative sign is introduced because the heat is
being transferred in the y direction.) Since the thermal conductivity K is defined by
_ ¼ KAdT=dy, the thermal conductivity becomes
Fourier’s first law, Q
ncV
vl CV vl
K¼ ¼ , (17:28)
3 3
where G is the reciprocal lattice vector (2p=a for a 1-D lattice). The collision is elastic in the
normal or N-process and the thermal energy is carried by phonon k3. However, if the vector
sum of the two original phonons is greater than half reciprocal lattice vector G=2 such that
the resultant vector is outside the first Brillouin zone, the resulting momentum is Bragg
reflected by G so that the phonon momentum k3 remains inside the first Brillouin zone as
shown in Figure 17.5. This event is referred to as an Umklapp (German for ‘‘turn over’’) or a
U-process and the G portion of momentum goes into what is called the crystal momentum as
an inelastic collision which results in heating the crystal rather than conducting heat.
If both k1 and k2 < p=2a, the resultant vector will always remain in the first Brillouin
zone and the collision will be an N-process. However, as they become larger than p=2a,
a U-process becomes more likely. Using the Debye model, the frequency associated with a
phonon whose k ¼ p=2a is v ¼ v0p=2a. The ratio of this frequency to the Debye frequency is
v v0 p 1
¼ ¼ 0:403: (17:31)
vD 2a v0 ð6p2 N=V Þ1=3
k1
k1
k3 G
k3
k2 k2
FIGURE 17.5
Illustration of an N-process in which the collision is elastic and a U-process in which the collision is inelastic. If the
resulting vector from a phonon–phonon collision falls outside the first Brillouin zone (the square box), the
momentum is Bragg reflected back into the first Brillouin zone with a transfer of G -h momentum to the lattice
in the form of crystal momentum.
330 Introduction to the Physics and Chemistry of Materials
Thus the U-processes can be expected to become important when the temperature
approaches 0.4QD and will limit the conductivity as the temperature increases.
Since thermal conduction of nonmetals is limited at low temperatures by the falloff of the
heat capacity and at high temperatures by increased Umklapp scattering, a peak in thermal
conductivity would be expected near 0.4uD.
Ceramics are generally poor thermal conductors because of the scattering from impur-
ities in the form of ions of different sizes and from grain boundaries as well as other
defects. Room temperature thermal conductivities are on the order of 10–50 W=m K.
However, pure single crystals can have very high thermal conductivities, especially those
systems such as diamond that have large elastic constants which result in high sound
velocities as shown in Figure 17.6. Room temperature conductivities for diamond can
range from 2000 to 2500 W=m K, several times higher than metals. Graphite also has a
very high thermal conductivity in its basal plane. Commercially available annealed
pyrolytic graphite has a conductivity of 1700 W=m K and is used to remove heat in
laptop computers as well as in other thermal management applications. High-conductiv-
ity graphite fibers can be aligned in a polymeric matrix to make a strong, highly
conductive sheet useful in space radiator panels. Such graphite fibers can have conduct-
ivities up to 1100 W=m K.
Glasses, on the other hand, are very poor thermal conductors because of their lack of a
periodic structure, which makes for very efficient phonon scattering. Their conductivities
are on the order of 1 W=m K. Polymers are even better thermal insulators with conduct-
ivities ranging from 0.1 to 0.5 W=m K. Woods are also good insulators with conductivities
ranging from 0.1 to 0.4 depending on the type and moisture content. Still air is an
extremely poor conductor so porous ceramics are very good insulators. The best insulator
is silica aerogel. An aerogel is a gel in which the liquid has been replaced by a gas. Silica
aerogels have been made with densities as low as 0.001 g=cm3 and with thermal conduct-
ivities as low as 0.003 W=m K.
140 10
Thermal conductivity (W/cm · K)
120
8
100
6
80
60
4
40
2
20
0 0
1 10 100 1000 104 1 10 100 1000
(a) Temperature (K) (b) Temperature (K)
FIGURE 17.6
Thermal conductivities of (a) diamond and (b) cubic SiC. The conductivity of diamond reaches a peak of
12,000 W=m K at 100 K, 30 times greater than Cu. The thermal conductivities are limited by heat capacity
below 100 K and by phonon collisions above 100 K. (Data taken from Adachi, Handbook of Physical Properties
of Semiconductors.)
Thermal Properties of Solids 331
1 4nk2 T
K ¼ n(3=2)kv2 t ¼ t, (17:33)
3 mp
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
since
v ¼ (8kT)=(pm) for an ideal classical gas and t is the average time between collisions.
Now if the correct value for the electronic heat capacity given by Equation 17.22 is
put into Equation 17.28 along with v ¼ vF and l ¼ vFt, where nF2 ¼ 2«F=m, the thermal
conductivity of the electrons becomes
1 1 p2 nk2 T 2 p2 nk2 T
K ¼ Cvl ¼ vF t ¼ t: (17:34)
3 3 2 «F 3 m
Note that the smaller heat capacity of the electron gas is compensated for by the higher
Fermi velocity and an expression close to the classical result is obtained but for totally
different reasons.
L(T) ¼ L0 ð1 þ aT Þ, (17:35)
where a is the coefficient of linear thermal expansion (CTE) which is O(106 K1). A similar
expression can be written for volume expansion and it is easy to show that since a is small,
the coefficient of volumetric expansion is 3a.
Thermal expansion comes about through the asymmetry of the bond energy function.
This asymmetry produces anharmonic oscillations of the atoms in the lattice; i.e., instead of
oscillating with equal amplitude on both sides of their equilibrium position, they can swing
farther to the soft side of the bond energy function than toward the steep side. As a result,
the average positions of the atoms are displaced from the equilibrium position at the
bottom of the bond–energy curve, which corresponds to the 0 K location. The average
displacement of the lattice atoms can be found from
ð1
xebUðxÞ dx
h xi ¼ ð1
1 , (17:36)
ebUðxÞ dx
1
where U(x) is the bond energy function as a function of the displacement x from the
equilibrium and b ¼ 1=kT.
The bond–energy curve can be approximated by U(x) ¼ Cx2 Dx3 in which C repre-
sents the harmonic part of the potential and D is the anharmonic part. It is assumed that
332 Introduction to the Physics and Chemistry of Materials
bDx3 << 1 so that exp(bDx3) can be expanded as 1 þ bDx3 þ . Keeping only terms
through second order, the integrals can then be written as
ð1
2
ebCx x þ bDx4 þ dx
3DkT
hxi ¼ ð1
1 ¼ : (17:37)
4C2
ebCx 1 þ bDx3 þ dx
2
1
This model was used with the bond–energy relations introduced in Chapter 7 to estimate
the thermal expansion of ionic compounds.
0 1 0 10 1
JQ c11 c12 c13 rT
@ JM A ¼ @ c21 c22 c23 A@rC A, (17:38)
JE c31 c32 c33 rV
where JQ, JM, and JE are the heat, mass, and electric current fluxes, respectively, that
respond to the temperature, composition, and electric potential gradient. The diagonal
terms can be immediately identified as the thermal conductivity K, the chemical diffusivity
D, and the electrical conductivity, s. In addition, the off-diagonal terms represent coupling
that occurs because of the interaction between electrons, phonons, and the ion cores. This
coupling produces some interesting and useful effects.
c21
JM ¼ c21 rT ¼ JQ ; rV ¼ 0: (17:39)
K
Equation 17.39 predicts a mass flux driven by a thermal gradient, a phenomenon known as
the Soret–Ludwig effect or sometimes as thermomigration or thermal diffusion. One can
think of the mass transport arising from the interactions of the phonons carrying the heat
flux (JQ) with the different constituents of the material. Components with positive heat of
transport migrate to the hot end while those with negative heat of transport migrate to the
cold end. The effect is more pronounced in liquids and gases where the mobility of the
particles is higher. Compositions of the melt have been altered in directional solidification
experiments by Soret diffusion, although such effects are often masked by convective
transport. Thermal diffusion has also been used in Clausius–Dickel columns for isotope
separation. The Soret effect for solids is generally small but can become significant when
there are high temperature gradients, such as those found in fuel pellets and in the
cladding on fuel rods in nuclear reactors.
Thermal Properties of Solids 333
The complement of the Soret effect is the Dufour effect in which a concentration gradient
causes a heat flow according to
c12
JQ ¼ c12 rC ¼ JM ; rV ¼ 0 (17:40)
D
by interactions between the diffusion current (JM) and phonons in the solid. Given the fact
that diffusion is a very slow process in solids, there is little diffusion current to interact with
the phonons, which makes the Dufour effect virtually undetectable.
17.6.2 Electrotransport
The c23 term is associated with electrotransport in which the electrons in the current flow
interact with impurities and tend to drag them along. From the matrix Equation 17.38,
c23
JM ¼ c23 rV ¼ JE ; rT ¼ 0: (17:41)
s
Under most circumstances, electrotransport (or electromigration as it is commonly called)
is a small effect. However, in integrated circuits, as the interconnects become smaller
and smaller, the current densities JE can become large enough to cause structural
changes in the Al interconnects, which caused reliability problems in some of the early
VLSI circuits. Naturally, this problem prompted a great deal of research into electrotran-
sport, which resulted in methods for preventing the degradation of interconnects from
its effects.
The complement of electrotransport predicts a current flow resulting from a concentra-
tion gradient as the diffusion current of atoms or ions drag electrons along
c32
JE ¼ c32 rC ¼ JM : (17:42)
D
Again because solid state diffusion is so slow, the effect is negligible.
The Seebeck coefficient is defined as S(T) ¼ dV=dT and we can identify c31 ¼ S(T)s. The
Seebeck effect arises partly from the small temperature dependence of the Fermi
level, which causes electrons to migrate from the region of higher Fermi level (higher
chemical potential) to the lower. But the dominant effect comes from the interaction of the
phonons with the electrons as they move through the solid, dragging the electrons along
with them.
Next consider the case of a homogenous conductor with an applied electric field. From
the matrix Equation 17.38 we see that
From Ohm’s law, rV ¼ JE=s, which combined with Equation 17.44 gives
JE
JQ ¼ c13 ¼ P JE , (17:45)
s
where P is the Peltier coefficient. The electron flow interacts with phonons to move heat
along with the current flow. The Peltier effect has a number of practical applications,
primarily as a solid-state electronic heat pump which is the heart of a variety of thermo-
electric devices (TEDs), which will be discussed later.
17.7 Applications
17.7.1 Thermocouples
The Seebeck voltage can only be measured in terms of the difference in Seebeck
coefficients between two conductors as illustrated in Figure 17.7. Junctions are formed by
spot welding wires with compositions A and B at each end. Wire B may be broken and
connected to a voltmeter with leads of a different composition as long as these
two connections are held at the same temperature. The voltage read by the voltmeter is
given by
ð
T2
where SA and SB are the Seebeck coefficients of the two wires. These coefficients are
generally dependent on temperature and can be approximated in the form a þ bx þ .
Integrating Equation 17.46, we see the Seebeck voltage will be a cubic (or higher order
polynomial) of the difference in temperature between the two junctions. There will also be
a contact potential at the junctions between the wire B and the meter leads, but if these two
junctions are at the same temperature, there will be no potential difference and the effect
from these junctions will cancel.
Before electronic compensated reference junctions became available, the reference
junction T1 was immersed in ice water and the temperature of the other junction was
determined by reading the microvoltmeter and either looking up the temperature
corresponding to that voltage in tables or evaluating a polynomial representing the
calibration data. Nowadays, the two thermocouple leads can be connected directly to a
A=D converter whose contacts are compensated for contact potential and resistance
electronically and the digital voltage is converted to temperature using a built-in
algorithm.
Metal A
T2 T1
Metal B Metal B
FIGURE 17.7
Typical thermocouple arrangement with junctions held at T1 and
T2. The Seebeck voltage is read by a microvolt meter in series with μV
one of the wires.
Thermal Properties of Solids 335
TABLE 17.1
Thermocouple Types and Their Properties
Types Materials Range (8C) EMF=deg. (mV=8C)
Thermocouples are widely used in measuring temperatures and are particularly useful
in the processing of materials because of their high temperature range. The most com-
monly used thermocouples are listed in Table 17.1.
Thermocouples are also used in the safety shut-off valves in gas appliances. The current
developed by a thermocouple inserted in the pilot flame flows though a solenoid that holds
the gas supply open. If the pilot flame were to go out, the valve would automatically close,
preventing the release of gas. It may seem surprising that such a little voltage could
produce enough current to activate a solenoid valve, but there is practically no internal
resistance in the thermocouple junction. This means that the current that flows is the
Seebeck voltage divided by the resistance in the wires of the solenoid, which can be a
small fraction of an ohm.
The Seebeck voltage generated at a moving solidification interface in a directionally
solidified alloy has been used to measure the kinetic undercooling and to observe the
incipient breakdown of a plane interface to dendritic growth (see Section 13.2.6).
Load
Cold
Hot
FIGURE 17.8
Schematic of a thermopile or thermoelectric generator. Alternating
pairs of thermocouple wires are joined so that the hot junctions are
embedded in the hot plate and the cold junctions are embedded in
a cold plate capable of rejecting heat. Current will flow through
B the load as long as a temperature difference is maintained between
the hot and the cold plates.
336 Introduction to the Physics and Chemistry of Materials
space where solar cells do not receive enough sunlight. The hot plate is heated by the
radioactive decay of Pu-238 and the cold plate radiates to space. Such a device is called a
radioisotope thermoelectric generator or RTG. Similar devices are used on earth to make
electricity from low-level process heat that would otherwise be wasted.
17.8 Summary
Classical thermodynamics using the equipartition of energy principle predicts that the
lattice molar heat capacity will be given by 1=2R for each of the six degrees of freedom in a
solid (three kinetic and three potential energy) for a total of 3R. As the temperature is
increased, the observed heat capacity of materials approaches this value, which is known
as the Dulong–Petit limit. However, at low temperatures, the observed heat capacity
approaches zero as T3.
Using the Planck distribution, Debye constructed a model in which the lattice heat
capacity goes to zero as T3. The Debye model assumes a linear dispersion relation charac-
teristic of a continuous medium rather than a chain of discrete atoms and cuts off the
distribution at a frequency such that the number of normal modes is equal to 3 the number
of atoms. This frequency, known as the Debye frequency, is given by vD ¼ v0(6p2N=V)1=3,
where v0 is the velocity of sound in the medium and N=V is the atoms per unit
volume. A Debye temperature QD is defined in terms of the Debye frequency QD ¼ -hvD=k.
For T >> QD, the Debye model approaches the classical Dulong–Petit limit.
Even though the Debye model uses an unrealistic assumption for the dispersion relation,
its primary justification comes from its excellent agreement with the observed heat cap-
acity, especially at the lower temperatures. Apparently, the heat capacity is not particularly
sensitive to the exact form of the dispersion relationship.
Thermal Properties of Solids 337
The small electronic contribution to the heat capacity of metals can be understood by
applying Fermi–Dirac statistics to the free electron gas. Since the ground state energy is
much higher than ambient thermal energy, only those electrons at the top of the Fermi sea
can be thermally excited. As a result the electronic heat capacity is given by C ¼ (p2=2)=
(RT=TF) instead of the classical value of 3=2R. At very low temperatures, the electronic
contribution, which has a linear temperature dependence, can exceed the lattice contribu-
tion to heat capacity, which has a T3 temperature dependence.
The thermal conductivity can be calculated for phonons as well as electrons using the
ideal gas model derived from kinetic theory, K ¼ Cvv l=3, where l is the mean free path of
electrons or phonons. The thermal conductivity of the lattice is low at low temperatures
because the lattice heat capacity goes to zero as the temperature approaches 0 K. The lattice
thermal conductivity is limited by scattering from lattices with different sized ions and
from defects such as grain boundaries, which is why ceramics are generally poor thermal
conductors. Glasses are especially poor conductors because of their lack of a periodic
structure. However, single crystal ceramics can be very good thermal conductors, espe-
cially diamond and sapphire whose stiff lattices give them high sound velocities. Aligned
pyrolytic graphite has thermal conductivity seven times higher than aluminum. Thermal
conductivity tends to decrease beyond 0.4 times the Debye temperature because of
inelastic phonon–phonon collisions (U-processes) that become more prevalent at higher
temperatures.
Metals generally have higher thermal conductivity than nonmetals because of the
presence of free electrons, but the electrons do not contribute to the heat capacity as they
would be expected to form classical considerations. The low heat capacity is compensated
by the high Fermi velocity, so the conductivity calculated using quantum mechanics is
almost the same as the classical result.
Thermal expansion is a result of the asymmetry of the energy–bond curve, which
produces anharmonic oscillations on the part of the atoms in the lattice. Thus the time–
average position of the atoms is displaced from the equilibrium position at the bottom of
the bond–energy well. This displacement increases with temperature, giving rise to the
coefficient of linear expansion.
Interactions between electrons, phonons, and ion cores can produce a number of inter-
esting coupling effects between heat flow, mass flow, and current flow. The most import-
ant of these coupling effects are the Seebeck effect, the production of current flow in
conjunction with heat flow, and the Peltier effect, the flow of heat in conjunction with
current flow. These two effects make possible thermocouples for temperature measure-
ment, thermoelectric generators and refrigerators that have no moving parts.
Bibliography
Adachi’s Handbook of Physical Properties of Semiconductors, Springer, New York, 2004.
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia, 1976.
Barsoum, M.W., Fundamentals of Ceramics, Institute of Physics Publishing, Bristol and Philadelphia,
2003.
Dekker, A.J., Solid State Physics, Prentice Hall, Engelwood Cliffs, New Jersey, 1962.
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley & Sons, New York, 1966.
Omega Temperature Handbook, available online at www.omegaeng.cz=temperature=Z=zsection.asp.
338 Introduction to the Physics and Chemistry of Materials
Problems
1. Show that the density of states for a 2-D ideal gas is a constant that does not contain E.
2. Show that the energy of a 2-D ideal gas is kT.
3. Adjust the density of states we developed,
2
12N sin1 (v=v0 )
C(v) ¼ 2 1=2
p v20 v2
V ¼ IR: (18:1)
(Actually this not really Ohm’s law, but the definition of the resistance R.) The constant of
proportionality R is called the resistance and is measured in ohms (1 V ¼ 1 V=A) and
depends not only on the metal but also on its size and configuration. This is what is known
339
340 Introduction to the Physics and Chemistry of Materials
as an extensive property since it depends of the size or extent of the material. R, however,
can be shown to be directly proportional to the length of the conductor and inversely
proportional to the cross-sectional area; thus, for a particular metal
rl l
R¼ ¼ , (18:2)
A sA
where
r is the resistivity, V m
s is the conductivity, (V m)1 or mho=m
Both resistivity and conductivity are intensive properties that depend only on the metal
itself. (The International unit of conductivity is now siemens per meter (S=m), where S is
defined as 1 A=V.) If we rewrite Equation 18.1 and substitute the above for R,
V VsA
I¼ ¼ ¼ EsA (18:3)
R l
and then divide by the area (A), we get the general form of Ohm’s law:
J ¼ sE, (18:4)
where
J is the current density
E is the applied electric field
The essential feature of Ohm’s law is that J is directly proportional to the applied field E
and s, being a property of the material, is independent of the field. Note that J and E are
vector quantities while s is a scalar (tensor of rank zero) for an isotropic media; however,
it will be a tensor of rank 2 for an anisotropic material such as a single crystal.
Treating the electrons as a classical ideal gas, the average speed would be
rffiffiffiffiffiffiffiffi
8kT
vth ¼ : (18:5)
pm
This velocity is random, so there are just as many electrons moving to the left as to the right
resulting in no net current flow. To cause a current to flow, this thermal velocity distribu-
tion must be biased by applying an electric field to produce a drift velocity that carries the
current.
The current density J can be expressed in terms of the drift velocity as
J ¼ nevd , (18:6)
where
n is the electron density
e is the magnitude of the electronic charge (the drift velocity for electrons will be
negative as well as the charge carried by the electrons so the current density will
be positive).
Another intensive property of the material is the mobility of the electrons m, which is
defined for electrons by the relation
Free Electrons in Metals 341
vd
m , (18:7)
E
where vd is the average or drift velocity of the electron when an electric field E is applied.
Like the conductivity, mobility can be a tensor of rank 2 and is a scalar only for an
isotropic media. (The sign of the mobility is always taken to be positive for charge carriers
of either sign.)
Since
s is related to m by
s ¼ nem (18:9)
x ¼ eE,
m€ (18:10)
ðe=mÞEt2
_ ¼ ðe=mÞEt
x(t) and x(t) ¼ : (18:11)
2
Assume a collision occurs at time t. The average drift velocity can be found from
xð t Þ ðe=mÞEt
vd ¼ hx_ ðt Þi ¼ ¼ : (18:12)
t 2
However, this model assumes that the electrons come to a complete stop after each
collision. A better model can be obtained by writing Newton’s law with a drag term,
1 e
v_ d vd ¼ E: (18:13)
t m
The solution is
et
vd ¼ (1 et=t ) E: (18:14)
m
Here it may be seen that t is the e-folding response time to a change in the E field and
that the steady-state velocity is the terminal velocity at which the drag force, vd=t is
balanced by the force exerted by the field. The t can also be interpreted as the average
time between collisions. The electrons give up some of their energy in these collisions
which create phonons, adding to the thermal energy of the lattice. The result is known as
ohmic heating.
342 Introduction to the Physics and Chemistry of Materials
Putting this vd into Equation 18.7, the mobility and conductivity can be written as
e ne2 t
m¼ t and s¼ : (18:15)
m m
Note that since mobility and conductivity, which are material properties, depend on the
collision time, then this collision time must be independent of the applied field if Ohm’s
law is to be obeyed. Why should the collision time be a material property? The mean free
path for electrons to travel before collisions would seem to be a much more reasonable
material property since it is determined by the size and number density of atoms in the
structure. But the mean free path l is the velocity times the collision time t and since vd is
directly proportional to tE, the collision time would be inversely proportional to the square
root of the field, in violation of Ohm’s law. This paradox can be resolved (at least
temporarily) by assuming that the thermal velocities of the electrons are much higher
than vd so that l ¼ t(vth vd) tvth. Let us now check to see if this assumption is valid.
The thermal velocity of electrons at 300 K from Equation 18.5 is 1.07105 m=s. For Cu,
the measured conductivity is 6 107 s=m. Assuming each atom contributes one free
electron to the electron gas, the number density is
8:96 g=cm3
n¼ 6:023 1023 atoms=mol ¼ 8:5 1022 electrons=cm3 ¼ 8:5 1028 =m3 :
63:55 g=mol
4I 4(20)
J¼ ¼ ¼ 2:5 107 A=m2
pd2 p(0:001)2
J 2:5 107
vd ¼ ¼ ¼ 0:00184 m=s:
ne (8:5 1028 )(1:6 1019 )
Clearly the drift velocity is much smaller than the thermal velocity of the electrons, so
we can take t as l=vth and Ohm’s law is obeyed. Of course we know that it does not
take this long for the current to run through the wire. The fact that a light turns on
the instant the switch is thrown can be explained by the marble-in-the-tube analogy.
If the tube is full of marbles and you insert a new marble in one end, a marble will come
out of the other end almost instantly, even though the average velocity of the marbles in
the tube is very slow.
Using the measured value for the conductivity of Cu, its mobility must be
s 6 107
m¼ ¼ ¼ 4:4 103 m2 =V s:
ne 8:5 1028 1:6 1019
The collision time, which is also the response time to a changing field, is much shorter than
the period of electronic frequencies, so that Ohm’s law can be expected to be obeyed for
alternating currents as well as for direct currents. However, we will find in later chapters
that deal with the interaction of light with metals that modifications will have to be made
to Ohm’s law.
where ri, rd, and rth are respectively the impurity, defect, and thermal contributions to
the resistivity. The impurity and defect contributions can be combined into r0, which is the
extrapolated value at 0 K.
12
10
tion
lid solu
u so
8 BeC d
ene
Resistivity (μΩ cm)
e hard
u ag
BeC
6 ss
Bra
4
ned
nh arde
2 rsio FIGURE 18.1
e
Disp e Cu
Pur Resistivity of Cu as a function of temperature. The
dashed line above pure Cu represents work-hardened
0 Cu. (From Askeland, D.R. and Phule, P.P., The Science
0 100 200 300 400 500 600 700 and Engineering of Materials, 5th edn., Thompson,
Temperature (K) Canada, 2006.)
344 Introduction to the Physics and Chemistry of Materials
decreases the resistivity of solid solution Be-Cu because the Be atoms coalesce to form
dispersed Be clusters rather than a random distribution of Be atoms. The added resistivity
contributed by a mole fraction x of a solid solution impurity atom can be described by
Nordheim’s rule:
where A is the defect resistivity coefficient that is a function of the host and impurity
atoms and can be quite large. For example, Cu-60 At % Ni has a resistivity that
is 30 times higher than pure Cu. In systems that exhibit order–disorder transitions,
the resistivity of the ordered phase is less than the disordered phase, as might be
expected.
Dispersion or age hardening has less effect than solid solution hardening as seen in
Figure 18.1 because the atoms are coalesced as a second phase. Also note that dispersion
hardening with 7% Al2O3 particles produces only a modest increase in resistivity. The
presence of a second phase follows the rule of mixtures, i.e.,
r0 (Vb ) ¼ ra Va þ rb Vb ¼ ra (1 Vb ) þ rb Vb , (18:18)
where the Vs are the volume fractions of the two phases, respectively. This relation also
applies to the conductivity of eutectics.
Dislocations and grain boundaries are also much less effective in scattering electrons
because they leave the vast bulk of atoms in the lattice undisturbed; therefore, work
hardening and grain refining have relatively little effect on resistivity. This is fortunate
because Cu house wiring tends to work harden from bending as electrical outlets are being
installed. If this resulted in a large resistance change, hot spots would develop and fires
could result.
0.6
0.4
r(T )/r(ΘD)
0.2
0 FIGURE 18.2
0 0.2 0.4 0.6
Universal reduced temperature plot for metals. It may be seen that
T/ΘD
the resistivity becomes linear with temperature for T 0.20 QD.
5 QDð=T
T x5 dx
r ðT Þ ¼ A , (18:19)
QD ðe 1Þð1 ex Þ
x
0
where A is a constant for a particular metal. However, if we divide r(T) by r(QD) and plot
this reduced resistivity against the reduced temperature T=uD, we obtain a universal plot
that is valid for all metals as shown in Figure 18.2.
where the 1=t0 ¼ vth=l0 is the contribution from impurities and defects and the 1=tth ¼ vth=lth
is the temperature-dependent contribution from the ion cores. Since the collision cross
section is dependent on the square of the amplitude of the vibrating ion cores and the
vibrational amplitude is proportional to the square root of the temperature (assuming T >
the 20% of the Debye temperature), lth must be inversely proportional to T as we argued
before. Therefore, in order for the resistivity to increase linearly with temperature, vth
would have to be independent of temperature. But vth is proportional to T1=2, so assuming
that the electrons in a metal behave an ideal gas in thermal equilibrium would lead to a
temperature dependence of resistivity in the form of r(T) / AT 1=2 þ BT 3=2, which is contra-
dictory to the observed thermal behavior of the conductivities of metals.
best conductors. What about the mean free path of poor conductors whose resistivity
can be several orders of magnitude higher than Cu? Their mean free paths would
have to be much shorter than the distance between the atoms. What could cause this
scattering?
p rffiffiffiffiffiffiffi
2«F
jvF j ¼ F ¼ : (18:22)
m m
For Cu, we had found n ¼ 8.5 1028 electrons=m3, which yields a Fermi energy of
1.131 1018 J (or 7.058 eV) and a Fermi velocity vF ¼ 1.57 106 m=s, which is an order
of magnitude larger than the thermal velocity. Since the Fermi velocity is independent of
the applied field and only a very weak function of temperature, substituting vF for vth
resolves the problems with the temperature dependence of resistivity described in
Section 18.3.1. Now the mean free path for Cu is on the order of 108 m, which corresponds
to 100 lattice parameters. (We will find a more sophisticated way of calculating resistance
in Chapter 19, which explains the large variations of resistivity among the transition
metals.)
Free Electrons in Metals 347
Ex
py py FIGURE 18.3
Fermi sphere in the absence of an applied field
(left). When an electric field is applied in the x
_
_ direction, each electron on the surface of the Fermi
_
_
_ sphere gains momentum in the þx direction and
px px moves to a higher energy state. This frees up
_
_ energy states just below the Fermi surface allowing
__ the other electrons to move to higher energy states,
_
the result being that the entire Fermi sphere is
shifted to the right.
One can visualize the distribution of momenta in momentum space in the absence of
an applied electric field as a sphere with radius pF centered about the origin as shown in
Figure 18.3. This we call the Fermi sphere and its surface is called the Fermi surface. Later,
we show that the Fermi surface is distorted by Bragg reflections of electrons near bound-
aries of the Brillouin zones. The electrons below the Fermi energy are in filled states and
cannot respond directly to the applied electric field. Only those electrons at the Fermi
energy can be accelerated by the applied field to a velocity vd þ vF. However, this creates
new states at the Fermi surface in which electrons just below the Fermi level can move into,
and so on. The net result is that each electron is able to increase its velocity by vd.
The drift velocity when an electric field is applied is still given by vd ¼ etE=m, so the
entire Fermi sphere is shifted by an amount dp ¼ etE. In the Drude model, all of
the electrons were assumed to have random velocities with an average speed nth but the
average velocity would be zero in the absence of an applied field. An applied field would
increase each electron’s average velocity by vd so the current is carried by all of the
electrons with a velocity vd.
However, the electrons in a Fermi gas do not have random velocities. Those on the
leading edge of the Fermi sphere have velocities vFx and those on the back edge have
velocities vFx. While it is true that the applied field increases each electron’s velocity by
vd, it is not true that the current is carried by each electron moving at vd. To illustrate this,
consider a simple one-dimensional Fermi model with 11 particles in which the Fermi
velocity is 5 m=s and quantum states are separated by 1 m=s as shown below.
5 4 3 2 1 0 1 2 3 4 5
4 3 2 1 0 1 2 3 4 5 6
Note that the particle with velocities from 4 to 4 m=s cancel each other and carry no net
charge. All the charge is carried by the two particles with velocities 5 and 6 m=s. So instead
of 11 particles moving at 1 m=s, as would be the case with the Drude model, we have two
particles moving with an average speed of 5.5 m=s to carry the current. The sum of the
products of the charge carriers and their velocities is the same in either model.
A rough estimate of the number of electrons in the Fermi sphere that are able
to participate in carrying charge when an electric field is applied is nvd=vF. So instead of
n electrons moving at vd, we have n(vd=vF) electrons moving at vF. Therefore for simple
metals in which the Fermi surface is essentially spherical, the conductivity and mobility
given by Equation 18.15, m ¼ (e=m)t and s ¼ ne2t=m are still valid, but the transition metals
348 Introduction to the Physics and Chemistry of Materials
in which the d-electrons play a larger role required a more rigorous analysis, as discussed
in Chapter 19.
v e
v_ þ ¼ ðE þ v BÞ: (18:23)
t m
For a direct current in the xy plane in the presence of a magnetic field B in the z-direction,
the steady-state solution is
et
vx ¼ Ex þ Bz vy
m (18:24)
et
vy ¼ Ey Bz vx :
m
Assume the applied field is in the x-direction causing a current jx to flow. The presence
of the field will deflect electrons in the y-direction to the walls of the conductor. Since
the current cannot flow out through the walls, a field Ey ¼ Bzvx will develop such that
vy ¼ 0. Putting this into the first of the above equations gives
eBz t
Ey ¼ Ex : (18:25)
m
The Hall coefficient is defined as RH ¼ Ey=jxBz. Putting jx ¼ sEx into Equation 18.24,
we find
for electrons. Note that measurement of the Hall coefficient determines not only the carrier
concentration n but the sign of the carriers—in this case, electrons. (Note: This n is the total
number of charge carriers, not just those at the Fermi level. This relationship holds for both
the Drude and the Fermi model as long as the conductivity can be written as s ¼ ne2t=m).
We shall see in later chapters that it is possible in semiconductors (and in some metals)
for holes in the electrons band structure to act as positively charged carriers, Had the
carriers been holes, the signs in the equation of motion would have been reversed resulting
in a positive Hall coefficient. (Note: This simplified derivation of the Hall effect is only
Free Electrons in Metals 349
+++++ −−−−−
e− VH
p+
B B B B
I
c
d
FIGURE 18.4
Schematic of a Hall measurement. Both positive and negative carriers will be swept to the same side. The
magnitude and polarity of the Hall voltage measures the number density and sign of the majority carriers.
correct if there is a single carrier. If there are both positive and negative carriers, a more
complicated derivation is required as discussed in Chapter 20.)
The Hall coefficient is measured by flowing a known current I through a rectangular bar
of the material of interest. Let the width of the bar perpendicular to the B-field be c and its
height d as shown in Figure 18.4. The Ey field produces a potential difference VH across the
bar, which is given by Eyc. The current density jx ¼ I=cd. Therefore,
Ey VH d
RH ¼ ¼ : (18:27)
jx B z IBz
Had the carriers been holes, the vx would have the opposite sign, they would be deflected
to the same side of the conductor as the electrons, but would have produced an opposite
E-field and opposite Hall voltage VH. Thus the measurement of the Hall coefficient is a
powerful tool that gives both the sign and the number density of the majority carriers in
metals as well as in semiconductors.
Recall that s ¼ nem; therefore, m ¼ RHs. Thus if s is known or measured separately, the
Hall coefficient also gives a measure of the majority carrier mobility.
The measured Hall coefficients for various metals are given in Table 18.1 below together
with the carrier density inferred from the Hall coefficient and the valence electron density.
The agreement is close but not exact for reasons we will get into when we study the band
theory and introduce the effective mass of the electrons. But look what happens with Zn.
The Hall coefficient has a positive sign. Does this mean that the carriers in Zn are holes? or
electrons with negative mass? To make it even stranger, at high magnetic fields, Al has a
positive Hall coefficient (Ashcroft and Mermin, 1976). We will examine these phenomena
in the later chapters.
TABLE 18.1
Hall Coefficients of Simple Metals
Metal RH (SI Units) ncarriers (m3) nvalence (m3)
Small Hall-effect sensors are widely used in conjunction with small magnets as proxim-
ity detectors in door and window alarms in security systems and have replaced the old
breaker points in automotive ignition systems. They can also be used as ammeters to
measure current flow in wires.
