Sie sind auf Seite 1von 56

Introduction to Materials

This section will provide a basic introduction to materials and material fabrication
processing. It is important that NDT personnel have some background in material
science for a couple of reasons. First, nondestructive testing almost always
involves the interaction of energy of some type (mechanics, sound, electricity,
magnetism or radiation) with a material. To understand how energy interacts with
a material, it is necessary to know a little about the material. Secondly, NDT often
involves detecting manufacturing defects and service induced damage and,
therefore, it is necessary to understand how defects and damage occur.
This section will begin with an introduction to the four common types of
engineering materials. The structure of materials at the atomic level will then be
considered, along with some atomic level features that give materials their
characteristic properties. Some of the properties that are important for the
structural performance of a material and methods for modifying these properties
will also be covered.
In the second half of this text, methods used to shape and form materials into
useful shapes will be discussed. Some of the defects that can occur during the
manufacturing process, as well as service induced damage will be highlighted. This
section will conclude with a summary of the role that NDT plays in ensuring the
structural integrity of a component.
General Material Classifications
There are thousands of materials available for use
in engineering applications. Most materials fall into
one of three classes that are based on the atomic
bonding forces of a particular material. These three
classifications are metallic, ceramic and polymeric.
Additionally, different materials can be combined
to create a composite material. Within each of these
classifications, materials are often further organized
into groups based on their chemical composition or
certain physical or mechanical properties. Composite materials are often grouped
by the types of materials combined or the way the materials are arranged together.
Below is a list of some of the commonly classification of materials within these
four general groups of materials.
Metals
Ferrous metals and alloys (irons,
carbon steels, alloy steels,
stainless steels, tool and die
Polymeric
Thermoplastics plastics
Thermoset plastics
steels)
Nonferrous metals and alloys
(aluminum, copper, magnesium,
nickel, titanium, precious
metals, refractory metals,
superalloys)
Elastomers




Ceramics
Glasses
Glass ceramics
Graphite
Diamond

Composites
Reinforced plastics
Metal-matrix composites
Ceramic-matrix
composites
Sandwich structures
Concrete
Each of these general groups will be discussed in more detail in the following
pages.
Metals
Metals account for about two thirds of all the elements and about 24% of the mass
of the planet. Metals have useful properties including strength, ductility, high
melting points, thermal and electrical conductivity, and toughness. From the
periodic table, it can be seen that a large number of the elements are classified as
being a metal. A few of the common metals and their typical uses are presented
below.
Common Metallic Materials
Iron/Steel - Steel alloys are used for strength critical applications
Aluminum - Aluminum and its alloys are used because they are easy to
form, readily available, inexpensive, and recyclable.
Copper - Copper and copper alloys have a number of properties that make
them useful, including high electrical and thermal conductivity, high
ductility, and good corrosion resistance.
Titanium - Titanium alloys are used for strength in higher temperature
(~1000 F) application, when component weight is a concern, or when good
corrosion resistance is required
Nickel - Nickel alloys are used for still higher temperatures (~1500-2000 F)
applications or when good corrosion resistance is required.
Refractory materials are used for the highest temperature (> 2000 F)
applications.

The key feature that distinguishes metals from non-metals is their bonding.
Metallic materials have free electrons that are free to move easily from one atom to
the next. The existence of these free electrons has a number of profound
consequences for the properties of metallic materials. For example, metallic
materials tend to be good electrical conductors because the free electrons can move
around within the metal so freely. More on the structure of metals will be discussed
later.

Ceramics
A ceramic has traditionally been defined as an inorganic, nonmetallic solid that is
prepared from powdered materials, is fabricated into products through the
application of heat, and displays such characteristic properties as hardness,
strength, low electrical conductivity, and brittleness." The word ceramic comes the
from Greek word "keramikos", which means "pottery." They are typically
crystalline in nature and are compounds formed between metallic and nonmetallic
elements such as aluminum and oxygen (alumina-Al
2
O
3
), calcium and oxygen
(calcia - CaO), and silicon and nitrogen (silicon nitride-Si
3
N
4
).
Depending on their method of formation,
ceramics can be dense or lightweight. Typically,
they will demonstrate excellent strength and
hardness properties; however, they are often
brittle in nature. Ceramics can also be formed to
serve as electrically conductive materials or
insulators. Some ceramics, like superconductors,
also display magnetic properties. They are also more resistant to high temperatures
and harsh environments than metals and polymers. Due to ceramic materials wide
range of properties, they are used for a multitude of applications.
The broad categories or segments that make up the ceramic industry can be
classified as:
Structural clay products (brick, sewer pipe, roofing and wall tile, flue
linings, etc.)
Whitewares (dinnerware, floor and wall tile, electrical porcelain, etc.)
Refractories (brick and monolithic products used in metal, glass, cements,
ceramics, energy conversion, petroleum, and chemicals industries)
Glasses (flat glass (windows), container glass (bottles), pressed and blown
glass (dinnerware), glass fibers (home insulation), and advanced/specialty
glass (optical fibers))
Abrasives (natural (garnet, diamond, etc.) and synthetic (silicon carbide,
diamond, fused alumina, etc.) abrasives are used for grinding, cutting,
polishing, lapping, or pressure blasting of materials)
Cements (for roads, bridges, buildings, dams, and etc.)
Advanced ceramics
o Structural (wear parts, bioceramics, cutting tools, and engine
components)
o Electrical (capacitors, insulators, substrates, integrated circuit
packages, piezoelectrics, magnets and superconductors)
o Coatings (engine components, cutting tools, and industrial wear parts)
o Chemical and environmental (filters, membranes, catalysts, and
catalyst supports)
The atoms in ceramic materials are held together by a chemical bond which will be
discussed a bit later. Briefly though, the two most common chemical bonds for
ceramic materials are covalent and ionic. Covalent and ionic bonds are much
stronger than in metallic bonds and, generally speaking, this is why ceramics are
brittle and metals are ductile.

Polymers
A polymeric solid can be thought of as a material that contains many chemically
bonded parts or units which themselves are bonded together to form a solid. The
word polymer literally means "many parts." Two industrially important polymeric
materials are plastics and elastomers. Plastics are a large and varied group of
synthetic materials which are processed by forming or molding into shape. Just as
there are many types of metals such as aluminum and copper, there are many types
of plastics, such as polyethylene and nylon. Elastomers or rubbers can be
elastically deformed a large amount when a force is applied to them and can return
to their original shape (or almost) when the force is released.
Polymers have many properties that make them attractive to use in certain
conditions. Many polymers:
are less dense than metals or ceramics,
resist atmospheric and other forms of corrosion,
offer good compatibility with human tissue, or
exhibit excellent resistance to the conduction of electrical current.
The polymer plastics can be divided into two
classes, thermoplastics and thermosetting
plastics, depending on how they are structurally
and chemically bonded. Thermoplastic polymers
comprise the four most important commodity
materials polyethylene, polypropylene,
polystyrene and polyvinyl chloride. There are
also a number of specialized engineering
polymers. The term thermoplastic indicates
that these materials melt on heating and may be
processed by a variety of molding and extrusion
techniques. Alternately, thermosetting
polymers can not be melted or remelted.
Thermosetting polymers include alkyds, amino
and phenolic resins, epoxies, polyurethanes, and unsaturated polyesters.
Rubber is a natural occurring polymer. However, most polymers are created by
engineering the combination of hydrogen and carbon atoms and the arrangement of
the chains they form. The polymer molecule is a long chain of covalent-bonded
atoms and secondary bonds then hold groups of polymer chains together to form
the polymeric material. Polymers are primarily produced from petroleum or natural
gas raw products but the use of organic substances is growing. The super-material
known as Kevlar is a man-made polymer. Kevlar is used in bullet-proof vests,
strong/lightweight frames, and underwater cables that are 20 times stronger than
steel.
Composites
A composite is commonly defined as a combination of two or more distinct
materials, each of which retains its own distinctive properties, to create a new
material with properties that cannot be achieved by any of the components acting
alone. Using this definition, it can be determined that a wide range of engineering
materials fall into this category. For example, concrete is a composite because it is
a mixture of Portland cement and aggregate. Fiberglass sheet is a composite since
it is made of glass fibers imbedded in a polymer.
Composite materials are said to have two phases. The reinforcing phase is the
fibers, sheets, or particles that are embedded in the matrix phase. The reinforcing
material and the matrix material can be metal, ceramic, or polymer. Typically,
reinforcing materials are strong with low densities while the matrix is usually a
ductile, or tough, material.
Some of the common classifications of
composites are:
Reinforced plastics
Metal-matrix composites
Ceramic-matrix composites
Sandwich structures
Concrete
Composite materials can take many
forms but they can be separated into three
categories based on the strengthening
mechanism. These categories are
dispersion strengthened, particle
reinforced and fiber reinforced.
Dispersion strengthened composites have a fine distribution of secondary particles
in the matrix of the material. These particles impede the mechanisms that allow a
material to deform. (These mechanisms include dislocation movement and slip,
which will be discussed later). Many metal-matrix composites would fall into the
dispersion strengthened composite category. Particle reinforced composites have a
large volume fraction of particle dispersed in the matrix and the load is shared by
the particles and the matrix. Most commercial ceramics and many filled polymers
are particle-reinforced composites. In fiber-reinforced composites, the fiber is the
primary load-bearing component. Fiberglass and carbon fiber composites are
examples of fiber-reinforced composites.
If the composite is designed and fabricated correctly, it combines the strength of
the reinforcement with the toughness of the matrix to achieve a combination of
desirable properties not available in any single conventional material. Some
composites also offer the advantage of being tailorable so that properties, such as
strength and stiffness, can easily be changed by changing amount or orientation of
the reinforcement material. The downside is that such composites are often more
expensive than conventional materials.
Structure of Materials
It should be clear that all matter is made of atoms. From the periodic table, it can
be seen that there are only about 100 different kinds of atoms in the entire
Universe. These same 100 atoms form thousands of different substances ranging
from the air we breathe to the metal used to support tall buildings. Metals behave
differently than ceramics, and ceramics behave differently than polymers. The
properties of matter depend on which atoms are used and how they are bonded
together.

