You are on page 1of 12

STRUCTURE OF ENGINEERING MATERIALS 5

1. S TRUCTURE OF E NGINEERING M ATERIALS

As a consequence of internal factors (nature and magnitude of interaction forces) and external
factors (temperature, pressure, nature and magnitude of fields) a lot of atoms can exist in one
of the following states of aggregation: solid, liquid, gas, plasma and hyper-dense nonradiative
material (black holes).

Function of their constitutive element arrangement degree, the solid bodies have three
structures: amorphous, mezzomorphous and crystals. Different by the others, the crystalline
structure is an aggregation of atoms or molecules with a regular internal structure and the
external form of a solid enclosed by symmetrically arranged plane faces. The amorphous
structure is obtained from super-cooling of liquids and it is not an equilibrium state because
they gradually degenerate in crystalline bodies. The mezzomorphous state is between the
others, like liquid crystals. According to these reasons only the crystal bodies can be named
solid bodies.

1.1. Crystal Structure of Metals

For the chemist, an element is a metal if its oxide dissolved in water produces an alkaline
solution. For the physicist an element is a metal if it displays good electrical conductivity, which
decreases with increasing of temperature. The metallurgist, being primarily concerned with the
mechanical, electrical, and magnetic properties of materials, generally regards an element as a
metal if it contains the physicist's requirements and also can be plastically deformed to some
extent. In fact, the metallurgist's principal business is to find out how to modify these metallic
properties through control of the composition and structure of metals. An engineer, who
understands the degree to which metallic properties can be altered, can utilize metals and
alloys for structural and other uses more intelligently than by relying solely on handbooks.

The term unalloyed metals is chosen here rather than pure metals since purity of
metals is never absolute and varies from 99,999% in some refined metals to 98,0% or less of
the element in some metals not subjected to expensive and tedious refining processes.

Because they have relatively high densities, metals must consist of atoms, which are
packed very closely together. From the first approximation it is permissible to think of the atoms
of a metal as round, hard spheres that exert attractive forces in all directions. Given a number
of these hard-sphere atoms, they can be arranged to be close-packed and to occupy a
minimum volume.

Sometimes it is more convenient to single out a small group of atoms in the stack and
then describe the atoms' arrangement in this group. The group of atoms chosen for this
purpose is called the unit cell of the structure. One is at liberty to choose one of a number of
6 TECHNICAL ENGINEERING

equally accurate ways of representing the unit cell, but experience has shown that for each
structure there is one cell which is most easily visualized and which best shows the properties
of the material.

The great majority of the metals crystallize in one of the systems:

• Body-Centered Cubic (BCC);

• Face-Centered Cubic (FCC);

• Hexagonal Close-Packed (HCP).

Many of the important metals, including iron α, chromium and tungsten, have the BCC
structure (Fig. 1.1). They have a high tensile strength and a moderate plasticity. The unit cell of
BCC structure contains two atoms.

Figure 1.1. The body-centered cubic (BCC) structure

The face-centered cubic (FCC) structure (Fig. 1.2) is proper to Cu, Ag, Au, Al, Pb, Ni,
Feγ, which have great ductility and malleability. The unit cell of FCC structure contains four
atoms.

The hexagonal close-packed (HCP) structure (Fig. 1.3) gives a reduced plasticity. Be,
Mg, Zn, Cd, Ti are examples of HCP metals.

The structures of a number of metals deviate more markedly from close packing.
Bismuth, antimony, and gallium have an atomic arrangement with the symmetry of a
rhombohedron rather than cube. These metals have so-called open structures, meaning that

Figure 1.2. The face-centered cubic (FCC) structure


STRUCTURE OF ENGINEERING MATERIALS 7

Figure 1.3. The hexagonal close-packed (HCP) structure

there is a substantial amount of empty space surrounding each atom, considered as a hard
sphere. It is interesting in this connection that these metals decrease in volume when they melt,
whereas most metals have a greater specific volume in the liquid state than in the solid state.

The term lattice is frequently used in the description of crystal structures. A lattice is an
array of points repeated regularly throughout space. One important characteristic of a metal is
its lattice parameters, the dimensions of its unit cell. In the metals having cubic symmetry the
size of the lattice is fixed when the length of the edge of the cubic unit cell is given. Cubic
metals have therefore only one lattice parameter. The lattice parameter of a metal can be
measured by observing the diffraction of an x-ray beam passed through the metal.