The constant of proportionality L ¼ k=sT is called the Lorenz number. Putting in the
appropriate values for the Boltzmann constant and the electronic charge, we get for the
Lorenz number
2
K 4 k 2 4 1:38 1023
L¼ ¼ ¼ ¼ 0:94 108 WV:
sT p e p 1:602 1019
Observed values for the Lorenz number range from 2.06 108 for Cd to 3.13 108 for W.
However, when the corrected electronic conductivity K ¼ (p2=3) (nk2T=m)t (Equation
17.36) is used to compute the Lorenz number,
K p2 k 2
L¼ ¼ ¼ 2:445 108 WV, (18:29)
sT 3 e
being good conductors of electricity. In fact, polymers were among the best insulators
known with conductivities <1010 S=cm. Then came the discovery of ceramic supercon-
ductors by Muller and Bednortz in the late 1960s and the discovery of conductive polymers
by Heegar, MacDiarmid, and Shirakawa in the 1970s. The later discovery is bound to
have far more technological impact because, unlike ceramics, polymers are more versatile
and are conductive at ambient rather than at cryogenic temperatures. More recently,
researchers have found ways to engineer semiconducting properties into polymers result-
ing in materials with controllable bandgaps in which the charge carriers can be either
positive or negative. Such polymers can now fill the full range of electronic and photonic
applications like their silicon and compound semiconductor cousins. Because of their low
cost, ease of fabrication, and flexible nature, photonic polymers stand to revolutionize flat
panel display technology.
18.8 Summary
The Drude theory treats the electrons in a metal as a classical monatomic gas. This simple
theory is able to explain many observed properties of metals such as why they are good
352 Introduction to the Physics and Chemistry of Materials
electrical and thermal conductors and reflectors of light. The electrical conduction
obeys Ohm’s law J ¼ sE, where J is the current density and E is the applied electric
field. The conductivity s is a property of the material. Another material property is the
mobility m of the charge carriers defined as their average velocity when subjected to
an applied electric field. The mobility is related to the conductivity by s ¼ nem, where n
is the electron density and e is the electronic charge. Using Newton’s law we find the drift
velocity of electrons being accelerated by the electric field to be vd ¼ etE=m and the
conductivity to be given by s ¼ ne2t=m, where t is the relaxation time or the time between
collisions.
To satisfy Ohm’s law, this collision time must be independent of the field. If we
assume that these collisions are between the electrons traveling over some mean free
path l that is associated with the lattice, the average speed of the electrons would have
to be much larger than the drift speed produced by the applied electric field. This
would be the case if the electrons behaved as an ideal gas in thermal equilibrium with
the lattice.
The resistivity r ¼ 1=s is found to have a linear relationship with temperature in the
form r(T) ¼ r0 þ aT (Matthiessen’s rule), where r0 is due to collisions with structural
and impurity imperfections and the temperature dependence comes about from collisions
with the lattice ions whose cross sections increase linearly with temperature. Impurity
atoms, such as found in solid solution alloys, produce a much larger increase in resistivity
than structural defects such as dislocations and grain boundaries or condensed second
phases because they are more widely dispersed. Also there is a departure from the linear
temperature dependence of the resistivity at low temperatures because all of the
phonon modes are active. Grüneisen used the Debye theory to develop a universal
relationship between reduced resistivity and reduced temperature that holds for all
metals.
The observed first power relation between resistivity and temperature implies
that the collision velocity must not only be independent of the applied field and be
much greater than the drift velocity but must also be independent of temperature,
which rules out the thermal velocity as the motion responsible for the observed collision
rate. With the discovery of quantum mechanics, the difficulties encountered by the
classical electron gas model are resolved by having the electrons travel at the Fermi
velocity.
But only the electrons at the surface of the Fermi sphere can move with the Fermi
velocity. So instead of all of the electrons moving at the drift speed, we now have the
current carried by much fewer but much faster electrons at the Fermi surface that
can respond to the applied field. This new Fermi model requires a reformulation of the
conductivity and it will be shown later that the Drude formula for conductivity still holds
for simple metals but for wrong reason.
Charges moving through a conductor are deflected to the sides of the conductor by
a magnetic field causing a voltage difference called the Hall voltage to appear across
the lateral surfaces. Measurement of the Hall voltage for a given magnetic field and
conductor geometry determines the sign and number density of the charge carriers in the
conductor. If the conductivity is also known, the mobility of the charge carriers can be
determined. Again, for simple metals, the number density of charge carriers measured
from the Hall coefficient is the total number of free electrons, not just those actually
carrying the current.
The ratio of the thermal conductivity to the electrical conductivity times the absolute
temperature is known as the Wiedemann and Franz ratio and involves only universal
physical constants. Therefore, this ratio should be the same for any metal (provided that
the heat is predominately carried by the electrons). This relationship depends on the fact
Free Electrons in Metals 353
that the collision times that inhibit the flow of the electrons is the same whether they are
driven by an applied field of a thermal gradient.
Bibliography
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia, 1976.
Askeland, D.R. and Phule, P.P., The Science and Engineering of Materials, 5th edn., Thompson, Canada,
2006.
Hush, N.S., An overview of the first half-century of molecular electronics, Ann. N.Y. Acad. Sci., 100,
2003, 1–20.
Ibach, H. and Lüth, H., Solid State Physics, 3rd edn., Springer-Verlag, New York, 1990.
Kittel, C., Introduction to Solid State Physics, 7th edn., John & and Sons, New York, 1966.
Srivastava, C.M. and Srinivasan, C., Science of Engineering Materials, Wiley Eastern, New Delhi, India,
1987.
Problem
1. The Earth’s magnetic field is 0.5 G (1 G ¼ 104 T). Do a feasibility analysis on the
possibility of building a magnetometer using the Hall effect. Assume you can detect a
Hall voltage as small as 1 mV. Also you want to minimize the power consumption. What
material would you want to use? What dimensions would be required? What voltages
and currents would be required?
19
Band Theory of Metals
In spite of the success of the free electron theory after the introduction of the Fermi energy
and Fermi velocity to straighten out some of the inconsistencies in the original Drude
theory, the free electron theory still does not account for all that is known. It is necessary to
introduce a modification to the free electron theory to account for the interaction of
electrons with the charged ion cores. There are two ways to go about this. One could
start with the empty lattice (free electrons) and add the interactions of the core ions as a
perturbation to the energy of the electrons (nearly free electron model). Or one could start
with the core atoms and describe the energy bands that form as the electron wavefunctions
overlap (tight-binding approximation). These models give a rough approximation to the
actual band structure and serve only to give the reader some insight into the importance of
the band structure and how it originates. The actual calculation of the energy band
diagrams is enormously complicated and is beyond the scope of this chapter.
where
nx ¼ 0, 1, 2, . . .
ny ¼ 0, 1, 2, . . .
nz ¼ 0, 1, 2, . . .
355
356 Introduction to the Physics and Chemistry of Materials
One can easily see that this solution satisfies the boundary conditions
3=2
1
cðx þ L, y, zÞ ¼ ð1=LÞ3=2 e2piðnx (xþL)þny yþnz zÞ=L ¼ e2pinx e2piðnx xþny yþnz zÞ=L ¼ cðx, y, zÞ:
L
Recall in the previous treatment where we allowed only standing waves, we found energy
eigenvalues given by
h2 p2 2
En ¼ n where nx ¼ 1, 2, 3, . . . ; ny ¼ 1, 2, 3, . . . ; nz ¼ 1, 2, 3, . . . : (19:3)
2m L
Clearly, employing periodic boundary conditions produces a different set of energy levels;
however, it will also produce a different density of states. Since E ¼h2 k2=2m, ~
k ¼ (2p=L) ~
n.
Thus the volume associated with each energy state in k-space is (2p=L)3. The minimum
energy configuration is a Fermi sphere with radius kmax. The volume of this sphere in
3
k-space is (4p=3)kmax . The number of electrons that can be accommodated is
3
(4p=3)kmax L3 2m EF 3=2
Nelectrons ¼ 2 ¼ 2 (19:4)
ð2p=LÞ3 3p h2
h2 2 2=3
EF ¼ 3p ne , (19:5)
2m
where ne is the number of electrons per unit volume. This is the same as we calculated from
the previous model that admitted only standing waves. The reason for this agreement is
that in the previous model, the volume per state in k-space was (p=L)3 which is only 1=8 as
large as in the present model. But since the present model allows both plus and minus
values of n, we now integrate over the entire sphere in n-space (or k-space) rather than over
just one octant. Thus in either case, the relationship between the minimum or Fermi energy
and number of states or electron density is the same.
If one considers a one-dimensional (1-D) line of atoms in the limit of no electron
interaction (free electron approximation), the energy dispersion curve (E vs. k) is parabolic
since E k2 as shown in Figure 19.1. Actually, the curve is not continuous, but a locus of
FIGURE 19.1
Energy versus k for a 1-D line of atoms with spacing a
assuming no interaction with the ion cores (free electron
approximation). These curves are actually the locus of the −3π/a −2π/a −π/a 0 π/a 2π/a 3π/a
energy states spaced at 2p=L. Reciprocal lattice vector (k)
Band Theory of Metals 357
yy *
−
FIGURE 19.2
Two standing wave solutions of the Schrödinger equation.
Note that the sin2 solution (dashed curve) places the electrons
between the ion cores, which gives it a higher energy than
−π/a 0 π/a the cos2 solution (solid curve) that puts the electrons on the
k lattice ions.
points spaced at 2p=L along the k-axis which represent the energy states. As waves, the
electrons are subject to Bragg reflections from the lattice planes just as x-rays. Since
k ¼ 2p=l, the Bragg criteria, 2a sin u ¼ nl, is satisfied for u ¼ 908 when k ¼ np=a. For these
values of k, the electrons are reflected constructively to form standing waves. The first such
reflection occurs at n ¼ 1 or k ¼ p=a, which corresponds to the first Brillouin zone for a
simple cubic lattice.
Recall that a standing wave can be decomposed into two traveling waves moving in
opposite direction, or k ¼ p=a. There are two possible reflection modes
FIGURE 19.3
Energy versus k for a 1-D line of atoms in which
interactions with ion cores are considered. The elec-
Energy gap 2 tron wavefunctions are Bragg-reflected at the Brillouin
zone boundaries (np=a). As seen in Figure 19.2, the
resulting standing waves then have two possible ener-
Energy gap 1 gies, depending on whether the maximum CC* sits on
the ion cores (lowest energy solution) or in between
−3π/a −2π/a −π/a 0 π/a 2π/a 3π/a the ion cores (highest energy solution). This difference
in energy causes energy gaps Eg1 and Eg2 to appear at
Reciprocal lattice vector (k) the Brillouin zone boundaries.
358 Introduction to the Physics and Chemistry of Materials
If these conditions are not met, the two elements will have limited solid solubility or may
form either compounds or new phases. The ratio of Zn=Cu atomic radii is 1.04, so the first
condition is satisfied. Cu crystallizes in the fcc configuration; whereas, Zn is hcp. Also, Cu
is univalent and Zn is divalent, so the last two conditions are not met.
Note that a-brass, that forms up to 38 At% Zn, is fcc. For higher concentration of Zn,
the system undergoes a solid-phase transformation to b-brass which is bcc up to 48 At%.
At this point it becomes g (a complex cubic unit cell containing 52 atoms). The addition of
still more Zn produces the « and h phases which are hcp with different c=a ratios. Cu–Al,
Cu–Sn, Ag–Zn, Ag–Al, Ag–Cd and other mixed valence systems show similar phase
behavior.
1100
900
b
Temperature (°C)
700 a Cu bcc
fcc g
d
complex
500
e
fcc Zn
300
100
0 10 20 30 40 50 60 70 80 90 100
Atomic % Zn
FIGURE 19.4
Cu–Zn phase diagram. The a phase is fcc and extends to 36 At% Zn. The b phase is bcc and extends to 48 At%.
(From Massalski, T.B., Senior Editor, Handbook of Binary Alloy Phase Diagrams, Vols. 1–3, American Society for
Metals, 1990. Reprinted with permission of ASM International. All rights reserved.)
Band Theory of Metals 359
This behavior can be understood from Brillouin zone theory. The number of electrons
contained in a Fermi sphere is
2(4p=3)kF3 kF3 L3
Ne ¼ ¼ (19:8)
ð2p=LÞ3 3p2
or
FIGURE 19.5
Fermi surfaces for the noble metals (Cu, Ag, and
Au) fit inside the first Brillouin zone (however,
see Figure 19.15). The volume inside the Fermi
surface represents the filled minimum energy
states. The space between the Fermi surface
and the polygon are the unfilled higher energy
states.
360 Introduction to the Physics and Chemistry of Materials
It turns out that there is. A bcc direct lattice has a fcc reciprocal
pffiffiffi lattice and it can be
shown that the largest kF that will just fit inside is kF ¼ p 2=a (again proof left to
the student). This is shorter than the maximum radius for the previous case, but the bcc
lattice has only two atoms per unit cell. Therefore Equation 19.10 can be written as
pffiffiffi
2n p3 2 2
kF3 ¼ 3p 3 ¼
2
(19:11)
a a3
pffiffiffi
and n ¼ p 2=3 ¼ 1:48. which implies that when x > 0.48, the system will try to lower its
energy by seeking other structures that can accommodate more electrons in the first
Brillouin zone. As it turns, out the g phase is a complex cubic phase with 52 atoms in the
unit cell and the « and h phases are hcp with different c=a ratios.
where uk(r) has the periodicity of the lattice, i.e., uk(r þ a) ¼ uk(r) (see a standard solid-state
physics text for proof). The theorem simply states that the wavefunction is modulated by
the lattice periodicity. One of the consequences is that a wavefunction is not changed after
being displaced by G, i.e.,
(see Appendix). Putting these Bloch wavefunctions into the Schrödinger equation,
HcK (r) ¼ HcKþG (r) ¼ E(K)cK (r) ¼ EðK þ GÞcKþG (r) (19:14)
with the result that E(k) ¼ E(k þ G), which may be generalized to E(k) ¼ E(k þ nG). This
important result says that possible energy states are not restricted to a single parabola in
k-space, but can be found equally well on parabolas shifted by the integral values of the
reciprocal lattice vector G. Therefore, it is not necessary to plot the E versus k curves for k
beyond the first Brillouin zone; the higher energy states can be represented by simply shifting
k back into the first Brillouin zone by adding (or subtracting) multiples of G. Another way of
saying this is, for a given value of k, there are multiple solutions of the Schrödinger equation in
which the energy increases by (k þ nG)2. The result is shown in Figure 19.6 for the 1-D line of
atoms. Recall that the G for a simple cubic lattice is 2p=a, so all of the energy states can be
shown by shifting the energy values in the expanded zone scheme we saw in Figure 19.6 by
integral values of 2p=a so they appear in the first Brillouin zone. This is called the reduced
band scheme. Figure 19.7a shows the bands for electrons in the absence of core interactions;
Figure 19.7b shows the effects of interactions with the core electrons. Recall that we saw
something like this in Chapter 16 in dealing with the acoustic dispersion relations.
Recall that the spacing between states is 2p=L along the k-axis. Therefore, the number of
states between p=a and p=a is just the number of ion cores since a is the spacing between the
Band Theory of Metals 361
FIGURE 19.6
Bloch solution for E versus k for periodic lattice in the limit
of vanishing interaction between electrons and ion cores.
The different parabolas represent the energy values for k
−3π/a −2π/a −π/a 0 π/a 2π/a 3π/a
shifted by integral values of G which, for a simple cubic
Reciprocal lattice vector (k) lattice, is 2p=a.
Energy gap 2
Energy gap 1
FIGURE 19.7
Reduced zone representation for a 1-D line of atoms in the empty lattice model (a) and with core interactions (b)
which give rise to energy gaps Eg1 and Eg2.
atoms in a simple cubic lattice. If each atom contributes one electron, the first zone will be half
full; two electrons per atom would fill the band. Similarly, four electrons per atom would fill
the second band. Thus group IV semiconductors (as well as III–V and II–VI compounds), have
filled valence bands. The empty band above this is the conduction band. However, as we shall
see later, there is more to this story. Given this picture, one would expect elements with even
valences would be semiconductors or insulators. Indeed, this would be the case if it were not
for band overlap, which allows group II elements to be conductive.
field, but they are also acted upon by the lattice. The equation of motion can be written
@vg
m ¼ Fext þ Flattice , (19:15)
@t
where Flattice is generally not known. The vg is the group velocity which, by definition,
@v 1 @E
vg ¼ ¼ : (19:16)
@k h @k
Therefore
@vg 1 @ @E 1 @ 2 E @k
¼ ¼ : (19:17)
@t h @t @k h @k2 @t
362 Introduction to the Physics and Chemistry of Materials
The external force should equal the time rate change in momentum,
@p @k
Fext ¼ ¼ h , (19:18)
@t @t
@k h2 @vg
Fext ¼ h ¼ 2 : (19:19)
@t @ E=@k @t
2
We can put this in the form of Newton’s law and include the effects of Flattice by assigning
the coefficient of the time derivative of the velocity term as an effective mass, i.e.,
h2
m* : (19:20)
@ 2 E=@k2
h2 k2 @ 2 E h
E¼ ; ¼ ; and m* ¼ m: (19:21)
2m @k2 m
Notice that the effective mass varies as the curvature of the band. In the conduction band
in the vicinity of k ¼ 0 where the band is curved upward, the effective mass of an electron is
positive. However, near an inflection point the effective mass becomes infinite, and then
becomes negative in the regions where the band curves downward. It follows that elec-
trons in the second or valence band near k ¼ 0, have negative mass.
19.3.3 Holes
If an electron is removed from the valence band, it leaves a hole. A hole can be thought of
as a particle with positive charge, but with negative electron mass. Therefore, holes near
the top of a valence band with negative curvature will have positive mass and positive
charge and will move in the direction of an applied electric field. An electron in the
conduction band with positive curvature will move opposite to the field because of its
negative charge and positive mass.
dv v e
þ ¼ ðE þ v BÞ: (19:22)
dt t m*
Let the magnetic field be in the z-direction and the right-hand circularly polarized radiation
be given by Ex ¼ E0eivt and Ey ¼ iE0eivt. The solution to Equation 19.22 can be written as
vx ¼ v0eivt and vy ¼ iv0eivt. Inserting these quantities into Equation 19.22 and identifying
the cyclotron frequency as vc ¼ eBz=m*, the v0 can be written as
Band Theory of Metals 363
et E0 ½1 iðv vc Þt
v0 ¼ : (19:23)
m* 1 þ ðv vc Þ2 t 2
and
Þ
The power density is P ¼ cycle ReðJÞ ReðEÞdðvtÞ which becomes
sE20
P¼ : (19:26)
1 þ ðv vc Þ2 t 2
The resonant or cyclotron frequency is detected by absorption of power from the micro-
wave beam which becomes a maximum when v ¼ vc, as can be seen from Equation 19.26.
The effective electron mass can then be determined from m* ¼ eBz=vc. The effective mass of
holes can be found in a similar manner using left-hand polarized microwave radiation.
In conductive metals, it is difficult to penetrate to much depth with microwave radiation.
The magnetic field and the microwave beam are oriented along the surface of the con-
ductor. The circulating electrons will only feel the field when they are within the skin depth
of the radiation, so when the applied frequency is some integer times the resonance
frequency, absorption will occur. This technique is known as the Azbel–Kramer cyclotron
resonance (AKCR) method.
Fermi level
Energy
Γ X W L Γ K X
FIGURE 19.8
Band structure for Al along directions of high symmetry designated by X, W, L, and K as shown in Figure 6.7(a).
(From Segal, B., Phys. Rev., 125, 115, 1962. With permission.)
Energy
4s
Fermi level
3d
Γ X W L Γ K X
FIGURE 19.9
Band structure of Cu. Transitions of the 3d electron states (flat lines just below the Fermi level to 4s states at the
Fermi level absorb blue light giving Cu is reddish color. (From Segal, B., Phys. Rev., 125, 114, 1962. With
permission.)
where
v(k) is the Fermi velocity
S(k) is the density of states near the Fermi surface.
When an electric field, Eap, is applied in the x-direction, the symmetry is disturbed and we
must expand the integrand
@SðkÞ
vðkÞSðkÞ ¼ vðkÞS0 ðkÞ vx Dkx : (19:28)
@kx
The first term of the expansion integrates to zero by symmetry. We can write the incremental
change in momentum in the x-direction resulting from the applied field as Dkx ¼ mvd=h ¼ et
Eap=h. We can also write @SðkÞ=@kx ¼ hvx dSðkÞ=dE and djkj ¼ dE=hjvj. Putting all of this
back into Equation 19.27, we can write the conductivity as
ðð 2
Jx e2 vx
s¼ ¼ tdS(k): (19:29)
Eap h j vj
EF
The square of the Fermi velocity v2 ¼ v2x þ v2y þ v2z , but since v2x ¼ v2y ¼ v2z , v2 ¼ 3v2x , and
v2x =jvj ¼ (v=3) ¼ hkF =3m* where hkF is the Fermi momentum. (Here we use the effective
mass to be more general.) Assuming t is not a function of k, Equation 19.29 can now be
written as
ðð
e2 t
s¼ kF dS(k): (19:30)
3m*
EF
For a simple metal, such as Cu where the s-electrons are the primary carries, the Fermi
surface is nearly spherical and the integral over its surface is 4pkF2 . The total volume
in k-space included in the integral is 4pkF3 . If the system has volume L3, the number
of states is therefore 4pk3=(2p=L)3 which can accommodate 8pk3=(2p=L)3 electrons or
8pk3=(2p)3 electrons per unit volume. Remembering from Equation 19.4 kF3 =3p2 ¼ ne , the
electron density of the Fermi sphere, Equation 19.30 can be written for simple metals as
e2 t 8pkF3 e2 t kF3 e2 t
s¼ 3
¼ ¼ ne , (19:31)
3m* ð2pÞ m* 3p2 m*
which is identical to the Drude model except that the effective mass of the electrons
replaces the normal mass. What is happening here is the smaller number of electrons
available to carry the current is compensated for by multiplying their number by the Fermi
momentum in Equation 19.31. This gives the same volume in k-space as would be obtained
by multiplying all of the electrons in the Fermi sphere by the drift velocity, as was
illustrated in the 1-D model in Chapter 18.
Being able to compute the conductivity from band theory gives new insight into the
widely differing conductivities that are found among the various metals. For example, Cu,
which is a good conductor, has its Fermi level in the region where the s-bands are steep (see
Figure 19.12), meaning the electrons have high velocity (vg ¼ (1=h)@E=@k). Fe, on the other
hand, has its Fermi level lower in the bands populated by the d-electrons where, despite the
larger density of states, the bands are flat making the electrons less mobile because they are
partially involved in forming covalent bonds. Consequently, Fe is not as good a conductor
of electricity than Cu.
366 Introduction to the Physics and Chemistry of Materials
Vacuum level
FIGURE 19.10
Energy well of an individual atom in which A, B, and
C are bound states. Radius
f EF
FIGURE 19.11
Illustration of a periodic potential with a band structure of bound electrons (A and B) and nearly free electrons in
the Fermi region (C).
Band Theory of Metals 367
where f(r – rj) is the wavefunction of the jth isolated atom at distance rj. Recall we used this
technique to describe the binding process when the wavefunctions of two H atoms began
were brought together. The coefficients Ckj are assumed to be given by
to overlap as they
N 1=2 exp ik rj to put the function in the Bloch form since it is periodic in rj. Putting these
coefficients into the original wavefunction gives
X
ck ðrÞ ¼ N 1=2 exp ik rj f(r rj ): (19:33)
j
We now compute the energy of the electrons under the influence of the other atoms in the
crystal.
XX
E ¼ hkjHjki ¼ N 1 exp ik (rj rm ) hfm jHjfj i: (19:34)
j m
Since all j atoms are equivalent, we pick the j ¼ 0 atom, set rj ¼ 0, and multiply by the N
atoms in the ensemble.
X X
E¼ expðik rm Þhfm jHjf0 i ¼ hf0 jHjf0 i þ expðik rm Þhfm jHjf0 i: (19:35)
m¼0 m¼1
where hf0jHjf0i ¼ E0 is the binding energy of the electron in the j ¼ 0 atom due to the
potential from its nucleus, and hfmjHjf0i ¼ E1 is the influence of the nearest neighbor
atoms on the electron. The binding energy E0 will be slightly lower than that of an
isolated atom due to a perturbation on the potential from its nucleus from the overlap of
nearest neighbor wavefunctions.
For a simple cubic lattice, rm ¼ (a, 0, 0), (0, a, 0), (0, 0, a). The dispersion relations are
given by
A bcc structure has nearest neighbors at rm ¼ (a=2, a=2, a=2). The dispersion relations
are given by
An fcc structure has nearest neighbors at rm ¼ (a=2, a=2, 0), (a=2, 0, a=2), (0a=2,
a=2). The aspersion relations are given by
E ¼ E0 4E1 cos ky a=2 cosðkz a=2Þ þ cosðkz a=2Þ cosðkx a=2Þ
The band structure for a simple cubic structure computed in the k100 and the k110 directions
is shown in Figure 19.12. The Fermi surface can be determined from Equations 19.37, 19.38,
or 19.39 by setting E(k) ¼ EF and solving for the values of kx, ky, and kz. An example of
368 Introduction to the Physics and Chemistry of Materials
EF
Energy
FIGURE 19.12
Energy of a simple cubic system computed
from Equation 19.37. E0 was set to 3 and [110] [100]
E1 to 1. The EF was taken to be 3. 21/2π/a Γ π/a
such a Fermi surface is for a simple cubic lattice shown as a contour plot in Figure 19.13
and in three dimension in Figure 19.14.
The approximations used in the tight-bonding approach are only valid for non-
degenerate s-electron (spherically symmetric) wavefunctions. It is introduced here only
FIGURE 19.13
A contour map of the Fermi surface of the simple cubic system in the first Brillouin zone that was described
in Figure 19.12. The contours go in steps of 0.1 kz with the outer contour corresponding to kz ¼ 0 and the inner
to kz ¼ 1.
Band Theory of Metals 369
FIGURE 19.14
3-D rendition of the Fermi surface of a simple cubic system described in Figures 19.12 and 19.13.
to illustrate that energy bands do indeed exist even in systems with no free electrons.
Actual band calculations require much more sophisticated analytical and numerical
methods (Ashcroft and Mermin, 1976).
The effect of the energy gap in the vicinity of the Brillouin zone boundary is to distort the
Fermi surface such that a face will be formed at the zone interface as we saw for the case of
the simple cubic system. This schematic in Figure 19.15 illustrates a better approximation of
the actual Fermi surface of the noble metals as calculated from Equation 19.39 than that
shown in Figure 19.5.
FIGURE 19.15
Sketch of the actual Fermi surface of the noble metals.
EF
FIGURE 19.16
Simple 2-D model to explain conductiv-
Energy
ð
1 ð
1
g(E)dE
N¼ CE1=2 e(EEF )=kT : (19:40)
e(Em)=kB T þ 1
EF þf EF þf
Accounting for the momenta of the electrons traveling in the direction of the surface, the
thermionic current is given by the Richardson equation,
4pem
J¼ ðkTÞ2 ef=kT : (19:41)
h3
The equation has the Arrhenius form so plotting log J against 1=kT offers an alternative
method for obtaining the work function. Also, it is clear that efficient cathode materials
must have a small work function.
372 Introduction to the Physics and Chemistry of Materials
19.7 Summary
Despite the success with the free electron theory, especially when quantum effects are
taken into consideration, it must be recognized that electrons are not entirely free; they still
interact with the core ions. We introduce periodic boundary conditions to describe travel-
ing waves of electrons moving through a conductor. These periodic boundary conditions
produce a different set of quantum states than the stationary waves we determined by
requiring the wavefunctions to vanish at the edge of the conductor. The density of states
found for traveling waves is also different, but when the new energy states are weighted by
the new density of states and summed over the occupied states, the same value for the
Fermi energy is obtained as was found by summing over the standing wave states.
These traveling waves are diffracted at the Brillouin zones to form two sets of standing
waves. One set of standing waves places the electrons between the ion cores, which
produces a higher energy (less negative) than the wave that places the electrons on top
of the ion cores. There are no energy states between the two diffracted waves at the
Brillouin zone boundary, hence an energy gap is formed at the Brillouin zone boundary.
The Fermi sphere for atoms with one valence electron per atom can easily fit in a bcc
reciprocal lattice (fcc direct lattice). If a divalent element is added, the additional electrons
will increase the diameter of the Fermi sphere. When the Fermi sphere begins to touch the
Brillouin zone, the additional electrons must either go to the unfilled states in the corners of
the first Brillouin zone, which are higher energy states, go across the Brillion zone, which
costs energy because of the energy gap, or undergo a solid-phase transition to a new
reciprocal lattice with a Brillouin that can accommodate more electrons. This model
explains the solid-phase transformations of many mixed-valency binary alloys in which
the phases progress from fcc to bcc to hcp as more divalent component is added.
The Bloch theorem states that the wavefunction of a single electron moving in a periodic
potential is modulated by the potential according to ck(r) ¼ uk(r)eikr. As a consequence,
cKþG(r) ¼ cK(r), which means wavefunctions are not changed after being displaced by the
reciprocal lattice vector G. It therefore follows that E(k) ¼ E(k þ nG), which means that
energy bands on an E versus k plot are the same if displaced along the k-axis by G. It
also follows that all of the bands can be shifted back into the first Brillouin zone by adding
or subtracting G.
As the electrons move through the lattice under the influence of an applied field, they
interact with the lattice and transfer momentum to the lattice (or vice versa). This momen-
tum transfer is accounted for by assigning the electron an effective mass that is defined as
m* h2 (@ 2 E=@k2 )1 . Since the effective mass is inversely proportional to the curvature of
the energy band, bands that curve upward have positive effective mass and those that
pjwstk|402064|1435428786
curve downward have negative effective mass. Electrons occupying sharply curved bands
have small effective mass and tend to be very mobile, while those occupying flat bands are
heavy and sluggish.
The band structure in real systems is more complex than the simple models we have
developed in trying to understand the electronic properties of metals. This is especially true
in the transition metals where the d-electrons contribute significantly to the band structure.
We see that conductivity is a function of the Fermi surface and the density of states in the
vicinity of the Fermi level. By examining the calculated or measured band structure we can
obtain much more insight into interband transitions, which account for the color of metals,
and why some metals are much better conductors than others.
One method for approximating the band structure that is more suitable for the descrip-
tion of the bands in semiconductors and insulators is the tight-binding method. This
method treats the perturbations on the bound electrons due the overlap of the nearest
Band Theory of Metals 373
neighbor electrons which give rise to the partial delocalization of these bound electrons
and the formation of electronic bands similar to those formed by the nearly free
electrons. The top of these bands is called the valence band. If these bands are completely
full, no conduction can take place. The antibonding states form a conduction band and the
separation between the highest state in the valence band and the lowest antibonding state
is the bandgap energy. Materials with bandgap energies <3.1 eV are considered semicon-
ductors and those with bandgap energies >3.1 eV are considered insulators (3.1 eV is the
top of the visible spectrum).
The presence of energy gaps near the Brillouin zone boundaries distort the Fermi surface
causing it to penetrate into the zone interface. The tight-binding approximation allows a
better approximation of the actual shape of the Fermi for simple cubic, bcc, and fcc
structures for simple metals with only s-electrons.
Even though energy gaps appear in bands at the Brillouin zone, the Brillouin zones occur
at different values of k, depending on the direction taken. Since the bands are symmetrical
about k ¼ 0, it is common practice to plot the bands in one k-direction on one side of the
origin and use the other half of the plot to show the bands in a different k-direction. If the
band in one direction overlaps an energy gap in another direction, electrons have new
energy states to move into and no actual gap exists. This explains the conduction in
divalent metals whose two electrons per atom would fill the first energy band. The
Fermi energy in these circumstances lies in the overlap region causing holes to form in
the first band. Conduction takes place by both electrons and holes, but if the holes
dominate, the metal will exhibit a positive Hall coefficient.
Appendix
A.19.1 Proof That a Wavefunction Displaced by G Is Unchanged
Since uk has the periodicity of the lattice, it can be expressed in a Fourier series as
X
uk ¼ An einkr , (A:19:2)
n
Since G ¼ 2p=a for a simple cubic lattice, exp (inG r) ¼ exp (iG r) and Equation A19.3
can be written as
X
ukþG ¼ eiGr An einkr ¼ eiGr uk : (A:19:4)
n
374 Introduction to the Physics and Chemistry of Materials
Putting this back into Equation A.19.1 and using the Bloch theorem,
Bibliography
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Belmont, MA, 1976.
Hummel, R.E., Electronic Properties of Materials, 3rd edn. Springer-Verlag, New York, 2000.
Ibach, H. and Lüth, H., Solid State Physics, 3rd edn. Springer-Verlag, New York, 1990.
Kittel, C., Introduction to Solid State Physics, 7th edn. John Wiley & Sons, New York, 1966.
Levy, R.A., Principles of Solid State Physics, Academic Press, New York, 1968.
Livingston, J.D., Electronic Properties of Engineering Materials, Wiley, MIT series in Materials Science
and Engineering, John Wiley & Sons, New York, 1999.
Massalski, T.B., Senior Editor, Handbook of Binary Alloy Phase Diagrams, Vols. 1–3, American Society
for Metals, Materials Park, Ohio, 1990.
Omar, M.A., Elementary Solid State Physics, Addison-Wesley, Reading, MA, 1975.
Segal, B., Fermi surface and energy bands of copper, Phys. Rev., 125(1),115, 1962.
Problems
1. Show that the largest Fermi spherepffiffiffithat can just fit inside the first Brillouin zone for a fcc
direct lattice has a radius kF ¼ p 3=a.
2. Show that the largest Fermi spherepffiffithat
ffi can just fit inside the first Brillouin zone for a bcc
direct lattice has a radius kF ¼ p 2=a.
3. Make a scale drawing of the first two Brillouin zones for a fcc direct (bcc reciprocal)
lattice in the (110) plane shown in Figure 6.7a. (Hint: plot the lattice points on graph
paper and construct perpendicular bisectors.) Now draw a set of circles representing the
Fermi spheres for simple metals containing 1, 2, and 3 electrons per atom.
4. Repeat question 3 for a bcc direct (fcc reciprocal) lattice shown in Figure 6.7.
20
Semiconductors
Semiconductors form the core of modern electronic and photonic devices (photonics refers
to the generation, manipulation, and detection of photons). Unlike metals, the conductivity
of pure (intrinsic) semiconductors increases with temperature, which cannot be explained
by the Drude theory. The formal definition of a semiconductor is a material whose
conductivity lies between that of a metal and an insulator. Perhaps a better definition
would be a material with a bandgap greater than zero and less than some arbitrary value
such as 3 eV, which is the highest energy of a visible photon. At least this definition
alludes to the photonic applications of semiconductors. What makes semiconductors
useful is the ability to control both the sign and the number of charge carriers by adding
impurity atoms to form extrinsic semiconductors.