The structure of materials can be classified by the general magnitude of various
features being considered. The three most common major classification of
structural, listed generally in increasing size, are:
Atomic structure, which includes features that cannot be seen, such as the types of
bonding between the atoms, and the way the atoms are arranged.
Microstructure, which includes features that can be seen using a microscope, but
seldom with the naked eye.
Macrostructure, which includes features that can be seen with the naked eye)
The atomic structure primarily affects the chemical, physical, thermal, electrical,
magnetic, and optical properties. The microstructure and macrostructure can also
affect these properties but they generally have a larger effect on mechanical
properties and on the rate of chemical reaction. The properties of a material offer
clues as to the structure of the material. The strength of metals suggests that these
atoms are held together by strong bonds. However, these bonds must also allow
atoms to move since metals are also usually formable. To understand the structure
of a material, the type of atoms present, and how the atoms are arranged and
bonded must be known. Lets first look at atomic bonding.
Atomic Bonding
(Metallic, Ionic, Covalent, and van
der Waals Bonds)
From elementary chemistry it is known that the
atomic structure of any element is made up of a
positively charged nucleus surrounded by
electrons revolving around it. An elements
atomic number indicates the number of positively
charged protons in the nucleus. The atomic
weight of an atom indicates how many protons
and neutrons in the nucleus. To determine the
number of neutrons in an atom, the atomic
number is simply subtracted from the atomic
weight.
Atoms like to have a balanced electrical charge. Therefore, they usually have
negatively charged electrons surrounding the nucleus in numbers equal to the
number of protons. It is also known that electrons are present with different
energies and it is convenient to consider these electrons surrounding the nucleus in
energy shells. For example, magnesium, with an atomic number of 12, has two
electrons in the inner shell, eight in the second shell and two in the outer shell.
All chemical bonds involve electrons. Atoms will stay close together if they have a
shared interest in one or more electrons. Atoms are at their most stable when they
have no partially-filled electron shells. If an atom has only a few electrons in a
shell, it will tend to lose them to empty the shell. These elements are metals. When
metal atoms bond, a metallic bond occurs. When an atom has a nearly full electron
shell, it will try to find electrons from another atom so that it can fill its outer shell.
These elements are usually described as nonmetals. The bond between two
nonmetal atoms is usually a covalent bond. Where metal and nonmetal atom come
together an ionic bond occurs. There are also other, less common, types of bond
but the details are beyond the scope of this material. On the next few pages, the
Metallic, Covalent and Ionic bonds will be covered in more detail.
Ionic Bonds
Ionic bonding occurs between charged particles. These may be atoms or groups of
atoms, but this discuss will be conducted in terms of single atoms. Ionic bonding
occurs between metal atoms and nonmetal atoms. Metals usually have 1, 2, or 3
electrons in their outermost shell. Nonmetals have 5, 6, or 7 electrons in their outer
shell. Atoms with outer shells that are only partially filled are unstable. To become
stable, the metal atom wants to get rid of one or more electrons in its outer shell.
Losing electrons will either result in an empty outer shell or get it closer to having
an empty outer shell. It would like to have an empty outer shell because the next
lower energy shell is a stable shell with eight electrons.


Since electrons have a negative charge, the atom that gains electrons becomes a
negatively charged ions (aka anion) because it now has more electrons than
protons. Alternately, an atom that loses electrons becomes a positively charged ion
(aka cations). The particles in an ionic compound are held together because there
are oppositely charged particles that are attracted to one another.
The images above schematically show the process that takes place during the
formation of an ionic bond between sodium and chlorine atoms. Note that sodium
has one valence electron that it would like to give up so that it would become
stable with a full outer shell of eight. Also note that chlorine has seven valence
electrons and it would like to gain an electron in order to have a full shell of eight.
The transfer of the electron causes the previously neutral sodium atom to become a
positively charged ion (cation), and the previously neutral chlorine atom to become
a negatively charged ion (anion). The attraction for the cation and the anion is
called the ionic bond.
Generally, solid materials with ionic bonds:
are hard because particles cannot easily slide past one another.
are good insulators because there are no free electrons or ions (unless
dissolved or melted).
are transparent because their electrons are not moving from atom to atom
and less likely to interact with light photons.
are brittle and tend to cleave rather than deform because bonds are strong.
have high melting point because ionic bonds are relatively strong.
Covalent Bonding
Where a compound only contains nonmetal atoms, a covalent bond is formed by
atoms sharing two or more electrons. Nonmetals have 4 or more electrons in their
outer shells (except boron). With this many electrons in the outer shell, it would
require more energy to remove the electrons than would be gained by making new
bonds. Therefore, both the atoms involved share a pair of electrons. Each atom
gives one of its outer electrons to the electron pair, which then spends some time
with each atom. Consequently, both atoms are held near each other since both
atoms have a share in the electrons.


More than one electron pair can be formed with half of the electrons coming from
one atom and the rest from the other atom. An important feature of this bond is that
the electrons are tightly held and equally shared by the participating atoms. The
atoms can be of the same element or different elements. In each molecule, the
bonds between the atoms are strong but the bonds between molecules are usually
weak. This makes many solid materials with covalent bonds brittle. Many ceramic
materials have covalent bonds.
Compounds with covalent bonds may be solid, liquid or gas at room temperature
depending on the number of atoms in the compound. The more atoms in each
molecule, the higher a compounds melting and boiling temperature will be. Since
most covalent compounds contain only a few atoms and the forces between
molecules are weak, most covalent compounds have low melting and boiling
points. However, some, like carbon compounds, can be very large. An example is
the diamond in which carbon atoms each share four electrons to form giant lattices.
Some Common Features of Materials with Covalent Bonds:
Hard
Good insulators
Transparent
Brittle or cleave rather than deform
Metallic Bonding
A common characteristic of metallic elements is they contain
only one to three electrons in the outer shell. When an
element has only one, two or three valence electrons (i.e.
electrons in the outer shell), the bond between these electrons
and the nucleus is relatively weak. So, for example, when
aluminum atoms are grouped together in a block of metal, the
outer electrons leave individual atoms to become part of
common electron cloud. In this arrangement, the valence
electrons have considerable mobility and are able to conduct heat and electricity
easily. Also, the delocalized nature of the bonds, make it possible for the atoms to
slide past each other when the metal is deformed instead of fracturing like glass or
other brittle material.

Since the aluminum atoms lose two electrons, they end up having a positive charge
and are designated Al
3+
ions (cations). These ions repel each other but are held
together in the block because the negative electrons are attracted to the positively
charged ions. A result of the sharing of electrons is the cations arrange themselves
in a regular pattern. This regular pattern of atoms is the crystalline structure of
metals. In the crystal lattice, atoms are packed closely together to maximize the
strength of the bonds. An actual piece of metal consists of many tiny crystals called
grains that touch at grain boundaries.
Some Common Features of Materials with Metallic Bonds:
Good electrical and thermal conductors due to their free valence electrons
Opaque
Relatively ductile
Van der Waals Bond
The van der Waal bonds occur to some extent in all materials but are
particularly important in plastics and polymers. These materials are made up
of a long string molecules consisting of carbon atoms covalently bonded
with other atoms, such as hydrogen, nitrogen, oxygen, fluorine. The
covalent bonds within the
molecules are very strong and
rupture only under extreme
conditions. The bonds between
the molecules that allow
sliding and rupture to occur are
called van der Waal forces.
When ionic and covalent bonds
are present, there is some
imbalance in the electrical
charge of the molecule. Take water as an example. Research has determined
the hydrogen atoms are bonded to the oxygen atoms at an angle of 104.5.
This angle produces a positive polarity at the hydrogen-rich end of the
molecule and a negative polarity at the other end. A result of this charge
imbalance is that water molecules are attracted to each other. This is the
force that holds the molecules together in a drop of water.
This same concept can be carried on to plastics, except that as molecules
become larger, the van der Waal forces between molecules also increases.
For example, in polyethylene the molecules are composed of hydrogen and
carbon atoms in the same ratio as ethylene gas. But there are more of each
type of atom in the polyethylene molecules and as the number of atoms in a
molecule increases, the matter passes from a gas to a liquid and finally to a
solid.
Polymers are often classified as being either a thermoplastic or a
thermosetting material. Thermoplastic materials can be easily remelted for
forming or recycling and thermosetting material cannot be easily remelted.
In thermoplastic materials consist of long chainlike molecules. Heat can be
used to break the van der Waal forces between the molecules and change the
form of the material from a solid to a liquid. By contrast, thermosetting
materials have a three-dimensional network of covalent bonds. These bonds
cannot be easily broken by heating and, therefore, can not be remelted and
formed as easily as thermoplastics.
Solid State Structure
In the previous pages, some of the mechanisms that bond together the
multitude of individual atoms or molecules of a solid material were
discussed. These forces may be primary chemical bonds, as in metals and
ionic solids, or they may be secondary van der Waals forces of solids, such
as in ice, paraffin wax and most polymers. In solids, the way the atoms or
molecules arrange themselves contributes to the appearance and the
properties of the materials.
Atoms can be gathered together as an aggregate through a number of
different processes, including condensation, pressurization, chemical
reaction, electrodeposition, and melting. The process usually determines, at
least initially, whether the collection of atoms will take to form of a gas,
liquid or solid. The state usually changes as its temperature or pressure is
changed. Melting is the process most often used to form an aggregate of
atoms. When the temperature of a melt is lowered to a certain point, the
liquid will form either a crystalline solid or and amorphous solid.
Amorphous Solids
A solid substance with its atoms held apart at equilibrium spacing, but with
no long-range periodicity in atom location in its structure is an amorphous
solid. Examples of amorphous solids are glass and some types of plastic.
They are sometimes described as supercooled liquids because their
molecules are arranged in a random manner some what as in the liquid state.
For example, glass is commonly made from silicon dioxide or quartz sand,
which has a crystalline structure. When the sand is melted and the liquid is
cooled rapidly enough to avoid crystallization, an amorphous solid called a
glass is formed. Amorphous solids do not show a sharp phase change from
solid to liquid at a definite melting point, but rather soften gradually when
they are heated. The physical properties of amorphous solids are identical in
all directions along any axis so they are said to have isotropic properties,
which will be discussed in more detail later
.
Crystalline Solids
More than 90% of naturally occurring and artificially prepared solids are
crystalline. Minerals, sand, clay, limestone, metals, carbon (diamond and
graphite), salts ( NaCl, KCl etc.), all have crystalline structures. A crystal is
a regular, repeating arrangement of atoms or molecules. The majority of
solids, including all metals, adopt a crystalline arrangement because the
amount of stabilization achieved by anchoring interactions between
neighboring particles is at its greatest when the particles adopt regular
(rather than random) arrangements. In the crystalline arrangement, the
particles pack efficiently together to minimize the total intermolecular
energy.
The regular repeating pattern that the atoms arrange in is called the
crystalline lattice. The scanning tunneling microscope (STM) makes it
possible to image the electron cloud associated individual atoms at the
surface of a material. Below is an STM image of a platinum surface
showing the regular alignment of atoms.

Courtesy: IBM Research, Almaden Research Center.
Crystal Structure
Crystal structures may be conveniently specified by describing the
arrangement within the solid of a small representative group of atoms or
molecules, called the unit cell. By multiplying identical unit cells in three
directions, the location of all the particles in the crystal is determined. In
nature, 14 different types of crystal structures or lattices are found. The
simplest crystalline unit cell to picture is the cubic, where the atoms are
lined up in a square, 3D grid. The unit cell is simply a box with an atom at
each corner. Simple cubic crystals are relatively rare, mostly because they
tend to easily distort. However, many crystals form body-centered-cubic
(bcc) or face-centered-cubic (fcc) structures, which are cubic with either an
extra atom centered in the cube or centered in each face of the cube. Most
metals form bcc, fcc or Hexagonal Close Packed (hpc) structures; however,
the structure can change depending on temperature. These three structures
will be discussed in more detail on the following page.
Crystalline structure is important because it contributes to the properties of a
material. For example, it is easier for planes of atoms to slide by each other
if those planes are closely packed. Therefore, lattice structures with closely
packed planes allow more plastic deformation than those that are not closely
packed. Additionally, cubic lattice structures allow slippage to occur more
easily than non-cubic lattices. This is because their symmetry provides
closely packed planes in several directions. A face-centered cubic crystal
structure will exhibit more ductility (deform more readily under load before
breaking) than a body-centered cubic structure. The bcc lattice, although
cubic, is not closely packed and forms strong metals. Alpha-iron and
tungsten have the bcc form. The fcc lattice is both cubic and closely packed
and forms more ductile materials. Gamma-iron, silver, gold, and lead have
fcc structures. Finally, HCP lattices are closely packed, but not cubic. HCP
metals like cobalt and zinc are not as ductile as the fcc metals.
Primary Metallic Crystalline Structures
(BCC, FCC, HCP)
As pointed out on the previous page, there are 14
different types of crystal unit cell structures or lattices
are found in nature. However most metals and many
other solids have unit cell structures described as body
center cubic (bcc), face centered cubic (fcc) or
Hexagonal Close Packed (hcp). Since these structures
are most common, they will be discussed in more detail.
Body-Centered Cubic (BCC) Structure
The body-centered cubic unit cell has atoms at each of the eight corners of a
cube (like the cubic unit cell) plus one atom in the center of the cube (left
image below). Each of the corner atoms is the corner of another cube so the
corner atoms are shared among eight unit cells. It is said to have a
coordination number of 8. The bcc unit cell consists of a net total of two
atoms; one in the center and eight eighths from corners atoms as shown in
the middle image below (middle image below). The image below highlights
a unit cell in a larger section of the lattice.