When the unit cell does not have cubic symmetry, more than one lattice parameter has
to be specified. In hexagonal crystals these are, first, the distance, a, between neighboring
lattice points in the close-packed planes (or basal planes, as they are often called) and,
second, the distance from the top to the bottom of the unit cell, c.

If the lattice parameters of a metal and the mass of its individual atoms are known (from
measurements with a mass spectrometer or calculated as the atomic weight over Avogadro's
number), it should be an easy matter to calculate its density. The first step in carrying out such
a calculation is to find the number of atoms contained in a volume equal to that of the unit cell.
In simple structures this number can be found by inspection: it is necessary only to imagine the
unit cell displaced slightly so that it contains the whole lattice points and then count the number
of points in it. An alternative method, which does not tax the observer's three-dimensional
visualization so much, is illustrated in the following examples. Consider a BCC unit cell. Each of
the corner atoms is shared between eight cells, and so the total number of atoms in the cell is

1 center atom + 8 × 1/8 corner atoms = 2 atoms per unit cell.

For the FCC unit cell one finds in the same way that

6 × 1/2 face atoms + 8 × 1/8 corner atoms = 4 atoms per unit cell.
8 TECHNICAL ENGINEERING

To the extent that the atoms of a metal can be considered to be hard spheres, they can
have assigned a definite diameter. The atomic diameter can be calculated when the crystal
structure and lattice parameters of a metal are known. The atoms of a crystal structure are in
contact along certain directions in the unit cell known as the close-packed directions.
Inspection of Fig. 1.2 shows, for example, that the close-packed directions in the FCC structure
are along the diagonals of the cube faces. If a is the lattice parameter, the length of a face
1
diagonal is a 2 and the atomic diameter must be a 2.
2
The hard-sphere model of the atoms of a metal is only a first approximation. It is found,
in fact, that the atoms of a given metal may not always appear to have the same diameter. For
example, iron atoms behave as if they had one diameter when in pure iron metal, a slightly
different one when in an iron-nickel alloy, and still another in ferric chloride crystals. Neverthe-
less the concept of atomic diameter has proved to be a useful one in metallurgy and plays an
important role in understanding the formation of alloys.

The adjacent atoms to a given atom along the close-packed directions of a crystal are
called the nearest neighbors of the given atom. Each atom in the FCC structure has 12 nearest
neighbors, as can be seen from Fig. 1.2. So does each atom in the HCP structure. But in the
BCC structure each atom has only 8 nearest neighbors. The number of nearest neighbors
found in a given crystal structure is named the coordination number of the structure. Thus BCC
iron has a coordination number of 8 and copper has a coordination number of 12.

The points of a crystal lattice define an array of crystal planes. Some of these planes
are easily visualized. For example, the planes which mark out the unit cells of the FCC and
BCC lattices shown in Figs. 1.1 and 1.2 constitute a set of crystal planes known as the cube
planes. Each of these planes is repeated indefinitely throughout the lattice as an array of
parallel planes separated by a distance equal to the lattice parameter in the simple cubic
structure. Each set of crystal planes has its own characteristic interplanar distance.

Because a given type of crystal plane is repeated more or less indefinitely by the lattice,
the location of a specific plane is not of much interest in crystallography. The orientation of a
plane or a set of planes, relative to the edges of the unit cell is important, however. In order to
specify this orientation, it is convenient to establish as a set of crystal axes one set of the
edges of one of the unit cells of the structure. For the FCC and BCC structures, the crystal
axes are a set of rectangular cartesian coordinates. The intercepts of a crystal plane on these
axes could be measured in units of centimeters or angstroms, but it proves to be more useful in
crystallography to use the lattice parameters themselves as the units of measure on the
crystallographic axes. Planes that are in the same orientation will then have the same
intercepts in different crystals, even though the lattice parameters are different.

Some metals exist in more than one crystalline form. Iron is BCC at temperatures
ranging up to 910°C (Fig. 2.1), but at this temperature it undergoes an allotropic transformation
and becomes FCC. The FCC phase is stable up to 1390°C, when temperature transforms it
back to the BCC structure and remains that way up to the melting point. The different
crystalline forms of a metal are named allotropic forms. Many metals such as aluminum and
STRUCTURE OF ENGINEERING MATERIALS 9

copper have only one crystal structure, but many others, particularly among the elements in the
transition series of the periodic table, have two or more allotropic forms.