375
376 Introduction to the Physics and Chemistry of Materials
4
p
6
Energy (eV)
Eg
2 s
FIGURE 20.1 4
Band scheme illustrating the transition from
aSn
metal-like behavior for the larger atoms, to semi-
conductor and insulator behavior in the smaller Ge
diameter atoms that are closer together. The band- Si
C
gaps for C, Si, Ge, and a-Sn as well as their lattice
parameters are shown to scale. The number of 0.8
0 0.4 1.00
electrons that can be accommodated in each
band is indicated by the numeral in the band. Lattice parameter (nm)
TABLE 20.1
Cohesive Energies and Bandgap Energies
for Group IV Semiconductors
C Si Ge Sn
Energy gap 2
energy states that lie in the bandgap making the system electronically impure. These
dangling bonds can be tied up by exposing Si to H2 at high temperatures to form Si–H
bonds. This allows amorphous Si to replace single crystal Si in large-scale applications such
as solar cells, which is desirable because amorphous Si can be produced in large scale at a
small fraction of the cost of single crystal Si.
g(EC )
Pe ðEC ,TÞ ¼ (20:1)
exp½(EC EF )=kT þ 1
and the probability of a hole being formed in the valence band is 1 probability of its
remaining in the valence band or
gðEV Þ
Pp ðEV ,T Þ ¼ 1 : (20:2)
exp½(EV EF )=kT þ 1
gðEC Þ gðEV Þ
¼1 : (20:3)
exp½(EC EF )=kT þ 1 exp½(EV EF )=kT þ 1
If the density of states below the valence band g(EV) is the same as the density of states
above the conduction band g(EC), it may be seen that the Fermi level is EF ¼ (EC þ EV)=2
Semiconductors 379
TABLE 20.2
Lattice Parameters and Bandgap Energies for Selected
Semiconductors
System Lattice Parameter (Å) Bandgap (eV)
C 3.567 5.3
Cubic-SiC 4.3596 6.27
Si 5.4307 1.107
Ge 5.657 0.67
a-Sn 6.4912 0.08
Zn-S 5.409 3.54
ZnSe 5.667 2.58
ZnTe 6.101 2.26
CdS 5.5818 2.42
CdSe 6.05 1.74
CdTe 6.477 1.44
HgSe 6.084 0.30
HgTe 6.460 0.15
Cubic-AlN 4.38 6.28
AlP 5.467 2.48
AlAs 5.6622 2.2
AlSb 6.1355 1.6
a-GaN 3.189=5.184 3.42
Cubic-GaN 4.52 3.23
GaP 5.4505 2.24
GaAs 5.6315 1.35
GaSb 6.0854 0.67
InN 3.548=5.760 1.95
InP 5.8687 1.27
InAs 6.0583 0.36
InSb 6.4788 0.165
Sources: Data taken from Gersten, J.I. and Smith, F.W. The
Physics and Chemistry of Materials, John Wiley & Sons,
New York, 2001; Kittel, C., Introduction to Solid State
Physics, 7th edn. John Wiley & Sons, New York, 1966;
Navon, D.H., Electronic Materials and Devices,
Houghton Mifflin, Boston, MA, 1975.
and lies in the middle of the energy gap. (Actually the density of states in the
conduction band is not necessarily the same as in the valence band, but this will be
addressed in Section 20.2.2.)
ð
1
The density of states g(EC) is the same as our previous result (Equation 15.28) with the
effective electron mass m* substituted for m and E EC substituted for E, or
380 Introduction to the Physics and Chemistry of Materials
V 2me* 3=2
gð E C Þ ¼ 2 2
ðE EC Þ1=2 : (20:5)
2p h
Since kT 0.026 eV and (E EF) 1 eV, the Fermi function may be approximated as
1 1
FðE,T Þ ¼ ¼ eðEF EÞ=kT : (20:6)
eðEEF Þ=kT þ1 eðEEF Þ=kT
The number of electrons per volume residing in the conduction band is therefore,
ð
1
Ne V 2me 3=2 EF =kT
n¼ ¼ 2 e ðE EC Þ1=2 eE=kT dE: (20:7)
V 2p h2
EC
where g(EV) is the density of states below the valence band, i.e., it will contain the
factor (EV E)1=2 and m will be replaced by the effective hole mass m*. The F0 (E,T) is
given by
1 e(EEF )=kT 1
F0 ðE,T Þ ¼ 1 ¼ ¼ (E E)=kT e(EF E)=kT , (20:10)
e(EEF )=kT þ 1 e (EEF )=kT þ1 e F þ1
where Peff is the effective density of states for holes. A useful number to remember is
2pme kT 3=2
2 ¼ 2:476 1025 electrons (or holes)=m3 at T ¼ 3008K: (20:12)
h2
Thus the Neff ¼ 2:476 1025 ðme*=me Þ3=2 ðT=300Þ3=2 and Peff ¼ 2:476 1025 ðmp*=mp Þ3=2
ðT=300Þ3=2 .
The electron–hole product is an important quantity that is given by
np ¼ Neff Peff e(EC EV )=kT ¼ Neff Peff eðEg =kTÞ : (20:13)
Semiconductors 381
This is sometimes referred to as the law of mass action. Note that the number of charge
carriers depends only on the effective masses, the temperature, and the energy gap. Also
note that the Fermi level does not appear in this product, which means that no assumptions
have been made concerning the origin of the electrons and holes. Therefore Equation 20.13
applies to electrons that have been extrinsically generated (from donor or collector states)
as well as to intrinsic electrons and holes. As more electrons are added, some will go to
annihilating holes until the equilibrium (Equation 20.13) is reached.
If the material is intrinsic (no electrical active impurities), at equilibrium there must be a
hole for every electron promoted to the conduction band. Therefore, ni ¼ pi and
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi E =2kT
ni ¼ p i ¼ Neff Peff e g (20:14)
and the Fermi level is given by setting n ¼ p and solving Equations 20.8 and 20.11 for EF,
EC þ EV kT Peff
EF ¼ þ ln : (20:15)
2 2 Neff
0
Energy (eV)
−2
−4
−6
−8
FIGURE 20.3
−10 Band structure of Si. Note that the minimum in the
conduction band does not coincide with maximum in
−12
the valence band. (From Cheilkowski, J.R. and Cohen,
L Γ X K Γ M.L., Phys. Rev., B 14=2, 559, 1976. With permission.)
382 Introduction to the Physics and Chemistry of Materials
Energy (eV)
−2
−4
−6
−8
FIGURE 20.4 −10
Band structure of GaAs. Here the minimum in the conduc-
tion band does coincide with maximum in the valence −12
band. (From Herman, F. and Spicer, W.E., Phys. Rev.,
174=3, 906, 1968. With permission.) L Γ X K Γ
in the valence band in the GaAs band diagram, a direct transition is possible at the
bandgap energy.
In an indirect bandgap material such as Si and Ge, a transition at the bandgap energy
must be accompanied by a phonon to supply the needed momentum to reach the conduc-
tion band at its lowest point, as illustrated in Figure 20.5. Such transitions are possible but
are not as likely as the direct transitions possible in a direct bandgap material. This
difference is reflected in the absorbance at the band edge as seen in Section 20.6.
Photons may induce direct or vertical transitions from any occupied band to a higher
energy band that is not completely full. Such transitions are called interband transitions
and contribute to the absorption spectra. So even in an indirect bandgap material, valence
electrons can be promoted vertically to the conduction band by adsorbing photons of
sufficient energy. Once in the conduction band, the hot electrons become thermalized
through collisions and will eventually move to the lowest point in the band.
Electrons may also undergo intraband transitions in which they are promoted to a
higher energy within their own band by absorbing a photon. But since such transitions
must be assisted by a phonon, they, like other indirect transitions, are not as likely to occur
as interband transitions.
In a direct bandgap system such as GaAs, an electron may also make a transition from
the conduction band directly to the valence band by giving off a photon with energy Eg.
However, such a transition is not likely in an indirect bandgap system because the
minimum in the conduction band, where the electrons are likely to reside, does not occur
at k ¼ 0 (at G). In this case the transition, called a nonradiative transition, must involve a
phonon and the energy Eg eventually goes into lattice vibrations or heat. Thus either
FIGURE 20.5
Schematic of direct and indirect bandgap transi- Eg Eg
tion. The vertical transitions are allowed direct
transitions. The indirect transition from the k ¼ 0
to the minimum of the conduction band requires
k=0 k=0
the photon to combine with a phonon in order to
conserve momentum. Direct bandgap Indirect bandgap
Semiconductors 383
system may be used as a photodetector, but only systems with direct bandgaps are useful
for light-emitting diodes (LEDs) or solid-state lasers.
TABLE 20.3
Ionization Energies (eV) for Donors
and Acceptors
Dopant Si Ge
P 0.045 0.012
As 0.049 0.0127
Sb 0.039 0.010
B 0.045 0.0104
Al 0.057 0.0102
Ga 0.065 0.0108
In 0.16 0.0112
Therefore, we see that adding group V impurities creates states in the energy gap close to
the conduction band which are called donor states because they can donate electrons to the
conduction band. In this case, the carriers are electrons and we call the material n-type
because the carriers have a negative charge. Adding group III impurities creates states in
the energy gap close to the valence band which are called acceptor states because they can
accept electrons from the valence band. In this case, the carriers are holes and we call the
material p-type because the carriers have a positive charge. It is important to remember
that in either case the material does not have a net charge; the extra electrons (or holes) are
balanced by the extra (or deficit) positive charge in the impurity ion cores. If equal number
of donor and acceptor impurities are added, the extra electrons from the donor states
occupy the acceptor sites (or you could say that the holes annihilated the electrons by
recombining with them) and the material becomes semi-insulating again, similar to the
intrinsic material. This is called compensation doping.
The ionization energies of the donor and acceptor states for Si and for Ge are given in
Table 20.3. The energies are referred to the nearest band edge (conduction band for donors;
valence band for acceptors).
where ED is the energy of the donor state. Recall from Equation 20.8
n2 DE=kT
nþ e ¼ ND (20:20)
Neff
or
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi!1
ND DE=kT
n ¼ 2ND 1þ 1þ4 e , (20:21)
Neff
where DE ¼ EC ED.
Let us examine some limiting cases:
At very low temperatures, such that 4 (ND=Neff) expDE=kT 1,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n ND Neff eDE=2kT : (20:22)
When ln(n) is plotted against 1=T, the slope is DE=2k. This is characteristic of what is
known as the freeze-out region and the observed slope can be used to determine the
ionization energy, DE.
At high temperatures, where 4 (ND=Neff) exp(DE=kT)
1,
ND DE=kT ND DE=kT
n ¼ ND 1 e þ ND provided e
1: (20:23)
Neff Neff
1028
1024
1024
1020
1018
1016
Intrinsic region 1015
Saturation region
Freeze-out region
FIGURE 20.6 1012
Computed carrier concentration in 0.0 0.01 0.02 0.03 0.04 0.05
n-doped Si as a function of reciprocal
temperature at various doping levels. 1/T (K−1)
n
EF ¼ EC þ kT ln , (20:24)
Neff
where n is the total number of electrons. Now consider the limiting cases for the number of
ionized donors ffi obtained from Equation 20.21. First consider the freeze-out region, where
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
nþ ND Neff exp ( ðDE=2kT)Þ. Putting this into Equation 20.24,
DE kT ND
EF ¼ EC þ ln : (20:25)
2 2 Neff
Notice that at low temperatures, the Fermi level moves to between EC and ED which allows
a large number of donors to be ionized even if kT
DE.
Next consider the saturated region n nþ ¼ ND. Using Equation 20.23, Equation 20.24
becomes
"
#
ND 1 (ND =Neff )eDE=kT þ ND
EF ¼ EC þ kT ln EC þ kT ln : (20:26)
Neff Neff
One can see that adding donors raises the Fermi level. We see from Equation 20.24 that it is
possible to raise the EF above the conduction band in order to make the material ohmic
(metal-like) by making n > Neff. One is tempted to think that this could be accomplished
by setting ND ¼ Neff. However, exp(DE=kT) ¼ 6.66 at 300 K for Si. So as ND ! Neff, the
approximation used in Equation 20.26 is no longer valid and the Fermi level must be found
by using Equation 20.21 with Equation 20.24.
What happens when acceptors are added? Here we have to go back to Equation 20.11,
EF ¼ EV kT ln (p=Peff). For pi, recall that EF was between the conduction and the valence
Semiconductors 387
Eg 1.2
1026
ED 1.0 1024
1021
0.8 1015
Band energy (eV)
0.6 Intrinsic
FIGURE 20.7
0.4 Fermi level as a function of temperature for various
n-doping levels. At low temperatures, the Fermi level
0.2 lies between the donor level and the conduction band
energy which allows the donor states to be ionized at
lower temperatures. At higher temperatures, the Fermi
0 level falls to the intrinsic level. At 1026 electrons=m3, the
101 101.5 102 102.5 103
Fermi level is above the conduction band and the material
Temperature (K) becomes ohmic.
band. We see that increasing the hole concentration pushes the EF toward the valence
band and that by sufficient doping, it is also possible to make p-type material ohmic.
As the temperature is increased, the ni and pi increases while the number of donors
(or acceptors) remains constant (assuming they are all ionized), which drives EF back
toward EiF as seen in Figure 20.7.
J ¼ nmn eðE þ vn BÞ þ pmp e E þ vp B , (20:27)
where the subscripts n and p refer to the electrons and holes, respectively. Setting vy ¼ 0,
Jx ¼ eEx nmn þ pmp (20:28)
and
Jy ¼ eEy nmn þ pmp eBz nmn vnx þ pmp vpx : (20:29)
The velocities of the electrons are vnx ¼ mnEx and the holes are vpx ¼ mpEx. Putting these
into Equation 20.29,
Jy ¼ eEy nmn þ pmp eBz Ex pm2p nm2n : (20:30)
388 Introduction to the Physics and Chemistry of Materials
One can easily see that if n p, RH reduces to 1=ne. Hall measurements are easiest to
interpret in doped materials when either n p or n
p. Otherwise one is faced with four
unknowns, which require other measurements to resolve. For example, except for the
difference between electron and hole mobilities, the Hall effect would be zero for intrinsic
materials. One can also see that doing Hall measurements as a function of temperature
offers a means of determining the occupancy number and energy levels of the various
impurity states in the freeze-out region through Equation 20.22.
or
where e is the magnitude of the charge of an electron. The mobilities of the electrons and
holes are inhibited by scattering from lattice vibrations (phonon scattering) as well as from
impurity atoms and from lattice defects.
Recall that for metals, we argued that vdrift
vth < vF, and since the Fermi velocity was
independent of the electric field and temperature, v was therefore independent of tem-
perature. But in metals, vF > vth because of the fact that every atom contributed at least free
one electron to the Fermi distribution. (We do not count the bound electrons in the inner
core of the atoms.) Similarly, the electrons in the valence band of a semiconductor are
bound and do not contribute to the Fermi level as long as they are in this bound state.
Recall also that we placed the Fermi level between the valence and the conduction bands of
an intrinsic semiconductor so that the probability of promoting an electron to the conduc-
tion band would equal the probability of creating a hole in the valence band. So for
semiconductors, the Fermi velocity will be substantially less than the thermal velocity
since the number of free electrons will be much less than for a metal. Therefore, we can
assume for semiconductors that the average electron velocity v T1=2.
At very low temperatures (<10 K) scattering from lattice defects dominates in semi-
conductors, just as in metals. However, in the case of charged defects such as ionized donor
Semiconductors 389
106
1024
104
1021
102
Conductivity (S/m)
100 1018
10−2
1015
FIGURE 20.8
10−4 Calculated conductivity for n-doped Si as a
function of reciprocal temperature. The
electron and hole mobilities were taken to
10−6 be 0.14 and 0.048 m2=V s, respectively at
0.0 0.01 0.02 0.03 0.04 0.05 300 K and were assumed to fall as the 3=2
power of temperature. In the intrinsic
1/T (K−1) region, s exp (Eg=2kT).
gleaned from the analysis of the absorption edge in which the absorption coefficient a is
measured as a function of v (or l). The a can be obtained from a transmission measure-
ment using Beer’s law, I=I0 ¼ exp(ax) where the transmitted intensity I=I0 through a thin
slab of material as a function of frequency (or l). For very large values of a, it becomes no
longer feasible to measure the transmitted intensity and similar information can be
obtained from reflectivity measurements. Materials that absorb strongly also reflect
strongly for reasons that are discussed in Chapter 24. The structure of the absorption
edge will depend on whether the semiconductor has a direct or indirect bandgap.
20.6.3 Excitons
Although we tend to think of semiconductors being transparent to frequencies below the
bandgap energy, it is possible for a photon with less than the bandgap energy to lift an
electron from the valence band, but instead of going all the way to the conduction band,
the electron becomes attracted to the hole it created and forms a bound pair. Such a bound
electron–hole pair is called an exciton and can be described by hydrogen-like wavefunc-
tions. Since excitons can have several energy states, they are observed as small absorption
102
EgGaAs
Absorption coefficient (mm−1)
101
EgSi
100
10−1
10−2
FIGURE 20.9
Absorption coefficients for Si and GaAs. The band- 10−3
gap edges are located by vertical lines. (Absorption
coefficients were computed from Palik, E.D., Ed., 10−4 1.0
1.5 2.0 2.5 3.0
Handbook of Optical Constants of Solids, Vols. 1–3,
Academic Press, 1997.) Photon energy (eV)
Semiconductors 391
0.8
Reflectance
0.6
0.4
0.2
FIGURE 20.10
Reflectance spectrum of Si. Peaks at 3.4, 4.5, and
5.4 eV are direct interband transitions. (Reflectivity
0 was computed from Palik, E.D., Ed., Handbook of
0 E 5 10 15 20
g Optical Constants of Solids, Vols. 1–3, Academic Press,
Photon energy (eV) 1997.)
lines at energies just short of the absorption edge (or at wavelengths just beyond the cut-on
wavelength). Excitons can move through the crystal carrying energy (but not charge) and
eventually decay giving up their energy as a photon.
dopants, such as the iodine that Heeger, MacDiarmid, and Shirakawa added to polyace-
tylene, are required to increase the conductivity of polymers with conjugated electrons.
Dopants such as Na, K, Li, Ca, and tetrabutyl ammonium act as donors for n-type
polymers while I2, PH6, BF4, Cl, AsH4 act as acceptors for p-type polymers. Dopants are
added by exposing the polymer (or its precursor) to vapors or solutions containing
the dopant or by electrochemical methods. Unlike semiconductor in which the dopants
take their place as substitutional defects, dopants in polymers take their places interstitially
among the polymer chains and donate to or accept charges from the polymer backbone
thus forming new three-dimensional structures. The original I2-doped polyacetylene deve-
loped by Heeger et al. had a conductivity of 100 S=cm but now polymer conductivities
have exceeded 105 S=cm, which exceeds many metals.
20.8 Summary
Most semiconductors used in electronic and photonic devices are group IV elements or
compounds formed by III–V or II–VI elements. They are characterized by an sp3 hybrid
bond in which an s-electron is promoted to a p-state in order to form a tetrahedral directed
diamond-like bond. The four valence electrons occupy the bonding states from which the
antibonding states, which form the conduction band, are separated by the energy gap Eg.
Semiconductors 393
The width of the energy gap decreases as the lattice parameter increases, or as the atoms
become larger and farther apart. Thus, diamond has the highest bandgap energy and the
bandgap decreases with Si, Ge, and disappears with Sn and Pb, which are metals. The
compound systems follow the same trend.
Even though the valence electrons are tied up in covalent bonds, they are not com-
pletely localized and have similar band structure and dispersion relationships (E vs. k) as
nearly free electrons. Thus the band structure becomes a powerful tool for understanding
the properties and behavior of semiconductors.
When an electron is promoted to the conduction band, it leaves a hole in the valence
band, which frees up states in the valence band so that the valence electrons can also
respond to an electrical field. The hole in the valence band can be treated as a particle
with positive mass and positive charge, while the electron in the conduction band
behaves as a particle with positive mass and negative charge. Thus in response to an
applied electrical field, electrons carry the charge in one direction while holes carry the
opposite charge in the opposite direction and both contribute to the conduction of
current.
Systems in which the minimum of the conduction band coincides with the maximum in
the valence band at k ¼ 0 are called direct bandgap materials. In such materials, electrons
may go directly from the valence band to the conduction by absorbing a photon with
energy equal or greater than Eg. GaAs is an example of a direct bandgap material.
Si and Ge are indirect bandgap materials in which the minimum energy gap is at some
distance k from 0. In such systems, a photon with energy Eg must interact with a
phonon with momentum hk to promote an electron to the conduction band, which has
a lower probability than a direct transition. Electrons can be promoted directly to a higher
conduction band at k ¼ 0 if they have sufficient energy.
In direct bandgap materials, electrons may go directly from the conduction band to the
valence band with a photon emission, which makes them suitable as LEDs or solid-state
lasers. Indirect transitions from the conduction to the valence band must be accompanied
by the emission of a phonon and are nonradiative transitions, which results in heating the
material.
Pure materials, or at least materials with no impurity states in the bandgap region, are
called intrinsic semiconductors. The creation of an electron leaves a hole; therefore, the
number of holes must equal the number of electrons in an intrinsic material. The electron–
hole product is directly proportional to the Boltzmann factor, exp (Eg =kT). The Fermi
level, the energy for which the probability of creating a conduction electron is the same as
creating a hole is shown to be somewhere in the bandgap.
Pentavalent atoms may be added to group IV semiconductors to form donor states
close to the conduction band. Electrons are loosely attached to these donor atoms and are
thermally ionized to produce conduction electron without producing holes. The number
of conduction electrons, hence the conductivity of the materials, can be controlled by the
doping level or the number of donor atoms added. Material in which the majority carriers
are electrons are referred to as n-type materials. Adding donor states moves the Fermi
level toward the conduction band.
Similarly, trivalent atoms may be added to group IV semiconductors to form acceptor
states close to the valence band. Atoms from the valence band may be easily thermally
ionized to replace the missing electrons in the covalent bond, leaving holes in the valence
band. Materials in which the majority carriers are holes are referred to as p-type materials.
Adding acceptor states moves the Fermi level toward the valence band.
Neither n- or p-type materials have a net charge. The excess electrons (or holes) are
balanced by the charge of the ion cores. Also the np product remains the same as that of
394 Introduction to the Physics and Chemistry of Materials
intrinsic materials regardless of the dopants added. Adding electrons reduces the number
of holes and vice versa.
At low temperatures, not all of the donors (or acceptors) are ionized and the number
of carriers n(or p) exp (DE=2kT), where DE is the ionization energy of the donor
(or acceptor). This energy can be determined by making Hall measurements in this
freeze-out region.
At ambient temperatures, most of the donors or acceptors will be ionized and the carrier
concentration will essentially be the number of donors or acceptors in the system. Most
devices are designed to operate in this plateau or saturation region so their performance
will not be affected by modest changes in temperature.
At still higher temperatures where kT Eg, the intrinsic electron concentration
becomes larger than the donor (or acceptor) concentration and the material becomes
intrinsic again.
The Hall coefficient is more complicated when both electrons and holes are
present because they tend to cancel each other out in the Hall current. The Hall meas-
urement is the easiest to interpret in doped materials when the minority carriers can be
ignored.
The conductivity of a semiconductor is given by s ¼ nemn þ pemp. Except for very low
temperatures, the mobilities of the charge carriers in a semiconductor T3=2; therefore
in the plateau region, where the carrier concentration is more or less constant, s T3=2.
However, in the freeze-out region and in the intrinsic region, the carrier concentration is
increasing with temperature faster than the mobility is decreasing and the conductivity
increases with temperature.
Bibliography
Adachi, S., Ed., Handbook on Physical Properties of Semiconductors, Vols. 1–3, Springer-Verlag,
New York, 2004.
Cheilkowski, J.R. and Cohen, M.L., Nonlocal pseudo calculations for the electronic structure of eleven
diamond and zinc-blende semiconductors, Phys. Rev., B 14=2, 559, 1976.
Gersten, J.I. and Smith, F.W., The Physics and Chemistry of Materials, John Wiley & Sons, New York,
2001.
Herman, F. and Spicer, W.E., Spectral analysis of photoemissive yields in GaAs and related crystals,
Phys. Rev., 174=3, 906, 1986.
Hummel, R.E., Electronic Properties of Materials, 3rd edn. Springer, New York, 2000.
Ibach, H. and Lüth, H., Solid State Physics, 3rd edn. Springer-Verlag, New York, 1990.
Kittel, C., Introduction to Solid State Physics, 7th edn. John Wiley & Sons, New York, 1966.
Livingston, J.D., Electronic Properties of Engineering Materials, Wiley, MIT series in Materials Science
and Engineering, John Wiley & Sons, New York, 1999.
Navon, D.H., Electronic Materials and Devices, Houghton Mifflin, Boston, MA, 1975.
Palik, E.D., Ed., Handbook of Optical Constants of Solids, Vols. 1–3, Academic Press, San Diego,
CA, 1997.
Sapoval, B. and Hermann, C., Physics of Semiconductors, Springer-Verlag, New York, 1995.
Problems
You are given a piece of unknown semiconductor material and are asked to charac-
terize it electrically. Using cyclotron resonance you have determined the effective
Semiconductors 395
masses of the holes to be 0.5me and the electrons to be 0.4me. You have made a series of
Hall measurements and conductivity measurements with the results given below:
Now that we have laid the groundwork for understanding how we can engineer various
carrier concentrations and Fermi levels into semiconducting materials, we can explore
some of the numerous devices that have become possible using these novel materials.
397
398 Introduction to the Physics and Chemistry of Materials
+ _
+ _
+ _ _
_ +
p + n _
_ + +
+ _ _
+ P n
+ _ + _
+ _
(a) (b)
FIGURE 21.1
(a) Forward-biased p–n junction. The applied field pushes the electrons and holes together at the junction where
they combine causing current I to flow. (b) Reverse-biased p–n junction. The applied field draws the charge
carriers away from the junction thus preventing their recombination and no current flows.
Vacuum level
cp
cp cn Ec
eV0 = EFn − EFp cn
Ec
EFn Ex = −grad V
EV
EFP
EV
p-material n-material p-material
n-material
Depletion zone
FIGURE 21.2
When a p-type material is brought into intimate contact with n-type material, electrons flow from the n- to the
p-material until their Fermi energies (chemical potentials) become equal, causing a shift in the conduction and
valence bands in both the p-side and the n-side, which creates a contact potential difference V0. The charge
separation results in a strong electric field E ¼ rV (right-to-left).
which this field acts is called the depletion region or the space charge region. In this region,
the electrons and holes have recombined; so there are no free carriers. Even though an
electron combines with a hole and is no longer a free carrier, it still deposits a charge. The
built-in electric field caused by the diffusion current is really the feature that makes p–n
junctions useful as we shall see.
21.1.1 Analysis
Now let us see if we can describe this process analytically. From Equation 20.18 we can see
that adding ND donor states will raise the Fermi level above the intrinsic Fermi level in the
n-material by
nn ND þ ni
EnF EiF ¼ kT ln ¼ kT ln (if all donors ionized): (21:1)
ni ni
Theory and Applications of Junctions 399
pp = NA E(x) nn = ND
p-type Depletion region n-type
nP = ni2/NA Pn = ni2/ND
Xp x=0 Xn
E(0)
FIGURE 21.3
Schematic of the electron and hole concentration and their associated electric field near an unbiased p–n junction.
The electric field as a function of x in the depletion zone is depicted underneath the junction.
Similarly, adding acceptor states lowers the Fermi level below the intrinsic Fermi level in
the p-material by
p pp NA þ ni
EiF EF ¼ kT ln ¼ kT ln (if all acceptors ionized): (21:2)
ni ni
Subtracting Equation 21.2 from Equation 21.1, the contact potential V0 is obtained as
p nn p p
eV0 ¼ EnF EF ¼ kT ln : (21:3)
n2i
If all donor and acceptor states are ionized and if NA ni and ND ni, the number of
electrons in the n-type material, nn ¼ ND, and the number of holes in the p-type material,
pp ¼ NA, and the contact potential V0 is given by
p ND NA
eV0 ¼ EnF EF ¼ kT ln : (21:4)
n2i
Recall that ni2 ¼ ni pi ¼ nn pn ¼ np pp, where pn and np are the minority carriers in the p-type
and n-type materials, respectively.
n n pp pp nn
eV0 ¼ kT ln ¼ kT ln ¼ kT ln : (21:5)
nn pn pn np
For NA ¼ ND ¼ 1024 electrons=m3 and ni np ¼ 1032=m3 at 300 K, the contact potential will be
48
10
V0 ¼ 0:026 ln ¼ 0:958 V:
1032
400 Introduction to the Physics and Chemistry of Materials
An alternative derivation of the above result can be obtained by writing the hole current as
the sum of the current produced by the electric field and the diffusion current,
@p
Jp ¼ e pmp Ex Dp : (21:6)
@x
@p @V
Dp ¼ pmp Ex ¼ pmp : (21:7)
@x @x
Since the hole gradient is negative, hole flow is positive or from the p- to n-type material.
This means the Ex must be negative or in the direction of the p-type material.
Integrating this expression with the boundary condition that V ¼ Vp and p ¼ pp in the
p-type material far removed from the junction, we get
pp
Dp ln ¼ mp V Vp : (21:8)
p
Similarly, if we take the boundary condition that V ¼ Vn and p ¼ pn in the n-type material
far away from the junction,
pn
Dp ln ¼ mp ðV Vn Þ: (21:9)
p
Subtracting Equation 21.9 from Equation 21.8, and using the Einstein relationship,
Dp ¼ kTmp=e, we get
pp
e(Vn Vp ) ¼ eV0 ¼ kT ln , (21:10)
pn
dEx eND
¼ from 0 to xn , (21:11)
dx «
where « is the dielectric constant. Integrating with the boundary condition that requires the
electric (E) field to vanish at xn, we get
eND
Ex ¼ (x xn ): (21:12)
«
Theory and Applications of Junctions 401
Integrating again with the boundary condition that requires the potential to vanish
at xn,
eND 2
V¼ (x x2 ): (21:13)
2« n
A similar argument can be made for the width of the depletion zone in the p-side. Note that
the width of the depletion zone varies as the inverse square root of the doping level. Using
the same doping level as above and taking the dielectric constant to be 12, the width of the
depletion zone into the n-side is
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 12 8:85 1012 0:985
xn ¼ ¼ 3:6 108 m:
1:6 1019 1024
Thus we see that the depletion layers are extremely thin, on the order of tens of nano-
meters. Putting this width into Equation 21.12, we obtain an electric field of 5.54 107 V=m.
pp Þ x p
¼ ee(V0 V)=kT ½e(V0 V)=kT , (21:15)
pn Þxn
Vacuum level
Ec
e (V0 − V )
eV
EF
Ev FIGURE 21.4
Forward-biased p–n junction. Applying a plus voltage to
the p-material reduces the contact potential and raises the
effective Fermi level allowing more electrons to diffuse
Depletion zone
from the n-side to the p-side.
402 Introduction to the Physics and Chemistry of Materials
p p = NA nn = ND
Xp x=0 Xn
FIGURE 21.5
Carrier distribution and related electric field in a forward-biased p–n junction. The number of holes that were
injected at the edge of the depletion zone on the n-side is given by pn exp (eV=kT) where pn is the hole density in the
n-material away from the junction region. Note also that the electric field has been reduced by 2V=xn.
where
pp)xn is the hole concentration in the p-type material at the edge of the space charge
region
pn)xn is the hole or minority carrier concentration in the n-type material at the edge of
the space charge region
pp
¼ ee(V0 )=kT , (21:16)
pn
then
This is known as the junction law and says that the minority carrier density at the edge of
the space charge region is enhanced by exp(eV=kT), where V is the applied forward bias.
The excess hole density is
This process of enhancing the minority carrier concentration at the edge of the space charge
region is called minority carrier injection. The holes that were injected into the n-type
material by the applied voltage will migrate toward the negative electrode and will be met
by electrons drifting from the negative electrode. The holes are annihilated by the electrons
so that J ¼ e(nmn þ pmp)E remains constant.
When the bias is reversed as shown in Figure 21.6, the equilibrium is disturbed in the
other direction. The increased potential forces minority holes from the n-side of the
depletion zone into the space charge region where they are swept into the p-material
resulting in a reverse current flow. However, in this case, the current is carried by the
minority carriers so the reverse-biased current will be much less than the forward-biased
current.
Theory and Applications of Junctions 403
Vacuum level
Ec
e (V0+V )
EFn
EV eV
FIGURE 21.6
Reversing the bias increases the electric field, which shuts
down the flow of majority carriers, but now allows minority
carriers to diffuse through the depletion zone carrying a small
Depletion zone
amount of leakage current in the direction of the reverse bias.
where I0 is the limit of the leakage current due to the minority carriers flowing down the
potential gradient under reverse bias. However, because minority carriers have a limited
lifetime t before they are annihilated by a majority carrier, their velocities are restricted to
their diffusion speeds given by D=L, where D is their diffusion coefficient and L is the
pffiffiffiffiffiffi
diffusion length which is given by Dt .
The leakage current I0 can be expressed as
0sffiffiffiffiffiffi sffiffiffiffiffiffi 1 0 sffiffiffiffiffiffi sffiffiffiffiffiffi 1
p p
Dnp D 1 D n
1 Dn A
I0 ¼ e@ A ¼ en2 @
n p
n p þ p p n þ p pn , (21:20)
t np tn i
NA t np ND t n
where
Dnp is the diffusion coefficient of holes in the n-material
p
Dn is the diffusion coefficient of electrons in the p-material
pjwstk|402064|1435428802
p
t np and t n are the lifetimes of the respective minority carriers
These lifetimes are extremely short because there are so many majority carriers to combine
with. Neither the diffusion coefficients nor the lifetimes are particularly temperature
dependent, but n2i exp (Eg =kT). The leakage current can be minimized by selecting a
high bandgap material and by heavy doping. Typical values for the leakage current
density J0 are O(106 A=m2) at ambient temperatures.
Equation 21.20 is known as the rectifier equation because of the fact that a large current
can flow if V is positive (p-side made positive or forward biased), but only a small current
results as V is made negative (reverse biased) as shown in Figure 21.7. Circuit elements that
utilize p–n junctions are known as diodes. They find use as rectifiers, blocking devices for
preventing unwanted reverse current flow, and even as logic devices.