The bcc arrangement does not allow the atoms to pack together as closely as
the fcc or hcp arrangements. The bcc structure is often the high temperature
form of metals that are close-packed at lower temperatures. The volume of
atoms in a cell per the total volume of a cell is called the packing factor.
The bcc unit cell has a packing factor of 0.68.
Some of the materials that have a bcc structure include lithium, sodium,
potassium, chromium, barium, vanadium, alpha-iron and tungsten. Metals
which have a bcc structure are usually harder and less malleable than close-
packed metals such as gold. When the metal is deformed, the planes of
atoms must slip over each other, and this is more difficult in the bcc
structure. It should be noted that there are other important mechanisms for
hardening materials, such as introducing impurities or defects which make
slipping more difficult. These hardening mechanisms will be discussed
latter.
Face Centered Cubic (FCC) Structure
The face centered cubic structure has atoms located at each of the corners
and the centers of all the cubic faces (left image below). Each of the corner
atoms is the corner of another cube so the corner atoms are shared among
eight unit cells. Additionally, each of its six face centered atoms is shared
with an adjacent atom. Since 12 of its atoms are shared, it is said to have a
coordination number of 12. The fcc unit cell consists of a net total of four
atoms; eight eighths from corners atoms and six halves of the face atoms as
shown in the middle image above. The image below highlights a unit cell in
a larger section of the lattice.


In the fcc structure (and the hcp structure) the atoms can pack closer
together than they can in the bcc structure. The atoms from one layer nest
themselves in the empty space between the atoms of the adjacent layer. To
picture packing arrangement, imagine a box filled with a layer of balls that
are aligned in columns and rows. When a few additional balls are tossed in
the box, they will not balance directly on top of the balls in the first layer but
instead will come to rest in the pocket created between four balls of the
bottom layer. As more balls are added they will pack together to fill up all
the pockets. The packing factor (the volume of atoms in a cell per the total
volume of a cell) is 0.74 for fcc crystals. Some of the metals that have the
fcc structure include aluminum, copper, gold, iridium, lead, nickel, platinum
and silver.
Hexagonal Close Packed (HPC) Structure
Another common close packed structure is the hexagonal close pack. The
hexagonal structure of alternating layers is shifted so its atoms are aligned to
the gaps of the preceding layer. The atoms from one layer nest themselves in
the empty space between the atoms of the adjacent layer just like in the fcc
structure. However, instead of being a cubic structure, the pattern is
hexagonal. (See image below.) The difference between the HPC and FCC
structure is discussed later in this section.

The hcp structure has three layers of atoms. In each the top and bottom
layer, there are six atoms that arrange themselves in the shape of a hexagon
and a seventh atom that sits in the middle of the hexagon. The middle layer
has three atoms nestle in the triangular "grooves" of the top and bottom
plane. Note that there are six of these "grooves" surrounding each atom in
the hexagonal plane, but only three of them can be filled by atoms.
As shown in the middle image above, there are six atoms in the hcp unit
cell. Each of the 12 atoms in the corners of the top and bottom layers
contribute 1/6 atom to the unit cell, the two atoms in the center of the
hexagon of both the top and bottom layers each contribute atom and each
of the three atom in the middle layer contribute 1 atom. The image on the
right above attempts to show several hcp unit cells in a larger lattice.
The coordination number of the atoms in this structure is 12. There are six
nearest neighbors in the same close packed layer, three in the layer above
and three in the layer below. The packing factor is 0.74, which is the same
as the fcc unit cell. The hcp structure is very common for elemental metals
and some examples include beryllium, cadmium, magnesium, titanium, zinc
and zirconium.
Similarities and Difference Between the
FCC and HCP Structure
The face centered cubic and hexagonal close packed structures both have a
packing factor of 0.74, consist of closely packed planes of atoms, and have a
coordination number of 12. The difference between the fcc and hcp is the
stacking sequence. The hcp layers cycle among the two equivalent shifted
positions whereas the fcc layers cycle between three positions. As can be
seen in the image, the hcp structure contains only two types of planes with
an alternating ABAB arrangement. Notice how the atoms of the third plane
are in exactly the same position as the atoms in the first plane. However, the
fcc structure contains three types of planes with a ABCABC arrangement.
Notice how the atoms in rows A and C are no longer aligned. Remember
that cubic lattice structures allow slippage to occur more easily than non-
cubic lattices, so hcp metals are not as ductile as the fcc metals.

The table below shows the stable room temperature crystal structures for
several elemental metals.
Metal Crystal Structure Atomic Radius (nm)
Aluminum FCC 0.1431
Cadmium HCP 0.1490
Chromium BCC 0.1249
Cobalt HCP 0.1253
Copper FCC 0.1278
Gold FCC 0.1442
Iron (Alpha) BCC 0.1241
Lead FCC 0.1750
Magnesium HCP 0.1599
Molybdenum BCC 0.1363
Nickel FCC 0.1246
Platinum FCC 0.1387
Silver FCC 0.1445
Tantalum BCC 0.1430
Titanium (Alpha) HCP 0.1445
Tungsten BCC 0.1371
Zinc HCP 0.1332
A nanometer (nm) equals 10
-9
meter or 10 Angstrom units.
Solidification
The crystallization of a large amount of material from a single point of nucleation
results in a single crystal. In engineering materials, single crystals are produced
only under carefully controlled conditions. The expense of producing single crystal
materials is only justified for special applications, such as turbine engine blades,
solar cells, and piezoelectric materials. Normally when a material begins to
solidify, multiple crystals begin to grow in the liquid and a polycrystalline (more
than one crystal) solid forms.
The moment a crystal begins to grow is know as nucleation and the point where it
occurs is the nucleation point. At the solidification temperature, atoms of a liquid,
such as melted metal, begin to bond together at the nucleation points and start to
form crystals. The final sizes of the individual crystals depend on the number of
nucleation points. The crystals increase in size by the progressive addition of atoms
and grow until they impinge upon adjacent growing crystal.

a) Nucleation of crystals, b) crystal growth, c) irregular grains form as crystals
grow together, d) grain boundaries as seen in a microscope.
In engineering materials, a crystal is usually referred to as a grain. A grain is
merely a crystal without smooth faces because its growth was impeded by contact
with another grain or a boundary surface. The interface formed between grains is
called a grain boundary. The atoms between the grains (at the grain boundaries)
have no crystalline structure and are said to be disordered.
Grains are sometimes large enough to be visible under an ordinary light
microscope or even to the unaided eye. The spangles that are seen on newly
galvanized metals are grains. Rapid cooling generally results in more nucleation
points and smaller grains (a fine grain structure). Slow cooling generally results in
larger grains which will have lower strength, hardness and ductility.
Dendrites
In metals, the crystals that form in the liquid
during freezing generally follow a pattern
consisting of a main branch with many
appendages. A crystal with this morphology
slightly resembles a pine tree and is called a
dendrite, which means branching. The formation
of dendrites occurs because crystals grow in
defined planes due to the crystal lattice they
create. The figure to the right shows how a cubic
crystal can grow in a melt in three dimensions,
which correspond to the six faces of the cube.
For clarity of illustration, the adding of unit cells
with continued solidification from the six faces is shown simply as lines.
Secondary dendrite arms branch off the primary arm, and tertiary arms off the
secondary arms and etcetera.
During freezing of a polycrystalline material,
many dendritic crystals form and grow until they
eventually become large enough to impinge
upon each other. Eventually, the
interdendriticspaces between the dendrite arms
crystallize to yield a more regular crystal. The
original dendritic pattern may not be apparent
when examining the microstructure of a material.
However, dendrites can often be seen in
solidification voids that sometimes occur in
castings or welds, as shown to the right..
Shrinkage
Most materials contract or shrink during
solidification and cooling. Shrinkage is the result
of:
Contraction of the liquid as it cools prior to its solidification
Contraction during phase change from a liquid to solid
Contraction of the solid as it continues to cool to ambient temperature.
Shrinkage can sometimes cause cracking to occur in component as it solidifies.
Since the coolest area of a volume of liquid is where it contacts a mold or die,
solidification usually begins first at this surface. As the crystals grow inward, the
material continues to shrink. If the solid surface is too rigid and will not deform to
accommodate the internal shrinkage, the stresses can become high enough to
exceed the tensile strength of the material and cause a crack to form. Shrinkage
cavitation sometimes occurs because as a material solidifies inward, shrinkage
occurred to such an extent that there is not enough atoms present to fill the
available space and a void is left.
Anisotropy and Isotropy
In a single crystal, the physical and mechanical properties often differ with
orientation. It can be seen from looking at our models of crystalline structure that
atoms should be able to slip over one another or distort in relation to one another
easier in some directions than others. When the properties of a material vary with
different crystallographic orientations, the material is said to be anisotropic.
Alternately, when the properties of a material are the same in all directions, the
material is said to be isotropic. For many polycrystalline materials the grain
orientations are random before any working (deformation) of the material is done.
Therefore, even if the individual grains are anisotropic, the property differences
tend to average out and, overall, the material is isotropic. When a material is
formed, the grains are usually distorted and elongated in one or more directions
which makes the material anisotropic. Material forming will be discussed later but
lets continue discussing crystalline structure at the atomic level.
Crystal Defects
A perfect crystal, with every atom of the same type in the correct position, does not
exist. All crystals have some defects. Defects contribute to the mechanical
properties of metals. In fact, using the term defect is sort of a misnomer since
these features are commonly intentionally used to manipulate the mechanical
properties of a material. Adding alloying elements to a metal is one way of
introducing a crystal defect. Nevertheless, the term defect will be used, just keep
in mind that crystalline defects are not always bad. There are basic classes of
crystal defects:
point defects, which are places where an atom is missing or irregularly
placed in the lattice structure. Point defects include lattice vacancies, self-
interstitial atoms, substitution impurity atoms, and interstitial impurity
atoms
linear defects, which are groups of atoms in irregular positions. Linear
defects are commonly called dislocations.
planar defects, which are interfaces between homogeneous regions of the
material. Planar defects include grain boundaries, stacking faults and
external surfaces.
It is important to note at this point that plastic deformation in a material occurs due
to the movement of dislocations (linear defects). Millions of dislocations result for
plastic forming operations such as rolling and extruding. It is also important to note
that any defect in the regular lattice structure disrupts the motion of dislocation,
which makes slip or plastic deformation more difficult. These defects not only
include the point and planer defects mentioned above, and also other dislocations.
Dislocation movement produces additional dislocations, and when dislocations run
into each other it often impedes movement of the dislocations. This drives up the
force needed to move the dislocation or, in other words, strengthens the material.
Each of the crystal defects will be discussed in more detail in the following pages.
Point Defects
Point defects are where an atom is missing
or is in an irregular place in the lattice
structure. Point defects include self
interstitial atoms, interstitial impurity
atoms, substitutional atoms and vacancies.
A self interstitial atom is an extra atom that
has crowded its way into an interstitial void
in the crystal structure. Self interstitial
atoms occur only in low concentrations in
metals because they distort and highly
stress the tightly packed lattice structure.
A substitutional impurity atom is an atom of
a different type than the bulk atoms, which
has replaced one of the bulk atoms in the
lattice. Substitutional impurity atoms are
usually close in size (within approximately
15%) to the bulk atom. An example of
substitutional impurity atoms is the zinc
atoms in brass. In brass, zinc atoms with a radius of 0.133 nm have replaced some
of the copper atoms, which have a radius of 0.128 nm.