Allotropic transformations occur because in certain temperature ranges one structure is


more stable than the others. It is found that the energy differences between the different
structures of a metal that undergoes allotropic transformation are very small. Thus small
changes in atomic forces can change the structure of a metal from one type of crystal to
another. It is also found that in many cases allotropic transformations are due to magnetic
interactions between the atoms, but in most cases a full understanding of these forces has not
yet been attained by theorists.

1.2. Grains and Boundaries

Metals are essentially always crystalline; there is no such thing as "molecules" of solid metals.
The unit cell of a metal crystal repeats itself in three dimensions (with occasional imperfections)
far from 103 to 108 atom diameters. For example, a rod of metal 25x200 mm, contains 1025
atoms, may be produced as a single crystal. However, most metals are polycrystalline, i.e.,
made up of many crystals, which typically will be from 0,1 to 0,01 mm in diameter. The
individual crystals of a pure metal are identical except for the orientation of their crystal axes
with respect to an external reference system. They are called grains, and their average
diameter is named the grain size. Each individual grain is in atomic contact with all of its
neighbors; there normally are no voids (Fig. 1.4).

When two crystalline grains of different orientation with respect to an external reference
come in contact, it is evident that in the interface, atoms cannot match up perfectly with both
crystal lattices. Between the two grains, there must be a transition layer where the structure is
that of neither one grain nor the other. Because
the atoms in this transition layer are not in their
proper places with respect to each grain, it is
expected that the grain boundary will have a
higher energy than the material in the neighboring
grains. Exposed to a chemical etching solution,
the grain boundary will be dissolved away faster
than the surrounding material and, in an etched
microstructure will appear as a groove in an
otherwise flat surface. Even when the etching
solution used does not differently color the
differently oriented grains, the grain boundaries of
a microstructure will be revealed as black lines.

Evidence available today indicates that the


Figure 1.4. Iron carbon alloy (0,5% C).
transition layer of irregular structure in a grain
Heat treatment: 850°C 15 min/water
boundary is only a few atom layers thick. It is
+715°C 96h/air+730°C 250 s/water.
Etched 25s in 1% alcoholic nitric acid
10 TECHNICAL ENGINEERING

permissible, then, to think of the boundary as a plane interface

In addition to grain shape, the grain size in a sample is often important. Grain size can
range from large enough to be easily seen by the naked eye to so small as to be barely
resolvable with the most powerful optical microscope. It can be measured from a micrograph
taken at known magnification by finding the number of grains per unit area. The ASTM grain-
size number N is defined to be such that 2N-1 equals the number of grains per square inch at
linear magnification of 100 times.

1.3. Structure of Nonmetallic Solids

The crystal structures of the common metals, discussed above, are exceedingly simple in
comparison with most of the structures found among the species of the mineral kingdom.

A number of different schemes can be used to classify substances when their structures
are described. In mineralogy, for example, a classification based on chemical composition is
generally used, and crystals are arranged according to whether they are elements, oxides,
sulfides, silicates, etc. A classification based on the type of bonding forces acting between the
atoms of the crystal will be used here. On this basis four main classes of crystals may be
recognized. They are:

1. Metallic, already described above;

2. Ionic, in which the interatomic forces are those between electrically charged ions
such as Na+ and Cl-;

3. Covalent, in which the atoms are bound together with chemical covalent bonds;

4. Molecular, which includes those crystals made up of chemically saturated molecules


held together by van der Waals forces.

Ionic Crystals One of the most common of the ionic crystal structures is that of sodium
chloride (Fig. 1.5). This structure is based on the FCC lattice, but instead of having one atom
associated with each lattice point as in most metals, there are two atoms per lattice point, one
1
centered on the point and another at distance a away
2
along the edge of the FCC unit cell.

The "atoms" in this case are actually ions, and from


the pair associated with a lattice point, one will be Na+ and
the other Cl-. Each positive ion is surrounded by six
negatively charged, nearest neighbors and, conversely,
each negative ion has six positive ions as nearest
neighbors, as would be expected, since unlike charges
attract.