404 Introduction to the Physics and Chemistry of Materials
60
50
40
Current (μA)
30
20
10
Vacuum level
χn fn Metal n-type
fm Ec
EFn eV0 = fm − fn
Ec
EFm EF
Metal EV
n-type
EV
Depletion
zone
FIGURE 21.8
Band structure at the interface of an n-type semiconductor and a metal with a lower Fermi level. Electrons will
flow from the conduction band into the metal giving it a negative charge, which results in a contact potential V0.
The resulting electric field limits the electron diffusion just as in a p–n junction.
proportional to J exp ½eðV0 V Þ=kT . Since J ¼ 0 when V ¼ 0, the current flow can be
written as
J eeðV0 V Þ=kT eeV0 =kT ¼ J0 eeV=kT 1 , (21:21)
Vaccum level
Metal n-type
fm cn fn
Ec
EFm EF
Ec
EFn
Metal n-type Ev
Ev
Depletion
zone
FIGURE 21.9
Ohmic contact at an n-metal interface in which fm < fn. Here electrons can flow directly from the metal into the
n-semiconductor and vice versa and no rectification takes place.
brought together, electrons will flow from the metal to the semiconductor creating a layer
of negative charge at the semiconductor interface. The resulting field bends the conduction
band down allowing the electrons from the metal to flow directly into the conduction band
as illustrated in Figure 21.9. Similarly, ohmic junctions can be made in p-type semicon-
ductors by selecting metals in which fm > fp.
Of course it is not always possible to select metals for the interconnects that have the
desired Fermi levels. As an alternative, heavy doping moves the Fermi level closer to
the conduction band in n-type semiconductors as we saw in Chapter 20. This reduces the
Schottky barrier as well as the width of the depletion region so that electrons can tunnel
through the barrier. An even better alternative, if possible, is to dope heavily enough to
move the Fermi level into the conduction band. Then the semiconductor becomes metal-
like and the barrier is eliminated. Similar arguments apply to doping p-type semiconduct-
ors so that the Fermi level moves into the valance band.
V kT V 1 kT kT
Rð V Þ ¼ ¼ ð0:025 V=2Þ, <V< : (21:22)
I0 ðeeV=kT 1Þ eI0 2I0 I0 e e
VeV=0:025
R ðV Þ ¼ , (21:23)
I0
RL
Voltage
Time
FIGURE 21.10
Simple half-wave rectifier circuit (upper). The symbol for a diode is
an isosceles triangle with a line at its vertex. The alternating voltage
Voltage
RL
Voltage
FIGURE 21.11
Time Full-wave rectifier using a bridge circuit. The bridge circuit pro-
vides positive voltage to the load regardless of the polarity of the
source voltage.
408 Introduction to the Physics and Chemistry of Materials
FIGURE 21.12
Diagram of a simple crystal receiver for AM broadcasts. The carrier
frequency is picked up at the antenna and appears across the L–C tank
circuit. The diode clips the negative half of the carrier waveform and the
headphones respond to the amplitude of the audio waveform that was
impressed on the carrier waveform.
A B V
A B Out A B Out
0 0 0 0 0 0
1 0 1 1 0 0
0 1 1 A B 0 1 0
Out Out
1 1 1 1 1 1
FIGURE 21.13
Logical OR circuit (left) and AND circuit (right) with their corresponding truth tables. The applied voltage V is
typically 5 V and a logical ‘‘1’’ is 5 V and a logical ‘‘0’’ is ground.
Theory and Applications of Junctions 409
Binary output
1 2 4 8
2
3
Decimal input
4
5
6
7
8 FIGURE 21.14
9 Decimal to binary converter matrix using diodes as a ROM.
For example, an input on the ‘‘6’’ line will produce ‘‘0110’’ on
the output lines.
EC
EV
EF EF
EC
n-type
(a) EV
EC
p-type
EV EF
EF EC
(b) n-type
EV
EC
FIGURE 21.15
Unbiased tunnel diode (a). Small for- p-type EC
ward bias (b) causes tunneling current n-type
EV
flow. When forward bias raises EC in EF
EF
the n-material above EV in the p-material
(c), tunneling stops and the device EV
becomes an ordinary diode. (c)
Gunn-effect oscillators (or Gunn diodes as they are frequently called, although technic-
ally, they are not diodes) in a resonant cavity can be used to make very simple low power
microwave oscillators. Handheld radar speed detectors use Gunn-effect oscillators as their
microwave source.
60
50
40
30
Current (μA)
b
20
10
c
0 a
−10
−20
FIGURE 21.16
I–V curve for a tunnel diode. Tunneling current
−30
increases with forward bias from a to b. When tunnel-
ing stops, the current drops to c. The region b–c − 0.1 − 0.05 0 0.05 0.1
exhibits negative resistance. Voltage (V)
Theory and Applications of Junctions 411
0
Energy (eV)
−2
FIGURE 21.17
Band structure of GaAs in the vicinity of the energy gap
(adapted from Figure 20.4). Note the shallow secondary
valley (X) to the right of the steep valley centered at the
direct bandgap (G). Increasing the applied voltage moves
high mobility electrons through intraband transitions
from G to X where their mobilites are lower. Thus,
increase in the voltage results in decreased current flow
L Γ X K Γ and negative effective resistance.
lasers for CD and DVD will allow four times the storage of information because of the
smaller diffraction-limited spot size.
the formation of triplet excitons is favored by 3:1, so the quantum efficiency of an OLED
material appeared to be inherently limited to less than 25%.
An important breakthrough came with the discovery that with the addition of phos-
phorescent dopant emitters such as Pt, Ir, Os, etc., the energy could be transferred from
the triplet states to the dopant molecules which could decay radiatively. As a result, the
internal quantum efficiency (the number of photons generated inside the device per
electron–hole pair injected) of these phosphorescent organic light-emitting diodes (PHO-
LEDs) (PHOLED is a trademark of the Universal Display Corporation, Ewing, New Jersey)
can approach a quantum efficiency 100% at a luminescence of 100 cd=sr.
Despite the gains in internal quantum efficiency, the overall efficiency is still less than
19%. As much as 80% of the photons produced cannot get out because of the high index of
refraction of the polymer material. Approaches such as texturing the surface of the glass,
index matching or using low index substrate materials, and the use of arrays of microlenses
have made increases in the extraction efficiency.
OLED flat and flexible panel display technology is advancing rapidly and full color
displays are currently being used in cell phones. Sony recently announced a 2.5 in. flexible
screen TV that is only 0.3 mm thick and now has an 11 in. OLED TV in production.
Samsung also announced a prototype 17 in. high definition (1600 1200 pixels) active
matrix OLED display panel.
Flexible PLED electro luminescent (EL) panels are now available from several manufac-
turers. Research is still underway to improve the color, durability, and efficiency of the EL
panels. A replacement for the ITO-coated glass or plastic, which presently is the only
practical hole injecting material, is being sought. ITO is brittle and can be easily cracked.
The ITO glass has a transmission of only 0.85, so 15% of the light is lost at the window.
However, the biggest problem with ITO is that its conductivity is too low to deliver power
to large areas.
21.5 Photodiode
Photodiodes convert photons into electrical current and are used as power sources as well
as photodetectors. Recall that the internal fields setup in the depletion region sweep holes
toward the p-region and electrons toward the n-region. Photons above the energy gap
absorbed in or near the depletion region produce the electron–hole pairs, which are swept
414 Introduction to the Physics and Chemistry of Materials
Ec
p-type
EV
EF EF
Ec
n-type
EV
(a)
Ec
Ec
p-type hn eV n-type
EV
EF EF
EV
(b)
FIGURE 21.18
Schematic showing the operation of a solid state laser. In the unbiased state (a) the Fermi level lies above the
conduction band in the n-side and below the valence band on the p-side. (b) Biasing the junction lifts the filled
states in the conduction band in the n-region above the above empty states below the valence band in the p-side,
thus inverting the electron population.
to their respective regions where they may be collected by an external circuit. As opposed
to photoconductive detectors whose conductance depends on the photon flux, photovol-
taic (p.v.) detectors produce a voltage that is proportional to photon flux.
Photons absorbed
in the depletion
layer
Conductor
FIGURE 21.19 p-type
Schematic of the structure of a typical solar cell.
Depletion region
Electrons and holes created by the photons are
swept by the internal field in the depletion region n-type −−−−−−−
toward the negative and positive electrodes
respectively, thus creating a source of current flow. Metal backing
Theory and Applications of Junctions 415
I0(exp(V/kT )−1)
+ −
Icircuit
Iphoto
FIGURE 21.20
Equivalent circuit of a solar cell as a current source in parallel
RL
with a forward-biased diode.
where
I0 is the leakage current
A is the exposed area of the cell
G is the rate per volume at which photons produce electron–hole pairs
Lp and Lh are the diffusion lengths of the minority carriers.
Some texts define a photodiode current Ipd that flows in the direction of the current in the
diode or Ipd ¼ Icircuit as shown in Figure 21.21. The operating point is in the lower right
hand corner and the negative I–V product signifies power being generated.
In the absence of a load, the photocurrent increases the voltage across the cell until it is
balanced by the current flowing through the forward-biased diode. The open circuit voltage
kT Iphoto
Voc ¼ ln þ1 : (21:25)
e I0
2.0
1.5
Photodiode current (mA)
1.0
0.5 Dark
Voc
−0.5 Isc
Illuminated
FIGURE 21.21
−0.6
I–V curve for a solar cell in the dark and when illuminated.
The operating point of a solar battery is in the fourth quad-
−6 −4 −2 0 2 4
rant where the negative I–V product indicator power being
Voltage (V) generated.
416 Introduction to the Physics and Chemistry of Materials
matched load. Single crystal Si is expensive to produce since wafers must be sawed from
Czochralski-grown ingots. It is much more cost effective to use polycrystalline or amorph-
ous Si and accept the reduction in efficiency due to the shortened carrier lifetimes in these
cheaper materials. Polycrystalline cells have efficiencies that range from 13% to 15% and
amorphous cells are only 5%–7% efficient.
The primary factor that reduces the efficiency of a solar cell is the fact that photons with
energies less than the bandgap cannot produce electron–hole pairs. We saw in Figure 20.9
that even though the absorption edge for Si is 1.1 eV, the absorptivity increases fairly
slowly until around 3–4 eV because of the indirect bandgap, thus 30%–50% of the
available spectrum is lost to begin with. Materials with lower bandgaps can be used, but
they are not as efficient in capturing the shorter wavelength radiation.
One solution, although by no means a cheap alternative, is to make compound cells.
A stack of cells starts with a high bandgap cell on the sun side to tap the higher energy
photons. Photons with energies below the bandgap of the high bandgap cell are not
absorbed and go on the second and third layers where the lower energy photons get
absorbed. For example, Spectrolab, Inc. manufactures a triple junction cell for use in
spacecraft that sandwiches p–n GaInP2=GaAs=Ge cells in a stack. The ISC ¼ 170 A=m2
and the VOC ¼ 2.665 V. The power delivered to a matched load is 392 W=m2 for an
efficiency of 28% (assuming the full unattenuated solar input of 1400 W=m2) is incident
on the cell. As one might imagine, such cells are extremely expensive to manufacture and
are generally used only in space applications where weight and transportation costs can
justify the manufacturing cost of such cells.
Another alternative is to use mirrors or to concentrate the sun on the cells. Recently a
team led by Allen Barnett (University of Delaware) announced an efficiency of 42.8%
obtained by concentrating 20 suns onto a beam splitter that divided the solar input into
three spectral bands and imaged each of these bands onto a cell optimized for that band.
21.6 Summary
Since the Fermi level is moved closer to the conduction band in n-type semiconductors and
closer to the valence band in p-type semiconductors, bringing these materials into intimate
contact will cause electrons to flow from the n-type into the p-type material in order to
equalize their Fermi levels. The electrons that diffuse into the p-type material create a
depletion zone in the vicinity of the junction where there are no free carriers. The annihi-
lation of an electron–hole pair leaves a negative charge in the p-material and a positive
charge in the n-material. This charge separation creates a contact potential and an electric
field in the depletion zone.
Equilibrium is reached at the contact potential V0 when the field in the depletion zone
balances the diffusion current. The field is in the direction of the p-type material, so any
hole that happens to be in the depletion region is swept toward the p-type material.
Likewise any electron that drifts into the depletion region will be swept into the n-type
material.
Reducing the contact potential by applying a forward bias or positive voltage to the
p-material upsets the equilibrium and increases the flow of holes, resulting in current flow
through the circuit. The current flow requires electrons to leave the n-type material and
flow through the junction, leaving excess holes in the vicinity of the depletion zone. These
holes will migrate toward the negative electrode until they recombine with the electrons
provided by the current source, which also replenishes the holes in the p-type material.
Theory and Applications of Junctions 417
This process is called minority carrier injection and the time before these minority carriers
recombine is the minority carrier lifetime.
When the bias is reversed, the minority carriers are driven into the depletion region by
the applied potential resulting in a reverse current, but since the reverse current consists of
the minority carriers, it will be much smaller than the forward-biased current. The circuit
response is described by the rectifier equation, I ¼ I0 ½expðeV=kTÞ 1, where I0 is the
leakage current from the minority carriers driven by the reverse bias potential. This non-
linear response is useful for a number of circuit applications such as rectifiers to convert AC
to DC.
Diodes may be constructed to break down at a given reverse potential allowing them to
be used as voltage regulators. This type of diode is known as a Zener diode.
Diodes can also be formed by a metal–semiconductor junction. These diodes are called
Schottky diodes. In order to make metal to semiconductor contacts that do not act as
diodes, it is necessary to make the semiconducting material ohmic or metal-like in the
vicinity of the contact by raising the doping level to put the Fermi energy in the conduction
(or valence) band.
Using the negative resistance characteristics of tunnel diodes and Gunn diodes, simple
low-cost oscillators are possible that have found their way into handheld radar units and
other microwave applications.
LEDs can produce infrared or visible light from the recombination of carriers in the
junction regions. By inverting the electron population and providing a resonant cavity,
stimulated emission may be produced making solid-state lasers possible.
It is also possible for diodes to capture photons near the junction and produce voltage
and current, thus making p.v. detectors and solar cells possible.
Bibliography
Hummel, R.E., Electronic Properties of Materials, 3rd edn., Springer-Verlag, New York, 2000.
Navon, D.H., Electronic Materials and Devices, Houghton Mifflin, Boston, MA, 1975.
Kwong, R.C. et al., High efficiency light emitters, Org. Electron., 4, 2003, 155–164.
Livingston, J.D., Electronic Properties of Engineering Materials, Wiley, MIT series in Materials Science
and Engineering, John Wiley & Sons, 1999.
Sapoval, B. and Hermann, C., Physics of Semiconductors, Springer-Verlag, New York, 1995.
Tang, C.W. and Van Slyke, S.A., Organic electroluminescent diode, Appl. Phys. Lett., 51, 1987, 914.
Problems
Spectrolab makes a GaAs=Ge single junction solar cell that has an open circuit voltage
VOC ¼ 1.025 V and a short-circuit current JSC ¼ 30.5 mA=cm2 in full sun (0 air mass).
1. Find J0.
2. What is the maximum power a 1 m2 array of these cells could deliver to a resistive
load?
3. What would the load resistance need to be in order to obtain maximum power?
4. Solar constant at zero air mass is 1354 W=m2. What is the efficiency of this cell?
22
Transistors, Quantum Wells, and Superlattices
Before the invention of the transistor in 1947 at Bell Laboratories by John Bardeen, Walter
H. Brattain, and William B. Shockley, vacuum tubes were the workhorse of electronics.
Vacuum tubes were bulky and fragile, required high voltage (>45 V) and a heater current
to operate, and had lifetimes comparable to electric light bulbs. It is certainly no exagger-
ation to say that the transistor made possible modern solid-state electronics. Transistors
serve much the same function as the old vacuum tubes but are much smaller, rugged,
operate on low voltages, and have virtually infinite lifetimes.
Once the purification and processing technology required to make transistors was
developed, a virtual explosion of new concepts became possible resulting in a vast array
of new electronic and photonic devices, a few of which will be described in this chapter.
The active mode is of most interest if the transistor is to be used as an amplifier. The
saturated and cutoff modes are of interest if the transistor is to be used as a switch.
The inverse mode is used only in very special applications and will not be considered here.
In the active mode, minority carriers are injected into the base region by forward biasing
the emitter, just as in a forward-biased diode. Because of the narrow depletion zone
419
420 Introduction to the Physics and Chemistry of Materials
FIGURE 22.1
Schematic of a pnp transistor. An npn transistor is similar except
Ib
the n and p are interchanged.
+ _
(a) (b)
FIGURE 22.2
Standard symbols for (a) pnp and (b) npn. BJTs biased in the active or ‘‘on’’ state. The arrow is in the direction of
the forward bias of the emitter–base portion.
between the emitter and base and aided by the reverse bias between the base and collector,
most of the carriers are swept into the collector before they can recombine and contribute to
the base current.
where a is the fraction of the minority carriers that are swept to the collector. Since Vc is
negative and I0c aIe, Ic aIe.
Ie Ic
+ RL VL
FIGURE 22.3 Ve
Common base amplifier circuit using a pnp transistor. An equiva-
lent circuit can be constructed using an npn transistor by reversing
the above polarities. +
Transistors, Quantum Wells, and Superlattices 421
If a load is placed in series with the source providing the reverse bias to the collector
circuit, the voltage across the load is
VL ¼ aIe RL , (22:3)
Let RL ¼ 5 KV, Ie ¼ 5 mA, kT=e ¼ 0.026 V, a 1, the voltage gain is 1000. In this case the
voltage gain is determined by the load resistor and can be virtually any value.
Ic aIe
¼ ¼ b, (22:6)
Ib ð1 aÞIe
where the short circuit current gain b is defined as a=(1 a). Recall that a is the fraction of
the injected minority carriers that make it to the collector and is a function of the design of
the transistor, i.e., the width of the base region and the dopant levels that determine the
width of the depletion zones. Therefore, the b is also a function of the transistor design and
is used to characterize the transistor.
+
+
Ib
Ie
FIGURE 22.4
_ Common emitter current amplifier using an npn transistor.
422 Introduction to the Physics and Chemistry of Materials
−V00
VC1 VC2
T1 T2
FIGURE 22.5
Basic flip-flop circuit using pnp BJTs. If T1 is conducting, T2 is
turned off and vice versa. A negative pulse at VC1 will turn on T2
and turn off T1. This type of circuit is the basic element in
counters and serves as the memory element in a SRAM.
shutting it off. Conversely, if T2 is conducting, T1 is shut off. So the circuit is bistable: either
one or the other transistor is ‘‘on’’ and the other is ‘‘off.’’
If T1 happens to be on, the state may be changed by applying a negative pulse to VC1.
This will cause T2 to start conducting, shutting off T1.
Flip-flop circuits can be used as data storage by assigning the state in which one transistor
is conducting a logical 1 and the state when the other transistor is conducting a logical 0.
They are also useful as binary counters. A series of flip-flops is wired in sequence so that
when the first is cycled, it changes the state of the second flip-flop so that it counts every
second pulse. The second flip-flop, in turn, changes the state of the next flip-flop so that it
counts every fourth pulse, and so on. Thus n flip-flops can count 2n1 pulses.
FIGURE 22.6 Ip
Typical vacuum tube circuit. A high-voltage source (Bþ ¼ 100 V) is
RL
connected to the plate of the tube through a load resistor, RL. Elec- VL
trons boil off of the cathode when heated by the heater filament and
are attracted to the plate by the high plate voltage. The plate current, B+
Ip, can be regulated by the grid voltage. A negative bias of a few volts
on the grid, Vg, repels the electrons trying to get to the anode and Vg
shuts off the plate current. The changes in plate current resulting Heater
from a small signal imposed on the grid is seen as a much larger circuit
voltage appearing across the load resistor.
Transistors, Quantum Wells, and Superlattices 423
Source Gate
(metal) Drain
(metal) S D
(metal)
SiO2
(insulator) n p+ n
p+
G
Black
(metal)
FIGURE 22.9
Schematic of an n-channel JFET. The circuit symbol to the right indicates a p-type gate for this n-channel device.
The arrow is reversed for a p-channel JFET.
Transistors, Quantum Wells, and Superlattices 425
floating gate. The control gate is connected to the bit line. A logical ‘‘1’’ is written into the
floating gate by applying a negative high voltage to the control gate which forces electrons
to tunnel through the oxide layer. When the voltage is removed, the electrons are not able
to escape so they simply stay there, causing the FET to either turn ‘‘on’’ or ‘‘off,’’ depend-
ing on whether it is an enhancement or depletion type. The charge of the floating gate may
be electronically erased by reversing the bias on the control gate.
FIGURE 22.10
Schematic of a DRAM. Information is stored in the form of charge on the capacitors. Information is stored as
charges on the capacitors.
discussed previously. Metal contacts grown on a thin oxide (or other insulator) back of an
n-type silicon crystal are held at a negative potential, which attracts minority holes to the
region next to the metal contact, forming a depletion layer when equilibrium is reached.
Each metal contact represents a pixel of information. Light incident on the silicon will
produce additional electron–hole pairs and the holes will be swept by the applied field into
the depletion layer where they are stored as excess positive charge, the magnitude of the
charge being proportional to the integrated number of photons incident since the last time
the system was read out.
The readout of the system consists of manipulating the voltage on the metal contacts in
order to shift the accumulated charges at the silicon–insulator interface along each row to
the edge of the focal plane array where they are collected by a charge-sensitive amplifier.
The mechanism for performing this operation is called a shift register. A schematic shift
register readout is shown in Figure 22.11.
+
6 5 4 3 2 1
n-silicon
(a)
6 5 4 3 2 1
n-silicon
(b)
++
6 5 4 3 2 1
+++ + +++++
++++
n-silicon
(c)
6 5 4 3 2 1
+++ +
+++++ ++++
n-silicon
(d) ++++
6 5 4 3 2 1
+++
+ +++++
n-silicon
(e)
FIGURE 22.11
Schematic of a shift register readout of charges gathered under the pixels of an Si CCD video camera. The amount
of charge is proportional to the light received. In (a) the charge under electrode 1 has been read. Making electrode
1 more negative transfers the charge under electrode 2 to electrode 1 (b). Charge from electrode 2 is read out in
(c) and charge from electrode 3 is transferred to electrode 2. The charge from electrode 4 is transferred to electrode
3 and the charge from electrode 2 is transferred to electrode 1 in (d) where it gets read out in (e). In this way each
of the charges collected can be clocked out serially to form an analog video stream for each row of pixels.
428 Introduction to the Physics and Chemistry of Materials
22.6 Heterojunctions
With the use of epitaxial growth methods such as molecular beam epitaxy (MBE) or metal
organic vapor deposition (MOCVD), it is possible to grow single crystalline ternary alloy
systems such as AlxGa1xAs or quaternary systems such as GaxIn1xAsyP1y with con-
trolled composition as well as to form heterostructures by growing one compound semi-
conductor epitaxially on top of another compound semiconductor. (Epitaxy means that the
lattice periodicity is maintained across the growth interface.)
By varying x in AlxGa1xAs, one can adjust the bandgap from that of AlAs (2.2 eV) to
GaAs (1.35 eV). At x ¼ 0.45, the alloy changes from a direct bandgap system like GaAs to
an indirect bandgap system like AlAs. Figure 22.12 plots bandgap energy against lattice
constant. By choosing systems with a minimum lattice mismatch, the lattice strain is
reduced and defect density is lowered.
It is now possible to form heterostructures in which the transition from one system to the
other takes place over one atomic layer. When two such semiconductors are brought into
intimate contact, the electron affinities (energy required to remove an electron from the
bottom of the conduction band to vacuum state) must line up, and at thermal equilibrium,
the Fermi levels must also line up. The result is band bending with discontinuities in both
the valence (DEV) and conduction band (DEC) at the interface as shown in Figure 22.13.
(Recall that in homojunctions between n- and p-type material, the vacuum levels already
lined up because both sides of the junction were the same material so no such discontinu-
ities appeared.)
7
AlN
6
5
Bandgap energy (eV)
4
ZnS
GaN
3
AlP CdS ZnTe
GaP AlAs
2 InN AlSb
GaAs InP CdTe
1 Si
Ge GaSb
InSb HgTe
0
4.0 4.5 5.0 5.5 6.0 6.5 7.0
Lattice parameter (Å)
FIGURE 22.12
Bandgap energy versus lattice constant for various elemental and compound semiconductors. The lines connect-
ing the different compounds indicate the range of energy gaps that can be obtained by alloying the different
components with similar lattice constants.
Transistors, Quantum Wells, and Superlattices 429
c2 c2
c1
ΔEC EF2 Eg1
EF
Eg2
Eg1
EF1 Eg2
ΔEV
FIGURE 22.13
Two different semiconductor materials with different energy gaps before forming a junction (left). The differences
between the two conduction bands and between the two valence bands relative to the vacuum energy must be
preserved at the junction. This condition causes discontinuities in the bands when a junction is formed and the
Fermi levels are matched (right).
++
−−− ++
−− +
EF
FIGURE 22.14
Heterojunction between heavily n-doped AlGaAs and
lightly n-doped GaAs. The higher concentration of electrons
in the AlGaAs causes them to diffuse across the junction into
the nearly intrinsic GaAs where they congregate in the well
GaAs n+AlGaAs between the conduction band and the Fermi level.
The ionized donors in the heavily doped AlGaAs system produce a high concentration
of electrons which diffuse across the junction and wind up above the conduction band of
the GaAs, where they remain as free carriers. Thus these electrons are removed from the
material containing the donors that created them and are transported to a material with
few ionized donors or acceptors. Rutherford scattering from ionized impurities (donors) is
a mobility-limiting mechanism at low temperatures. This separation of free electrons from
the donors that produced them is call modulation doping and results in a dramatic increase
in mobility at low temperatures. It should be understood that the high mobility electrons
are confined to plane parallel to the heterojunction.
n+ AlGaAs
Gate
Drain
Source
FIGURE 22.15
Schematic of a HEMT. The highly doped AlGaAs
layer forces its electrons into the semi-insulating
_____________
GaAs where they have very high mobility. Changing
the bias on the gate alters the number of electrons in Semi-insulating GaAs
the n-channel.
EF
--------
Ec
22.7 Superlattices
With the development of molecular beam technology (MBE, MOCVD), it is possible to
not only make heterojunctions, but epitaxial layers can now be grown with thicknesses that
can be controlled to atomic dimensions. By sandwiching a lower bandgap material
Transistors, Quantum Wells, and Superlattices 431
between two layers of higher bandgap materials, electrons in the lower bandgap material
are confined in 1-D very much as in the electron-in-a-box model. Instead of being in a
conduction or valence band, these electrons are in discrete energy states whose levels can
be controlled by the width of the lower bandgap material. Such a structure is called a single
quantum well. A series of periodic quantum wells can also be constructed to form multiple
quantum wells (MQWs) which constitute a compositional superlattice. Such structures can
lead to some very novel electronic and photonic devices as illustrated by a few examples in
the following sections.
E1
pjwstk|402064|1435428838
E1
EF
EC
AlGaAs
AlGaAs
AlGaAs
GaAs
AlGaAs
n+ GaAs
FIGURE 22.17
RTA: A thin layer of intrinsic GaAs is sandwiched between two thin layers of intrinsic AlGaAs, which has a higher
bandgap than GaAs, to form a quantum well. The two AlGaAs layers are sandwiched between two layers of
heavily doped nþ GaAs. In the unbiased state (left), a bound state energy of the quantum well E1 is above the
Fermi energy and only a few of the electrons from the nþ GaAs can tunnel through. As the bias is increased, the
tunnel current increases until it reaches its peak (right), when the bound state is in resonance with the Fermi level.
A further increase in voltage produces decreasing current because there are no available states in the well.
432 Introduction to the Physics and Chemistry of Materials
hn
Ip.c
Transparent n-region p-region n-region
contact EC
e− e−
np p+
+ EV
Metal contact
FIGURE 22.18
Schematic of a nipi structure (left). Very thin alternating layers of n- and p-doped GaAs form a sandwich-like
structure. When an incident photon creates an electron–hole pair, the holes are swept into the p-layer and the
electrons are swept into the n-layer by the fields created by the junctions. Since there are few minority carriers to
recombine with, a larger fraction of these charges can be collected.
An interesting feature of this structure is that the free electrons will be found in the
minima of the conduction bands while the holes reside in the maxima of the valence band,
thus are physically separated from each other. This provides extremely long recombination
times for photo-induced electron–hole pairs. Thus a photocurrent can exist much longer in
a nipi structure than in a homogeneously doped semiconductor.
Increasing the minority carrier lifetime is very important in photonic devices. In a
heavily doped material, the majority carriers far out number the minority carriers under
equilibrium conditions because of the law of mass action that requires the np product to
be constant. However, absorption of photons produces equal numbers of electron–hole
pairs. Under this nonequilibrium situation, the majority carriers are increased by only a
small fraction; whereas, the minority carriers are increased by a very large fraction. Since
there are a very large number of majority carriers to annihilate them as the system returns
to equilibrium, it is necessary for the minority carrier lifetime to be long enough for
them to reach an electrode in order to be counted. By using a nipi structure to capture
and separate the electrons and holes, highly efficient photoconductive devices can
be fabricated.
AlInAs
FIGURE 22.19
E3 Schematic of a QC laser. A single electron stays in
the conduction band of the two semiconductors as
E2
it traverses the superlattice. It loses energy when it
E1 e− hn = E3 − E2 is in the lower bandgap material by dropping
e− from a higher quantum level to a lower level,
emitting photons as it drops. Thus a single elec-
GaInAs tron can produce as many photons as there are
stages in the lattice.
‘‘electronic water fall’’ as it was described by its inventor, Frederico Capasso, which is the
basis for the term QC describing the process (Faist et al., 1996).
QC lasers are unique in many ways. Rather than requiring a specific bandgap material to
laze at a particular wavelength, the wavelength may be selected by the dimensions of the
InGaAs layers that determine the energy levels in that region. Since they can be tuned to
look for specific molecular bands in the 3.4–17 mm, this makes such devices very useful for
molecular spectroscopy in the mid- and far IR region. The fact that a single electron can
produce up to 75 photons makes them very efficient and powerful (as solid-state lasers go),
up to 0.5 W at room temperature in a pulse mode, and 0.2 W at LN2 temperature in a CW
mode. As a result, these devices have found use in trace gas detection, air pollution
monitoring, industrial process control, and breath analysis for early detection of disease.
22.9 Summary
The first transistors, which were BJTs, were essentially back-to-back diodes with a
control gate between them. In the active mode, the transistor could be turned on by
forward biasing the emitter–gate diode. Most of the minority holes that were injected
into the gate region are swept by the internal fields into the collector. The a of the transistor
is the fraction of the injected minority carriers that leave the emitter and get to the
collector. The fraction of the minority carries that go to the gate is 1 a, so the current
gain, the ratio of the collector current to the gate current, is called the b and is given by
b ¼ a=(1 a). BJTs are low impedance devices, meaning they draw current, and are
primarily used as current amplifiers, using the small gate current to control a larger
collector current.
FETs control the flow of current from their source to drain by a voltage applied to
their gate. The current that flows in a FET can either be electrons (n-channel) or holes
(p-channel) and they can either be normally off when unbiased (enhancement type)
or normally on (depletion type). Metal–insulator–semiconductor FETs (MISFETs or
MOSFETs) have insulated gates and draw practically no current. They are used in appli-
cations such as flip-flop counters in digital watches and in SRAMs where current flow must
be kept to a minimum. DRAMs store digital information as a charge or no-charge on a
capacitor that is controlled by a single transistor. Since a DRAM only requires one
transistor per bit of information compared to six in a SRAM, DRAM memory is smaller
and cheaper than SRAM, although DRAM memory has to be continually refreshed as the
charge leaks off of the capacitor.
Both the SRAM and DRAM memories are volatile, meaning information is lost if the
power is removed. EEPROMs utilize a floating insulated gate on a FET. The floating gate
can be charged or discharged by tunneling from a control electrode and will remain in that
state until changed by a voltage pulse on the control electrode. EEPROMs are used in
today’s memory sticks and digital cameras.
Photoelectrons or holes can be collected under charged metal electrodes that form the
pixels in a CCD video camera. By manipulating the voltages on the electrodes, these
charges may be transferred serially in a shift register to a charge-sensitive amplifier to
produce an analog video output.
In 1965 Gordon Moore observed that the number of transistors per area of an integrated
circuit had doubled every year since its invention. This trend has continued and has
become known as Moore’s law.
With the development of sophisticated growth techniques that can lay down epitaxial
layers of different materials with atomic precision and form heterojunctions of different
bandgap materials, new dimensions in device technology have opened up. Thus one can
engineer discrete energy levels of the electrons confined in a low bandgap material
between two higher bandgap materials. The 1-D quantum wells can be used to make
highly efficient double heterojunction lasers and RTDs that can operate in the THz region.
By repeating these layered structures to make superlattices, cascade quantum well lasers
can be made in which a single electron can generate many photons. Carrying the
quantum confinement to 2-D and 3-D, quantum wire and quantum dots have been
produced in which their electrical and optical properties can be designed into their
structure.
Transistors, Quantum Wells, and Superlattices 435
Bibliography
Bhattacharya, P., Properties of III-V Quantum Wells and Superlattices, Institution of Electrical Engineers,
now the Institution of Engineering and Technology, U.K., 1996.
Faist, Capasso, et al., Appl. Phys. Lett., 68, 1996, 3680–3682.
Faist, J., Capasso, F., Sivco, D.L., et al., Quantum cascade laser, Science, 264, 1994, 553.
Hummel, R.E., Electronic Properties of Materials, 3rd edn., Springer, New York, 2000.
Ibach, H. and Lüth, H., Solid State Physics, 3rd edn., Springer-Verlag, New York, 1990.
Kastner, M.A., The single electron transistor and artificial atoms, Ann. Phys. (Leipzig), 9(11–12), 2000,
885–894.
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley & Sons, New York, 1966.
Kouklin, et al., Appl. Phys. Lett., 76, 2000, 406.
Livingston, J.D., Electronic Properties of Engineering Materials, Wiley, MIT series in Materials Science
and Engineering, John Wiley & Sons, New York, 1999.
Navon, D.H., Electronic Materials and Devices, Houghton Mifflin, Boston, 1975.
Percival, et al., GaAs quantum wire lasers grown on v-grooved substrates isolated by self-aligned ion
implantation, IEEE Trans. Electron Devices, 47, 2000, 1769–1772.