Interstitial impurity atoms are much smaller than the atoms in the bulk matrix.
Interstitial impurity atoms fit into the open space between the bulk atoms of the
lattice structure. An example of interstitial impurity atoms is the carbon atoms that
are added to iron to make steel. Carbon atoms, with a radius of 0.071 nm, fit nicely
in the open spaces between the larger (0.124 nm) iron atoms.
Vacancies are empty spaces where an atom should be, but is missing. They are
common, especially at high temperatures when atoms are frequently and randomly
change their positions leaving behind empty lattice sites. In most cases diffusion
(mass transport by atomic motion) can only occur because of vacancies.
Linear Defects - Dislocations
Dislocations are another type of defect in crystals. Dislocations are areas were the
atoms are out of position in the crystal structure. Dislocations are generated and
move when a stress is applied. The motion of dislocations allows slip plastic
deformation to occur.
Before the discovery of the dislocation by Taylor, Orowan and Polyani in 1934, no
one could figure out how the plastic deformation properties of a metal could be
greatly changed by solely by forming (without changing the chemical
composition). This became even bigger mystery when in the early 1900s scientists
estimated that metals undergo plastic deformation at forces much smaller than the
theoretical strength of the forces that are holding the metal atoms together. Many
metallurgists remained skeptical of the dislocation theory until the development of
the transmission electron microscope in the late 1950s. The TEM allowed
experimental evidence to be collected that showed that the strength and ductility of
metals are controlled by dislocations.

There are two basic types of dislocations, the edge dislocation and the screw
dislocation. Actually, edge and screw dislocations are just extreme forms of the
possible dislocation structures that can occur. Most dislocations are probably a
hybrid of the edge and screw forms but this discussion will be limited to these two
types.
Edge Dislocations
The edge defect can be easily visualized as an extra half-plane of atoms in a lattice.
The dislocation is called a line defect because the locus of defective points
produced in the lattice by the dislocation lie along a line. This line runs along the
top of the extra half-plane. The inter-atomic bonds are significantly distorted only
in the immediate vicinity of the dislocation line.

Understanding the movement of a dislocation is key to understanding why
dislocations allow deformation to occur at much lower stress than in a perfect
crystal. Dislocation motion is analogous to movement of a caterpillar. The
caterpillar would have to exert a large force to move its entire body at once.
Instead it moves the rear portion of its body forward a small amount and creates a
hump. The hump then moves forward and eventual moves all of the body forward
by a small amount.

As shown in the set of images above, the dislocation moves similarly moves a
small amount at a time. The dislocation in the top half of the crystal is slipping one
plane at a time as it moves to the right from its position in image (a) to its position
in image (b) and finally image (c). In the process of slipping one plane at a time the
dislocation propagates across the crystal. The movement of the dislocation across
the plane eventually causes the top half of the crystal to move with respect to the
bottom half. However, only a small fraction of the bonds are broken at any given
time. Movement in this manner requires a much smaller force than breaking all the
bonds across the middle plane simultaneously.
Screw Dislocations
There is a second basic type of
dislocation, called screw dislocation.
The screw dislocation is slightly more
difficult to visualize. The motion of a
screw dislocation is also a result of
shear stress, but the defect line
movement is perpendicular to direction
of the stress and the atom displacement,
rather than parallel. To visualize a screw
dislocation, imagine a block of metal
with a shear stress applied across one
end so that the metal begins to rip. This
is shown in the upper right image. The
lower right image shows the plane of
atoms just above the rip. The atoms
represented by the blue circles have not
yet moved from their original position.
The atoms represented by the red circles
have moved to their new position in the
lattice and have reestablished metallic
bonds. The atoms represented by the
green circles are in the process of
moving. It can be seen that only a
portion of the bonds are broke at any given time. As was the case with the edge
dislocation, movement in this manner requires a much smaller force than breaking
all the bonds across the middle plane simultaneously.
If the shear force is increased, the atoms will continue to slip to the right. A row of
the green atoms will find there way back into a proper spot in the lattice (and
become red) and a row of the blue atoms will slip out of position (and become
green). In this way, the screw dislocation will move upward in the image, which is
perpendicular to direction of the stress. Recall that the edge dislocation moves
parallel to the direction of stress. As shown in the image below, the net plastic
deformation of both edge and screw dislocations is the same, however.

The dislocations move along the densest planes of atoms in a material, because the
stress needed to move the dislocation increases with the spacing between the
planes. FCC and BCC metals have many dense planes, so dislocations move
relatively easy and these materials have high ductility. Metals are strengthened by
making it more difficult for dislocations to move. This may involve the
introduction of obstacles, such as interstitial atoms or grain boundaries, to pin
the dislocations. Also, as a material plastically deforms, more dislocations are
produced and they will get into each others way and impede movement. This is
why strain or work hardening occurs.
In ionically bonded materials, the ion must move past an area with a repulsive
charge in order to get to the next location of the same charge. Therefore, slip is
difficult and the materials are brittle. Likewise, the low density packing of covalent
materials makes them generally more brittle than metals.
Planar Defects
Stacking Faults and Twin Boundaries
A disruption of the long-range stacking sequence can produce two other common
types of crystal defects: 1) a stacking fault and 2) a twin region. A change in the
stacking sequence over a few atomic spacings produces a stacking fault whereas a
change over many atomic spacings produces a twin region.

A stacking fault is a one or two layer interruption in the stacking sequence of atom
planes. Stacking faults occur in a number of crystal structures, but it is easiest to
see how they occur in close packed structures. For example, it is know from a
previous discussion that face centered cubic (fcc) structures differ from hexagonal
close packed (hcp) structures only in their stacking order. For hcp and fcc
structures, the first two layers arrange themselves identically, and are said to have
an AB arrangement. If the third layer is placed so that its atoms are directly above
those of the first (A) layer, the stacking will be ABA. This is the hcp structure, and
it continues ABABABAB. However it is possible for the third layer atoms to
arrange themselves so that they are in line with the first layer to produce an ABC
arrangement which is that of the fcc structure. So, if the hcp structure is going
along as ABABAB and suddenly switches to ABABABCABAB, there is a
stacking fault present.
Alternately, in the fcc arrangement the pattern is ABCABCABC. A stacking fault
in an fcc structure would appear as one of the C planes missing. In other words the
pattern would become ABCABCAB_ABCABC.
If a stacking fault does not corrects itself immediately but continues over some
number of atomic spacings, it will produce a second stacking fault that is the twin
of the first one. For example if the stacking pattern is ABABABAB but switches to
ABCABCABC for a period of time before switching back to ABABABAB, a pair
of twin stacking faults is produced. The red region in the stacking sequence that
goes ABCABCACBACBABCABC is the twin plane and the twin boundaries are
the A planes on each end of the highlighted region.
Grain Boundaries in Polycrystals
Another type of planer defect is the grain boundary. Up to this point, the discussion
has focused on defects of single crystals. However, solids generally consist of a
number of crystallites or grains. Grains can range in size from nanometers to
millimeters across and their orientations are usually rotated with respect to
neighboring grains. Where one grain stops and another begins is know as a grain
boundary. Grain boundaries limit the lengths and motions of dislocations.
Therefore, having smaller grains (more grain boundary surface area) strengthens a
material. The size of the grains can be controlled by the cooling rate when the
material cast or heat treated. Generally, rapid cooling produces smaller grains
whereas slow cooling result in larger grains. For more information, refer to the
discussion on solidification.
Bulk Defects
Bulk defects occur on a much bigger
scale than the rest of the crystal defects
discussed in this section. However, for
the sake of completeness and since they
do affect the movement of dislocations, a
few of the more common bulk defects
will be mentioned. Voids are regions
where there are a large number of atoms
missing from the lattice. The image to the
right is a void in a piece of metal The image was acquired using a Scanning
Electron Microscope (SEM). Voids can occur for a number of reasons. When voids
occur due to air bubbles becoming trapped when a material solidifies, it is
commonly called porosity. When a void occurs due to the shrinkage of a material
as it solidifies, it is called cavitation.
Another type of bulk defect occurs when impurity atoms cluster together to form
small regions of a different phase. The term phase refers to that region of space
occupied by a physically homogeneous material. These regions are often called
precipitates. Phases and precipitates will be discussed in more detail latter.
Elastic/Plastic Deformation
When a sufficient load is applied to a metal or other structural material, it will
cause the material to change shape. This change in shape is called deformation. A
temporary shape change that is self-reversing after the force is removed, so that the
object returns to its original shape, is called elastic deformation. In other words,
elastic deformation is a change in shape of a material at low stress that is
recoverable after the stress is removed. This type of deformation involves
stretching of the bonds, but the atoms do not slip past each other.
When the stress is sufficient to
permanently deform the metal, it is
called plastic deformation. As
discussed in the section on crystal
defects, plastic deformation
involves the breaking of a limited
number of atomic bonds by the
movement of dislocations. Recall
that the force needed to break the
bonds of all the atoms in a crystal
plane all at once is very great.
However, the movement of
dislocations allows atoms in crystal
planes to slip past one another at a
much lower stress levels. Since the
energy required to move is lowest
along the densest planes of atoms,
dislocations have a preferred
direction of travel within a grain of
the material. This results in slip
that occurs along parallel planes within the grain. These parallel slip planes group
together to form slip bands, which can be seen with an optical microscope. A slip
band appears as a single line under the microscope, but it is in fact made up of
closely spaced parallel slip planes as shown in the image.
Fatigue Crack Initiation
While on the subject of dislocations, it is appropriate to
briefly discuss fatigue. Fatigue is one of the primary
reasons for the failure of structural components. The life of
a fatigue crack has two parts, initiation and propagation.
Dislocations play a major role in the fatigue crack initiation
phase. It has been observed in laboratory testing that after a
large number of loading cycles dislocations pile up and
form structures called persistent slip bands (PSB). An
example of a PSB is shown in the micrograph image to the
right.
PSBs are areas that rise above (extrusion) or fall below
(intrusion) the surface of the component due to movement
of material along slip planes. This leaves tiny steps in the
surface that serve as stress risers where fatigue cracks can
initiate. A crack at the edge of a PSB is shown in the
image below taken with a scanning electron microscope
(SEM).





