ALL Slides
ALL Slides
Technology
Course Instructor
Dr. Raghukiran, N.
School of Mechanical Engineering
VIT University – Chennai campus
The Materials Science and Engineering Tetrahedron
Processing:
Processing:
Processing:
Structure:
Structure:
Structure:
Structure:
Properties:
Properties:
Properties:
Properties:
Performance:
Performance:
Performance:
Ceramics
Metals
Polymers
• Metals
• Ceramics
• Polymers
• Composites
• Advanced materials
• Semiconductors
• Biomaterials
• Smart Materials
• Nanomaterials
• Lattice
a three-dimensional array of points coinciding with atom positions (or
sphere centers).
For the FCC structure, the atomic packing factor is 0.74, which is the
maximum packing possible for spheres all having the same diameter.
Metals typically have relatively large atomic packing factors to
maximize the shielding provided by the free electron cloud.
Solution
Try Yourself:
Show for the body-centered cubic crystal structure
that the unit cell edge length a and the atomic radius
R are related through
• The BCC unit cell length a and atomic radius R are related
through
Using the triangle NOP
or
Try Yourself:
Show that the atomic packing factor for BCC is 0.68
Where
n = number of atoms associated with each unit cell
A = atomic weight
VC = volume of the unit cell
NA = Avogadro’s number (6.022x1023 atoms/mol)
Example:
Try Yourself:
Copper has an atomic radius of 0.128 nm, an FCC
crystal structure, and an atomic weight of 63.5 g/mol.
Compute its theoretical density and compare the
answer with its measured density.
Answer:
Answer:
Where,
primed indices are associated with the three-index scheme and
unprimed with the new Miller–Bravais four-index system.
Example problem
Now convert these indices into an index set referenced to the four-axis scheme. This
requires the use of equations mentioned in the earlier slides.
• For cubic crystals, the angle, θ between two planes, (h1 k1 l1) and (h2 k2 l2) is given by:
Determine the Miller indices for the plane shown in the below sketch
Solution:
1. Because the plane passes through the selected origin O, a new
origin must be chosen at the corner of an adjacent unit cell,
taken as O’ and shown in sketch on the right side..
Solution (continued..)
2. This plane is parallel to the x axis, and the intercept may be
taken as ∞a. The y and z axes’ intersections, referenced to the
new origin O’, are -b and c/2, respectively.
3. Thus, in terms of the lattice parameters a, b, and c, these
intersections are ∞, -1 and ½ respectively.
4. The reciprocals of these numbers are 0, -1, and 2; because all are
integers, no further reduction is necessary.
_
5. Finally, enclosure in parentheses yields (012).
Solution (continued..)
(110) atomic plane for FCC crystal structure (110) atomic plane for BCC crystal structure
Example:
# Although six atoms have centers that lie on
this plane, only one-quarter of each of atoms
A, C, D, and F, and one-half of atoms B and
E, for a total equivalence of just 2 atoms, are
on that plane.
# Furthermore, the area of this rectangular
section is equal to,
# and the planar density is determined as
(110) atomic plane for FCC crystal structure follows:
Single crystals
For a crystalline solid, when the periodic and repeated arrangement of
atoms is perfect or extends throughout the entirety of the specimen
without interruption, the result is a single crystal.
There exists some atomic mismatch within the region where two
grains meet; this area, called a grain boundary.
Grain 1 Grain 2
Grain
boundary Grain 3
Grain
boundary
Anisotropy
• The physical properties of single crystals of some substances
depend on the crystallographic direction in which measurements
are taken.
• For example, the elastic modulus, the electrical conductivity, and
the index of refraction may have different values in the [100] and
[111] directions.
• This directionality of properties is termed anisotropy, and it is
associated with the variance of atomic or ionic spacing with
crystallographic direction.
• Substances in which measured properties are independent of
the direction of measurement are isotropic.
Non-crystalline materials
noncrystalline solids lack a systematic and regular arrangement
of atoms over relatively large atomic distances. Sometimes such
materials are also called amorphous (meaning literally “without
form”) materials, or supercooled liquids, inasmuch as their
atomic structure resembles that of a liquid.
Crystalline Non-crystalline
MEE1005: Materials Engineering and Technology - Dr. Raghukiran 98
Imperfection in solids
Example:
Two-dimensional representations of a
vacancy and a self-interstitial.
Self-interstitial
A self-interstitial is an atom from the crystal that is crowded into an
interstitial site, a small void space that under ordinary circumstances
is not occupied.
Dislocation line: the line that is defined along the end of the extra
half-plane of atoms (perpendicular to the plane of the page).
MEE1005: Materials Engineering and Technology - Dr. Raghukiran 112
Edge dislocation
• Within the region around the edge dislocation line, there is some
localized lattice distortion.