Sapoval, B. and Hermann, C., Physics of Semiconductors, Springer-Verlag, New York, 1995.
23
Dielectrics and the Dielectric Function
Dielectric materials in the form of ceramics or polymers are characterized by the fact that
they contain virtually no free charges and therefore are electrical insulators. The lack of free
charge implies a bandgap energy that is at least large enough to prevent thermal ionization
of electrons and is generally high enough for them to be transparent in the visible
spectrum. However, just because their conductivity is small, the electrical and optical
properties of dielectrics are by no means uninteresting. The presence of an electric field
distorts the atoms and molecules so that the center of their positive charge is displaced
from the center of their negative charge and electric dipoles are formed. The electrical and
optical properties are determined primarily by the dielectric material’s ability to form
electric dipoles in the presence of an electric field. But, since one of the dielectric’s primary
functions is that of an electrical insulator, we shall consider this attribute first.
437
438 Introduction to the Physics and Chemistry of Materials
given by the Einstein relationship Di ¼ kT mi=e (see Chapter 21). Therefore, the ionic current
is given by
ni z2i e2 Di
Ji ¼ E: (23:1)
kT
For Ni ¼ 1028=m3 and Di 1012 m2=s, the conductivity s ¼ nizi2e2Di=kT 105 S=m.
23.2.1 Capacitors
To illustrate how dielectrics interact electrically, we first consider the capacitor, a device for
storing electric charge. Basically, a capacitor is simply a sandwich consisting of a dielectric
surrounded by two conductive surfaces. When connected to a source, electrons will
accumulate on one plate which will induce an equal but opposite charge on the opposite
plate as shown in Figure 23.1. Once the potential difference between the plates become
equal to the source, current ceases to flow in the circuit and the capacitor is fully charged.
The capacitance C is defined as the charge stored divided by the voltage applied, or
C ¼ Q=V. The unit of capacitance is the farad (from Michael Faraday) which, by definition,
is 1 Coulomb=V. The stored charge is equal to the product of the area A, the electric field,
and the dielectric constant of the material between the plates, or Q ¼ «EA. From this
relationship, we see that the permittivity « must have the units of Coulombs=Volt-m or
Coulombs2=Nt-m2. Since the electric displacement D ¼ «E, we see that D is the charge
displaced per area, or the surface charge on the capacitor. The electric field is the potential
V divided by the distance d between the plates or E ¼ V=d. Therefore, C ¼ «A=d.
Now if there were nothing in the space between the two capacitor plates, the permittivity
« would just be the permittivity of free space «0. But if we place a dielectric material
+ + + + + + + +++++++++++++
+ − − − − − − −
+ + + + + + +
− − − − − − −−−−−−−−−−−−−
−
(a) (b)
FIGURE 23.1
A fully charged capacitor with no dielectric between the plates (a). When a dielectric is inserted between the plates
(b), the induced dipoles in the dielectric line up with their positive charge pointing toward the negative plate. The
presence of these positive charges attracts more electrons from the battery increasing the negative charge on the
plate. Similarly, more positive charge collects on the opposite plate, which increases the capacitance.
440 Introduction to the Physics and Chemistry of Materials
between the plates, the electric field, which runs from plus to minus, aligns the electric
dipoles in the dielectric which produces a negative surface charge on the side of the
dielectric facing the positive capacitor plate and a positive surface charge on the side facing
the negative capacitor plate (think in terms of the charges on the nose and tail of the
dipoles). These surface charges induce more positive charges to collect on the positive plate
and more negative charges to collect on the negative plate, which increases the charge
stored by DQ. The total charge can be written as Q ¼ Q0 þ DQ ¼ DA ¼ «0E þ P, where P is
the polarization of the dielectric. Thus we see that the additional charge on the capacitor
plates was brought about by the polarization of the dielectric and the capacitance is
C ¼ «0«rA=d is directly related to the relative dielectric constant, «r.
In an alternating current circuit, the current usually reverses before the capacitor is fully
charged, thus the capacitor is continually being charged and discharged with reversing
polarity. Thereby an alternating current can flow through a circuit containing a series
capacitor whereas direct current cannot. As a result, capacitors are frequently used in
electronic circuits to block unwanted DC in an AC circuit or to shunt unwanted AC ripple
to ground in a DC circuit.
If a capacitor is discharged into a resistive load, the time to discharge 1=e of the charge is
the product of the resistance and capacitance RC. This RC time constant can be set by
choice of resistance and capacitance and is used in many timing circuits.
Another use of capacitors is to store energy that can be recovered quickly, much faster
than from a battery. Dumping a large capacitor bank into a low-resistance load can result
in a very large current flow and this technique is used for exploding wire experiments and
other types of plasma research. The energy stored in a capacitor is (1=2)CV2 ¼ (1=2)«AV2=d.
One can see the need of materials with high dielectric constants and high dielectric
strengths for high energy capacitor banks.
Recently a new type of capacitor called supercapacitor has been developed by Maxwell
Technologies (San Diego, California). These supercapacitors store their charge in an electro-
lytic double layer and have capacitance measured in farads rather than microfarads, which
is the typical capacitance of conventional capacitors. Unfortunately, they can only operate at
low voltages (2.5 V); therefore the energy storage is limited to 1–10 W h=kg, about 1=10
that of a nickel-metal hydride battery. Supercapacitors find uses in memory backups for
short power interruptions, and because they can be charged and discharged much faster
than batteries, they can be used for energy recovery system in electric and hybrid vehicles.
p
_ +
FIGURE 23.2
Electronic polarization. The presence of an electric field E dis-
torts the electron’s orbit and displaces its center from the posi-
tively charged nucleus creating a dipole moment p.
The magnitude of the displacement caused by an oscillating electric field can be obtained
by treating the system as a damped harmonic oscillator in which the restoring force is
characterized by v0, the resonant frequency of the system.
eE(v) ivt
€
x þ gx_ þ v20 x ¼ e : (23:2)
m
x(t) ¼ ~
x(v)eivt , (23:3)
where ~
x(v) is the complex amplitude given by
e=m
~
x(v) ¼ E(v), (23:4)
v20 v2 igv
p e~x(v) e2 =m
a(v) ¼ ¼ 2 : (23:5)
E E(v) v0 v2 igv
The resonant frequency v0 is material dependent and is typically 1015–1016 rad=s (ultra-
violet). The damping constant is generally small compared to v0 and is only important near
resonance (v v0).
For v < v0 (visible range),
This is the contribution from a single electron. For a many electron atom, we need the
effective ai for each ion species so that we may sum over the various ionic species to obtain
the dielectric function
P(v) 1 X
«r (v) ¼ 1 þ ¼1þ ni ai (v): (23:6)
«0 E «0 i
442 Introduction to the Physics and Chemistry of Materials
It is customary to denote the static electronic dielectric function (v v0) as «(1), which
seems strange. The only justification seems to be that «(0) is reserved for the total static
dielectric constant (ionic þ electronic) and the infinity simply means at frequencies far
above those where the ionic contributions are no longer significant, but are still much
smaller than v0.
− − +
+ − −
FIGURE 23.3
Ionic polarization. With no applied field (a), +
+ − + + −
the dipoles between the anions and the cat-
ions all cancel each other. In the presence of
an electric field (b), the cations are displaced
in the direction of the field and the anions +
− + − − −
are displaced in the opposite direction, giv-
ing a net dipole moment in the direction of
the field. (a) (b)
Dielectrics and the Dielectric Function 443
In the vicinity of k ¼ 0, (eika þ eika) ¼ 2. We now have a set of simultaneous equations that
can be solved for the complex amplitudes j and h,
However, the quantity of interest is (j h) since this represents the oscillating change in the
position of the two charges, hence, the induced oscillating dipole moment is e(j h).
Multiply the top equation by M, the bottom by m, subtract and identify the reduced
mass m ¼ mM=(m þ M). After rearranging terms, we get the particular solution,
2b eE0
v2 ðj hÞ ¼ : (23:13)
m m
The complimentary solution is obtained by setting the driving term eE0 to 0 in Equation
23.12. The terms may be rearranged to give the same indicial equation as Equation 16.16
(Chapter 16) for which the solution at k ¼ 0 was found to be
2bðM þ mÞ 2b
v2T ¼ ¼ : (23:14)
Mm m
We denote this natural frequency as vT, symbolizing this as the frequency of the transverse
phonon spectrum in the optical branch. Putting this back into the particular solution
(Equation 23.13),
eE0
v2T v2 ðj hÞ ¼ : (23:15)
m
Now the induced dipole is just the relative displacement between the ions (j h) times the
unit of charge and the ionic contribution to the polarization obtained from Equation 23.15
(with damping terms omitted) is just
Nions e2 E0
Pionic ¼ Nions ðj hÞe ¼ 2 : (23:16)
m vT v2
It is customary to add the electronic contribution to the ionic contribution and write
Nions e2
«r (v) ¼ «(1) þ : (23:17)
«0 m v2T v2
For v vT, the static dielectric constant «(0) is real and positive and is given by
Nions e2
«(0) ¼ þ «(1): (23:18)
«0 mv2T
444 Introduction to the Physics and Chemistry of Materials
ð«(0) «(1)Þv2T
«(v) ¼ «(1) þ : (23:19)
v2T v2
We see in Figure 23.4 that «(v) blows up at vT, then goes negative and eventually comes
back up to «(1). We define vL as that value of v when it crosses back into positive territory
or the value for v that makes «(vL) ¼ 0 so that
ð«(0) «(1)Þv2T
¼ «(1) (23:20)
v2T v2L
or
ð«(0) «(1)Þ «(0)
v2L ¼ v2T 1 ¼ v2T : (23:21)
«(1) «(1)
This is the LST relationship, which relates the ratio of the squares of the longitudinal to the
transverse resonance frequencies to the ratio of the static to electronic dielectric function.
The behavior of the total dielectric function (ionic þ electronic) in the vicinity of vT is
shown in Figure 23.4. The vT was taken to be 1013=s and the static ionic contribution,
«(0) «(1) was set to 5. The g was set to 0.05 as before.
60
40
20
0
er
−20
−40
FIGURE 23.4
Ionic dielectric function in the vicinity of vT.
−60
The vL is defined as the value for v when 12.4 13 wL 13.6
the dielectric function becomes positive
again. log w (rad/s)
Dielectrics and the Dielectric Function 445
E
+ F
d Θ FIGURE 23.5
E Forces on an electric dipole in an electric field E. The dipole
moment p is the product of the charge and the separation or qd.
The E-field produces a torque given by p E ¼ pE sin u that
F tends to align the dipole with the field. The work required to
-
rotate the dipole through angle u is the integral of the torque
E from 0 to u, or pE cos u (assuming U ¼ 0 at u ¼ p=2).
The presence of an electric field tends to align the individual dipole moments in the
material, but thermal effects work to destroy this alignment as will be shown below. Using
Maxwell–Boltzmann statistics, we can calculate the average dipole moment by setting the
energy required to move from the equilibrium to a higher energy equal to pEcos u.
At thermal equilibrium, the average dipole moment will be given by
Ðp
(p cos u)epE cos u=kT 2p sin udu
0
hpi ¼ Ðp ¼ pL(x), (23:22)
epE cos u=kT 2p sin udu
0
where x ¼ pE=kT and the Langevin function L(x) ¼ coth x 1=x. At ambient temperatures,
kT pE. Since coth x ¼ 1=x þ x=3 x3=45 þ , for small x, L(x) ! x=3. For very low
temperatures, x 1, and L(x) ! 1. At ambient temperatures (x 1)
Np2 E C
P¼ ¼ E: (23:24)
3kT T
This is similar to the Curie law of paramagnetism, which relates the magnetic susceptibility
to the Curie constant for a given material times the applied magnetic field divided by the
absolute temperature (see Chapter 25). This type of behavior in dielectrics is known as
paraelectric behavior. In both paramagnetic and paraelectric materials, the applied field
tends to order the alignment of the dipole moments of individual atoms or molecules while
thermal energy tends to randomize them. Both cases follow the Curie law in which the
magnitude of the induced dipole moment is directly proportional to the applied field and
inversely proportional to the temperature. The constant of proportionality, or Curie con-
stant, is a material property.
where n is a jump frequency. Similarly, the rate at which dipoles flip from right to left is
dNL
¼ nNR eðfþpEÞ=kT : (23:26)
dt
NL
¼ e2pE=kT : (23:27)
NR
In this case, the average polarizability is three times greater than that for a gas or a liquid
because the dipoles can only be either parallel or antiparallel. It is interesting to note
that the barrier height cancels out and does not appear in the final result. This result
assumes that equilibrium has been reached, but does not say how long it takes to come to
equilibrium.
With no field present, the rate of change of the polarization is
P_ ¼ p N_ R N_ L ¼ pnef=kT ðNL NR Þ ¼ P(t)nef=kT : (23:29)
The solution is
P(t) «0 x0
P_ ðtÞ þ ¼ E0 eivt : (23:31)
t t
x þ x~=t ¼ x0 =t
iv~ (23:32)
from which
x0 x0
x~ ¼ ¼ ð1 ivt Þ: (23:33)
1 þ ivt 1 þ v2 t 2
x0 x 0 vt
x 0 (v) ¼ Reðx(v)Þ ¼ ; x 00 (v) ¼ Imðx(v)Þ ¼ , (23:34)
1 þ v2 t 2 1 þ v2 t 2
Dielectrics and the Dielectric Function 447
Np2
x ð 0Þ ¼ : (23:35)
«0 kT
Notice that the susceptibility contributions from the permanent dipoles vanish for
v 1=t.
A similar result was obtained by Debye for liquids and gases in which there is no
potential barrier to inhibit the dipole motion. In this case the relaxation time results from
the viscosity of the medium and is given by t ¼ 4phR3=kT, where R is the radius of the
molecule and the susceptibility is obtained from Equation 23.24 as
Np2
x ð 0Þ ¼ : (23:36)
3«0 kT
Let E(t) ¼ E0cos(vt) and PðtÞ ¼ «0 x~E0 cosðvtÞ. The complex susceptibility can be written as
x~ ¼ xð0Þeif , where x0 ¼ xð0Þ cosðfÞ and x 00 ¼ x ð0Þ sinðfÞ. The ratio x00 =x0 is called the loss
tangent. P(t) becomes «0 x ð0ÞE0 cosðvt fÞ as the polarization lags behind the field by the
phase angle f. The product PY(t)E(t) becomes
P_ ðtÞEðtÞ ¼ v«0 xð0ÞE20 sinðvtÞ cosðvtÞ cosðfÞ þ cos2 ðvtÞ sinðfÞ : (23:38)
1 1
Q ¼ «0 xð0ÞE20 sinðfÞ ¼ «0 E20 x00 ðvÞ: (23:39)
2 2
1 vt
Q ¼ «0 E20 xð0Þ : (23:40)
2 1 þ v2 t 2
++++++++++++++
+
+
++ + + ++ f+ +
+P + + +
E − Eloc − − −
−− −− −− −−
−- −
FIGURE 23.6
Geometry for obtaining the local field using the Clausius–
Mozzotti model. −−−−−−−−−−−−−−−
Consider the local field inside a dielectric medium in an applied field E0 illustrated in
Figure 23.6.
If we cut out a small spherical cavity around the point in question, the field in the center
of this sphere will be given by the sum of the applied field E, the field due to the charges on
the inside surface of the sphere, the field due to the dipoles outside the sphere and the field
due to the dipoles inside the sphere. For materials with cubic symmetry, the latter contri-
bution vanishes by symmetry, as does the field from the dipoles outside of the cavity. The
charge density on the surface of the cavity is given by P cos f. The field due to the surface
charges inside the cavity is then given by
ðp
P cos f P
Ein ¼ 2p R2 sin fdf cos f ¼ : (23:41)
4p«0 R2 3«0
0
P NaEloc
Eloc ¼ E þ ¼Eþ : (23:42)
3«0 3«0
E
Eloc ¼ : (23:43)
1 Na=3«0
NaE
P¼ : (23:44)
ð1 Na=3«0 Þ
P Na
x¼ ¼ : (23:45)
«0 E «0 ð1 Na=3«0 Þ
Na 1 þ 2Na=3«0
«r ¼ 1 þ x ¼ 1 þ ¼ , (23:46)
«0 ð1 Na=3«0 Þ 1 Na=3«0
Dielectrics and the Dielectric Function 449
«R 1 Na
¼ : (23:47)
«R þ 2 3«0
P(v) ELoc X
«r (v) ¼ 1 þ ¼1þ ni ai (v): (23:48)
«0 E «0 E i
Values for a are measured for various ionic species and are typically O(1040) at optical
frequencies for the various ionic species (see Chapter 24), hence
na 1029 1040
¼ ¼ O(1):
«0 9:9 1012
It is possible by treating the atom as a sphere of uniform charge and using Gauss’ law to
find the displacement between the nucleus and the center of the charge resulting from an
applied electric field. Using this model the static electronic polarizability a(0) can be
estimated to be 4p«0R3, where R is the atomic radius.
So we see that the effect of electronic polarization in the visible region is to increase the
dielectric constant by a quantity that is on the order of 1. For the more general case, the
solution for a(v) given by Equation 23.5 can be separated into real and imaginary parts by
multiplying top and bottom by the complex conjugate,
e2 =m 2
a(v) ¼ 2
v0 v2 þ igv : (23:50)
v20 v2 þg2 v2
As v ! v0, the Re(a) gets larger until v ¼ v0, at which point it reaches its peak value,
limited only by the damping term. As v > v0, the sign of the Re(a) becomes negative and
then eventually rises and approaches 1, the free space value. The imaginary part is only
significant in the region where the real part goes through zero.
To illustrate the effect of the local field correction, the dielectric function is plotted using
Equation 23.6 (uncorrected) and Equation 23.49 (corrected) in Figure 23.7. The quantity Na
(0)=«0 was taken to be 1.01 and v0 was set to 1016=s. The g in Equation 23.50 was set to 0.05
s to provide some damping to keep the dielectric function from blowing up.
There is a small difference in the «r for v v0, 2.523 for the corrected versus 2.01 for the
uncorrected. This difference will be greater for larger values of Na(0)=«0. Also note the
450 Introduction to the Physics and Chemistry of Materials
15
10
er
−5
corrected function blows up before the resonant frequency. This can be understood from
the following analysis. The real part of the electronic function from Equation 23.50,
ignoring the damping term, can be written as
a(0)v20
að v Þ ¼ : (23:51)
v20 v2
From Equation 23.49. the dielectric function blows up when Na(v)=3«0 ¼ 1. (This is some-
times referred to as the polarization catastrophe.) The denominator in Equation 23.49 can
be written using Equation 23.51 as
N a(0)
1 ¼0 (23:52)
3«0 1 v21 =v20
and solved for v1, the actual frequency where the dielectric function becomes singular,
to obtain
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Na(0)
v1 ¼ v0 1 : (23:53)
3«0
Thus, we see that the effect of the local field is to reduce the natural frequency of the
oscillations of the nucleus and the electron shell.
There appears to be some controversy in the literature over how to (or if to) correct the
ionic dielectric function for the local field. This issue will be discussed in the Appendix.
30
20
10
0
er
−10
Vis
−20
hf uhf Microwave Infrared UV X-ray
−30
109 1011 1013 1015 1017
w (rad/s)
FIGURE 23.8
Total dielectric function for a hypothetical material with a static dielectric constant of 20, t ¼ 109 s, «(0) ¼ 5 and «(1) ¼
2.523. The damping constant g was set to 0.1 s. Note the resonances in the far infrared and in the near ultraviolet.
xperm (0)ð«(0) «ð1ÞÞ 1 v2 =v2T 1 þ (2=3«0 )NaðvÞ
«r (v) ¼ þ 2 þ , (23:54)
1 þ v2 t 2 1 v =vT þg =v
2 2 2 2 1 (1=3«0 )NaðvÞ
where the a(v) is given by Equation 23.50. The total dielectric function is displayed in
Figure 23.8.
23.4 Ferroelectrics
Some of dielectric materials have a spontaneous dipole moment that does not require the
presence of an applied electric field, just as ferromagnetic materials have a spontaneous
magnetic field in the absence of an applied magnetic field. Such materials are called
ferroelectrics because they exhibit several similarities to ferromagnetic materials in that
they form domains of like polarization and they exhibit hysteresis in response to
alternating electric fields as illustrated in Figure 23.9a and b. Also as in ferromagnets,
pjwstk|402064|1435428843
D D
e0E e 0E Psat
Psat
Pret
Ecoer
E E
(a) (b)
FIGURE 23.9
(a) Hysteresis curve for a dipolar material for vt 1. The depolarization curve simply retraces the polarization
curve and the polarization vanishes at E ¼ 0. The «0E is the component of D after all the dipoles become aligned
(Psat). If vt 1, a hysteresis will be observed. (b) Hysteresis curve for a ferroelectric material. Pret is the
o
polarization retained when the field is removed. Ecoer is the coercive field required to destroy the polarization.
The area enclosed is lost to heat during each cycle.
because their dipole moment will change with temperature (due to thermal expansion)
which induces an additional charge on their surfaces. Some pyroelectrics have a permanent
dipole moment that can be reversed by applying a strong electric field allowing them to
exhibit a hysteresis effect. This class of materials is the ferroelectrics. Pyroelectrics that are
not ferroelectrics may have a coercive field greater than their breakdown field or their
Curie temperature may be above their melting temperature. Of course, all ferroelectrics are
also pyroelectrics, but not the other way around.
In addition to the 10 point groups mentioned above there are 10 more point groups that
lack a center of inversion but are not polar. These are c3h, d2,3,4,6,2d,3h, s4, t and td. or 6, 222,
32, 422, 622,
42m,
62m,
4, 23, and 43m. Crystals in this class, when distorted, may produce a
dipole and are called piezoelectrics. Therefore, all ferroelectrics and pyroelectrics are also
piezoelectrics, but not all piezoelectrics are pyroelectrics or ferroelectrics.
The first ferroelectric material to be discovered was some sodium–potassium tartrate tetra-
hydrate crystals prepared in 1672 by a pharmacist, Elie Seignette, who lived in La Rochelle,
France—hence the name Rochelle salts. However, their piezoelectric properties were not
discovered until 1880 and their ferroelectric properties not until the 1920s. This material is
only a ferroelectric between the temperatures of 188C and 238C where it transforms from
orthorhombic to monoclinic and back to orthorhombic. The Curie temperature as well as its
polarization is increased when the hydrogen is replaced with deuterium.
Later ferroelectricity was discovered in KH2PO4 (KDP) below its Curie temperature of
123 K where it transforms from its piezoelectric tetragonal d2d (42m) form to the ferroelec-
tric c2v (2 mm) symmetry. RbH2PO4 and KH2AsO4 (KDA) also exhibit ferroelectricity at
low temperatures. Another member of this class is triglycene sulfate (TGS). It transforms
from its monoclinic form c2 (2) to its centrosymetric form c2h (2=m) at its Curie temperature
of 49.78C. The Curie temperatures of these systems are also increased in their dueterated
Dielectrics and the Dielectric Function 453
P
(a) (b)
FIGURE 23.10
BaTiO3 molecule above the 1208C transition temperature (a). The thermal expansion of the lattice provides
sufficient space for the Ti4þ ion the remain in the center of the cubic unit cell with the perovskite structure
(Pm3m). Below 1208C (b) the thermal expansion can no longer maintain the cubic symmetry and a solid state
transition to a tetragonal lattice (P4 mm) occurs. The four O2 ions squeeze the Ti4þ to one side or the other
creating a permanent electric dipole.
forms. These water soluble molecules have hydrogen-bonded PO3 at the corners of the unit
cell. The fact that replacement of the H by D strongly suggests that these bonds play an
important role in the ferroelectric behavior of these compounds, perhaps in the position of
the H atom in the H-bond. The fact that the ferroelectric transition appears to be second-
order also suggests an order–disorder transition in the position of these H (or D) atoms.
Another class of ferroelectrics is the perovskites such as the titanates (BaTiO3, PbTi03,
SrTiO3, CaTiO3), the niobates (KNbO3, NaNbO3), the ilmentites, (LiNbO3, LiTaO3)),
the ternary lead zirconate titanate (PZT), and the quaternary lead lanthanum zirconate
titanate (PLZT). These materials undergo a displacive-type phase transformation at their
Curie temperature.
For example, the oxygen perovskites such as BaTiO3 transforms from cubic perovskite to
tetragonal at 1208C. In the perovskite unit cell, the Ti4þ sits in the middle of the cube, the
Ba2þ on the four corners, and the O2 on the six sides as illustrated in Figure 23.10. The
point group is Oh or mm and because of the cubic symmetry, there is no electric dipole.
Below 1208C, however, the structure collapses into tetragonal as the bond lengths shorten
at the lower temperatures. There is not enough room for the Ti4þ as it is being squeezed by
the O2 on the sides, so it is pushed in one direction while the four O2 ions are pushed in
the other direction. The resulting crystal must then be described by C4v in the Schoenflies
notation or by 4 mm in the international notation indicating a tetragonal lattice with
fourfold symmetry about the c-axis and two sets of vertical mirror planes. It should be
mentioned that BaTiO3 also undergoes two other ferroelectric transitions, one at 58C
where it becomes orthorhombic, and another at 908C where it becomes tetragonal.
The order–disorder transitions typical of the water-soluble hydrogen bonded ferroelec-
trics are second-order transformations as illustrated in Figure 23.11, while the displacive
transformations are first-order with some second-order tendencies (a softening of the
spontaneous polarization and a rise in heat capacity).
1
Displacive
P(T)/P(0)
Order/disorder
FIGURE 23.11
Typical behavior of the retained polarization
as a function of temperature. Order–disorder
transitions vanish continuously at the Curie
temperature and are second-order transitions.
Displacive transitions vanish discontinuously 0
at the Curie temperature and are first-order 0 1
T/TC
transitions.
TC ¼ 393 K, Psat ¼ 0.26 C=m2 (from Ashcroft and Merman), a ¼ 0.401 nm, «r(TC) ¼ 35000
(from Fischer-Cripps).
The number of dipoles N ¼ a3 ¼ 1.551 1028=m3. The dipole moment p ¼ Psat=N ¼ 1.67
1029 C m (accepted value 7.68 1030 from Fischer-Cripps). Since the charge on the Ti
ion is þ4, the displacement d ¼ p=4e ¼ 0.026 nm (accepted value 0.012 nm from Fischer-
Cripps). These numbers are not entirely consistent and seem to suggest a smaller value for
N, but the value 1.551 1028=m3 is consistent with the value 1.555 1028=m3 computed
from a molecular weight of 233.192 and a density of 6.02 g=cm3.
From the Arrhenius slope of the paraelectric curve taken from Kittel, the Curie constant
C is found to be 1.24 105 K and the dielectric constant can be expressed as «r ¼ C=(T T0),
where T0 is the Curie point, not the Curie temperature. The Curie point ¼ 389 K is chosen so
that the «r matches the measured value at TC shown in Figure 23.12. This is done to avoid
the singularity at T ¼ TC.
40,000
Susceptibility
20,000
0
0 1 2 3
T/TC
FIGURE 23.12
Typical dielectric susceptibility of a displacive ferroelectric transition versus the ratio of temperature to the Curie
temperature.
Dielectrics and the Dielectric Function 455
where TC ¼ gNp2=«0k. Thus we have the susceptibility in the form of the Curie–Weiss law
where the Curie constant is C ¼ Np2=«0k, which is the same as the previous result. Letting
the applied field go to zero, we go back to Equation 23.28 and write
which can be solved for P. Since tanh(x) can be no greater than 1, as T goes to TC, g
Pp=«0kTC ¼ P=Np if there is to be a solution. From this
«0 kTC
g¼ (23:57)
Np2
and
P P P TC
¼ ¼ tanh : (23:58)
Np PSat PSat T
P=PC
T=TC ¼ : (23:59)
a tanhðP=PC Þ
23.4.4 Antiferroelectrics
Continuing the analogy between ferroelectricity and ferromagnetism, there also exists
materials such as ammonium dihydrogen phosphate NH4PO4 (ADP), PdZrP03, and
NaNbO3 that behave as antiferroelectrics. Like the ferroelectrics, they form spontaneous
dipoles below their transition temperature, but the dipoles are antiparallel, thus they cancel
out their net dipole moment. Their structure resists being polarized by an applied electric
456 Introduction to the Physics and Chemistry of Materials
P(T )/Psat
0
FIGURE 23.13 0 1
Polarization versus temperature predicted using the Curie–
T/TC
Weiss model for spontaneous polarization.
field until a transition temperature is reached at which point their dipoles become mobile
and the material becomes paraelectric.
23.4.5 Piezoelectrics
Noncentrosymmetric materials, when distorted, will form dipole moments as the center of
one charge is moved away from the center of the opposite charge. These dipoles will
produce a surface charge and the voltage between the two surfaces will be given by this
charge divided by the capacitance. Conversely, applying a voltage across the material will
cause the material to expand or contract, a property called electrostriction. Such materials
are called piezoelectrics (‘‘piezo’’ meaning push) and convert mechanical energy directly to
electrical energy and vice versa.
As discussed previously, many materials are piezoelectric without being ferroelectric,
quartz being a prime example (see Figure 23.14). But since ferroelectrics are also piezo-
electric and since the piezoelectric property is stronger in some of these ferroelectrics such
as barium titanate, PZT, and PLZT, these materials are primarily used as piezoelectrics.
Composites of PZT imbedded in polymers are used to make flexible transducers that can
be curved to suit the application.
−
−
4+ 4−
4+ p
− − − − − −
FIGURE 23.14
The SiO4
4 building block of quartz. When unstressed, the center of charge resides in center of the Si
4þ
ion. When
compressed, the bonds to the bottom three oxygen ions bend, shifting the center of negative charge above the
center of the Si4þ ion, thus creating a dipole p.
Dielectrics and the Dielectric Function 457
23.5 Applications
23.5.1 Applications of Ferroelectrics
Because of the large dielectric constant of barium titanate at ambient temperature, it finds
application in ceramic disc capacitors. The large dielectric constant allows higher capaci-
tance with less volume.
Because ferroelectrics can retain their polarization indefinitely until their polarity is
switched by the application of an electric field, they can serve as nonvolatile memory elements
or ferroelectric RAMs (FRAMs). The storage capacitors in dynamic RAMs (DRAMs) can be
replaced by ferroelectrics and eliminate the need to continually refresh the memory.
Thin ferroelectric films are finding applications as optical waveguides to carry light
along a substrate. PLZT is particularly interesting because it is transparent and has a
high electro-optical coefficient which makes it a candidate for applications for optical
switching, optical memories, and display devices.
Quartz, since it is a piezoelectric and not a ferroelectric, has no hysteresis loss when it
oscillates, thus quartz crystal oscillators are widely used as frequency control devices in
radios, computers, and watches. Since the frequency is a function of the mass of the crystal,
they can serve as deposition monitors (quartz crystal microbalances) with sensitivities of
less than 1 ng. By functionalizing the surface to absorb specific gases, they can also act as
chemical sensors. The temperature sensitivity of a quartz crystal oscillator can be minim-
ized by choosing the cut of the crystal relative to the optical axis, which is necessary for its
use as a frequency standard. On the other hand, a cut can be chosen to maximize the
frequency dependence on temperature and quartz crystal thermometers with millikelvin
resolution are available.
23.5.4 Electrets
Electrets are materials that can retain a high surface charge for a long period of time.
Electrets can be made from ferroelectric materials that have been poled (heated above their
Curie temperatures in the presence of a strong electric field to align all of the dipoles and
then quenched). Or, charge may be sprayed on a dielectric by a corona discharge or by
electron bombardment. Electrets are made in this fashion from polymers such as Mylar,
Teflon, or carnauba wax. Obviously, the material must be a good insulator (s for Teflon is
1022=V cm compared to 1014=V cm for quartz). With their frozen-in electric field,
electrets find applications in speakers and microphones, in hearing aids, in particle
removal from air, and in electrostatic printing processes such as xerography.
23.6 Summary
As nonconductors of electricity, dielectrics interact with electric fields through their ability
to form dipoles, which is called polarization. There are three sources of polarization:
electric, ionic, and dipolar.
Electronic polarization results from the orbits of the electrons being displaced from the
nuclear charge by the electric field. The dielectric constant from the electronic polarization
is denoted as «(1). The resonance frequency occurs in the ultraviolet.
Ionic polarization occurs in ionically and covalently bonded materials on which there is
some charge separation. The ionic dielectric constant plus the electronic dielectric constant
is denoted by «(0). The resonance frequency for ionic polarization is in the infrared.
Polar liquids and gases possess permanent dipoles which can be aligned by an applied
electric field. The aligning energy provided by the field is in competition with thermal
motion which tends to randomize the orientation of the dipoles. Such materials are called
paraelectrics. When the field is removed, the dipoles relax with a time constant t, which is
on the order of 1011 s for liquids such as water where the dipoles are free to rotate. These
dipoles can couple with microwave radiation to cause heating of the material, which is the
principle on which microwave ovens operate.
In dipolar solids, frozen polar liquids or solids with built-in dipole moments, the dipoles
can also be aligned by an external field and are not free to rotate. In this case, the relaxation
time is proportional to exp(f=kT), where f is the potential barrier that must be overcome
for the dipole to rotate. If the potential barrier is on the order of the chemical bond, the
relaxation time becomes virtually infinite. Such solids are said to have spontaneous or
permanent dipole moments and are called pyroelectrics. If the dipole moment can be
reversed by an electric field, the material is also a ferroelectric.
Dielectrics and the Dielectric Function 459
where ˆ is the natural lattice frequency. For v ˆ (but v0) the ionic term drops out
and the electronic contribution is seen to be
P
ni ai «(1) 1
¼ , (A:23:2)
3«0 «(1) þ 2
«(0) «(1)
« ð vÞ ¼ « ð 1 Þ þ , (A:23:5)
1 v2 =v2T
where vT is the observed frequency where the ionic contribution causes «(v) to become
singular. Equation A.23.5 is consistent with 23.A4 if
«(1) þ 2
v2T ¼ ˆ2 (proof left to the reader): (A:23:6)
«(0) þ 2
Since «(1) < «(0), the effect of the internal field is to lower the resonance frequency.
The final result, Equation A.23.5 is the same as the real part of the electronic and ionic
dielectric function found in Equation 23.19. The question is whether it is more accurate to
calculate the ionic contribution with or without the correction for the local field (see
Problems 6–8).
Bibliography
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia, 1976.
Barsoum, M.W., Fundamentals of Ceramics, Institute of Physic Publishing, Bristol and Philadelphia,
2003.