Diffusion
Diffusion is the migration of atoms from a region of high
concentration to a region of low concentration. In a
homogeneous material, atoms are routinely moving around
but the movement is random (i.e. there is always an equal
number of atoms moving in all directions). In an
inhomogeneous material, all the atoms are moving near
randomly, but there is a migration of atoms to areas where
their concentrations are lower. In other words, there is a net
diffusion.
Atom diffusion can occur by the motion of host or
substitutional atoms to vacancies (vacancy diffusion), or
interstitial impurities atoms to different interstitial positions
(interstitial diffusion). In order to move, an atom must overcome the bond energy
due to nearby atoms. This is more easily achieved at high temperatures when the
atoms are vibrating strongly. Carburizing, which will be discussed later, is an
example of diffusion is used.
Property Modification
Many structural metals undergo some special treatment to modify their properties
so that they will perform better for their intended use. This treatment can include
mechanical working, such as rolling or forging, alloying and/or thermal treatments.
Consider aluminum as an example. Commercially pure aluminum (1100) has a
tensile strength of around 13,000 psi, which limits its usefulness in structural
applications. However, by cold-working aluminum, its strength can be
approximately doubled. Also, strength increases are obtained by adding alloying
metals such as manganese, silicon, copper, magnesium and zinc. Further, many
aluminum alloys are strengthened by heat treatment. Some heat-treatable
aluminum alloys obtain tensile strengths that can exceed 100,000 psi.
Strengthening/Hardening Mechanisms
As discussed in the previous section, the ability of a crystalline material to
plastically deform largely depends on the ability for dislocation to move within a
material. Therefore, impeding the movement of dislocations will result in the
strengthening of the material. There are a number of ways to impede dislocation
movement, which include:
controlling the grain size (reducing continuity of atomic planes)
strain hardening (creating and tangling dislocations)
alloying (introducing point defects and
more grains to pin dislocation)
Control of Grain Size
The size of the grains within a material also has
an effect on the strength of the material. The
boundary between grains acts as a barrier to
dislocation movement and the resulting slip
because adjacent grains have different
orientations. Since the atom alignment is
different and slip planes are discontinuous between grains. The smaller the grains,
the shorter the distance atoms can move along a particular slip plane. Therefore,
smaller grains improve the strength of a material. The size and number of grains
within a material is controlled by the rate of solidification from the liquid phase.
Strain Hardening
Strain hardening (also called work-hardening or cold-working) is the process of
making a metal harder and stronger through plastic deformation. When a metal is
plastically deformed, dislocations move and additional dislocations are generated.
The more dislocations within a material, the more they will interact and become
pinned or tangled. This will result in a decrease in the mobility of the dislocations
and a strengthening of the material. This type of strengthening is commonly called
cold-working. It is called cold-working because the plastic deformation must
occurs at a temperature low enough that atoms cannot rearrange themselves. When
a metal is worked at higher temperatures (hot-working) the dislocations can
rearrange and little strengthening is achieved.
Strain hardening can be easily demonstrated with piece of wire or a paper clip.
Bend a straight section back and forth several times. Notice that it is more difficult
to bend the metal at the same place. In the strain hardened area dislocations have
formed and become tangled, increasing the strength of the material. Continued
bending will eventually cause the wire to break at the bend due to fatigue cracking.
(After a large number of bending cycles, dislocations form structures called
Persistent Slip Bands (PSB). PSBs are basically tiny areas where the dislocations
have piled up and moved the material surface out leave steps in the surface that act
as stress risers or crack initiation points.)

It should be understood, however, that increasing
the strength by cold-working will also result in a
reduction in ductility. The graph to the right
shows the yield strength and the percent
elongation as a function of percent cold-work for
a few example materials. Notice that for each
material, a small amount of cold-working results in a significant reduction in
ductility.

Effects of Elevated Temperature on Strain Hardened Materials
When strain hardened materials are exposed to elevated temperatures, the
strengthening that resulted from the plastic deformation can be lost. This can be a
bad thing if the strengthening is needed to support a load. However, strengthening
due to strain hardening is not always desirable, especially if the material is being
heavily formed since ductility will be lowered.
Heat treatment can be used to remove the effects of strain hardening. Three things
can occur during heat treatment:
1. Recovery
2. Recrystallization
3. Grain growth
Recovery
When a stain hardened material
is held at an elevated temperature
an increase in atomic diffusion
occurs that relieves some of the
internal strain energy. Remember
that atoms are not fixed in
position but can move around
when they have enough energy to
break their bonds. Diffusion
increases rapidly with rising
temperature and this allows
atoms in severely strained
regions to move to unstrained
positions. In other words, atoms are freer to move around and recover a normal
position in the lattice structure. This is known as the recovery phase and it results
in an adjustment of strain on a microscopic scale. Internal residual stresses are
lowered due to a reduction in the dislocation density and a movement of
dislocation to lower-energy positions. The tangles of dislocations condense into
sharp two-dimensional boundaries and the dislocation density within these areas
decrease. These areas are called subgrains. There is no appreciable reduction in the
strength and hardness of the material but corrosion resistance often improves.

Recrystallization
At a higher temperature, new, strain-free grains nucleate and grow inside the old
distorted grains and at the grain boundaries. These new grains grow to replace the
deformed grains produced by the strain hardening. With recrystallization, the
mechanical properties return to their original weaker and more ductile states.
Recrystallization depends on the temperature, the amount of time at this
temperature and also the amount of strain hardening that the material experienced.
The more strain hardening, the lower the temperature will be at which
recrystallization occurs. Also, a minimum amount (typically 2-20%) of cold work
is necessary for any amount of recrystallization to occur. The size the new grains is
also partially dependant on the amount of strain hardening. The greater the stain
hardening, the more nuclei for the new grains, and the resulting grain size will be
smaller (at least initially).

Grain Growth
If a specimen is left at the high temperature beyond the time needed for complete
recrystallization, the grains begin to grow in size. This occurs because diffusion
occurs across the grain boundaries and larger grains have less grain boundary
surface area per unit of volume. Therefore, the larger grains lose fewer atoms and
grow at the expense of the smaller grains. Larger grains will reduce the strength
and toughness of the material.
Alloying
Only a few elements are widely used commercially in their pure form. Generally,
other elements are present to produce greater strength, to improve corrosion
resistance, or simply as impurities left over from the refining process. The addition
of other elements into a metal is called alloying and the resulting metal is called an
alloy. Even if the added elements are nonmetals, alloys may still have metallic
properties.
Copper alloys were produced very early in our history. Bronze, an alloy of copper
and tin, was the first alloy known. It was easy to produce by simply adding tin to
molten copper. Tools and weapons made of this alloy were stronger than pure
copper ones. The typical alloying elements in some common metals are presented
in the table below.
Alloy Composition
Brass Copper, Zinc
Bronze Copper, Zinc, Tin
Pewter Tin, Copper, Bismuth, Antimony
Cast Iron Iron, Carbon, Manganese, Silicon
Steel Iron, Carbon (plus small amounts of other elements)
Stainless Steel Iron, Chromium, Nickel
The properties of alloys can be manipulated by varying composition. For example
steel formed from iron and carbon can vary substantially in hardness depending on
the amount of carbon added and the way in which it was processed.
When a second element is added, two basically different structural changes are
possible:
1. Solid solution strengthening occurs when the atoms of the new element form
a solid solution with the original element, but there is still only one phase.
Recall that the term phase refers to that region of space occupied by a
physically homogeneous material.
2. The atoms of the new elements form a new second phase. The entire
microstructure may change to this new phase or two phases may be present.
Solid Solution Strengthening
Solid solution strengthening involves the addition of other metallic elements that
will dissolve in the parent lattice and cause distortions because of the difference in
atom size between the parent metal and the solute metal. Recall from the section on
crystal point defects that it is possible to have substitutional impurity atoms, and
interstitial impurity atoms. A substitutional impurity atom is an atom of a different
type than the bulk atoms, which has replaced one of the bulk atoms in the lattice.
Substitutional impurity atoms are usually close in size (within approximately 15%)
to the bulk atom. Interstitial impurity atoms are much smaller than the atoms in the
bulk matrix. Interstitial impurity atoms fit into the open space between the bulk
atoms of the lattice structure.
Since the impurity atoms are smaller or larger than the surrounding atoms they
introduce tensile or compressive lattice strains. They disrupt the regular
arrangement of ions and make it more difficult for the layers to slide over each
other. This makes the alloy stronger and less ductile than the pure metal. For
example, an alloy of 30% nickel raises the cast tensile strength of copper from
25,000 PSI to 55,000 PSI.
Multiphase Metals
Still another method of strengthening the metal is adding elements that have no or
partial solubility in the parent metal. This will result in the appearance of a second
phase distributed throughout the crystal or between crystals. These secondary
phases can raise or reduce the strength of an alloy. For example, the addition of tin,
zinc, or aluminum to copper will result in an alloy with increased strength, but
alloying with lead or bismuth with result in a lower strength alloy. The properties
of a polyphase (two of more phase) material depend on the nature, amount, size,
shape, distribution, and orientation of the phases. Greek letters are commonly used
to distinguish the different solid phases in a given alloy.

Phases can be seen on a microscopic scale with an optical microscope after the
surface has been properly polished and etched. Below is a micrograph take at 125x
of lead-tin alloy composed of two phases. The light colored regions are a tin-rich
phase and the dark colored regions are a lead-rich phase.

Alloying (continued)
Phase Diagrams
As previously stated, the phase diagram is simply a map showing the structure of
phases present as the temperature and overall composition of the alloy are varied.
It is a very useful tool for understanding and controlling the structures of polyphase
materials. A binary phase diagram shows the phases formed in differing mixtures
of two elements over a range of temperatures. When an alloy exhibits more than
two phases, a different type of phase diagram must be used, such as a ternary
diagram for three phase alloys. This discussion will focus on the binary phase
diagram.
On the binary phase diagram,
compositions run from 100% Element A
on the left, through all possible mixtures,
to 100% Element B on the right. The
composition of an alloy is given in the
form A - x%B. For example, Cu - 20%Al
is 80% copper and 20% aluminum.
Weight percentages are often used to
specify the proportions of the alloying
elements, but atomic percent are
sometimes used. Weight percentages will be used throughout this text.
Alloys generally do not have a single
melting point, but instead melt (or
alternately solidify) over a range of
temperatures. At each end of the phase
diagram only one of the elements is
present (100% A or 100% B) so a
specific melting point does exists.
Additionally, there is sometimes a
mixture of the constituent elements which produces melting at a single temperature
like a pure element. This is called the eutectic point.
At compositions other than at the pure A,
pure B and the eutectic points, when the
alloy is cooled from a high temperature it
will begin to solidify at a certain
temperature but will remain in a mushy
(liquid plus solid) condition over a range
of temperatures. If experiments are
conducted over a range of compositions
to determine the temperature at which
the alloys start to solidify, this data can
be potted on the phase diagram to produce a curve. This start of solidification
curve will join the three single solidification points and is called the liquidus line.
Up to a few percent of composition, it is
possible for one element to remain
dissolve in another while both are in the
solid state. This is called solid solubility
and the solubility limit normally changes
with temperature. The extent of the solid
solubility region can be plotted onto the
phase diagram. In this example, the alpha
phase is the region of solid solution
where some of B atoms have dissolved in
a matrix of A atoms. The beta phase is the region where a small percentage of A
atoms have dissolved in a matrix of B atoms. It is important to note that some
elements have zero solid solubility in other elements. An example is
aluminum/silicon alloys, where
aluminum has zero solid solubility in
silicon.
If an alloy's composition does not place it
within the alpha or beta solid solution
regions, the alloy will become fully solid
at the eutectic temperature. The
eutectic line on the phase diagram
indicates where this transformation will
occur over the range of compositions. At
alloy compositions and temperatures
between the liquidus temperature and the eutectic temperature, a mushy mix of
either alpha or beta phase will exist as solid masses within a liquid mixture of A
and B. These are the alpha plus liquid and the beta plus liquid areas on the phase
diagram. The region below the eutectic line, and outside the solid solution region,
will be a solid mixture of alpha and beta.
Alloying (continued)
Tie and Lever Rules
Simply by looking at a phase diagram it is possible to tell what phase or phases an
alloy will have at a given temperature. But, it is also possible to get quantitative
information from the diagram. Consider the alloy at the temperature shown on the
phase diagram. It is easy to see that at
this temperature, it is a mixture of alpha
and liquid phases. Using a tie line it is
also possible to determine the
composition of the phases at this
temperature. A tie line is an isothermal
(constant temperature) line drawn
through the alloy's position on the phase
diagram when it is in a two phase field.
The points where the ends of the tie line
intersect the two adjacent solubility
curves indicate the compositions of the two phases that exist in equilibrium at this
temperature. In this example, the tie line shows that the alpha phase is 5.2%B and
the liquid phase is 34.5%B at this temperature. It is important to keep in mind that
the tie rule addresses the determination of the compositions of the constituent
phases within the sample and it does not address the overall chemical composition
of the sample, which remains unchanged.
It is also possible to determine how much of each phase exists at the given
temperature using the lever rule. It is important to know the amounts of each phase
present because the properties of the alloy depend on the amount of each phase
present. The lever rule uses the tie line and the basic scientific principle of the
conservation of mass to determine the ratio of the two phases present. The tie-line
gives the chemical compositions of each of the two phases, and the combined
amounts of these two compositions must add up to the alloy's overall composition
(C
o
), which is known. In other words, C
o
must be composed of the appropriate
amount of at composition C

and of liquid at C
liq
. So basically, the proportions of
the phases present are given by the relative lengths of the two sections of the tie
line.
The fraction of alpha phase present is the given by the ratio of the C
o
to C
liq
portion
of the tie line and the total length of the tie line (C
liq
to C