Figure 1.5. The crystal structure


of NaCl
STRUCTURE OF ENGINEERING MATERIALS 11

Figure 1.6. The crystal structure Figure 1.7. The tetrahedral array
of diamond of SiO4-4

Covalent Crystals Diamond (Fig. 1.6) is a typical covalent (shared electrons) crystal
which, as can be seen from the drawing, is based on an FCC lattice. The lines drawn between
the atoms in the figure represent the covalent bonds, four for each carbon atom as in the
molecule of CH4. In addition to diamond, the elements silicon, germanium, and an allotropic
form of tin have this so-called diamond cubic structure. The network of covalent bonds
extending throughout the structure is responsible for the great hardness of these elements.

The silicates represent a very important and very large class of covalent crystals. The
basic unit from which these structures are built is the tetrahedral array of one Si and four O
-4
atoms shown in Fig. 1.7. Because these SiO4 tetrahedra can be linked up in many different
ways with each other and with various positive ions, a great variety of different silicate
structures is possible.

Glasses and Polymers These substances are not crystalline; they have no regular,
periodic arrangement of atoms. Most glasses consist primarily of silicate ions, SiO2, to which an
appreciable number of large-sized atoms such as Na and Ca have been added. The added
atoms, since they do not fit into the silicate structure very well, make it more difficult for
crystallization to occur when the melt is cooled. Glass is then a super-cooled liquid having a
very high viscosity.

Polymers, such as rubber or lucite, consist of very long organic molecules twisted up
and intertwined with each other. Some polymers do, however, show a degree of crystallinity: in
some regions of the polymer the long molecules are all lined up parallel to themselves so that,
within this region, there is a fairly regular pattern of molecular arrangement. Such polymers are
sometimes named crystalline, although they do not have a regular crystal structure in the
proper sense.

1.4. Mechanical Properties of Materials

Understanding the mechanical properties of materials is one of the most difficult tasks faced by
the solid-state physicist, and progress in this direction has not been so great as in the case of
the electrical, thermal, and magnetic properties. For purposes of discussion it is more
12 TECHNICAL ENGINEERING

convenient to classify mechanical properties as elastic or plastic and as structure-sensitive or


structure-insensitive. Elastic properties are the appropriate elastic constants, while plastic
properties include the strength properties and the creep, fatigue, and fracture characteristics
(since fracture, except in ideally brittle materials, always involves some plastic flow). Structure-
sensitive properties are those that depend in a sensitive way on the extent and distribution of
atomic-scale imperfections in the material and so can vary widely from sample to sample of the
same material. The yield strength of a pure metal depends, for example, on the past thermal
and mechanical history of the particular sample tested and can vary by 100% or more from
sample to sample. Structure-insensitive properties such as the elastic and optical constants of
a metal are pretty much the same from sample to sample.

Elastic Properties The elastic constants of a crystal are essentially a measure of the
force required to displace the atoms of the crystal relative to their initial position. Figure 1.8
shows how the interactions force between atoms as a function of the interatomic distance. The
interatomic equilibrium distance is that corresponding to the minimum in the curve and so would
be at point a. Applying a tensile stress to the crystal along the x direction causes the
interatomic distance to increase and the interactions force to rise.

Because the elastic constants are so closely related to the interatomic forces, less can
be done to alter them in any given material. Moderate amounts of them can be changed by
alloying, but are relatively insensitive to cold work, irradiation, or other treatments that can be
applied to a material.

Amorphous materials are elastically isotropic: the Young's modulus and shear modulus
are the same, regardless of the direction in which the material is stressed. Even cubic crystals,
however, are elastically anisotropic. Young's modulus in copper, for example, varies from 194
MPa in the [111] direction to 68000 MPa in the
[100] direction. While an isotropic material has
two independent elastic constants, a cubic
crystal has three and the calculation of
Young's modulus in an arbitrary
crystallographic direction requires the use of
all three constants. A polycrystalline metal,
when the grains are at random orientations
with respect to each other, behaves as if it
were an isotropic elastic material with elastic
constants, which are averages of those of the
individual grains.