Dekker, A.J., Solid State Physics, Prentice-Hall, Englewood Cliffs, NJ, 1962.
Fischer-Cripps, A.C., The Materials Physics Companion, Taylor & Francis, New York, 2008.
Frohlich, H., Dielectric properties of dipolar solids, Proc. Royal Soc. Lond. A., 185=1003, 399–414, 1946.
Hummel, R.E., Electronic Properties of Materials, 3rd edn., Springer, New York, 2000.
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley & Sons, New York, 1966.
Naranjo, O.B., Gimzewski, J.K., and Putterman, S., Observation of nuclear fusion driven by a
pyroelectric crystal. Nature, April 28, 2005.
Omar, M.A., Elementary Solid State Physics, Addison-Wesley, Reading, MA, 1975.
Yacoby, Y. and Girshberg, Y., Theory of ferroelectric phase transitions in pure and mixed perovskites,
Mat. Res. Soc. Symp. Proc., Vol. 718, Materials Research Society, Warrendale, Pennsylvania, 2002.
Problems
1. How would you define the dielectric constant of a ferroelectric material? How would
you go about measuring it?
2. The relaxation time for liquid water is 1011 s. What microwave frequency would cause
maximum heating? Actual microwave ovens operate at 2.45 GHz, somewhat less than
the frequency from maximum heating. Why do you suppose this is?
3. Can ferroelectric behavior be described in terms of the polarization catastrophe in the
Clausius–Mossotti equation? Use Equation 23.42 to relate the Clausius–Mossotti local
field to the polarization in the medium. Then use Equation 23.28 to relate the tempera-
ture to the polarization assuming E as the local field. Now use a procedure similar to
that used to develop the Curie–Weiss law (Equation 23.55) to obtain the Curie tempera-
ture and the Curie constant for BaTiO3. Does the result agree with observed behavior?
Dielectrics and the Dielectric Function 461
TABLE A.23.1
Static and Electronic Dielectric Constants
for Selected Ionic Compounds
Material «(0) obs «(1) obs
4. The ‘‘g’’ piezoelectric coefficient is the electric field produced by an applied stress in
units of volts meter per Newton. Derive an expression relating the electric displacement,
D, to the strain, ex, produced in the crystal by the applied stress.
5. Show that substituting v2T ¼ ˆ2 («(1) þ 2)=(«(0) þ 2) into Equation A.23.4 results in
Equation A.23.5.
6. Use the calculated values for the lattice frequency vT in Table 16.1 to compute the ionic
contribution to the dielectric function without the Clausius–Mossotti correction for
internal fields (Equation 23.18) and compare with observed measurements in Table
A.23.1. Now do the same using the correction developed by Ashcroft and Mermin.
Does the correction improve the result?
7. Repeat Problem 6 using the observed values for vT.
8. Now use the «(0) and «(1) values in Table A.23.1 with Equation A.23.6 to find ˆ. Then
use Equation A.23.6 to to reduce ˆ to vT and compare with the observed values. What
can you conclude about the need to correct the ionic contribution to the dielectric
function for internal fields with the Clausius–Mossotti equation?
24
Optical Properties of Materials
Having developed the dielectric function for dielectric materials in Chapter 23, we are now
in position to explore the optical properties of nonconductive materials. We will then
modify the dielectric function to account for free electrons and will explore the optical
properties of metals and other conductive media.
r.D ¼ r r E ¼ B_
(24:1)
r.B ¼ 0 _
r H ¼ J þ D,
where D ¼ «E, B ¼ mH, J ¼ sE. This set of equations describes all known electromagnetic
phenomena (which is quite remarkable considering that the electron had yet to be
discovered). The r . D ¼ r equation is a differential form of Gauss’s law, from which
Coulomb’s law can be derived as a special case. The r E ¼ B_ equation is a generalized
form of Faraday’s law and r H ¼ J þ D _ is a generalized form of Ampere’s law. The
r . B ¼ 0 describes the fact that magnetic field lines are closed loops, i.e., they do not
start and end on charges as do electric field lines because isolated magnetic charges
(magnetic monopoles) do not seem to exist. If one is discovered, this equation would
have to be modified to add a magnetic charge density like the electric charge density
in the r . D ¼ r equation and a magnetic current term would have to be added to the
r E equation to match the J in the r H equation. It is this added symmetry to
the equations that has prompted many researchers to look for magnetic monopoles, but
nobody has found one yet.
463
464 Introduction to the Physics and Chemistry of Materials
r.D ¼ 0 r E ¼ B_
(24:2)
r.B ¼ 0 _
r H ¼ D:
r (r E) ¼ r B_ (24:3)
€ ¼ «0Ë, therefore,
_ ¼D
But r . E ¼ 0 and r H
€¼ 1
E r2 E, (24:5)
m0 «0
which we recognize as a three-dimension vector wave equation for the E-vector. A similar
equation can be derived for the H-vector. For an electromagnetic wave traveling in the
x-direction and oscillating in the y-direction, the wave equation is
@ 2 Ey ðx, tÞ 1 @ 2 Ey ðx tÞ
¼ , (24:6)
@t 2 m0 «0 @x2
@ 2 Ey ðx, tÞ @ 2 E y ðx tÞ
¼ E0 v2 ei(kxþvt) and ¼ E0 k2 eiðkxþvtÞ : (24:8)
@t2 @x2
Therefore, putting this back into the wave equation requires that
k
v ¼ pffiffiffiffiffiffiffiffiffiffi : (24:9)
m0 «0
Since v ¼ 2pn, k ¼ 2p=l, and since the wavelength of a wave is given by its speed divided
by its frequency, the speed of an electromagnetic wave in vacuum must be
1
c ¼ pffiffiffiffiffiffiffiffiffiffi : (24:10)
m0 «0
Optical Properties of Materials 465
This is a truly remarkable result. We not only see that the speed of light c is a universal
constant that is interrelated to the universal constants m0 and «0 but that this speed is
independent of any reference system. You may recall that this was the key assertion in
Einstein’s theory of special relativity. Here this was predicted from Maxwell’s equations
some 30 years before.
1
n ¼ pffiffiffiffiffiffi : (24:11)
m«
This relative speed is called the index of refraction, which is given by the square root of the
dielectric constant.
@Hz @Ey
¼« : (24:14)
@x @t
Now take the indicated derivatives of the solutions to the wave equations, Ey ¼ E0ei(kx þ vt)
and Hz ¼ H0ei(kx þ vt) and put them into the Equations 24.13 and 24.14 above,
ikE0 ¼ ivmH0
(24:15)
ikH0 ¼ iv«E0 :
Multiply ikE0 ¼ ivmH0 by «E0, the ikH0 ¼ iv«E0 by mH0 and adding both, we get
«E20 ¼ mH02 . Thus we see that the amplitudes of E and H are related by
rffiffiffiffi
jEj m
¼ : (24:16)
jHj «
466 Introduction to the Physics and Chemistry of Materials
The energy flux propagated by the electromagnetic wave is give by P, the Poynting vector
defined as P ¼ E H, or
rffiffiffiffi rffiffiffiffi
« m
jPj ¼ E2 ¼ H2 : (24:17)
m «
Now let us consider what happens when an electromagnetic wave traveling through
vacuum encounters a material with different m and «. For simplicity, we consider only
the case of normal incidence. The E and H vectors must be conserved so that, if part of the
incident wave is reflected, the remainder must be transmitted. This can be expressed as
Ei ER ¼ ET and Hi HR ¼ HT : (24:18)
which can be used to replace the components of H in the above conservation relationship
to give
rffiffiffiffiffiffi rffiffiffiffiffiffi rffiffiffiffi rffiffiffiffi
«0 «0 « «
Ei þ ER ¼ ET ¼ ðEi ER Þ , (24:20)
m0 m0 m m
from which
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffi pffiffiffiffiffi
ER «=m «0 =m0 « «0 n 1
¼ pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi !m!m0 pffiffiffi pffiffiffiffiffi ¼ : (24:21)
Ei «=m þ «0 =m0 « þ «0 n þ 1
Since the reflected intensity is jERj2, the reflection coefficient can be written as
2
ER n 12 (n 1)2
R¼ ¼ ¼ : (24:22)
Ei n þ 1 (n þ 1)2
This can be generalized to account for an electromagnetic wave in medium with n1 incident
on medium n2 by writing
n1 n2 2
R¼ : (24:23)
n1 þ n2
part. These quantities, which are sometimes called the optical constants of a material
(although they do depend on v), characterize the optical properties of any material.
(Usually, in the literature, you will see these as lowercase n and k, but we will use upper
case here to keep from confusing these quantities with the other quantities we use n and k
to denote.) Therefore, we must replace the values for n in the above equations for R and T
with the complex ñ.
n ~2 2
~1 n ~1 n
n ~2 n ~*2
~*1 n N 2 þ K12 þ N22 þ K22 2N1 N2 2K1 K2
R ¼ ¼ ¼ 12 , (24:25)
n~1 þ n
~2 ~1 þ n
n ~2 ~1* þ n
n ~2* N1 þ K12 þ N22 þ K22 þ 2N1 N2 þ 2K1 K2
where the * denotes the complex conjugate. If either the incident or the emerging wave is in
air, for which n ¼ 1.0007, this equation reduces to
(N 1)2 þ K2
R¼ : (24:26)
(N þ 1)2 þ K2
So we see that k is also complex. Now the solution to the wave equation becomes
This is the equation for a damped wave or an exponentially decaying wave. Thus the
presence of an imaginary component in the complex index of refraction is to cause the
propagating wave to lose energy. Thus the K is sometimes called the extinction coefficient.
Since the intensity of the radiation I EE*, we can write
which is known as Beer’s law and where we have identified the absorption coefficient a as
2vK
a¼ : (24:30)
c
Having said that, it should be understood that dielectrics are transparent only as single
crystals or in the amorphous state (glasses). In a polycrystalline material, the light is lost
due to multiple scattering at the grain boundaries unless the wavelength is much greater
than the average crystal size (e.g., polycrystalline dielectric materials with crystals on the
order of micrometers are transparent to microwave (lranges from millimeter to centimeter)
radiation, but not to visible radiation).
In the previous section we showed (Equation 24.12) that the index of refraction was
simply the square root of the dielectric constant and, since the dielectric constant is a
function of frequency and can become negative, the index of refraction is a function of
frequency and may be complex. Summarizing,
pffiffiffiffiffiffiffiffiffiffiffi
~ðvÞ ¼ N ðvÞ þ iKðvÞ ¼
n «r ðvÞ: (24:31)
Having spent a fair amount of effort in Chapter 23 determining the dielectric function of
dielectric materials, we are now in a position to obtain their N and K values at various
wavelengths, which describe their optical properties.
and the polarizability to be given by Equation 23.50 (we can ignore the imaginary part for
the time being)
e2 =m
a(v) ¼ : (24:33)
v20 v2
The a values are given for v v0, so v0 can be found from Equation 24.33.
We can now determine the optical properties of a material in the visible range. But first
we have to know the polarizabilities of the components. Fortunately, these have been
determined and the values for some common optical materials are given in Table 24.1. By
assuming the electron density surrounding the nucleus is uniform, one can show using
Gauss law that the polarization should be given by 4p«0r3, where r is the ion radius. This
value is also included in Table 24.1 for comparison with the measured polarizability.
One can see that using 4p«0r3 estimates the polarizability to within an order of magni-
tude, but is not accurate enough to make useful predictions of the optical properties of
materials.
As an example, we will plot the N and K values for the SiO2 molecule, the building
block of quartz and the silicate glasses. Using the values in Table 24.1, from Equation 24.33
we find the resonance frequencies for the Si4þ ion to be 1.251 1017 rad=s and for the
O2 to be 8.075 1015 rad=s. There are 2.205 1028 ion pairs=m3, one third of which are
Si and twopthird are
ffiffiffiffiffiffiffiffiffiffiffi O ions. Forming
pthe sums
ffiffiffiffiffiffiffiffiffiffiffi and inserting them in Equation 24.33, the
N ðvÞ ¼ Re «r ðvÞ and KðvÞ ¼ Im «r ðvÞ . The «(0) ¼ 1.946 and N(0) ¼ 1.395 and N at
3.543 1015 rad=s (532 nm) is 1.505. The measured index of refraction of crystalline quartz
at 532 nm is 1.55.
Optical Properties of Materials 469
TABLE 24.1
Polarizability of Selected Ions
Ion a (Measured) 4p«0r3
The N and K values are plotted over the visible and near UV spectrum in Figure 24.1.
The plot is separated to show the detail near resonance at 7.043 1015 rad=s and where the
N becomes greater than zero at 9.8 1015 rad=s. Between these frequencies, the N is zero
and from Equation 24.26 the reflectivity will have the value of 1. Therefore, the high
frequency limit of quartz is 7.043 1015 rad=s, which is equivalent to 267 nm (UV). For
very thin films (<267 nm), some radiation can penetrate, but it will be quickly absorbed
because of the very large K in this region. For v 9.8 1015 rad=s, the electronic vibrations
are no longer able to keep up with the oscillating E-field of the photons, thus there will little
250
200
150
N, K
100
50
0
7.034 7.04 7.05
(a) Frequency (1015 rad/s)
FIGURE 24.1
(a) N (solid line) and K (dashed line) values for SiO2 (quartz) in the vicinity of the resonant frequency. A small
damping factor was used in these calculations to keep the N finite.
(continued)
470 Introduction to the Physics and Chemistry of Materials
1.5
N, K
1.0
0.5
0
8 10 12 14
(b) Frequency (1015 rad/s)
interaction between the photons and the dielectric and the N will return to its vacuum
value of 1. Note that in this region N < 1, which implies the phase velocity ¼ c=N will be
greater than the velocity of light.
Notice that the N increases with frequency until the resonant frequency is reached. This
increasing index with frequency is called dispersion and is responsible for blue light being
refracted more than red light as it passes through a prism. The N decreases with frequency
in the region between the resonant frequency and the point where the dielectric function
becomes positive again. This decrease of N with increasing frequency is called anomalous
dispersion.
The dispersion in the visible is seen in Figure 24.2, where N is plotted against l. Notice the
N ¼ 1.505 at l ¼ 532 nm, very close to the observed value of 1.55 as mentioned previously.
2.2
2.0
Index of refraction
1.8
1.6
40
20
N, K
10
0
4 4.2 4.4 4.6 4.8 5 FIGURE 24.3
wT wL N (solid) and K (dashed) values for InAs.
The transverse and longitudinal frequencies
Frequency (1013 rad/s) are indicated.
½«ð0Þ «ð1Þv2T
« ð vÞ ¼ « ð 1 Þ þ : (24:34)
v2T v2
Taking the real and imaginary parts of the square root of Equation 24.34, a typical
plot of the optical constants for an ionic or mixed covalent=ionic compound is shown in
Figure 24.3. This particular plot was calculated for InAs using «(0) ¼ 14.9, «(1) ¼ 12.3, and
vT ¼ 4.1 1013 rad=s, which corresponds to a wave number of 218 cm1 (data from Kittel,
1996). Notice that the N > 0 at 4.5 1013 rad=s in accordance with the Lyddane, Sachs,
Teller (LST) relationship.
Figure 24.4 is a plot of the reflectivity of InAs calculated from Equation 24.26 compared
with the observed value. The observed reflectivity was computed from the N and K values
from Palik, 1997.
1.2
0.8
Reflectivity
0.6
0.4
«ð0Þv2T «ð1Þv2T v2T «ð0Þv2T
« ð vÞ ¼ « ð 1 Þ þ ¼ « ð1 Þ 1 þ : (24:35)
vT v2 vT v2
2 2 vT v2
2 v2T v2
2
v2 «ð1Þv2L vL v2
«ð v Þ ¼ «ð 1 Þ 2 þ 2 ¼ «ð 1 Þ 2 : (24:36)
vT v2 vT v2 vT v2
The index of refraction N ¼ c=v ¼ ckv ¼ «(v)1=2. We can write the dispersion relationship as
2
c2 k2 vL v2
¼ « ð 1 Þ : (24:37)
v2 v2T v2
Solving this equation for v and plotting v versus k provides a dispersion curve for NaCl
as shown in Figure 24.5. Notice that the curve has two branches; the upper branch is
the phonon dispersion relation for the longitudinal mode and approaches vL at k ¼ 0
while the lower branch is the dispersion curve for the transverse mode, which approaches
vT as k increases. No frequencies can propagate between vT and vL, which causes an
energy gap in this region.
The group velocity of light (@v=@k) is zero at vT and vL. The index of refraction N can be
found from Figure 24.5 by dividing the velocity of light by the phase velocity (v=k) and
plotting this against v. Since the v in the lower branch can never exceed vT as k increases,
the phase velocity gets smaller and the index of refraction increases dramatically as v
1 1014
w (rad/s)
wL
FIGURE 24.5
Dispersion relationship (v vs. k) for NaCl in the wT
vicinity of small k. The dashed diagonal line repre-
sents the speed of light in the medium with «(1)
and the dash-dot diagonal line represents the speed
of light in a medium with «(0). The horizontal lines 0
at vT and vL are the energy gap. The lattice cannot 0 5 105
propagate frequencies between these frequencies. k (m−1)
Optical Properties of Materials 473
approaches vT, as in Figure 24.3. In the upper branch, the phase velocity is infinite at k ¼ 0
and quickly approaches the velocity of light in a medium with dielectric constant «(1).
The lower curve in Figure 24.5 represents the coupling of photons with the transverse
lattice vibrations. In this case, the photon and phonon travel together with the same
frequency and wavelength. The quanta of this coupled wave are called polaritons. The
absorption edge is vL at 5 1013 rad=s (l ¼ 60.8 mm). The vT is 3.1 1013 rad=s (l ¼ 37.7 mm
in free space). A photon just below this frequency represented by the end of the lower
branch in Figure 24.5 would have a k of 5 105=m or l ¼ 12.5 mm in the medium giving it a
velocity of 6.7 107 m=s corresponding to N ¼ 4.8. One can see that the velocity of the
photon in the medium rapidly falls with increasing k as the N ¼ k=v increases.
Because there is an energy gap between vT and vT, no photon can enter and propagate in
the medium, the material becomes reflective as shown in Figure 24.4. This region is
sometimes called the reststrahlen (residual ray) band and can be used to obtain a narrow
band of infrared (IR) radiation by reflection from a blackbody source.
As mentioned in Chapter 16, a photon at normal incidence to the crystal can only excite a
transverse wave as illustrated in Figure 16.5. To excite a longitudinal wave along a line of
diatomic ions, there must be a component of the E-vector along the line as illustrated in
Figure 24.6. Vertically polarized photons impinging obliquely on a thin film of a material
such as LiF deposited on a reflective metal substrate such as Au or Ag will be strongly
absorbed at the longitudinal frequency while virtually no absorption is observed at the
transverse frequency or for horizontally polarized light at the longitudinal frequency
(see Chapter 10 in Kittel, 1996). The IR reflective film doubles the path and enhances the
absorption.
FIGURE 24.6
Vertically polarized light exciting a lon-
gitudinal vibration in a line of diatomic
ions. Undisturbed lattice (top) is excited
by a vertically polarized photon causing
the cations (black) and the anions (gray)
to oscillate back-and-forth relative to
each other.
474 Introduction to the Physics and Chemistry of Materials
x dip (0)
x~dip (v) ¼ ð1 ivt Þ, (24:38)
1 þ v2 t 2
where xdip(0) is given by Equation 23.36 as xdip(0) ¼ Ndip(p2dip =3«0 kT). Since x~dip(v) is
complex, the dielectric function will also be complex and can be written as the sum of a
real and imaginary part, «~(v) ¼ «0 (v) þ i«00 (v), where
These two equations are sometimes referred to as the Debye equations. Note that «00dip is
maximized when v ¼ 1=t and «0dip is reduced by ½, thus v ¼ 1=t could be considered the
damping frequency and plays a similar role as the plasma frequency in a conductive
medium.
When the dielectric
pffiffiffi function itself is complex, we must take the square root of a complex
~ ¼ «~ where the tildes represent complex quantities. We can write
number or n
and identify «0 ¼ N2 K2 and «00 ¼ 2NK. Solving these last two equations simultaneously for
N and K, we get
«0 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi «0 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
N2 ¼ þ «0 2 þ «00 2 and K2 ¼ þ «0 2 þ «00 2 : (24:41)
2 2 2 2
At very low frequencies (vt 1), N2 ! «0 ¼ «(0) þ xdip(0) because dipole rotation is able to
follow the changing E-vector and the polarization approaches the static value resulting in a
high index of refraction and high reflectivity. However, as v approaches 1=t, the damping
causes the extinction coefficient K to rise as the N falls, resulting in strong absorption of
radiation as shown in Figure 24.7.
10
6
N, K
2
FIGURE 24.7
N and K values for a dipolar medium with a
0
static dielectric constant of 80 and a relax-
109 1010 1011 1012
ation time t ¼ 1011 s, values typical of
water. Frequency (rad/s)
Optical Properties of Materials 475
Electron traps may also be created in other systems by disrupting the lattice by
energetic radiation bombardment. Diamonds and other gemstones are artificially colored
by exposing them to various forms of radiation followed by a partial anneal.
24.2.5.5 Excitons
If a photon with energy equal or greater than the bandgap energy is incident on a
semiconductor (or insulator), it will be absorbed by giving its energy to produce a hole
in the valence band and a free electron in the conduction band. Although we tend to think
of semiconductors being transparent to frequencies below the bandgap energy, it is
possible for a photon with less than the bandgap energy to lift an electron from the valence
band, but instead of going all the way to the conduction band, the electron is attracted to
the hole it created to form a bound pair. Such a bound electron–hole pair is called an
exciton and can be described by hydrogen-like wave functions. Since excitons can have
several energy states, they are observed as small absorption lines at wavelengths slightly
longer than the cut-on wavelength, the wavelength at which the semiconductor begins
transmitting light. Excitons can move through the crystal carrying energy (but not charge)
and eventually decay, giving up their energy as a photon.
silica. Also their losses are considerably higher than those in silica. Despite some excellent
fundamental research into these heavy metal glasses, they are a long way from replacing
silica optical fibers for telecommunications applications.
The chalcogenide fibers show promise for temperature sensors that carry blackbody
radiation to a detector, for IR imaging using coherent fiber bundles, and for chemical
sensors using evanescent wave spectroscopy (a sensing technique based on the inter-
action of the internally reflected light in an optical fiber with its chemical environment).
Hollow glass waveguides made from AgI-coated silica glass are showing the greatest
promise for laser power delivery (Harrington, J.A.).
24.2.7 Birefringence
Crystals with cubic symmetry have three equivalent axes that interact with light in
the same manner. This means that light passing through the crystal is refracted in the
same manner without change in polarization regardless of its orientation. Such crystals are
considered to be optically isotropic. The atoms or molecules in crystals with lower sym-
metry will not have the same nearest neighbor interactions in each direction and are
optically anisotropic. The c-axis in hexagonal, trigonal (rhombohedral), and tetragonal
crystals has a unique symmetry and is called the optical axis. Light propagating along
this axis behaves the same as in an isotropic crystal and is called the ordinary ray.
However, light that enters the crystal at an angle to the optical axis is split into two rays,
the ordinary ray with index of refraction nO and the extraordinary ray with index of
refraction nE. The measure of birefringence is the absolute value of the difference between
the two indices of refraction.
When the incident ray is split into the ordinary and extraordinary rays, the extraordinary
ray vibrates in the plane that includes the optical axis while the ordinary ray vibrates in a
perpendicular plane. Since the two rays traverse the crystal at different speeds and with
different angles, waves with a specific polarization can be selected in a Nicol prism by
removing one of the waves by internal reflection.
where the x(2) and x(3) are the nonlinear susceptibilities. These higher order terms in the
polarization can produce some very interesting and useful effects such as frequency
doubling (x(2)) and optical switching with the possibility of optical computing (x(3)).
Chi-2 (x(2)) materials are noncentrosymmetric, meaning they have no inversion symmetry.
Single crystals of potassium dihydrogen phosphate (KDP), triglycine sulfate (TGS), potas-
sium niobate (KNbO3), and lithium niobate (LiNbO3) are widely used for converting the IR
output from YAG or Ti:sapphire lasers to visible light. Organic materials have also been
found to possess high x(2) and x(3) values.
give energy to or receive energy from the system. In the case of molecules, some of the
photon energy can go into the excitation of vibrational modes, resulting in a decrease in
frequency of the photon. This is known as Stokes scattering. Conversely, an excited
molecule can give some of its energy to the photon, increasing its frequency, which is
called anti-Stokes scattering. The difference between the original photon frequency and
the Stokes or anti-Stokes frequency is the vibrational frequency of the molecule. These
effects were first reported by C.V. Raman in 1928 for which he received a Nobel Prize in
1930. Being able to observe molecular vibrational spectra by means of visible light is a
valuable tool for analyzing the composition of materials. Also, since the anti-Stokes line
depends on the number of molecules in the excited state, which in turn depends on the
temperature, the ratio of anti-Stokes to Stokes intensities can be used as a noncontact
temperature measurement, which is especially useful in combustion diagnostics. Raman
scattering is also used in remote sensing. A laser beam is used to illuminate an unknown
molecular solid, liquid, or vapor, which can be identified remotely from the Raman
scattered molecular spectra.
In solids, Raman scattering results from inelastic interactions between photons and
phonons in the optical mode. (Similar inelastic scattering from phonons in the acoustic
mode is called Brillouin scattering.) Solids have characteristic phonon modes, which can be
identified by Raman scattering.
v e
v_ þ ¼ E0 eivt , (24:43)
t m
eE0 t
v¼ eivt (24:44)
mð1 þ ivt Þ
ne2 E0 t
J ¼ nev ¼ eivt , (24:45)
mð1 þ ivt Þ
Optical Properties of Materials 479
Note that in the limit v ! 0, the real part is just the DC conductivity and the imaginary part
vanishes.
_ ¼ sE
rH¼JþD ~ þ «E_ sE _
~ þ «0 E: (24:47)
Here, we have ignored the polarization induced by the interaction of the wave with the
bound charge by setting « to «0.
Since both E and H eivt, the right hand side of this equation can be written as
(1 ivt) s0 s0 t
(i«0 v þ s)E
~ 0 ¼ i«0 v þ s0 E0 ¼ þ iv «0 E0 : (24:48)
ð1 þ v 2 t 2 Þ ð1 þ v2 t 2 Þ ð1 þ v 2 t 2 Þ
Now we want to represent (i«0v þ ~ ) in the above expression with iv«0«~, where «~ is a
complex dielectric constant that incorporates the effects of the free electrons in the medium.
We set «~¼ «0 þ i«00 where the prime and double prime are the real and imaginary parts,
respectively. Equating these to the real and imaginary parts of (i«0v þ ~ ) from above, we
obtain
0 00 s0 s0 t
iv«0 «~ ¼ iv«0 « v«0 « ¼ þ iv «0 , (24:49)
ð1 þ v 2 t 2 Þ ð1 þ v2 t 2 Þ
where we identify
s0 t s0
«0 ¼ 1 and «00 ¼ : (24:50)
«0 ð1 þ v2 t 2 Þ v«0 ð1 þ v2 t 2 Þ
ne2
v2p ¼ : (24:51)
m«0
The real and imaginary parts of the complex dielectric constant can be written in terms of
the plasma frequency:
480 Introduction to the Physics and Chemistry of Materials
TABLE 24.2
Properties of Various Conductive Media
Material n (Electrons=m3) s (S=m) vp (rad=s) t(s) vpt lp(mm)
14
Ag 5.85 1028
6.21 10
7
1.36 10 16
3.77 10 514.21 0.138
Cu 8.95 1028 5.88 107 1.69 1016 2.33 1014 393.63 0.112
Al 18.1 1028 3.64 107 2.40 1016 7.15 1015 171.54 0.079
Zn 13.1 1028 1.69 107 2.04 1016 4.58 1015 93.51 0.092
Inconel 8.95 1028 7.75 105 1.69 1016 3.07 1016 5.19 0.112
Si* 1.00 1025 1.60 102 5.64 1014 5.69 1013 101.46 10.56
Si** 1.00 1019 1.60 101 1.78 1011 5.69 1013 0.10 10565
GaAs* 1.00 1025 1.28 103 5.64 1014 4.55 1012 811.67 10.56
GaAs** 1.00 1019 1.28 100 1.78 1011 4.55 1012 0.81 10565
* Lightly doped.
** Heavily doped.
Before we proceed, it would be useful to examine the magnitudes of the parameters vp and
t that determine the optical constants. These are displayed in Table 24.2 for several metals
and semiconductors.
From Table 24.2, we see the vpt product is 1 for the metals. Taking vpt ¼ 100, we can
plot in Figure 24.8 the «0 and «00 as a function of l=lp, where lp is the wavelength
corresponding to the plasma frequency.
At short wavelengths, the real part of the dielectric approaches 1 and the imaginary part
approaches 0. Near the plasma frequency, the real part crosses zero and becomes increas-
ingly negative. The imaginary part starts to become increasingly negative.
0
Dielectric function
−5
−10
−15
−20
FIGURE 24.8
Real (solid) and imaginary (dashed) parts of the −25
0 1 2 3 4 5
dielectric function for a conductive media with
free electrons. l/lp
Optical Properties of Materials 481
100 102
101
10
100
1 10-1
10-2
K
N
0.1 10-3
10-4
0.01
10-5
0.001 10-6
0.1 1 10 100 0.01 1 10 100
(a) l/l p (b) l/l p
FIGURE 24.9
(a) Theoretical index of refraction N for a free electron metal as a function of l=lp and (b) the theoretical extinction
coefficient K for a free electron metal as a function of l=lp.
0.8
0.6
Reflectance
0.4
0.2
0
0.1 1 10 FIGURE 24.10
Theoretical reflection for a free electron metal as
l/l p a function of l=lp.
electromagnetic radiation. From Table 24.2, we find for metals this line is O(100 nm) in the
far-UV. One can gain more insight into the optics of conducting media by considering a
couple of limiting cases.
«00 vp t
2
N 2 K2 ¼ : (24:54)
2 2v
482 Introduction to the Physics and Chemistry of Materials
N 2 þ 2N þ 1 þ K 4N 4N
R¼ ¼1 2 : (24:55)
N 2 þ 2N þ 1 þ K N þ 2N þ 1 þ K
But since N ¼ K,
sffiffiffiffiffiffiffiffi
4N 2 2v
R¼1 1 ¼12 : (24:56)
2N 2 þ 2N þ 1 N v2p t
Using Equations 24.46 and 24.51 to express vp and t in more familiar terms, we can write
the reflectance as
rffiffiffiffiffiffiffiffiffiffi
2v«0
R¼12 : (24:57)
s
This expression for the reflectivity of metals at low frequencies is known as the Hagen–Rubens
relation.
and metals become transparent to radiation whose frequency is greater than their plasma
frequency vp. From Table 24.2, we see that plasma frequencies are O(1016 rad=s), which
corresponds to a wavelength of 188 nm in the UV region. Thus metals start to become
transparent in the UV and are quite transparent in the x-ray portion of the spectrum. (The
partial opacity of metals to x-rays is due to scattering of the x-rays from the electrons
bound to the ion cores, not from the interactions with the free electrons.)
100
10
1
N
0.1
0.01
0.001
0.1 1 10 100 1000
(a) l/l p
1000
100
10
1
K
0.1
frequency for Al is 2.40 1016 s1, which corresponds to 15.7 eV or lp ¼ 79 nm. There is a
strong interband transition at 1.5 eV, which corresponds to l=lp ¼ 10.12. The effect of this
transition shows up in both the N and K plot, but otherwise the measured values closely
follow the theoretical simple Drude free electron theory.
Similarly, the reflectance closely follows the Drude free electron model except for the
region in the vicinity of the interband transition as seen in Figure 24.12.
Despite the small loss in reflectance in the visible region, pure Al is the best reflector for the
far UV because its high electron density (three valence electrons per atom) extends its plasma
frequency to shorter wavelengths and because there are no additional interband transitions
in this region. Ag maintains its high reflectance throughout the visible spectrum, but
undergoes an interband transition at 4.0 eV (310 nm), which kills its UV reflectance.
However, we virtually never see a pure Al surface. Al has a high affinity for oxygen and
an oxide layer will quickly form when exposed to any O2 that may be in the atmosphere.
484 Introduction to the Physics and Chemistry of Materials
0.8
0.6
Reflectance
0.4
0.2
FIGURE 24.12
Measured reflectance of vapor deposited Al
under UHV (solid) compared with free electron 0
theory (dashed line). (The observed reflectance 0.1 1 10 100 1000
was computed from the measured N and K val-
ues shown in Figure 24.11a and b.) l/l p
The data shown in Figures 24.11 and 24.12 were taken under ultrahigh vacuum (UHV)
conditions (109 to 1010 Torr) and are presumed to be free of any contamination. The
oxide film becomes absorbent in the vacuum UV region (120 nm). Consequently, mirrors
and gratings for use in this spectral region must be coated with transparent protective
coatings such as LiF before they are removed from the UHV environment where they were
prepared. It is also worth mentioning that evaporated Al films have higher reflectance
than polished bulk Al because of the increased surface scattering created by the
polishing process.
Next we consider the observed optical spectrum of Cu, a metal known for its reddish
color. We see from Figure 24.13a and b that the N and K values approach the free electron
theory in the long wavelength limit but differ significantly in the vicinity of lp. These
departures are manifested in the reflectance spectrum shown in Figure 24.14a and b.
From Table 24.2, we see that lp ¼ 112 nm, and from Figure 24.14a, we see the reflectance
of Cu starts to fall off at around 8lp, which corresponds to the beginning of the visible
spectrum. Since the reflectance falls off more rapidly at the shorter wavelengths, the
reflectance of blue light is much less than that of red, hence the reddish color of Copper.
Adding Zn to the Cu shifts these transitions to longer wavelengths giving the alloy a
redder color. An experienced metallurgist can judge the amount of Zn in a brass alloy by its
color. Notice the bumps in the reflectance curve as the wavelength decreases. These are due
to additional interband transitions and can be better resolved by plotting the log of
reflectance against the log of the photon energy as shown in Figure 24.14b.
100
10
1
N
0.1
0.01
0.001
0.1 1 10 100
(a) l/l p
102
101
100
10−1
K
10−2
Gold’s yellow color is due to a similar mechanism. Silver has a band structure similar to
Cu except the d-bands are 4 eV below the Fermi level; hence the loss of reflectance occurs
in the UV region.