). Mathematically the
relationships can be written as f

<< (C
liq
C
o
)/(C
liq
- C

). The fraction of liquid


phase present is given by the ratio of the C
o
to C

portion of the tie line and the


total length of the tie line (C
liq
to C

). Mathematically this relationships can be


written as f
liq
<< (C
o
- C

)/(C
liq
- C

). Of course, the two values must total to equal


one.
Note that the right side of the tie line gives the proportion of the phase on the left
( phase in this example) and left side of the tie line gives the proportion of the
phase to the right (liquid phase in this example). It is easy to keep this relationship
straight by simply considering what the ratio would be near one of the tie line
intersect points. For example, if C
o
were near the liquidus line the ratio of the
liquid section of the line to the total length of the line will be nearly one.
Alloying (continued)
Composition, Microstructure, and the Phase Diagram
Lets finish this discussion on phase diagrams by briefly looking at three different
compositions of elements A and B, and how their microstructures will differ
because of their positions on the phase diagram. First a eutectic alloy, which is an
alloy with composition right at the eutectic point, will be considered. Then
compositions on both sides of the eutectic point will be discussed. An alloy with a
composition that lies to the left of the eutectic point on the phase diagram is called
a hypoeutectic alloy, and an alloy with a composition that lies to the right of the
eutectic point is called hypereutectic alloy. At this point, only the condition of slow
cooling, which will allow the alloy to solidify into it equilibrium condition, will be
considered. The microstructure can be controlled by manipulating the speed of
cooling the alloy, but this will be covered in the section on heat treatments.
Eutectic Alloys
First, consider the eutectic alloy of
elements A and B as it is cooled from a
temperature at location 1 to location 4 on
the phase diagram. At location 1, the
alloy is at a high enough temperature to
make the mixture fully liquid. The circles
below show a representation of the
alloy's microstructure at each of the
locations numbered on the phase
diagram.
At location 1, there is nothing of interest as the alloy is completely liquid. As the
alloy is slow cooled, it remains liquid until it reaches the eutectic temperature
(location 2) where it starts to solidify at any favorable nucleation sites. From the
microstructure image 2, it can be see that as the alloy solidifies it forms into
alternate layers of alpha and beta phase. This layered microstructure is known as
lamellar microstructure and the layers are often only of the order of 1 micron
across. The reason that a eutectic alloy forms in this way has to do with the
diffusion times required to form the solid.

The grains grow by adding alpha to alpha and beta to beta until they encounter
another grain (location 3). Further nucleation sites will also continue to form
within the liquid parts of the mixture. This solidification happens very rapidly as
any given volume of liquid in the melt reaches the eutectic temperature. Remember
that a eutectic composition solidifies at a single temperature like a pure element
and not over a temperature range.
As the now sold alloy cools to location 4, the composition of the layers of alpha
and beta continue to change as it cools. Atoms of A and B will diffuse between the
two phases to produce the equilibrium compositions of alpha and beta phase at a
given temperature. By drawing tie lines at various temperatures the eutectic point
on the phase diagram, it can be seen that the solubility of A in the beta phase and B
in the alpha phase decreases as the temperature decreases. Since this phase
composition change is due to diffusion, which is a relatively a slow process), it is
important that eutectic alloys be allowed to cool slowly to produce the correct
microstructure.
Hypoeutectic Alloys
Next, consider an alloy of A and B that
has an overall composition that places it
to the left of the eutectic point. When an
alloy falls to the left of the eutectic point
it is called a hypoeutectic alloy. At
location 1, the alloy is at a temperature
that is high enough to put it in a fully
liquid phase.
When the alloy is cooled, it remains in
the liquid state until it reaches the temperature where it crosses the liquidus line
(location 2). At this temperature, the alpha phase starts to solidify at any favorable
nucleation sites. The alpha solidifies as dendrites which grow to become grains of
alpha. The first solid phase to form is called the primary phase so, in this case,
primary alpha is formed.

As the alloy continues to cool (location 3) the existing nucleation sites will grow as
dendrites and further nucleation sites will form within the liquid part of the
mixture. The melt will have that mushy consistency of chunks in liquid while it is
in the alpha + liquid region of the phase diagram. Since the alpha phase is mostly
element A (with a small amount of B atoms in solid solution), the remaining liquid
becomes slightly richer in B as the liquid cools, which is indicated by the liquidus
line. The composition of the solid alpha phase also becomes slightly richer in B
atoms as the solid solution line shows.
This primary alpha phase growth and the accompanying phase composition shifts
continue until enough A atoms have been removed so that the remaining liquid is
of eutectic composition. This composition is achieved at the point where the
temperature crosses the eutectic line (location 4). At this point the primary alpha
phase stops forming. The remaining liquid starts to solidify into the lamellar
(alternating layers of alpha and beta phases) structure of a eutectic composition.
The eutectic structure will grow; adding alpha to the layers of alpha and beta to the
layers of beta in the eutectic regions, and new solidification sites will continue to
form. Remember that solidification occurs rapidly and without the need for a
further decrease in temperature once the liquid reaches the eutectic line. At this
point, the entire alloy has solidified into a mixture comprised of grains of alpha and
grains of eutectic mixture (alpha and beta). The microstructure from this point at
the eutectic line down to ambient temperature will look something like that shown
in micro 5.
Diffusion occurs as the alloy cools since the amount of element B in the alpha
phase changes with temperature. This occurs exactly like it did for the eutectic
alloy. Diffusion must also occur in the grains of pure alpha, as the composition of
alpha phase also changes with temperature.
Hypereutectic
Finally, consider an alloy of A and B that
has an overall composition that places it
to the right of the eutectic point. When
an alloy falls to the right of the eutectic
point it is called a hypereutectic alloy.
This alloy will solidify like the
hypoeutectic alloy did except it will pass
through the beta + liquid region of the
phase diagram rather than the alpha + liquid region. This will result in a
microstructure comprised of grains of beta and grains of eutectic mixture (alpha
and beta) rather than grains of alpha and grains of eutectic mixture (alpha and beta)
as the hypoeutectic alloy had.
At location 1, the alloy is at a temperature that is high enough to put it in a fully
liquid phase. When the alloy is cooled, it remains in the liquid state until it reaches
the temperature where it crosses the liquidus line (location 2). At this temperature,
the beta phase starts to solidify at any favorable nucleation sites. The beta solidifies
as dendrites which grow to become grains of beta. The first solid phase to form is
called the primary phase so, in this case, primary beta is formed.

As the alloy continues to cool (location 3) the existing nucleation sites will grow as
dendrites and further nucleation sites will form within the liquid part of the
mixture. Since the beta phase is mostly element B (with a small amount of A atoms
in solid solution), the remaining liquid becomes richer in A as the liquid cools,
which is indicated by the liquidus line. The composition of the solid beta phase
also becomes slightly richer in A atoms as the solid solution line shows.
This primary beta phase growth and the accompanying phase composition shifts
continue until enough B atoms have been removed so that the remaining liquid is
of eutectic composition. This composition is achieved at the point where the
temperature crosses the eutectic line (location 4). At this point the primary beta
phase stops forming. The remaining liquid starts to solidify into the lamellar
(alternating layers of alpha and beta phases) structure of a eutectic composition.
The eutectic structure will grow; adding alpha to the layers of alpha and beta to the
layers of beta in the eutectic regions, and new solidification sites will continue to
form. At this point, the entire alloy quickly solidifies into a mixture of beta grains
and eutectic mixture (alpha and beta) grains. The microstructure from this point at
the eutectic line down to ambient temperature will look something like that shown
in micro 5.
Diffusion occurs as the alloy cools since the amount of element B in the alpha
phase changes with temperature. This occurs exactly like it did for the eutectic
alloy. Diffusion must also occur in the grains of pure alpha, as the composition of
alpha phase also changes with temperature.

Thermal Treatments (Heat-Treating)
In the previous pages on the subjects of alloying and the binary phase diagram, the
microstructures of alloys that were allowed to solidify by slow cooling were
considered. It should also be known, however, that it is possible to modify the
microstructure of an alloy by subjecting it to various thermal treatments. Heat-
treating is a term used to describe all of the controlled heating and cooling
operations performed on a material in the solid state for the purpose of altering its
microstructure and/or properties. The focus of this discussion will be on metals but
is should be noted that heat-treatment is also used on ceramics and composites to
modify their properties.
The major objectives of the different kinds of thermal treatments are:
1. Soften the material for improved workability.
2. Increase the strength or hardness of the material.
3. Increase the toughness or resistance to fracture of the material.
4. Stabilize mechanical or physical properties against changes that might occur
during exposure to service environments.
5. Insure part dimensional stability.
6. Relieve undesirable residual stresses induced during part fabrication.
Different metals respond to treatment at different temperatures. Each metal has a
specific chemical composition, so changes in physical and structural properties
take place at different, critical temperatures. Even small percentages of elements in
the metal composition, such as carbon, will greatly determine the temperature,
time, method and rate of cooling that needs to be used in the heat treating process.
Depending on the thermal treatment used, the atomic structure and/or
microstructure of a material may change due to movement of dislocations, an
increase or decrease in solubility of atoms, an increase in grain size, the formation
of new grains of the same or different phase, a change in the crystal structure, and
others mechanisms.

Since there are so many ways in which metals are heat treated, it is not practical to
discuss them all. But, as an example, lets look at how heat treatment is used to
strengthen a copper aluminum alloy.
Precipitation Hardening
In designing alloys for strength, an approach often taken is to develop an alloy with
a structure that consists of particles (which impede dislocation movement)
dispersed in a ductile matrix. Such a dispersion can be obtained by choosing an
alloy that is a single phase at elevated temperature but on cooling will precipitate
another phase in the matrix. A thermal process is then developed to produce the
desired distribution of precipitate in the matrix. When the alloy is strengthened by
this thermal treatment, it is called precipitation strengthening or hardening.
Precipitation hardening consists of three main steps: solution treatment, quenching,
and aging. Solution treatment involves heating the alloy to a temperature that
allows the alloying atoms (called the solute) to dissolve into the solution. This
results in a homogeneous solid solution of one phase. Quenching rapidly cools the
solution and freezes the atoms in solution. In more technical terms, the quenching
cools the material so fast that the atoms of the alloying elements do not have time
to diffuse out of the solution. In the as-quenched condition, the solute is
supersaturated meaning that the lattice is overly stressed by the alloying atoms.
Aging is the process where the solute particles diffuse out of solution and into
clusters that distort and strengthen the material.
The precipitation hardening process for a copper-aluminum alloy is shown
graphically in the image below. On the right is phase diagram, which is a very
useful tool for understanding and controlling polyphase structures. The phase
diagram is simply a map showing the structure of phases present as the temperature
and overall composition of the alloy are varied. The images on the right in the
image show the resulting microstructure at each step in the process.