• The atoms above the dislocation line (shown in the previous
slide) are squeezed together, and those below are pulled apart
• The edge dislocation (shown in the previous slide) is represented
by the symbol Ʇ
• An edge dislocation may also be formed by an extra half-plane of
atoms that is included in the bottom portion of the crystal (as
shown below); its designation is Τ
• These defects exist in all solid materials that are much larger
than those heretofore discussed.
• These include pores, cracks, foreign inclusions, and other
phases.
• They are normally introduced during processing and fabrication
steps
In copper-zinc alloys containing more than 30% Zn, a second phase forms because of the
limited solubility of zinc in copper.
Under equilibrium
conditions, all
metals exhibit a
definite melting or
freezing point. If a
cooling curve is
plotted for a pure
metal. It will show a
horizontal line at
the melting or
freezing
temperature.
Cooling curve for a pure metal showing possible undercooling.
Cooling curve for a solid solution.
Phase diagram
Definition 1 Definition 2
Phase diagram is a graphical A phase diagram shows the
representation of the phases and their compositions at
physical states of a substance any combination of temperature
under different conditions of and alloy composition.
temperature and pressure.
• By removing the time axis from the curves and replacing it with composition, the cooling
curves indicate the temperatures of the solidus and liquidus for a given composition.
• This allows the solidus and liquidus to be plotted to produce the phase diagram:
Note that the denominator represents the total length of the tie line
and the numerator is the portion of the lever that is opposite the
composition of the solid we are trying to calculate.
1. Where L stands for liquid, and A and B are the two components
and α and β are two solid phases rich in A and B respectively.
2. The blue lines represent the liquidus and solidus lines.
eutectoid
Pearlite and
Cementine
Austenite
Ferrite
Pearlite and
Carbide
Pearlite
• FIGURE - The unit cell for (a) austenite, (b) ferrite, and (c) martensite. The effect
of the percentage of carbon (by weight) on the lattice dimensions for martensite
is shown in (d). Note the interstitial position of the carbon atoms and the increase
in dimension c with increasing carbon content. Thus, the unit cell of martensite is
in the shape of a rectangular prism.
Austenite
Ferrite
Pearlite
Pearlite
(High resolution)
Liquid + d ↔ austenite
The CCT diagram (solid lines) for a 1080 steel compared with the TTT diagram (dashed
lines).
(a) Surface hardening by localized heating. (b) Only the surface heats above the A1
temperature and is quenched to martensite.
249
Ferrous metals
250
Ferrous metals
251
Ferrous metals: classification
252
Classification of steels
253
Classification of steels
254
Low-Carbon Steels
255
Low-Carbon & High-strength, low-alloy (HSLA) steels
256
Low-Carbon & High-strength, low-alloy (HSLA) steels
257
Low-Carbon & High-strength, low-alloy (HSLA) steels
258
Medium-Carbon Steels
260
High-Carbon Steels
261
High-Carbon Steels
• Applications: These steels are used as cutting tools and dies for
forming and shaping materials, as well as in knives, razors,
hacksaw blades, springs, and high-strength wire.
262
Comparison
264
Applications for some of the tool steels
265
Stainless Steels
266
Stainless Steels
267
Stainless Steels
268
Stainless Steels
269
Stainless Steels
270
Effect of alloying elements on steel
271
Alloy steels
272
MEE1005: Materials Engineering and Technology - Dr. Raghukiran 273
Ferrous metals and alloys
274
Cast Iron
275
Fe-C True Equilibrium Diagram
T(°C)
1600
Graphite formation
1400 L Liquid +
promoted by
g +L Graphite
• Si > 1 wt% 1200 g 1153°C
Austenite 4.2 wt% C
• slow cooling
1000
a+g g + Graphite
800
740°C
0.65
600
a + Graphite
Adapted from Fig.
11.2,Callister 7e. (Fig. 11.2 400
adapted from Binary Alloy 0 1 2 3 4 90 100
Phase Diagrams, 2nd ed., (Fe) Co , wt% C
Vol. 1, T.B. Massalski (Ed.-
in-Chief), ASM International,
Materials Park, OH, 1990.)
276
Types of Cast Iron
Gray iron
1. graphite flakes
2. weak & brittle under tension
3. stronger under compression
4. excellent vibrational dampening
5. wear resistant Adapted from Fig. 11.3(a) & (b), Callister 7e.
White iron
1. <1wt% Si so harder but brittle
2. more cementite
Malleable iron
1. heat treat at 800-900ºC
2. graphite in rosettes
3. more ductile
278
Grey cast iron
Applications
1. Engines
a. Cylinder blocks, liners,
2. Brake drums, clutch plates
3. Pressure pipe fittings (AS2544)
4. Machinery beds
5. Furnace parts, ingot and glass moulds
285
Limitations of Ferrous Alloys
286
Strengthening mechanisms
of crystalline materials
• For high-angle grain boundaries, it is difficult for the dislocations to traverse grain
boundaries during deformation; rather, dislocations tend to “pile up” (or back up) at
grain boundaries.