24.3.9 Semiconductors
Intrinsic semiconductors or insulators whose bandgap is above the visible will have low
enough electron densities to place their plasma frequencies well below the visible region of
the electromagnetic spectrum. As shown in Chapter 20, such materials can be made
electrically conductive by adding impurities to create donor or acceptor states. Even at
doping levels as high as 1025 m3, the plasma frequency is well below the lowest part of the
visible spectrum. The ability to control the electron density by impurity doping makes it
possible to create reasonably good electrical conductors that are transparent to visible light.
486 Introduction to the Physics and Chemistry of Materials
0.8
0.6
Reflectance
0.4
0.2
0
0.1 1 10 100
(a) l/l p
0.1
Reflectance
0.01
0.001
0.1 1 10 100
(b) Photon energy (eV)
pjwstk|402064|1435428914
FIGURE 24.14
(a) Measured reflectance of vapor deposited Cu (solid) compared with free electron theory (dashed line). The
reflectance was computed from the measured N and K values shown in Figure 24.13a and b. (b) Log plot of the
measured reflectance of Cu versus photon energy. Peaks are associated with interband transitions.
Thin transparent conductive films such as indium tin oxide are crucial to the fabrication of
electroluminescent panels and to liquid crystal displays.
24.3.10 Plasmas
The sharp transition from transparency to reflection at the plasma frequency predicted
by the Drude free electron model can be better understood by considering a free
electron gas. The undamped motion of an electron responding to a time varying electric
field is m€x ¼ eE0eivt. The solution is x(v) ¼ eE0=mv2, the dipole moment of a single
Optical Properties of Materials 487
Energy
EF
FIGURE 24.15
Selected bands in Cu showing possible inter-
band transitions in the visible and near-UV.
Γ X W L Γ K X (Adapted from Figure 19.9)
electron is p(v) ¼ ex(v) ¼ e2E0=mv2 and the dipole moment per unit volume is P(v) ¼
ne2E0=mv2. The susceptibility becomes x(v) ¼ ne2=«0mv2 and the dielectric
function may be written as
ne2 v2p
«r ðvÞ ¼ 1 þ x ðvÞ ¼ 1 ¼ 1 , (24:60)
«0 mv2 v2
where vp is the frequency at which «r(v) ¼ 0. Notice that this is also the plasma frequency
we defined by Equation 24.51 and that Equation 24.60 is the high frequency limit of
Equation 24.52. For v < vp, the electrons behave individually and can oscillate with the
electric field and reflect the radiation as assumed by the Drude theory. However, when
the electron gas becomes dense enough, the electrons in the gas start to behave collectively
and begin to oscillate together in a longitudinal mode as shown in Figure 24.16. These
collective oscillations are quantized with energies given by hvp and are referred to as
pffiffiffiffiffiffiffiffiffiffiffi
plasmons. Since there is no damping, N ðvÞ ¼ «r ðvÞ and K ¼ 0. When the applied fre-
quency v is less than the plasma frequency, «r < 0 and N ¼ 0. For v > vp, N approaches 1 as
v increases.
Plasmons can be excited by photons entering the metal obliquely (Figure 24.6) so that
there is a component of the E-vector in the longitudinal direction or by an electron passing
through the metal. Plasmon energies are measured by measuring the energy lost by a beam
of electrons passing through a thin metal foil, a technique known as electron energy loss
spectroscopy (EELS).
The plasma frequency derived in Equation 24.60 also applies to electrons in the iono-
sphere. Radio waves below the plasma frequency (amplitude modulated [AM] and short-
wave broadcast) are reflected back to earth by the ionosphere and can carry over long
distances. Frequency modulated (FM) and TV frequencies (> 50 MHz) penetrate the iono-
sphere and reception is limited to line-of-sight distances. Spacecraft and direct broadcast
FIGURE 24.16
Plasmons consist of layers of electrons in a metal
that oscillate collectively at the plasma fre-
quency. The electrons interact with the ion
cores causing them to oscillate at their longitu-
dinal frequency vL.
488 Introduction to the Physics and Chemistry of Materials
satellites use UHF and microwave channels to penetrate the ionosphere. During spacecraft
reentry, the electron density gets so large because of the intense heat that even
these highest frequencies cannot penetrate the plasma, which causes the communication
blackout.
v2p c2 c2 k2
N2 ¼ 1 ¼ ¼ 2 , (24:61)
v2 n2 v
from which we obtain
c2 k2 ¼ v2 v2p : (24:62)
This dispersion relation shown in Figure 24.17 is similar to the upper or longitudinal
branch of Figure 24.5. No wave can propagate if v < vp and above vP, only the longitu-
dinal mode can be excited. Recall that we defined vL as the frequency at which the
dielectric constant became positive again. So we see that for conductive media, the plasma
frequency corresponds to the frequency of the longitudinal optical mode.
Even though the phase velocity v=k > c, it can be seen that the group velocity @v=@k
approaches c as v vp but is always less than c as can be verified by differentiating
Equation 24.62 with respect to k.
2
w/w p
FIGURE 24.17 0
Dispersion relation for radiation passing through a conductive 0 1 2 3
medium with frequency > vp. The dashed line represents the
kc/w p
velocity of light. No propagation is possible for v < vp.
Optical Properties of Materials 489
24.4 Summary
The optical properties of a material are characterized by the real and imaginary parts of the
index of refraction (N þ iK) as a function of wavelength. These optical parameters are
obtained from the square root of the dielectric function, which may also be complex.
Most dielectrics are transparent from the near IR to the near-UV. The short wavelength
limit is usually determined by electronic polarization unless the band edge is encountered
first. The band edge is related to the photon energy by l (mm) ¼ 1.24=E(eV). Since the
polarizability a(0) goes roughly as r3 and the resonant frequency v0 goes as a(0)1=2,
the lightest molecules such as LiF will have the shortest wavelength cutoff. The increase
in the N with frequency as the resonant frequency is approached is called dispersion and is
responsible for blue light being refracted more than red light in a prism.
The long wave limit for dielectrics is usually set by photons coupling with the optical
modes of the molecular vibrations as discussed in Chapter 16. The IR absorption edge goes
as the square root of the spring constant divided by the reduced mass of the system, so to
extend the IR cutoff, one looks for heavy molecules with a weak spring constant, or a
system with a low sound velocity since the velocity of sound goes as the square root of the
spring constant divided by the density.
Photons can couple with transverse optical phonons near k ¼ 0 in dielectric materials.
This quantized photon–phonon has energyhv and is call a polariton. As the energy of the
photon approaches the transverse resonant frequency, the N gets large and the velocity of
propagation approaches zero. The N drops suddenly to 0 at vT and remains 0 until vL and
no propagation can occur between these two frequencies. The crystal becomes almost a
perfect reflector between these two frequencies and this band is called the reststrahlen
region. Above vL only longitudinal optical modes can couple with photons. Longitudinal
modes can only be excited by photons traveling obliquely to the lattice planes, but this does
not mean that photons at normal incidence cannot propagate; they simply do not interact
with the phonons.
Media with dipolar molecules will have a large N at frequencies below their damping
frequency given by 1=t, where t is their relaxation time. Around this damping frequency,
the N drops and K increases as energy is being absorbed by the oscillating dipoles.
This absorption is the mechanism responsible for cooking foods in a microwave oven.
It is also responsible for the reflection of microwave radar from clouds. Above the damping
frequency, the dipoles cannot oscillate as fast as the oscillating E-field of the phonon and
there is no interaction.
The optical behavior of conductive media is characterized by the plasma frequency,
v2p ¼ ne2=m«0 which is a function only of the electron concentration. Plasma frequencies for
metals are O(1016=s) which corresponds to the vacuum UV. Below the plasma frequency,
the N and K values are large making the material highly reflective (except for inner band
transitions which are responsible for the color of Cu and Au). Above the plasma frequency,
metals start to becomes transmissive, but they will still absorb strongly until the v vP.
The plasma frequency of semiconductors is determined by the doping level. By selecting
an appropriate doping level, it is possible to make a conducting film that is still transparent
to visible radiation. Such films are widely used in optoelectronics devices.
The free electrons in the ionosphere will reflect short wave radio waves at frequencies
below their plasma frequency (50 MHz) but will transmit FM and TV frequencies. Space-
craft reentry creates a plasma whose frequency is much higher than that of the atmosphere,
which accounts for the communication blackout during this phase of a space mission.
490 Introduction to the Physics and Chemistry of Materials
Bibliography
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia, 1976.
Barsoum, M.W., Fundamentals of Ceramics, Institute of Physic Publishing, Bristol=Philadelphia, PA,
2003.
Gersten, J.I. and Smith, F.W., The Physics and Chemistry of Materials, John Wiley & Sons, New York,
2001.
Harrington, J.A., Infrared Fiber Optics, OSA Handbook, Vol. III, McGraw Hill, New York.
Henning, T.H., et al., Reference & Data Base of Optical Constants, Astron. Astrophys. Suppl., 136, 1999.
Available at www.astro.spbu.ru=JPDOC=f-dbase.html.
Hummel, R.E., Electronic Properties of Materials, 3rd edn., Springer-Verlag, New York, 2000.
Ibach, H. and Lüth, H., Solid State Physics, 3rd edn., Springer-Verlag, New York, 1990.
Kingery, W.D., Bowen, H.K. and Uhlmann, D.R., Introduction to Ceramics, 2nd edn., John Wiley &
Sons, New York, 1976.
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley & Sons, New York, 1966.
Palik, E.D., Ed., Handbook of Optical Constants of Solids, Vols. 1–3, Academic Press, San Diego,
California, 1997.
Wooten, F., Optical Properties of Solids, Academic Press, New York, 1972.
Problems
1. NaCl is often used as an IR window. What would you estimate its short wavelength
cutoff to be? a ¼ 3.0 1040 for Cl and 0.18 1040 for Naþ (MKS units).
2. Show that v2L v2Debye ðM þ mÞ2 =Mn.
3. You are called upon to design a solar blind filter for a far-UV space experiment. The sun
emits a strong L-alpha line at 121.6 nm, which you must block, but you would like to
transmit all wavelengths shorter than this as much as possible. You propose to use a
thin metallic foil as the filter. Consider an appropriate choice for the material. Plot the
transmission curve from 100 to 10 nm. (Better use log plots.)
4. Spectrophotometry is often used in qualitative and quantitative analysis to determine
the species and concentration of an unknown solute. A cuvette of known width x is
filled with the solution and white light is passed through the solution. The transmitted
light is analyzed with a monochromator to obtain the absorption spectra, which can be
used to identify the solute. The concentration of the solute can be determined by
measuring the absorbance A ¼ log10(Iref=Isample) at a wavelength where the solute
absorbs strongly. For dilute solutions, the absorptivity a is directly related to the
concentration so the measured A is directly related to the log of the concentration.
(a) Let the light with intensity I0 be partially reflected at the first air–cuvette interface with
reflectivity R and partially reflected with the same R as it emerges from the cuvette.
The index of refraction of the solution in the cuvette is matched to the cuvette so there
is no reflection at the solution–cuvette interfaces. Show that the intensity of the
emerging ray is given by I0(1 R)2exp(ax).
(b) Now trace the ray reflected from the back of the sample back through the solution
Nuntil it is reflected from the front surface, then back through the solution until it
emerges from the back. Show that the intensity of this ray is I0R2(1 R)2exp(3ax).
Continue to trace the internally reflected ray back-and-forth through the sample
and sum the total intensity of the transmitted ray. (After a few terms you should be
Optical Properties of Materials 491
P
able to express the sum in terms of a geometric series. 1 n¼0 s ¼ 1=(1 s) provided
n
s < 1.)
(c) Show that the transmitted intensity is I0 ð1 RÞ2 eax =1 R2 e2ax .
(d) This correction for multiple paths is generally ignored in spectrophotometric meas-
urements. Comment on when this is and is not justified. Does the fact that you are
comparing the transmitted light through the unknown sample to that of a reference
sample cancel out this source of error? Support your answer with relative error
calculations for various values of a.
(e) Using the procedure in (c), sum the total reflected intensity, the total absorbed
intensity, and show that the sum of transmitted, absorbed, and reflected intensities
is equal to the incident intensity.
5. You are given the task of designing a smooth absorptive coating to go on a reflective
metal surface. You are restricted to a thickness of one wavelength of the radiation you
want to absorb.
(a) What strategy would you adopt in looking for materials in terms of their N and K
(real and imaginary index of refraction) values to minimize the reflectance? You
have come up with a binder that has an index of refraction of N ¼ 1.2 and you have
found that you can control the K-value by doping it with carbon nanotubes.
(b) What value of K would you choose to minimize the total reflectivity? What would
be the value of this reflectivity?
(Trace a ray at normal incidence on the front surface. Let the transmitted fraction
of light enter the coating and reflect from the rear surface assuming R ¼ 1. Then let it
re-emerge through the front surface. The total reflectance is the portion of the
incident wave reflected from the front surface plus the ray that was reflected from
the rear surface. It is not necessary to consider multiple reflections within the
coating.)
6. You are asked to design a conductive coating that is transparent in the visible region. You
have chosen a semiconductor with a bandgap well above optical frequencies.
(a) What dopant level would you choose to maximize the conductivity and still be
transparent in the visible (>450 nm)?
(b) If the mobility m ¼ 0.1 m2=V s, what would the conductivity be? How does this
compare with Cu?
7. AM radio waves (540–1600 KHz) as well as shortwave broadcasts (up to 20 MHz) can be
heard over great distances because their waves are reflected back to Earth by the
ionosphere. However, TV (Channel 2 is 55 MHz) and FM broadcasts (88–108 MHz)
are limited to line of sight between the transmitter and receiver. With this information,
what limits can you place on the electron density of the ionosphere?
8. Several space shuttle mapping missions using L-band (20 cm) side-looking synthetic
aperture radar were able to see rock formations and water channels buried 3–4 m under
the sands of the Arabian Peninsula. Explain how this is possible and estimate the
conductivity of the sands that would allow such observations.
9. Relaxation time for H2O molecules is 1011 s. What is the frequency used in
a microwave oven to maximize the heating of foods containing water? What frequency
would you set your weather radar to in order to image clouds? Why does snow produce
less radar return than rain?
25
Magnetism and Magnetic Materials
493
494 Introduction to the Physics and Chemistry of Materials
Taking the path of integration at a constant radius R around the wire, B ¼ m0I=2pR or
H ¼ I=2pR (A=m) in free space. Similarly, the B inside of a long solenoid may be found by
taking the path of integration as a rectangle of length L parallel to the axis of the solenoid
that encloses one side of the windings. Since the field inside of a long solenoid is contained
inside the windings, only the side of the rectangle inside the solenoid will be along field
lines. Thus from Ampere’s law we find B ¼ mI(N=L), where I is the current flowing through
the wires and (N=L) is the number of turns per length.
A current flowing around a single loop with radius R constitutes a magnetic dipole with
a magnetic dipole moment with magnitude mM which equals the current times the area
enclosed by the current loop, or in this case, mM ¼ p IR2. The circuit form of Ampere’s law is
useful only in cases of high symmetry, which does not apply in this example. Thus to find
the resultant magnetic field, we must go to the more general law developed by Biot and
Savart, i.e.,
m0 I d‘ R
dB? ¼ , (25:2)
4p r3
where r is the distance from the element of length d‘ to the observation point. Integrating
this equation around a circular loop with radius R, the magnetic field along the axis can be
found to be
m0 IR2
B(x) ¼ 3=2
: (25:3)
2ðR2 þ x2 Þ
m0 IR2 m0 mM
B(x) ¼ ¼ , (25:4)
2x3 2px3
Electrons moving in their orbits in the atoms making up the material constitute current
loops, which should have magnetic moments (if they do not all cancel out). Classically, an
electron moving around the nuclei with velocity n represents a current qn=2pR, where q is
the electronic charge and 2pR is the circumference of the orbit. The magnetic moment of an
electron in orbit is then the product of this current and the area enclosed, or
qn qvR2 q qh
mM ¼ pR2 ¼ ¼ jLj ¼ me , (25:6)
2pR 2 2m 2m
where
L is the orbital angular momentum
me is the projection quantum number (the negative sign implies the mM is opposite to
the angular momentum)
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(Recall that jLj ¼ h ‘(‘ þ ‘) and ‘ ¼ 0, 1, 2 , n 1, and m‘ ¼ ‘, ‘ þ 1 þ ‘. L never
lines up with the magnetic field direction, but projects a component, hm‘ on the
B-direction.) The quantity
q
h
¼ mB ¼ 9:27 1024 A m2 (25:7)
2m
is called the Bohr magneton. The orbital contribution to the magnetic moment is in integral
units of Bohr magnetons. The electron spin also contributes to magnetic moment in units of
2mBms where ms ¼ 1=2.
q
mM ¼ ð2S þ LÞ: (25:8)
2m
q
mM ¼ gJ: (25:9)
2m
The appropriate expression for g can be found by taking dot product with J in Equations
25.8 and 25.9, which gives
From the law of cosines, J2 ¼ L2 þ S2 þ 2L . S which when put into the above equation yields
L2 þ 2S2 þ 3=2 J 2 L2 S2 J 2 þ S2 L2
g¼ 2
¼1þ , (25:11)
J 2J 2
496 Introduction to the Physics and Chemistry of Materials
where
J2 ¼ J0 (J0 þ 1)h
S2 ¼ S0 (S0 þ 1)h
L2 ¼ L0 (L0 þ 1)h
J0 , S0 , and L0 are the sums of the individual quantum number of the electrons making
up the atoms
These values are determined by Hund’s rules, which were empirically determined by
examining spectra.
25.3 Diamagnetism
Earnshaw’s theorem states that it is not possible to arrange a static combination of
electrostatic charges or magnets such that a charge or a magnet can be levitated in a stable
configuration (Earnshaw, 1842). However, diamagnetic materials are repelled by magnetic
fields and Earnshaw’s theorem does not apply. Superconductors are perfect diamagnets
(x ¼ 1) and can be levitated above permanent magnets. The diamagnetic moment of most
materials is too small to be of practical use and is often exceeded by the paramagnetic
susceptibility. But there are a few normal materials that have unusually high diamagnetic
susceptibilities that can be levitated by static magnetic fields.
magnetic moment (this is a consequence of Lenz’s law which states that the current
induced by a magnetic field creates an opposing field). Quantitatively, the new angular
frequency is
qB
v2 ¼ v20 v: (25:12)
m
Solving for v
qB
v ¼ v0 þ O B2 : (25:13)
2m
Had the B-field been opposite the mM, the effect would be to speed up the electron and
make the magnetic moment more negative and Equation 25.14 would be
qr2 qB q2 r2
m0M ¼ v0 þ ¼ mM B: (25:15)
2 2m 4m
If we sum over Z electrons per atom and N atoms=m3, the mM cancel out and the magnetic
susceptibility becomes
M m0 ZNDmm m ZNq2 r2
x¼ ¼ ¼ 0 : (25:16)
H B 4m
Since the susceptibility <0, the response is diamagnetic. We estimate the susceptibility by
taking Z ¼ 10, N ¼ 1028 m3, r ¼ 1010 m, m ¼ 1030 kg, and m0 ¼ 106, q 1019, x 105.
qB
vL ¼ : (25:17)
2m
This precession represents a shell of charge rotating about the B-direction. The resulting
magnetic moment is
qvL r2 q2 Br2
mM ¼ ¼ , (25:18)
2 4m
The effect is small, but all atoms have a small amount of diamagnetism. This includes
atoms in gases and liquids, atoms in van der Waals’ bonded crystals, ions in ionic salts, ion
cores in metals, inner core electrons in covalently bonded crystals, and in molecules of
living systems.
There are, however, materials in which the diamagnetic susceptibility is considerably
more than 105. Bismuth and pyrolytic graphite have high enough diamagnetic
susceptibilities that they can be floated above an array of small permanent magnets.
There are some interband effects in Bi and graphite that account for their unusually large
diamagnetic susceptibilities. See references for discussions of the unusually large dielectric
coefficients of water and graphite.
Water also has a diamagnetic susceptibility about 20 times smaller that Bi and graphite,
but high enough so that small living creatures can be levitated in the core of a large
superconducting magnet. A photograph of a frog being levitated in a 16 T magnet is
available at www.hfml.science.ru.nl=froglev.html. (The frog was not harmed.)
25.4 Paramagnetism
25.4.1 Langevin Paramagnetism (Classical Approach)
If we have unpaired electrons, it is possible to have a net magnetic moment. The energy of
a magnetic dipole in a B-field is given by U ¼ B . mM. The B-field will try to line up the
magnetic dipoles, but thermal motion will try to disrupt this ordering. At thermal equilib-
rium, the average dipole moment will be given by
Ðp
(mM cos u)emM B cos u=kT 2p sin udu
hmM i ¼ 0 Ðp ¼ mM L(a), (25:19)
emM B cos u=kT 2p sin udu
0
where a ¼ mMB=kT and the Langevin function L(a) ¼ coth a 1=a. At ambient temperat-
ures, kT mMB. Since coth a ¼ 1=a þ a=3 a3=45 þ , for small a, L(a) ! a=3. For very
low temperatures, a 1, and L(a) ! 1.
At ambient temperatures (a 1), hmMi ¼ mMa=3 ¼ m2M B=3kT and, since the magnetism
M is NhmMi,
Nm2M B C
M¼ ¼ H: (25:20)
3kT T
This is known as the Curie law of paramagnetism and C ¼ Nm0 m2M =3k is the Curie constant
for a given material. The magnetic susceptibility is
M Nm2M m0
x¼ ¼ , (25:21)
H 3kT
which is positive.
What materials would be expected to exhibit paramagnetism? Even-numbered elements
will always have electrons in pairs; hence, their magnetic moments will cancel each other.
Odd-numbered elements, could have unpaired electrons, but generally this odd electron
has formed either an ionic or covalent bond to form a full shell in which the electron pairs
Magnetism and Magnetic Materials 499
cancel themselves. Thus paramagnetic materials are generally salts of the transition elem-
ents with unfilled d-shells, or of the rare earth elements with unfilled f-shells.
However, there is one important exception—the O2 molecule. Recall from Figure 3.7,
the electron bonding scheme for N2. For O2, we need to place two more electrons and by
Hund’s rules they will have to go into the two py antibonding sites, giving us two unpaired
electron spins. Liquid O2 will cling to the poles of a strong magnet. Moreover, there are
commercially available process stream oxygen sensors based on the paramagnetic suscep-
tibility of the O2 molecule.
The large change in susceptibility with temperature of certain rare earth salts at cryo-
genic temperatures has been used to make highly sensitivity thermometers, especially with
the ability to detect minute changes in the magnetic field using superconducting quantum
interference detectors. See Chapter 26.
0
P
J
MJ gmB eMJ gmB B=kT
J 0
hmz i ¼ ¼ gmB J 0 BJ 0 (y), (25:22)
P
J0
eMJ gmB B=kT
J 0
Ng2 m2B J 0 ðJ 0 þ 1Þ
M¼ B, (25:24)
3kT
which is the classical result if m2M in Equation 25.21 is replaced by g2m2B J0 (J0 þ 1).
rare earth elements lie further inside the atom; hence, they are less perturbed and contrib-
ute according to the above theory.
Nm0 m2B
x¼ : (25:25)
3kT
where N(EF) is the density of states at the Fermi level. Since the Fermi energy N2=3, the
density of state can be written as N ðEF Þ ¼ dN=dEF ¼ 3N=2EF and Equation 25.26 becomes
This equation looks similar to the classical result except the inverse temperature has been
replaced with the Fermi temperature. Since TF=300 102, x 105, which is typical of the
observed values. Since the diamagnetic and paramagnetic susceptibilities are of the same
order of magnitude, metals can be either paramagnetic or diamagnetic, depending on which
susceptibility is the larger. Most metals are paramagnetic; Cu, Ag, and Au are diamagnetic.
25.5 Ferromagnetism
By far the most useful and fascinating aspect of magnetism is the mysterious ability
of certain materials to attract certain other materials with great force. Elements such as
Co, Fe, Ni, Gd, and Dy and compounds such as MnBi, MnSb, and MnAs exhibit spon-
taneous magnetization, a phenomena known as ferromagnetism. They maintain this
property until they reach a certain temperature known as their Curie temperature (TC),
at which point their magnetism is destroyed. Cooling below the Curie temperature does
not automatically restore their magnetism, but they can be remagnetized by placing them
in a strong magnetic field. How do we explain such behavior?
these spins aligned in adjacent atoms so that they remain aligned until they are disrupted
by thermal disordering at their Curie temperature, which in the case of Fe, is over 1000 K.
What could be the source of this ordering energy?
Weiss assumed that there was an internal H-field that was somehow proportional to the
magnetization M. Recall that in dielectrics, the local electric field was enhanced by the
electric dipoles in the surrounding media. Perhaps once the electron spins were lined up,
their magnetic field would be strong enough to keep them aligned. Weiss set this local field
B ¼ m0(H þ lM), where l is called the Weiss constant. Now consider the quantum equation
for paramagnetism (Equation 25.22). For electrons, MJ ¼ 1=2, 1=2 and the average value
for mz for H ¼ 0 is given by
mB emB B=kT emB B=kT
hmz i ¼ ¼ mB tanhðmB B=kTÞ ¼ mB tanhðmB m0 lM=kT Þ (25:28)
emB B=kT þ emB B=kT
which can be solved for M. (Note the similarity to Equation 23.28 for dipolar solids.) As T
goes to zero, the tanh goes to 1 and M ¼ Msat ¼ NeffmB where Neff is the effective number of
electrons with aligned spins. As T gets large so that m0mBlM=kT 1, Equation 25.29
becomes
and a solution to Equation 25.29 is possible only if T TC, the Curie temperature. We then
can use Equation 25.30 to find the Weiss constant l:
Putting this result into Equation 25.29, we can solve for the magnetization as a function of
temperature:
T M=Msat
¼ : (25:32)
TC atanhðM=Msat Þ
1.0
0.8
0.6
M/M sat
0.4
0.2
FIGURE 25.1 0
Universal magnetization versus reduced tem- 0 0.2 0.4 0.6 0.8 1.0
perature curve. The model gives very good agree-
T/T C
ment with measured data.
or
M C
x¼ ¼ ; T > TC , (25:35)
H T TC
where C ¼ Neffm0mB2=k is the Curie constant for the material. Note that for T > TC, the
spontaneous magnetism is destroyed and the magnetism is directly proportional to H,
which shows paramagnetic behavior.
For Fe, the saturation magnetization near 0 K is observed to be 1.75 106 A=m and the
Neff ¼ 1.75 106=mB ¼ 1.882 1029=m3. The number density of atoms is N ¼ 8.5 1028
atoms=m3. Therefore, the peff ¼ Neff=N ¼ 2.214. The Curie constant is 1.482 K and since
TC ¼ 1043 K, l ¼ 703. This would imply an internal M that is 700 times the saturation
magnetization. Clearly, this ordering energy cannot be magnetic in nature. The Curie–
Weiss law correctly describes paramagnetic behavior for T > TC; it gives a fairly accurate
description of the magnetization as a function of temperature, and it estimates the magni-
tude of the ordering energy required to the keep the electron spins aligned up to the TC. It
does not give a clue as to the source of the ordering energy or why some materials exhibit
spontaneous magnetization and others with unpaired spins do not. For these explanations,
we must turn to a quantum mechanical model.
It should also be mentioned that peff obtained from the saturation magnetism was
only 2.22, whereas the value for Fe2þ with four unpaired spins estimated from
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 S ðS0 þ 1Þ ¼ 4:90. There are several possible reasons for this lower value of peff in the
0
pure metal: spin–orbit interactions, interference with the conduction electrons, and the
Fermi level cutting off some of these higher energy states. Also some fraction of
the d-electrons participate in forming covalent bonds. It is not a perfect theory.
1500
Co
1000 Fe
Ni
TC (K)
500
Dy
0
Mn
Cr
−500
2.5 3.0 3.5 4.0 4.5 5.0 5.5
r a/r d
FIGURE 25.2
Bethe-Slater type curve in which the Curie temperature rather than the exchange energy is plotted against the ratio
of atomic radius to the radius of the d-electron shell (f-electron shell in the case of Dy). Negative temperatures for
Cr and Mn correspond to their Néel temperatures. (Data taken from Kittel, 7th edn., 1996).
using the Heisenberg exchange theory (the reader is referred to Ashcroft and Mermin [1976]
for details). Bethe and Slater showed that when the ratio of the distance between nearest
neighbors to the diameter of the orbits of the d-electrons > 3, the parallel spin state is
favored. Figure 25.2 is a plot of the Curie temperature, which is a measure of the exchange
energy, as a function of the radius of the atom ra divided by the radius of the d-electron
shell rd. The Néel temperatures of Cr and Mn were plotted as negative TC since these
elements are antiferromagnetic. The ra=rf for Dy was included to see if the model could be
extended to the rare earth series. Cu would fit on the curve past Ni, but it has a filled d-shell.
25.5.3 Antiferromagnetism
As mentioned in Section 25.1.1, certain materials exhibit what is called antiferromagnetism.
In these cases, the exchange energy favors antiparallel spins so that there is no net
magnetism. This negative ordering energy also resists paramagnetism until this energy is
destroyed by heating above the Néel temperature (sort of a negative Curie temperature).
The Curie–Weiss law can be written as
C
x¼ ; T > TN : (25:36)
T þ TN
25.5.4 Ferrimagnetism
Ceramic systems involving ferromagnetic elements may exhibit ferrimagnetism. Ferrimag-
netism occurs when the magnetic moments of the different constituents oppose one
another and cause a partial cancellation of the total magnetization. A good example of
this phenomenon is Fe3O4 (magnetite, the original lodestone used for navigation by ancient
504 Introduction to the Physics and Chemistry of Materials
mariners). Fe3O4 is actually a mixture of FeO and Fe2O3. The structure of this (and similar
cubic ferrites denoted by MO(Fe2O3) is inverse spinel in which the O atoms form a face-
centered cubic lattice with the Fe3þ occupying both the octahedral and tetrahedral sites,
and the Fe2þ occupying only tetrahedral sites. The O atom shares one electron with an Fe3þ
atom and the other with an adjacent Fe3þ atom. Since the spin arrangement of Fe3þ is
""""", the added electron cannot go in " without violating the Pauli principle, so it must go
in as """""#. But the other electron from the O atom is ", so it must go into the other Fe3þ
atom as "#####. Thus the magnetic moments of the Fe3þ ions all cancel out, leaving just the
Fe2þ electrons to contribute to the permanent magnetic moment.
Even though they do not form as strong a magnet as their ferromagnetic cousins, the
ferrites, as this class of ceramics is called, are very important technologically because they
are also insulators. Because they do not conduct electricity, they may be used as cores for
transformers or choke coils in high frequency radio or microwave applications where
losses from induced eddy currents must be avoided.
FIGURE 25.3
Magnetic domains in a demagnetized state (a). The
domains are arranged so that the magnetic field is
internal to the material in order to minimize the mag-
netic energy. When an external field is applied (b), the
domains that are favorably oriented grow at the expense
of those less favorably oriented by moving the Bloch
walls. (a) (b)
Magnetism and Magnetic Materials 505
The number of domains that can form is limited by the fact that it also costs energy to form
the domain wall. This energy is a combination of the anisotropy energy and the exchange
energy. In Fe, it is easier to magnetize along the [100] direction than along the [110] or [111]
directions. Therefore, it would be energetically favorable to change the orientation to the
magnetization in as short a distance as possible to minimize the number of orientations away
from the easy direction. However, a large change in orientation costs exchange energy since
the exchange force wants to keep the electron spins aligned. The thickness of the domain
wall is an optimization of the two competing energies and is generally O(0.1 mm). The
number of domains is a balance between the bulk magnetic energy saved by subdividing the
material into smaller domains and the energy cost of the additional wall required.
m0H Msat
B rem
Hcoer
H
(a)
m0H Msat
B rem
H coer
H
(b)
FIGURE 25.4
Examples of magnetic hysteresis loop for (a) a hard magnetic material and (b) a soft magnetic material. The flux
density remaining when the applied field is removed is the remanence, Brem. The reverse applied field required to
destroy the magnetism is the coercive strength, Hcoer. The area enclosed by the hysteresis loop represents energy
lost to heat during the magnetization–demagnetization cycle.
iron-neodymium-boron (Fe14Nd2B) magnets with a (BH)max ¼ 318 kJ=m3 for use in com-
pact brushless DC motors has found its way into the auto industry. These motors are used
for power windows and their smaller size made it possible to reduce the thickness of the
doors to give more interior room with the same exterior dimension. Their high efficiency
also makes them ideal motors for hybrid and electric drive vehicles.
Magnetic tapes require magnets with a high enough coercive strength to not be easily
demagnetized, but low enough for the recording head to easily change the state when
recording. Ferromagnetic CrO2 has largely replaced the ferrites used in earlier tapes
Magnetism and Magnetic Materials 507
because of its higher coercive strength. Saturated single domain grains are embedded in a
nonmagnetic binder and aligned with their long axes along the direction of the tape
motion. As they move under the recording head, their magnetization is altered according
to the amplitude of the signal to be recorded.
tensor of rank 4 for single crystals. However, for polycrystalline materials, the overall effect
can be averaged and the strain written as
@‘ 3l cos2 u 1=3
¼ , (25:37)
‘ 2
where
u is the angle between the direction of M and the direction the strain is measured in
l is the average of the saturated magnetostrictive constants l100 and l111 for the
material
This average may be positive or negative, meaning the material may contract in one
direction and expand in another, or the volume may contract or expand, depending on
the material. The 60 Hz hum associated with power transformers is a result of the
expansion and contraction of the laminations in the core.
The effect is small O(105) for the d-electron magnetic materials, but can be larger for
f-electron magnetic materials. The materials with the largest magnetostriction are the
Fe2RE alloys where RE ¼ Dy or Tb. However, it requires a large magnetic field to drive
the magnetostriction in these materials. Terfenol-D (TbxDy1xFe2) can produce a strain of
1600 mm=m at 2000 A=m and is the most commonly used material for magnetostrictive
transducers (see Clark, 1980).
Spontaneous magnetization below the Curie temperature is usually accompanied by an
increase in volume. Therefore, the magnetostriction that results as the spontaneous mag-
netization is destroyed by heating can be used to counteract the normal thermal expansion
and produce materials with a controlled coefficient of thermal expansion (CTE). Invar
(Fe64Ni36) has virtually zero CTE and is used for clock pendulums and other applications
where good dimensional stability is required. Fe58Ni42 matches the CTE of Si and FeNiCo
alloys (Kovar) can match the CTE of glasses. These Kovar alloys are used for glass–metal
seals in vacuum tubes.