Common Heat Treating Processes
A few of the more common terms used in heat treating are introduced below. It
should be noted that not all of the term are applicable to all alloys.
Age Hardening is a relatively low-temperature heat treatment process that
strengthens a material by causing the precipitation of components or phases of
alloy from a super-saturated solid solution condition.
Annealing is a softening process in which metals are heated and then allowed to
cool slowly. The purpose of annealing is to soften the material for improve
machinability, formability, and sometimes to control magnetic properties.
Normallizing is much like annealing, but the cooling process is much faster. This
results in increased strength but less ductility in the metal. Its purpose is to refine
grain structure, produce more uniform mechanical properties, and sometimes to
relieve internal and surface stresses.
Precipitation Heat Treatment is the three step process of solution treating,
quenching, and age hardening to increase the strength or hardness of an alloy.
Solution Heat Treatment involves heating the material to a temperature that puts
all the elements in solid solution and then cooling very rapidly to freeze the atoms
in place.
Stress Relieving is a low temperature heat treat process that is used to reduce the
level of residual stresses in a material.
Tempering involves gently heating a hardened metal and allowing it to cool
slowly will produce a metal that is still hard but also less brittle. This process is
known as tempering.
Quenching is the rapid cooling of a hot material. The medium used to quench the
material can vary from forced air, oil, water and others. Many steels are hardened
by heating and quenching. Quenching results in a metal that is very hard but also
brittle.
More information on heat treatment can be found in the material (ie aluminum,
steel, titanium, etc.) sections
Ceramic Structures
As discussed in the introduction, ceramics and related materials cover a wide range
of objects. Ceramics are a little more complex than metallic structures, which is
why metals were covered first. A ceramic has traditionally been defined as an
inorganic, nonmetallic solid that is prepared from powdered materials and is
fabricated into products through the application of heat. Most ceramics are made
up of two or more elements. This is called a compound. For example, alumina
(Al2O3) is a compound made up of aluminum atoms and oxygen atoms.
The two most common chemical bonds for ceramic materials are covalent and
ionic. The bonding of atoms together is much stronger in covalent and ionic
bonding than in metallic. This is why ceramics generally have the following
properties: high hardness, high compressive strength, and chemical inertness. This
strong bonding also accounts for the less attractive properties of ceramics, such as
low ductility and low tensile strength. The absence of free electrons is responsible
for making most ceramics poor conductors of electricity and heat.
However, it should be noted that the crystal structures of ceramics are many and
varied and this results in a very wide range of properties. For example, while
ceramics are perceived as electrical and thermal insulators, ceramic oxide (initially
based on Y-Ba-Cu-O) is the basis for high temperature superconductivity.
Diamond and silicon carbide have a higher thermal conductivity than aluminum or
copper. Control of the microstructure can overcome inherent stiffness to allow the
production of ceramic springs, and ceramic composites which have been produced
with a fracture toughness about half that of steel. Also, the atomic structures are
often of low symmetry that gives some ceramics interesting electromechanical
properties like piezoelectricity, which is used in sensors and transducers.
The structure of most ceramics varies from relatively simple to very complex. The
microstructure can be entirely glassy (glasses only); entirely crystalline; or a
combination of crystalline and glassy. In the latter case, the glassy phase usually
surrounds small crystals, bonding them together. The main compositional classes
of engineering ceramics are the oxides, nitrides and carbides.
Ceramic Structures (continued)
Ceramic Glass
Ceramics with an entirely glassy structure have certain properties that are quite
different from those of metals. Recall that when metal in the liquid state is cooled,
a crystalline solid precipitates when the melting freezing point is reached.
However, with a glassy material, as the liquid is cooled it becomes more and more
viscous. There is no sharp melting or freezing point. It goes from liquid to a soft
plastic solid and finally becomes hard and brittle. Because of this unique property,
it can be blown into shapes, in addition to being cast, rolled, drawn and otherwise
processed like a metal.
Glassy behavior is related to the atomic structure of the material. If pure silica
(SiO
2
) is fused together, a glass called vitreous silica is formed on cooling. The
basic unit structure of this glass is the silica tetrahedron, which is composed of a
single silicon atom surrounded by four equidistant oxygen atoms. The silicon
atoms occupy the openings (interstitials) between the oxygen atoms and share four
valence electrons with the oxygen atoms through covalent bonding. The silica
atom has four valence electrons and each of the oxygen atoms has two valence
electrons so the silica tetrahedron has four extra valence electrons to share with
adjacent tetrahedral. The silicate structures can link together by sharing the atoms
in two corners of the SiO
2
tetrahedrons, forming chain or ring structures. A
network of silica tetrahedral chains form, and at high temperatures these chains
easily slide past each other. As the melt cools, thermal vibrational energy decreases
and the chains can not move as easily so the structure becomes more rigid. Silica is
the most important constituent of glass, but other oxides are added to change
certain physical characteristics or to lower the melting point.
Ceramic Crystalline or Partially Crystalline Material
Most ceramics usually contain both metallic and nonmetallic elements with ionic
or covalent bonds. Therefore, the structure the metallic atoms, the structure of the
nonmetallic atoms, and the balance of charges produced by the valence electrons
must be considered. As with metals, the unit cell is used in describing the atomic
structure of ceramics. The cubic and the hexagonal cells are most common.
Additionally, the difference in radii between the metallic and nonmetallic ions
plays an important role in the arrangement of the unit cell.
In metals, the regular arrangement of atoms into densely packed planes led to the
occurrence of slip under stress, which gives metal their characteristic ductility. In
ceramics, brittle fracture rather than slip is common because both the arrangement
of the atoms and the type of bonding is different. The fracture or cleavage planes
of ceramics are the result of planes of regularly arranged atoms.
The building criteria for the crystal structure are:
maintain neutrality
charge balance dictates chemical formula
achieve closest packing
A few of the different types of ceramic materials outside of the glass family are
described below.
Silicate Ceramics
As mentioned previously, the silica structure is the
basic structure for many ceramics, as well as glass.
It has an internal arrangement consisting of
pyramid (tetrahedral or four-sided) units. Four
large oxygen (0) atoms surround each smaller
silicon (Si) atom. When silica tetrahedrons share
three corner atoms, they produce layered silicates
(talc, kaolinite clay, mica). Clay is the basic raw
material for many building products such as brick
and tile. When silica tetrahedrons share four comer
atoms, they produce framework silicates (quartz,
tridymite). Quartz is formed when the tetrahedra in this material are arranged in a
regular, orderly fashion. If silica in the molten state is cooled very slowly it
crystallizes at the freezing point. But if molten silica is cooled more rapidly, the
resulting solid is a disorderly arrangement which is glass.
Cement
Cement (Portland cement) is one of the main ingredients of concrete. There are a
number of different grades of cement but a typical Portland cement will contain 19
to 25% SiO
2
, 5 to 9% Al
2
O
3
, 60 to 64% CaO and 2 to 4% FeO. Cements are
prepared by grinding the clays and limestone in proper proportion, firing in a kiln,
and regrinding. When water is added, the minerals either decompose or combine
with water, and a new phase grows throughout the mass. The reaction is solution,
recrystallization, and precipitation of a silicate structure. It is usually important to
control the amount of water to prevent an excess that would not be part of the
structure and would weaken it. The heat of hydration (heat of reaction in the
adsorption of water) in setting of the cement can be large and can cause damage in
large structures.
Nitride Ceramics
Nitrides combine the superior hardness of
ceramics with high thermal and mechanical
stability, making them suitable for applications as
cutting tools, wear-resistant parts and structural
components at high temperatures. TiN has a cubic
structure which is perhaps the simplest and best
known of structure types. Cations and anions both
lie at the nodes of separate fcc lattices. The
structure is unchanged if the Ti and N atoms
(lattices) are interchanged.
Ferroelectric Ceramics
Depending on the crystal structure, in some crystal
lattices, the centers of the positive and negative
charges do not coincide even without the
application of external electric field. In this case, it
is said that there exists spontaneous polarization in
the crystal. When the polarization of the dielectric
can be altered by an electric field, it is called
ferroelectric. A typical ceramic ferroelectric is
barium titanate, BaTiO
3
. Ferroelectric materials,
especially polycrystalline ceramics, are very
promising for varieties of application fields such as
piezoelectric/electrostrictive transducers, and electrooptic.
Phase Diagram
The phase diagram is important in understanding the formation and control of the
microstructure of the microstructure of polyphase ceramics, just as it is with
polyphase metallic materials. Also, nonequilibrium structures are even more
prevalent in ceramics because the more
complex crystal structures are more
difficult to nucleate and to grow from
the melt.
Imperfections in Ceramics
Imperfections in ceramic crystals
include point defects and impurities like
in metals. However, in ceramics defect formation is strongly affected by the
condition of charge neutrality because the creation of areas of unbalanced charges
requires an expenditure of a large amount of energy. In ionic crystals, charge
neutrality often results in defects that come as pairs of ions with opposite charge or
several nearby point defects in which the sum of all charges is zero. Charge neutral
defects include the Frenkel and Schottky defects. A Frenkel-defect occurs when a
host atom moves into a nearby interstitial position to create a vacancy-interstitial
pair of cations. A Schottky-defect is a pair of nearby cation and anion vacancies.
Schottky defect occurs when a host atom leaves its position and moves to the
surface creating a vacancy-vacancy pair.
Sometimes, the composition may alter slightly to arrive at a more balanced atomic
charge. Solids such as SiO
2
, which have a well-defined chemical formula, are
called stoichiometric compounds. When the composition of a solid deviates from
the standard chemical formula, the resulting solid is said to be nonstoichiometric.
Nonstoichiometry and the existence of point defects in a solid are often closely
related. Anion vacancies are the source of the nonstoichiometry in SiO
2-x
,
Introduction of impurity atoms in the lattice is likely in conditions where the
charge is maintained. This is the case of electronegative impurities that substitute a
lattice anion or electropositive substitutional impurities. This is more likely for
similar ionic radii since this minimizes the energy required for lattice distortion.
Defects will appear if the charge of the impurities is not balanced.
Polymer Structure
Engineering polymers include natural materials such as rubber and synthetic
materials such as plastics and elastomers. Polymers are very useful materials
because their structures can be altered and tailored to produce materials 1) with a
range of mechanical properties 2) in a wide spectrum of colors and 3) with
different transparent properties.
Mers
A polymer is composed of many
simple molecules that are
repeating structural units called
monomers. A single polymer
molecule may consist of
hundreds to a million monomers
and may have a linear,
branched, or network structure.
Covalent bonds hold the atoms in the polymer molecules together and secondary
bonds then hold groups of polymer chains together to form the polymeric material.
Copolymers are polymers composed of two or more different types of monomers.
Mer
The repeating unit in a polymer chain
Monomer
A single mer unit (n=1)
Polymer
Many mer-units along a chain (n=10
3
or more)
Degree of Polymerization
The average number of mer-units in a chain.
Polymer Chains (Thermoplastics and Thermosets)
A polymer is an organic material and the backbone of every organic material is a
chain of carbon atoms. The carbon atom has four electrons in the outer shell. Each
of these valence electrons can form a covalent bond to another carbon atom or to a
foreign atom. The key to the polymer structure is that two carbon atoms can have
up to three common bonds and still bond with other atoms. The elements found
most frequently in polymers and their valence numbers are: H, F, Cl, Bf, and I with
1 valence electron; O and S with 2 valence electrons; n with 3 valence electrons
and C and Si with 4 valence electrons.
The ability for molecules to form long
chains is a vital to producing polymers.
Consider the material polyethylene,
which is made from ethane gas, C
2
H
6
.
Ethane gas has a two carbon atoms in
the chain and each of the two carbon
atoms share two valence electrons with
the other. If two molecules of ethane are
brought together, one of the carbon
bonds in each molecule can be broken
and the two molecules can be joined
with a carbon to carbon bond. After the two mers are joined, there are still two free
valence electrons at each end of the chain for joining other mers or polymer chains.
The process can continue liking more mers and polymers together until it is
stopped by the addition of anther chemical (a terminator), that fills the available
bond at each end of the molecule. This is called a linear polymer and is building
block for thermoplastic polymers.