Plastic Properties The plastic


properties of crystals are structure-sensitive,
i.e., dependent on the imperfections present in
a particular sample of material as well as on
Figure 1.8. The interactions -vs- interatomic the characteristics of the perfect crystal.
distance
STRUCTURE OF ENGINEERING MATERIALS 13

The mechanical properties of a material determinate its response to the application of


stress. Stresses may be tensions or shears and may be applied to a material singly or in
combinations. The description of the response of the material to a system of stresses can be
an exceedingly complex problem in engineering mechanics. In order to study the properties of
materials, the metallurgist relies on tests in which the stress pattern applied to a specimen is
purposely made simple, as in the tensile test. There is a danger here: the results obtained
under the idealized conditions of the laboratory may not be applicable in a real situation where
the stress pattern is not simple. Many spectacular failures of engineering structures was
resulted from inadequate attention to just this problem. Nevertheless, idealized laboratory tests
remain the starting point in any study of mechanical properties.

For over 75 years the breaking of tensile-test specimens has been a primary source of
information on the mechanical properties of metals, and it is not likely that this method of
testing will be supplanted for a long time to come. Tensile tests are of immediate practical
value, but the interpretation of tensile-test data is not so straightforward as is often assumed.
Unless certain limitations are realized, large numbers of useless data can be collected in
testing programs.

Young's Modulus is defined as the tensile stress divided by the tensile strain for
elastic deformation and so is the slope of the linear part of the stress-strain curve. In
engineering practice, the Young's modulus of a material is primarily of interest as a measure of
the amount of deflection which will occur in a structure under a given load; the lower the
modulus, the greater the elastic deflection will be for a given stress.

Yield Strength When a material under tension reaches the limit of its elastic strain and
begins to flow plastically, it is said to have yielded. The yield strength is then the stress at
which plastic flow starts. Some materials under certain well-defined test conditions show a
sharp yield point on their stress-strain curves, as illustrated in Fig. 1.9.

One of the most commonly used method of


defining the yield point is to construct a parallel line to
the elastic part of the stress-strain curve but displaced to
the right an amount equivalent to a strain of 0,20%. The
stress at which this line intersects the stress-strain curve
is the 0,20% offset yield strength. An offset of 0,10%
may sometimes be specified.

Two other quantities are sometimes used as


measures of the yield point. The proportional limit is
defined as the stress at which the stress-strain curve
first shows a measurable deviation from linearity, while
the elastic limit is the maximum stress to which a metal
may be subjected without permanent plastic
deformation. The difficulty with these quantities is that
they depend entirely on the sensitivity of the test
Figure 1.9. Tensile stress-strain
equipment used and so cannot be expected to be
curve
14 TECHNICAL ENGINEERING

reproducible in observations made by different observers. Also, the elastic limit can be found
only by repeated loading and unloading of the specimen, a tedious procedure at best. The
above-defined yield point is a much more useful property, and the use of the terms elastic limit
and proportional limit should be discouraged.

Tensile Strength is defined as the maximum load sustained by the specimen during
the tensile test, divided by its original cross-sectional area. It is sometimes called the ultimate
strength of the material. It will be readily appreciated that it is an entirely arbitrary quantity
having no real physical significance but useful as a metal quality indicator.

True Breaking Strength is the load on the specimen at the time of fracture divided by
the cross-sectional area at the fracture. It is therefore the actual stress that the material
sustained when it finally failed and it is a true physical quantity.

Reduction of Area is the quotient of the decrease in cross-sectional area at the plane
of fracture and the original area at that plane (times 100, to express as a percentage). Similarly
to percent elongation, this number, while related to ductility, does not differentiate between the
localized reduction in area associated with necking and the general reduction in area
throughout the gage length caused by uniform elongation.

Rate of Strain-hardening The slope of the stress-strain curve beyond the yield
strength shows the rate of increase of stress with increase in plastic strain. The steeper this
part of the curve, the greater is the hardening effect of the plastic deformation. While this
section of the stress-strain graph is curved, it ordinarily becomes a straight line in a log-log plot
and thus conforms to the relation σ = kε n , where n defines the slope of the early plastic
deformation part of the stress-strain plot and is called the strain-hardening extent. This useful
dimensionless quantity can be shown to be numerically equal to the strain at maximum load.
Metals with a low strain-hardening rate have less uniform elongation and start to neck locally
early in the tensile test, whereas metals with a very high rate of strain-hardening will tend to
have a large amount of uniform elongation and no localized necking under tension.