Faraday rotation also occurs in a free electron gas where the amount of rotation depends
on the product of B and the electron density integrated along the path. This is a tool used in
astrophysics for estimating the magnetic field on the sun, stars, and in the interstellar
medium. It also provides a method of determining the electron density in the ionosphere.
25.9.2 Magnetoresistance
25.9.2.1 Ordinary and Anisotropic Magnetoresistance
A change on the resistance of a conductor in a magnetic field was first noticed by Lord
Kelvin in 1856. The effect is caused by the Lorentz force which deflects the trajectory of the
current carriers as in the Hall effect (see Section 18.5) and is maximized when the current
flow is perpendicular to the magnetic field. The effect is called ordinary magnetoresistance
(OMR) or ordinary magnetoresistance.
In a nonmagnetic material, the number of conduction electrons with spin up will on the
average equal the number of spin-down electrons. However, in a ferromagnetic material,
there is a small energy difference between the two spin states so that we will have a
majority of spins that align themselves with the internal magnetic field. The scattering
times tup and tdown will be different and there will be a coupling of the resistivity with the
magnetic field. Because of the anisotropy of the magnetism, the coupling will be in the
form of a tensor and is called anisotropic magneto resistance (AMR).
Even though the effect is small, magnetoresistance is used for sensing magnetic fields
and for read heads on magnetic hard drives.
To incorporate this technology into a read head, one of the magnets is a hard permanent
magnet and the other is a soft magnetic material such as Supermalloy. The information is
coded on the disc as domains that are N or S for logical 1 or 0. As the disc moves under the
read head, the spin valve opens and closes depending on the field induced in the soft
magnetic head.
25.10 Summary
All materials interact to some degree with a magnetic field. Some materials are weakly
repelled by a magnetic field and are said to be diamagnetic. They have a negative
susceptibility (a measure of the ability to be magnetized). Other materials are weakly
attracted by a magnetic field and are said to be paramagnetic. Metals that produce a
spontaneous magnetic field are said to be ferromagnetic, and ceramics that produce
a spontaneous magnetic field are said to be ferrimagnetic. The magnetism in ferro- and
ferrimagnetic materials decreases with temperature and is lost completely at their Curie
temperature, which is a property of the material. Above the Curie temperature, ferro- and
ferrimagnetic materials become paramagnetic. Some metals are antiferromagnetic, which
means that they are not magnetic until they reach their Néel temperature where they
become paramagnetic.
Magnetism and Magnetic Materials 511
Diamagnetism occurs through the interaction of the core electrons with an applied
magnetic field, which causes the electrons to produce an opposing magnetic field. Para-
magnetism occurs when there are unpaired electrons, which the applied field tries to align,
while thermal motion tends to randomize their alignment. As a result, paramagnetism
decreases as the reciprocal of the absolute temperature. Classically, the free electrons in a
metal should exhibit strong paramagnetism, but the only electrons that are free to partici-
pate are those near the Fermi level. Both diamagnetism and paramagnetism are weak
effects. Diamagnetism occurs in all metals, but in some metals the paramagnetism is
greater than the diamagnetism and prevails.
Ferromagnetism occurs in the first row of the transition metals with unpaired 3d-electrons
(Fe, Co, Ni) and in the first row of the rare earths with unpaired f-electrons (Dy, Gd) when the
amount of overlap between the wavefunctions of the d- or f-electrons falls in a certain critical
range. A quantum mechanical exchange force between neighboring atoms keeps their
electronic spins aligned until the thermal energy becomes high enough at the Curie tem-
perature to destroy the alignment. If the overlap is greater, the exchange forces cause the
spins in the neighboring atoms to become antiparallel and the material is antiferromagnetic
(Cr, Mn). Ferrites are compounds of divalent metal oxides and ferromagnetic Fe2O3. In order
to form these compounds, some of the ferromagnetic atoms must go into the molecule with
their spins antiparallel and the material is said to be ferrimagnetic. Ferrimagnetic materials
behave like ferromagnetic materials, but are generally weaker magnetically because of the
partial cancellation of the ferromagnetic spins. However, since they are dielectrics and do
not conduct electricity, ferrites are suitable for tasks that ferromagnetics cannot do.
A ferromagnetic material is divided into a bunch of small regions called domains where
the spins are in the same direction within the domain, but may be in a different direction in
an adjacent domain. Domains are separated from each other by Bloch walls where the
electrons gradually change the directions of their spins. If the spins of the domains are in
random directions, there is no net magnetic moment. The material may be magnetized by
applying a strong magnetic H-field. Domains that are aligned with the field will grow at
the expense of the others by moving the domain walls. When all the domains are aligned
with the field, the material is saturated. When the field is removed, the remaining mag-
netism Brem is called the remanence. The Hcoer field required to demagnetize the material is
the coercivity. A plot of the magnetization–demagnetization cycle is the hysteresis. The
strength of a magnet is characterized by its energy product BremHcoer or by the area of the
largest B–H rectangle that can be drawn under the demagnetization curve.
Hard magnetic materials (materials with high energy products) are used for permanent
magnets. Materials can be made magnetically hard in much the same manner as metals are
hardened (e.g., alloying, precipitation hardening, and grain refining). If the grains are small
enough for only one domain to form, the demagnetizing field must be strong enough to
flip the spins in the domain rather than just move the domain wall. Soft magnetic materials
have low coercive strength and small BH product in order to minimize the energy loss in
the hysteresis loop. Materials are made magnetically soft by texturing the grains to allow
the Bloch walls to move easily in preferred directions or by using amorphous metals to
eliminate the grain boundaries.
Bibliography
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia, 1976.
Baibich, M.N., et al., Giant magnetoresistance of (001)Fe=(001)Cr magnetic superlattices, Phys. Rev.
Lett., 61=21, 1988, 2472–2475.
512 Introduction to the Physics and Chemistry of Materials
Binasch, G., et al., Enhanced magnetoresistance in layered magnetic structures with antiferromag-
netic interlayer exchange, Phys. Rev. B., 39=7, 1989, 4828–4830.
Christman, J.R., Fundamentals of Solid State Physics, John Wiley & Sons, 1988.
Clark, A.E., Ferromagnetic Materials, Vol. 1, ed. Wolfhart, E.P. North Holland, Amsterdam.
Das Sarma, S., Spintronics, Am. Sci., 89, 2001, 516–523.
Earnshaw, W., On the nature of the molecular forces which regulate the constitution of the luminif-
erous ether, Trans. Camb. Phil. Soc., 7, 1842, 97–112.
Gersten, J.I. and Smith, F.W., The Physics and Chemistry of Materials, John Wiley & Sons, 2001.
Hummel, R.E., Electronic Properties of Materials, 3rd edn., Springer-Verlag, New York, 2000.
Kittel, Introduction to Solid State Physics, John Wiley & Sons, New York, 1966.
Srivastava, C.M. and Srinivasan, C., Science of Engineering Materials, Wiley Eastern Ltd. New Delhi,
India, 1987.
Wohlfart, E.P., Ed., Ferromagnetic materials, in Handbook of Magnetic Materials, Vol. 1, North-Holland,
Amsterdam, 1980.
Problems
1. Show that (ZnOFe2O3)x(MnOFe2O3)1x has a magnetic moment of 7.5mB for x ¼ (1=2).
2. Would it be possible to directly measure the angular momentum of an electron? Do a
feasibility analysis for such an experiment. Assume you have a 1 g piece of Fe that is
magnetized to saturation that you can levitate in an electrostatic levitator. While the Fe
is levitated, heat the sample to above the Curie temperature. What will be the angular
momentum of the sample?
3. What geometry would you use for the sample in the above problem to maximize the
observed spin of the sample? What would be the expected angular rate? Would this be
observable? Would the ability to levitate a larger sample help?
4. You have two 1 cm cubic magnets with a B ¼ 1 T. What force would be required to pull
them apart? (Hint: use the principle of virtual work.)
5. Construct a model for the local magnetic field in a magnetic material similar to the
Clausius–Mossotti model for the local electric field in a dipolar field. Can such a model
explain the ordering in a ferromagnetic material? (Compare the predicted Weiss con-
stant against the observed Weiss constant for Fe.)
26
Superconductivity
513
514 Introduction to the Physics and Chemistry of Materials
TABLE 26.1
Elemental Superconductors
Group IVB Group VB Group VIB Group VIIB Group VIII Group VIII
Ti (0.39) V (5.39)
Zr (0.55) Nb (9.64) Mo (0.92) Tc (7.77) Ru (0.51)
Hf (0.12) Ta (4.48) W (0.01) Re (1.4) Re (0.66) Ir (0.14)
Group IIB Group IIIA Group IVA
Al (1.14)
Zn (0.88) Ga (1.09)
Cd (0.56) In (3.40) Sn (3.72)
Hg (4.15) Tl (2.39) Pb (7.19)
Source: Data from Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia,
1976.
These currents, once set in motion, persist for as long as the temperature is maintained
below the TC. Measurement of the magnetic moment from these circulating currents is one
of the ways of determining that the material is in a superconducting state. The persistent
currents are not just an effect of zero resistivity, but constitute a macroscopic quantum
state. Just as electrons, because they are in definite quantum states, can circulate in atoms
without radiating energy and can only transition to other allowed quantum states, one can
think of circulating supercurrents as being in some particular quantum state, in which they
will remain until they transition to another allowed state.
TABLE 26.2
Bandgaps and Transition Temperatures for Elemental
Superconductors
Element 2D0 (meV) TC 2D0=kTC
for different systems. (We will see later that the BCS theory predicts that the binding
energy of the Cooper pair is related to the transition temperature by 2D0 ¼ 3.53kTC.)
where
vD is the Debye frequency
U is the attractive potential between the two electrons that make up the Cooper pair
N(E) is the density of states at the Fermi level
4
2D0 ¼ kTC ¼ 3:51kTC : (26:4)
1:14
Superconductivity 517
m0 HC0
2
2N ðEÞðhvD Þ2 N ðEÞD0
GN GSC ¼ ¼ ð2=UN ðEÞÞ : (26:5)
2 e 1 2
The spike in the electronic heat spike at TC is given by
From Equation 26.3 the TC depends linearly on the Debye frequency, which predicts the
isotope effect since the Debye frequency is proportional to the inverse square root of
the mass of the vibrating atoms. We shall examine some of the other predictions in the
following section.
BðC ðC
H
m0 2
W¼ M dB ¼ m0 H dH ¼ H : (26:8)
2 C
0 0
The free energy of the superconducting state as a function of applied field can be written as
m0
GSC ðH,T Þ ¼ GSC (0,T) þ H ðT Þ2 : (26:9)
2
Since the free energy of the normal state is unaffected by the presence of a magnetic field
m0 2
GN ðH,T Þ ¼ GN ð0,T Þ ¼ GSC (0,T) þ H : (26:10)
2 C0
The difference in free energy between the normal and superconductive states is given by
subtracting Equation 26.9 from Equation 26.10
m0 2 m m
DGðH,TÞ ¼ H 0 H ðT Þ 2 ¼ 0 H C ðT Þ2 , (26:11)
2 C0 2 2
where HC ðT Þ2 ¼ HC02
H ðTÞ2 is the critical field required to destroy the supercon-
ducting state.
At the temperatures of interest, the heat capacity of the normal state is primarily due to
the electronic contribution gT. The entropy is therefore given by gT and the free energy is
518 Introduction to the Physics and Chemistry of Materials
Normal
−0.1
−0.2
G(T )/g T C2
Superconducting
−0.3
given by gT2=2. The free energies of the normal and superconducting state are shown in
Figure 26.1. Note that the free energies of the two states merge at T ¼ TC with identical
slopes, indicating a second-order phase transition.
Destroying the superconductivity by heating, i.e., raising the temperature above the
transition temperature, constitutes a second-order phase transition. There is no change in
the slope of the free energy, so the entropy is a continuous function as seen in Figure 26.3
and there is no latent heat involved. There will be a discontinuity in heat capacity, as will
be shown. Conversely, destroying the superconductivity by increasing the magnetic field
above the HC has the effect of raising the superconducting free energy curve until it crosses
that of the normal state. Now there will be a change in slopes of the free energy and the
phase transition would be first order.
The 2D0 represents the energy gap centered on the Fermi level and the 2kT is to account for
the two electrons forming the Cooper pair. The BCS theory (Equation 26.4) predicts that
this energy gap, 2D0 ¼ 3.51kTC and this relationship appears to be satisfied by a wide range
of superconductors. The constant can be determined by integrating CSC(T)=T from 0 to TC
and equating this to the entropy at TC which is gTC.
TðC ð1
CSC (T) e1:76=x
SSC ðTC Þ ¼ dT ¼ constant dx ¼ gTC : (26:13)
T x
0 0
The integral has the form of an exponential integral (Ei) and the constant thus becomes
gTC gTC
Constant ¼ ¼ : (26:14)
Ei(1:76) 0:067
The heat capacity may then be related to the critical temperature by
gTC 1:76
CSC (TC ) ¼ e ¼ 2:511gTC (26:15)
0:067
and the DC by
DC(TC ) ¼ 2:511gTC gTC ¼ 1:511gTC : (26:16)
This result shown in Figure 26.2 is close to but not identical to the DC ¼ 1.43 gTC (Equation
26.6) predicted by the BCS theory.
The entropy of the superconducting state can now be calculated from
ðT ðC
T=T
CSC (T) gTC2 e1:76=x
SSC ðT Þ ¼ dT ¼ dx (26:17)
T Ei(1:76) x
0 0
ð0 ð1 ðu 1:76=x
gTC2 gTC2 e
GSC ð0Þ ¼ GN (TC ) SSC (T)dT ¼ þ du dx ¼ 0:251gTC2 : (26:18)
2 Ei(1:76) x
TC 0 0
2 Superconducting
ΔC(TC)
C(T)/gT C
1.0
0.8
0.6
S(T)/gT C
al
m
or
0.4 N
g
c tin
ndu
0.2 co
r
pe
FIGURE 26.3 Su
The entropy in the superconducting state com- 0
puted from the model is compared to the entropy 0 0.2 0.4 0.6 0.8 1.0
in the normal state. The two must be the
T/T C
same at TC.
Equating this result to the work required to destroy the superconducting state at T ¼ 0
from Equation 26.8, we obtain a relation between the critical field and the transition
temperature at 0 K.
m0 HC0
2
¼ 0:251gTC2 , gTC2 ¼ 1:99m0 HC0
2
: (26:19)
2
The free energy at T between 0 and TC is obtained by integrating the entropy from 0 to T
and subtracting it from GSC(0),
ðT ðC
T=T ðu
gTC2 e1:76=x
GSC ðTÞ ¼ GSC (0) SSC (T)dT ¼ 0:251gTC2 du dx: (26:20)
Ei(1:76) x
0 0 0
Since the free energy of the electron gas in the normal state is gT2=2, the DG is
ðC
T=T ðu
gTC2 e1:76=x
DGSC ðT Þ ¼ GN (T) GSC (T) ¼ 0:5gT þ 2
0:251gTC2 þ du dx: (26:21)
Ei(1:76) x
0 0
The ratio of the critical magnetic field required to destroy the superconductive state at
temperature T to the critical field at T ¼ 0 is obtained from
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
HC (T) DGSC (T)
¼ : (26:22)
HC (0) DGSC (0)
The results are plotted in Figure 26.4 along with the Tuyn approximation, 1T2=TC2 , and
two asymptotic solutions suggested by the BCS theory.
Other predictions can be obtained from this thermodynamically consistent model. We
have already shown that CSC(TC) ¼ 2.511gTC (Equation 26.15) and that gTC2 ¼ 1.99 m0HC0 2
1.0
0.8
0.6
H C(T)/H C(0)
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1.0
T/T C
FIGURE 26.4
The ratio of the critical magnetic field at temperature T to the critical magnetic field at T ¼ 0 as a function of the
ratio of T=TC. The heavy line is the model prediction, the thin solid line slightly above it is the Tuyn approxima-
tion. The dashed line that falls below as T approaches TC is the low temperature BCS solution (1–1.06 T2=TC2 ), and
the straight dashed line above the curve is the BCS solution near TC (1.74 (1 T=TC)).
or
We now compare our model with data from various type-1 superconductors in Table 26.3.
TABLE 26.3
Critical Parameters for Elemental Superconductors
m0HC0
2
DCSC(TC) DCSC(TC)=
Element HC0 (Oe) TC (8K) g (mJ=mol K2) (mJ=mol) gTC2=m0 HC0
2
(mJ=mol K) gTC
The magnetic data was given in Oersteds, which are converted to ampere per meter by
multiplying by 79.6. The ratio gTC2 =m0HC02
ranges from 1.55 to 2.13 with an average value of
1.91, which is very close to the 1.99 predicted by Equation 26.19. Lacking data for DCSC(TC),
this value was calculated from Equation 26.24 and then used to calculate DCSC(TC)=gTC.
We see this ratio ranges from 1.41 to 2.11 with an average value of 1.61. This is reasonably
close to the model prediction of 1.51, which is closer to the experimental value than the BCS
prediction of 1.43 (Equation 26.6).
n eE
n_ þ ¼ , (26:25)
t m
d ns e2 E
J ¼ ens v_ ¼ , (26:26)
dt m
Equation 26.28 implies that either the term in the brackets is a constant or 0. For a
superconductor above the transition temperature, B can penetrate the material even
though jS ¼ 0 because nS ¼ 0. As the material is cooled below TC, jS 6¼ 0 and nS 6¼ 0;
therefore, B ¼ 0, which is required by the Meissner effect. We therefore assume the term
is zero because it is consistent with the observed behavior of superconductors and the
persistent supercurrents are given by
ns e2
r Js þ B ¼ 0, (26:29)
m
which is the second London equation.
r r B ¼ m 0 r Js (26:30)
m0 ns e2
r r B ¼ r(r B) r2 B ¼ B: (26:31)
m
Assume B is along the x-direction, parallel to the surface of the superconductor and z is
perpendicular to the surface directed inward. Since r B ¼ 0, the Equation 26.31 can be
written
d2 Bx m0 ns e2
Bx ¼ 0, (26:32)
dz2 m
which has the solution Bx(z) ¼ Bx(0)ez=lL, where lL is the London penetration depth
given by
1=2
m
lL ¼ : (26:33)
m0 ns e2
then
mDE
dk ¼ : (26:35)
2 k F
h
since DE ¼ 2D0. Pippard obtained a more precise result from BCS theory which is called the
Pippard coherence length given by
2hnF
j0 ¼ 1 mm (26:37)
pD0 e
for pure materials. The estimate from the uncertainty principle is remarkably close to
Pippard’s result.
524 Introduction to the Physics and Chemistry of Materials
In other words, it takes very little field to destroy the superconducting state as T
approaches TC. This is one of the reasons that superconducting systems must be operated
well below their transition temperature.
For alloy-type or impure systems (which are usually type-II superconductors), there are
two critical fields, HC1 and HC2. Like type-I superconductors, type-II superconductors
exhibit perfect diamagnetism up to HC1. The field starts to penetrate above HC1, but
superconductivity remains until HC2 is reached. Between HC1 and HC2, the system is in a
mixed state. Lines of magnetic flux start to penetrate in regions that have become normal
while the rest of the material remains superconducting. The bulk resistivity is still zero
because the current is carried by the superconducting regions.
For example, HC for pure Pb is 600 Gauss at 4 K. For Pb2W%In, HC1 ¼ 400 Gauss and HC2
3600 Gauss. The area under the M versus H curve is the same for both systems as
illustrated in Figure 26.5, but a type-II system can be operated at much higher fields.
Therefore, all practical superconductors will be type-II.
The factor that determines whether a material will by a type-I or type-II superconductor
is the ratio of the London penetration depth to the coherence length. If l=j < 0.71,
the material will be type-I; if l=j > 0.71, it will be a type-II. (The reason for this will be
discussed later.) Both penetration depth and coherence length are influenced by the
addition of impurities which disrupt the regularity of the lattice. The penetration depth
increases according to
1=2
j0
l ¼ lL þ1 (26:39)
‘
−M −M
HC HC H C2
(a) (b)
FIGURE 26.5
Schematic of the behavior of a type I superconductor (a) and a type-II superconductor (b). The type I exhibits
perfect diamagnetic behavior (M ¼ H) until HC1 where the M goes abruptly to zero. A type II semiconductor is
perfectly diamagnetic until HC1 at which point the magnetic field starts to penetrate. However the material
remains superconductive until the second critical field HC2 is reached. The area under the two curves is the same.
Superconductivity 525
1 1 1
þ , (26:40)
j j0 ‘
where ‘ is the mean free path between impurity atoms. Thus it is seen that pure materials
tend to be type-I superconductors, but in alloys, ‘ j0; hence they tend to be type-II
superconductors.
nP e
h 2nP e2
J¼ rf A: (26:42)
m m
(Note: If we take the curl of J in Equation 26.19, since r rf ¼ 0, we get r J ¼
(2nP e2 =m)r A ¼ (nS e2 =m) B, which is the second London equation (Equation 26.29).)
We integrate inside of a superconducting ring containing magnetic flux F at a sufficient
depth where J ¼ 0,
þ þ þ
nP e @f(L)
J dL ¼ h 2e A dL ¼ 0: (26:43)
m @L
h
F0 ¼ ¼ 2:06 1015 Tm2 : (26:44)
2e
2 kF dkc=m
h
2dE ¼ 2 ¼ 2D: (26:45)
m
So the critical current density at which the Cooper pair will break up is found from
ns eh ns eh mD ns eD
JC ¼ dkc=m ¼ ¼ , (26:46)
m m h2 kF hkF
where ns is the number of superconducting electrons in the Cooper pairs. Notice that D was
used here instead of D0, the binding energy per electron at T ¼ 0 K. The quantity D
diminishes with temperature and with the magnetic field as will be shown, and therefore,
so does JC.
Assume a wire with radius r. The field at the surface of the wire is
þ
H d‘ ¼ 2prH: (26:48)
1=2
HC m ns e2
JC ¼ ¼ HC 0 for l r: (26:49)
l m
We see that the limiting current is not affected by the wire size provided r
l. If the wire
with radius r < l is carrying current J, the critical field at the surface will be reached when
JC ¼ 2HC=r. In this case, the critical current density increases inversely with r, but since the
pjwstk|402064|1435428897
B0 F0 h
HC1 ¼ ¼ ¼ : (26:51)
m0 2pm0 l2 4pem0 l2
As the applied flux density increases, more flux vortices are created and their diameter
shrinks as they crowd together. Eventually their diameter shrinks to approximately the
coherence length and the supercurrents can no longer repel the field. The HC2 occurs when
the flux vortices fill up all of the available space as illustrated in Figure 26.6, which occurs
when the flux density equals F0=pj2,
F0 h
HC2 ¼ ¼ : (26:52)
pm0 j 2
2pem0 j2
Given these results, it is easy to see why the transition from type-I to type-II takes place
when l=j ¼ 0.71. In a type-I superconductor, the HC1 ¼ HC2 and this occurs when l2 ¼ j2=2.
Φ0 Φ0
N SC N
2½l
(a)
Φ0 Φ0 Φ0 Φ0 Φ0
(b)
FIGURE 26.6
Fluxoids at H just above HC1 (a) and just below HC2 (b). When the flux density reaches the point where each
quanta occupies a circle with radius ¼ j, there is no superconducting regions left and the material becomes normal.
528 Introduction to the Physics and Chemistry of Materials
TABLE 26.4
Layered Structure of the Various Ceramic Superconductors
Bi3þO2 Hg2þ
2 3þ 2þ 2 2 2þ
2O La Cu O O Sr O2Ba2þ
La3þO2þ O2Ba2þ Cu2þ2O2 Cu2þ2O2
4þ 2 3þ 2þ 2 2þ
Ti 2O O La Cu 2O Ca Ca2þ
2þ 2 2þ 2 3þ 2þ 2
Ba O Cu 2O Y Cu 2O Cu2þ2O2
4þ 2 3þ 2þ 2 2þ
Ti 2O O La Cu 2O Ca Ca2þ
La3þO2þ O2Ba2þ Cu2þ2O2 Cu2þ2O2
2 3þ 2þ 2 2 2þ
2O La Cu O O Sr O2Ba2þ
Bi3þO2 Hg2þ
BaTiO3 La2CuO4 YBa2Cu3O7 Bi2Sr2Ca2Cu3O10 HgBa2Ca2Cu3O8
Insulating TC ¼ 38 K TC ¼ 92 K TC ¼ 110 K TC ¼ 135 K
Superconductivity 529
the O7 meaning that on the average there are fewer than seven oxygen atoms in the unit
cell. The properties of the material can vary from nonconducting, to semiconducting, and
to superconducting, depending on the number of oxygen defects.
A major problem with the high TC materials is the fact that they are superconducting
only in the a–b planes. The initial preparations of YBa2Cu3O7 had very low critical
currents because of the weak coupling between the CuO2 planes. Techniques have since
then been developed to texture the material so that the c axes are aligned, which dramat-
ically improved the critical current density. Flux creep caused by the interaction of the
magnetic field with the fluxons (tubes of magnetic flux that penetrate type-II supercon-
ductors) was also a problem in these materials. One solution to the flux creep problem has
been to precipitate out a nonconductive phase such as the 211 phase in the case of
YBa2Cu3O7. The nonconducting 211 phase acts as pinning sites to keep the fluxons
from moving in the presence of a high magnetic field.
Permanent magnets using high TC materials with flux densities of several tesla are
currently in use serving as superconducting bearings and other functions such a maglev
trains. However, high TC materials have found little use in transmission lines and in large
bore superconducting magnets such as those used in MRI systems and similar applications
because of the difficulties in forming wires from these very hard and brittle ceramic
materials.
26.12 Applications
By far, the largest use of superconductors remains in superconducting magnets using the
Nb46.5W%Ti alloy although Nd3Sn is finding its way into these applications because of its
higher critical currents at LHe temperatures. Application for the high temperature super-
conducting ceramic systems have been slow to emerge because of their brittleness and
difficulties in forming them into cables with high current carrying capabilities. There are,
however, some rather exotic applications of superconductors based on their tunneling
properties, which are described in the following sections.
530 Introduction to the Physics and Chemistry of Materials
A
FIGURE 26.7
JJ formed by depositing a thin stripe of
metal on an insulator, coating this stripe
V
with a thin oxide insulating layer, and
then adding a second metallic stripe.
eh
J0 ¼ ns , (26:53)
ma
qV
FIGURE 26.8
Energy levels in the normal and super-
conducting state with no voltage applied
(left). When a potential qV is applied,
raising the Fermi level of the normal
metal to that top of the energy gap in Normal Superconductor Normal Superconductor
the superconductor, current will begin to
flow (right). (a) (b)
Superconductivity 531
Normal
Current
current
J0
FIGURE 26.9
Current–voltage for the DC Josephson effect. The supercur-
qΔ 0 rent can exist with no voltage drop until it reaches J0, at
Voltage which point the supercurrent switches to the normal current.
2eV
v¼ : (26:54)
h
For V ¼ 1 mV and v 3 109 rad=s or 500 MHz. This equation says that a photon of
energy hv ¼ 2eV is emitted every time a Cooper pair crosses the barrier. Since voltage and
frequency can be measured with a high degree of precision, this effect provides a very
precise measurement of e= h. Conversely, the NIST definition of a volt is now based on
frequency measurements and Equation 26.54.
I I
FIGURE 26.10
A superconducting quantum interference
detector formed by two parallel Josephson
junctions.
532 Introduction to the Physics and Chemistry of Materials
They are also used in magnetic resonance imaging (MRI), for geomagnetic mapping, and
for mineral prospecting. John Lipa and his team at Stanford University used SQUIDs to
measure temperature-induced changes in the magnetic moment of a paramagnetic salt in
order to build a thermometer that was capable of detecting temperature changes of 1 nK.
(This device was the heart of the Lambda point experiment that was carried out on a space
shuttle mission in 1992.)
SQUIDs are also being used in the Gravity Probe-B satellite to detect minute changes of
less than 0.5 milliarcseconds in the orientation of the superconducting gyroscope by
measuring its London moment. The London moment is the magnetic moment that is
generated by a rotating superconductor. The Gravity Probe-B experiment is a critical test
of Einstein’s general theory of relativity.
26.13 Summary
Superconductors are materials that conduct electricity with no resistance at temperatures
below their critical temperatures. A number of metallic elements, alloys, intermetallic
compounds, and ceramics exhibit this property, although few have transition temperatures
high enough for practical applications. The maximum current that can be carried by a
superconductor is limited by the magnetic field associated with the current that tends to
destroy the superconducting state. The critical magnetic field and thus the critical current
increase as the temperature is lowered, thus it is necessary to operate superconductors well
below their transition temperatures in most applications. Until the discovery of high
temperature ceramic superconductors in 1986, it was necessary to cool superconductors
with LHe, which is expensive, hard to keep, and nonrenewable. For this reason much effort
has gone into the search for materials with higher transition temperatures so they may be
cooled with less expensive fluids such as LN2.
No magnetic fields can exist in a superconductor. Any field that may have existed in the
normal state will be ejected when the material is cooled into the superconducting state by
supercurrents induced at the surface of the superconductor. The ejection of any magnetic
field is known as the Meissner effect. The supercurrents at the surface of the supercon-
ductor are persistent and will flow indefinitely. The superconducting state can be likened
to a macroscopic quantum state which is separated from the normal state by a small energy
gap 2D that is D below the Fermi level. Just as electrons in a atom may circulate indefinitely
in their ground state until they are kicked into a higher quantum state by absorbing energy,
supercurrents may flow in a superconductor until they are kicked into the normal state by
absorbing energy.
The theory of superconductivity was developed by Bardeen, Cooper, and Schrieffer in
1957 and is known as the BCS theory. Electrons interacting with the lattice ions form
loosely bound pairs known as Cooper pairs that can move through the lattice without
interference so long as the temperature and external magnetic field are low enough so as
exceed the binding energy of the pairs.
There are two types of superconductors. Type-I superconductors are perfectly diamag-
netic, meaning they reject the penetration of a magnetic field until they reach their critical
magnetic field, HC, where the superconductivity is destroyed. Type-II superconductors
behave like type-I superconductors up to their first critical field, HC1. Then they become a
Superconductivity 533
mixed state in which there are holes of normal material that have been penetrated by
magnetic flux lines called fluxoids or flux vortices within the superconducting material.
As the applied magnetic field increases, the number of these flux vortices increases until
eventually at HC2 there is no superconducting material left. Since type-II superconductors
can operate at much higher fields than type-I superconductors, they are used in most
practical applications.
Whether a material is type-I or type-II is determined by the ratio of the penetration depth l
(the distance into the surface a magnetic field can penetrate) to the coherencepffiffilength ffi j
(essentially the diameter of the Cooper pair). Type-II behavior occurs when l > 2j. Most
pure materials are type-I, but adding impurities such as alloying with another metal increases
l and decreases j, so most practical superconductors are alloys or intermetallic compounds.
The major application of superconductors is their use in superconducting magnets used
in medical imaging, nuclear magnetic resonance (NMR), particle accelerators, crystal
growth, and other industrial processes. Despite their great promise, the ceramic HTSCs
have been slow to find commercial applications, although they are finding limited use in
bulk permanent magnets for frictionless rotating bearings used for flywheel energy storage
systems and for maglev systems. Progress is being made in fabricating superconducting
tapes and cables for power transfer and demonstration projects are underway for their
use in the power grid. However, because of the high cost of fabricating these power
cables, their wide-spread use is presently not commercially viable. Other, more exotic
applications of superconductivity, lie in the ability to detect a single quantum of magnetic
flux using superconducting quantum interference detectors (SQUIDs). Such highly sensi-
tive devices open new possibilities in biomedicine, magnetic imaging, nondestructive
testing, prospecting, and surveillance.
Bibliography
Ashcroft, N.W. and Mermin, N.D., Solid State Physics, Brooks Cole, Philadelphia, PA, 1976.
Bardeen, J., Cooper, L.N., and Schrieffer, J.R., Theory of superconductivity, Phy. Rev., 108, 1175–1204,
1957.
Bednorz, J.G. and Müller, K.A., Possible high Tc superconductivity in the Ba–La–Cu–O system. Z.
Physik., B64, 189–193, 1986.
Christman, J.R., Fundamentals of Solid State Physics, John Wiley & Sons, New York, 1988.
Fujimoto, H., Developing a high-temperature superconducting bulk magnet for the Maglev Train of
the future, JOM, 50=10, 16–18, 1998.
Gersten, J.I. and Smith, F.W., The Physics and Chemistry of Materials, John Wiley & Sons, New York,
2001.
Ibach, H. and Lüth, H., Solid State Physics, 3rd edn., Springer-Verlag, New York, 1990.
Kittel, C., Introduction to Solid State Physics, 7th edn., John Wiley & Sons, New York, 1966.
Nagamatsu, J., et al., Superconductivity at 39 K in MgB2, Nature, 410, 63, 2001.
Fujimoto, H., Owens, F.J., and Poole, C.P. Jr., Progress in Superconductivity, The New Superconductors,
Springer US, 2002.
Post, R.F., New look at an old idea, the electromechanical battery, Sci. Technol. Rev., April 1996.
Tanaka, S., High-temperature superconductivity: Outlook and history, JSAP Int., 4, 17–22, 2001.
Wu, M.K., Ashburn, J.R., and Torng, C.J., Superconductivity at 93 K in a new mixed-phase
Y-Ba-Cu-O compound system at ambient pressure. Phys. Rev. Lett. 58(9), 908–910, 1987.
534 Introduction to the Physics and Chemistry of Materials
Problems
1. Equation 26.45 gives the critical current in terms of the binding energy of the Cooper
pair
ns eD
JC ¼ :
hkF
Equation 26.49 gives critical current in terms of magnetic field at the surface of a
cylindrical conductor
1=2
m0 ns e2
Jc ¼ H c :
m
PHYSICS and
CHEMISTRY
of MATERIALS
Providing the foundation needed for more advanced work in materials
science, Introduction to the Physics and Chemistry of Materials
discusses the structure and properties of materials and how these
materials are used in diverse applications. The text covers chemical
bonding, crystal structure, mechanical properties, phase transformations,
and materials processing. It also focuses on thermal, electronic,
photonic, optical, and magnetic properties of materials.
FEATURES
• Presents basic concepts of solid state chemistry, such as chemical
bonding and diffraction
• Relates mechanical, thermal, electronic, photonic, and magnetic
properties to the structure of organic and inorganic materials
• Introduces point and space groups as well as Strukturbericht
and Pearson notation
• Explores contemporary developments in photonics and
magnetoelectronics, including polymer light-emitting diodes,
flash memories, and magnetic storage
• Provides examples to evaluate how well simplified theories
predict the actual performance of various materials