The polymer chain is often shown in two dimensions, but it should be noted that
they have a three dimensional structure. Each bond is at 109 to the next and,
therefore, the carbon backbone extends through space like a twisted chain of
TinkerToys. When stress is applied, these chains stretch and the elongation of
polymers can be thousands of times greater than it is in crystalline structures.

The length of the polymer chain is very important. As the number of carbon atoms
in the chain is increased to beyond several hundred, the material will pass through
the liquid state and become a waxy solid. When the number of carbon atoms in the
chain is over 1,000, the solid material polyethylene, with its characteristics of
strength, flexibility and toughness, is obtained. The change in state occurs because
as the length of the molecules increases, the total binding forces between
molecules also increases.

It should also be noted that the molecules are not generally straight but are a
tangled mass. Thermoplastic materials, such as polyethylene, can be pictured as a
mass of intertwined worms randomly thrown into a pail. The binding forces are the
result of van der Waals forces between molecules and mechanical entanglement
between the chains. When thermoplastics are heated, there is more molecular
movement and the bonds between molecules can be easily broken. This is why
thermoplastic materials can be remelted.
There is another group of polymers in which a
single large network, instead of many molecules
is formed during polymerization. Since
polymerization is initially accomplished by
heating the raw materials and brining them
together, this group is called thermosetting
polymers or plastics. For this type of network structure to form, the mers must
have more than two places for boning to occur; otherwise, only a linear structure is
possible. These chains form jointed structures and rings, and may fold back and
forth to take on a partially crystalline structure.
Since these materials are essentially comprised of one giant molecule, there is no
movement between molecules once the mass has set. Thermosetting polymers are
more rigid and generally have higher strength than thermoplastic polymers. Also,
since there is no opportunity for motion between molecules in a thermosetting
polymer, they will not become plastic when heated.

Types of polymers
o Commodity plastics
PE = Polyethylene
PS = Polystyrene
PP = Polypropylene
PVC = Poly(vinyl chloride)
PET = Poly(ethylene terephthalate)
o Specialty or Engineering Plastics
Teflon (PTFE) = Poly(tetrafluoroethylene)
PC = Polycarbonate (Lexan)
Polyesters and Polyamides (Nylon)
Composite
Structures
A composite material is
basically a combination
of two or more materials,
each of which retains it
own distinctive
properties. Multiphase
metals are composite
materials on a micro scale, but generally the term composite is applied to
materials that are created by mechanically bonding two or more different
materials together. The resulting material has characteristics that are not
characteristic of the components in isolation. The concept of composite
materials is ancient. An example is adding straw to mud for building
stronger mud walls. Most commonly, composite materials have a bulk
phase, which is continuous, called the matrix; and a dispersed, non-
continuous, phase called the reinforcement. Some other examples of basic
composites include concrete (cement mixed with sand and aggregate),
reinforced concrete (steel rebar in concrete), and fiberglass (glass strands in
a resin matrix).
In about the mid 1960s, a
new group of composite
materials, called advanced
engineered composite
materials (aka advanced
composites), began to
emerge. Advanced
composites utilize a
combination of resins and
fibers, customarily
carbon/graphite, kevlar, or fiberglass with an epoxy resin. The fibers provide
the high stiffness, while the surrounding polymer resin matrix holds the
structure together. The fundamental design concept of composites is that the
bulk phase accepts the load over a large surface area, and transfers it to the
reinforcement material, which can carry a greater load. The significance
here lies in that there are numerous matrix materials and as many fiber
types, which can be combined in countless ways to produce just the desired
Components of Composite Materials
Matrix phase: bulk materials such as:
Metals Ceramics Polymers
Reinforcement: fibers and particulates such as:
Glass Carbon Kevlar
Silicon Carbide Boron Ceramic
Ceramic Metallic Aggregate
I nterface: area of mechanical
properties. These materials were first developed for use in the aerospace
industry because for certain application they have a higher stiffness to
weight or strength-to-weight ratio than metals. This means metal parts can
be replaced with lighter weight parts manufactured from advanced
composites. Generally, carbon-epoxy composites are two thirds the weight
of aluminum, and two and a half times as stiff. Composites are resistant to
fatigue damage and harsh environments, and are repairable.
Composites meeting the criteria of having mechanical bonding can also be
produced on a micro scale. For example, when tungsten carbide powder is
mixed with cobalt powder, and then pressed and sintered together, the
tungsten carbide retains its identity. The resulting material has a soft cobalt
matrix with tough tungsten carbide particles inside. This material is used to
produce carbide drill bits and is called a metal-matrix composite. A metal
matrix composite is a type of metal that is reinforced with another material
to improve strength, wear or some other characteristics.
Composite Structures (continued)
Classification of Composite Materials
Since the reinforcement material is of primary importance in the strengthening
mechanism of a composite, it is convenient to classify composites according to the
characteristics of the reinforcement. The following three categories are commonly
used.
1. Fiber Reinforced In this group of composites, the fiber is the primary load-
bearing component.
2. Dispersion Strengthened In this group, the matrix is the major load-
bearing component.
3. Particle Reinforced In this group, the load is shared by the matrix and the
particles.
Fiber Reinforced Composites
Fiberglass is likely the best know fiber reinforced composite but carbon-epoxy and
other advanced composites all fall into this category. The fibers can be in the form
of long continuous fibers, or they can be discontinuous fibers, particles, whiskers
and even weaved sheets. Fibers are usually combined with ductile matrix materials,
such as metals and polymers, to make them stiffer, while fibers are added to brittle
matrix materials like ceramics to increase toughness. The length-to diameter ratio
of the fiber, the strength of the bond between the fiber and the matrix, and the
amount of fiber are variables that affect the mechanical properties. It is important
to have a high length-to-diameter aspect ratio so that the applied load is effectively
transferred form the matrix to the fiber.
Fiber materials include:
Glass glass is the most common and inexpensive fiber and is usually use for the
reinforcement of polymer matrices. Glass has a high tensile strength and fairly low
density (2.5 g/cc).
Carbon-graphite - in advance composites, carbon fibers are the material of
choice. Carbon is a very light element, with a density of about 2.3 g/cc and its
stiffness is considerable higher than glass. Carbon fibers can have up to 3 times the
stiffness of steel and up to 15 times the strength of construction steel. The graphitic
structure is preferred over the diamond-like crystalline forms for making carbon
fiber because the graphitic structure is made of densely packed hexagonal layers,
stacked in a lamellar style. This structure results in mechanical and thermal
properties are highly anisotropic and this gives component designers the ability to
control the strength and stiffness of components by varying the orientation of the
fiber.
Polymer the strong covalent bonds of polymers can lead to impressive properties
when aligned along the fiber axis of high molecular weight chains. Kevlar is an
aramid (aromatic polyamide) composed of oriented aromatic chains, which makes
them rigid rod-like polymers. Its stiffness can be as high as 125 GPa and although
very strong in tension, it has very poor compression properties. Kevlar fibers are
mostly used to increase toughness in otherwise brittle matrices.
Ceramic fibers made from materials such as Alumina and SiC (Silicon carbide)
are advantageous in very high temperature applications, and also where
environmental attack is an issue. Ceramics have poor properties in tension and
shear, so most applications as reinforcement are in the particulate form.
Metallic - some metallic fibers have high strengths but since there density is very
high they are of little use in weight critical applications. Drawing very thin metallic
fibers (less than 100 micron) is also very expensive.
Dispersion Strengthen Composites
In dispersion strengthened composites, small particles on the order of 10-5 mm to
2.5 x 10-4 mm in diameter are added to the matrix material. These particles act to
help the matrix resist deformation. This makes the material harder and stronger.
Consider a metal matrix composite with a fine distribution of very hard and small
secondary particles. The matrix material is carrying most of the load and
deformation is accomplished by slip and dislocation movement. The secondary
particles impede slip and dislocation and, thereby, strengthen the material. The
mechanism is that same as precipitation hardening but effect is not quite as strong.
However, particles like oxides do not react with the matrix or go into solution at
high temperatures so the strengthening action is retained at elevated temperatures.
Particle Reinforced Composites
The particles in these composite are larger than in dispersion strengthened
composites. The particle diameter is typically on the order of a few microns. In this
case, the particles carry a major portion of the load. The particles are used to
increase the modulus and decrease the ductility of the matrix. An example of
particle reinforced composites is an automobile tire which has carbon black
particles in a matrix of polyisobutylene elastomeric polymer. Particle reinforced
composites are much easier and less costly than making fiber reinforced
composites. With polymeric matrices, the particles are simply added to the
polymer melt in an extruder or injection molder during polymer processing.
Similarly, reinforcing particles are added to a molten metal before it is cast.
Interface
1. The interface is a bounding surface or zone where a discontinuity occurs,
whether physical, mechanical, chemical etc.
2. The matrix material must "wet" the fiber. Coupling agents are frequently
used to improve wettability. Well "wetted" fibers increase the interface
surface area.
3. To obtain desirable properties in a composite, the applied load should be
effectively transferred from the matrix to the fibers via the interface. This
means that the interface must be large and exhibit strong adhesion between
fibers and matrix. Failure at the interface (called debonding) may or may not
be desirable. This will be explained later in fracture propagation modes.
4. Bonding with the matrix can be either weak van der Walls forces or strong
covalent bonds.
5. The internal surface area of the interface can go as high as 3000 cm2/cm3.
6. Interfacial strength is measured by simple tests that induce adhesive failure
between the fibers and the matrix. The most common is the Three-point
bend test or ILSS (interlaminar shear stress test)
We will consider the results of incorporating fibers in a matrix. The matrix, besides
holding the fibers together, has the important function of transferring the applied
load to the fibers. It is of great importance to be able to predict the properties of a
composite, given the component properties and their geometric arrangement.
Isotropy and Anisotropy in Composites
1. Fiber reinforced composite materials typically exhibit anisotropy. That is,
some properties vary depending upon which geometric axis or plane they
are measured along.
2. For a composite to be isotropic in a specific property, such as CTE or
Youngs modulus, all reinforcing elements, whether fibers or particles, have
to be randomly oriented. This is not easily achieved for discontinuous fibers,
since most processing methods tend to impart a certain orientation to the
fibers.
3. Continuous fibers in the form of sheets are usually used to deliberately make
the composite anisotropic in a particular direction that is known to be the
principally loaded axis or plane.

Das könnte Ihnen auch gefallen