The commonly used Charpy notched impact test is supposed to reveal the toughness
of metals under impact loading conditions. Actually, the test indicates the ability of metals to
absorb energy by local deformation. For example, the energy required to break an unnotched
bar in tension under impact loading (in the tensile impact test) correlates well with the area
under a true-stress-true-strain curve in a static tensile test. In both cases, the strong but brittle
metal and the weak but plastic metal show lower energy absorbed to fracture than a
moderately strong, moderately plastic specimen.

Creep is the flow of a material over a period of time when under a load too small to
produce any measurable plastic deformation at the time of application. The simplest type of
creep test is made by just hanging a weight on the test specimen and observing its elongation
as a function of time, by use of a measuring microscope or other sensitive detector of strain. In
more refined creep apparatus, a system of levers and cams may be employed to apply the load
in such a way that the stress on the specimen remains constant, the load being decreased as
STRUCTURE OF ENGINEERING MATERIALS 15

the cross-sectional area of the specimen decreases during its elongation. The rate at which a
material creeps is sensitive to the temperature of the test relative to the melting temperature.

Fatigue testing determines the ability of a material to withstand repeated applications of


a stress which in itself is too small to produce appreciable plastic deformation. Situations
involving fatigue frequently occur in engineering practice; a common example is a rotating
beam loaded transversely to the axis of rotation, e.g., an axle. The metal at the surface is
alternately stressed in tension and compression during each cycle of rotation. If the strains are
completely elastic, nothing happens. If, however, the stresses are high enough for some small
region of the metal to be even minutely deformed, permanent damage begins to build up in that
region. The high stresses can appear from overloading, from excessive vibration or loud noise
(sonic fatigue occurring in a jet aircraft), or from local stress concentrations at notches,
keyways, etc., when it was thought that the stresses were in a safe range. In any case, the
accumulation of permanent damage in a local region eventually results in the formation of a
crack, and once this occurs, stress concentration at its root is enormously increased. This
leads to the propagation of a crack through the bar and finally the occurrence of a "notch" type
of failure showing no evidence of ductility. In a corrosive environment the susceptibility of a
metal to fatigue failure may be enormously increased, a phenomenon known as corrosion
fatigue.

The hardness of a material can perhaps be best defined as that property which is
measured by a hardness tester. Actually, hardness is not a true physical property but depends
on a combination of different physical properties. There are a number of different methods of
measuring hardness and each method measures a different combination of the true properties
of the material. Consequently, hardness as measured by one test is not necessarily compara-
ble with hardness as measured by another test. The three principal types of hardness tests are
the scratch, the rebound and the penetration tests.

In the scratch test, the resistance of the surface is determined by comparison with a set
of standard materials ranging from talc (softest) to diamond (hardest); the test is used primarily
in testing minerals and refractories.

The rebound (scleroscope) test involves observing the height of rebound of a hardened
steel ball dropped from a given height on the test piece: the harder the material, the greater the
rebound. What is being measured in this case is the amount of energy absorbed in the material
during the impact. The rebound test is not extensively used in modern practice.

Almost all of the hardness testing of metals today are done by penetration tests. There
is a variety of these tests in use they all have the common feature that the amount of
penetration of a standard indenter into the metal under a known load is observed. The most
widely used penetration tests are the Brinell and the Rockwell. In the Brinell test, a hardened
steel ball is forced by a known load into the tested metal, and the diameter of the impression is
measured.

The Rockwell test is similar, except that steels balls of various diameters or a diamond
cone (Brale) can be used. The diameter of the impression made by the indenter is not
measured in the Rockwell test; instead, its depth is automatically indicated on a dial. In both
16 TECHNICAL ENGINEERING

these indentation tests, the results are affected not only by the original resistance of the metal
to deformation but also by the rate at which this resistance changes in the vicinity of the
indenter during the tests. The results are also affected by elastic properties of the metal, since
the diameter or depth of the indentation is measured after the load has been released, with an
accompanying elastic recovery or slight reversed dimensional change. Another variable that
should be recognized is the relative volumes of metal displaced by varying depths of
indentation of a sphere and a cone.

Hardness measurements are extremely useful as a quick and rough indication of the
mechanical properties of a metal. Multiplying the Brinell hardness number by 3,4 is obtaining
an approximation of the tensile strength in MPa.

You might also like