• A fine-grained material (one that has small grains) is harder and stronger than one that
is coarse grained, because the former has a greater total grain boundary area to impede
dislocation motion.
• For many materials, the yield strength σy varies with grain size according to
Hall–Petch equation
This expression is termed the Hall–Petch equation, where d is the average grain diameter,
and σ0 and ky are constants for a particular material.
• Note that the Hall–Petch equation is not valid for both very large (i.e., coarse) grain and
extremely fine grain polycrystalline materials.
• Grain size may be regulated by the rate of solidification from the liquid phase, and also
by plastic deformation followed by an appropriate heat treatment.
• grain size reduction improves not only strength, but also the toughness of many alloys.
• Boundaries between two different phases are also impediments to movements of
dislocations; this is important in the strengthening of more complex alloys.
• The sizes and shapes of the constituent phases significantly affect the mechanical
properties of multiphase alloys
(a) Representation of tensile lattice strains imposed on (a) Representation of compressive strains imposed on
host atoms by a smaller substitutional impurity atom. (b) host atoms by a larger substitutional impurity atom. (b)
Possible locations of smaller impurity atoms relative to Possible locations of larger impurity atoms relative to
an edge dislocation such that there is partial cancellation an edge dislocation such that there is partial
of impurity–dislocation lattice strains. cancellation of impurity–dislocation lattice strains.
• These solute atoms tend to diffuse to and segregate around dislocations in a way so as
to reduce the overall strain energy—that is, to cancel some of the strain in the lattice
surrounding a dislocation.
• To accomplish this, a smaller impurity atom is located where its tensile strain will
partially nullify some of the dislocation’s compressive strain.
• For the edge dislocation in Figure (previous slide), this would be adjacent to the
dislocation line and above the slip plane.
• A larger impurity atom would be situated as in Figure (previous slide).
• The resistance to slip is greater when impurity atoms are present because the overall
lattice strain must increase if a dislocation is torn away from them.
• a greater applied stress is necessary to first initiate and then continue plastic
deformation for solid-solution alloys, as opposed to pure metals; this is evidenced by
the enhancement of strength and hardness.
1. Strain hardening is the phenomenon whereby a ductile metal becomes harder and
stronger as it is plastically deformed.
2. Sometimes it is also called work hardening, or, because the temperature at which
deformation takes place is “cold” relative to the absolute melting temperature of the
metal, cold working.
3. Most metals strain harden at room temperature.
4. It is sometimes convenient to express the degree of plastic deformation as percent cold
work rather than as strain. Percent cold work (%CW) is defined as
where A0 is the original area of the cross section that experiences deformation and
Ad is the area after deformation.
MEE1005: Materials Engineering and Technology - Dr. Raghukiran 295
Strain Hardening
1. The strength and hardness of some metal alloys may be enhanced by the formation of
extremely small uniformly dispersed particles of a second phase within the original
phase matrix; this must be accomplished by phase transformations that are induced by
appropriate heat treatments.
2. The process is called precipitation hardening because the small particles of the new
phase are termed precipitates. Age hardening is also used to designate this procedure
because the strength develops with time, or as the alloy ages.
• Step 3: Age
• Finally, the supersaturated α is heated at a temperature below the solvus temperature. At
this aging temperature, atoms diffuse only short distances.
• Because the supersaturated is metastable, the extra copper atoms diffuse to numerous
nucleation sites and precipitates grow.
• When we go through the three steps described previously, we produce the phase in the
form of ultra-fine uniformly dispersed second-phase precipitate particles. This is what we
need for effective precipitation strengthening.
Tension Tests
Typical
engineering
stress–strain
behavior to fracture,
point F.
DIRECT STRESS !
1. When a force is applied to an elastic body, the body deforms. The way in which the body
deforms depends upon the type of force applied to it.
#$%&'( = ! =
" The
symbol
! called EPSILON
!
Strain has no unit’s since it is a ratio of length to length. Most engineering materials do not
stretch very mush before they become damages, so strain values are very small figures. It
is quite normal to change small numbers in to the exponent for 10-6( micro strain).
1. Elastic materials always spring back into shape when released. They also obey HOOKE’s
LAW.
2. This is the law of spring which states that deformation is directly proportional to the
force. F/x = stiffness = kN/m
• The stiffness is different for the different material and different sizes of the
material. We may eliminate the size by using stress and strain instead of force and
deformation:
• If F and x is refer to the direct stress and strain , then
A "" #A "
" = !! " = !! hence = and =
!" !
# !!
1. The stiffness is now in terms of stress and strain only and this constant is called the
MODULUS of ELASTICITY (E)
!= #A "
=
!" !
• A graph of stress against strain will be straight line with gradient of E. The
units of E are the same as the unit of stress.
Elastic behaviour
1. The curve is straight line trough out most of the region
2. Stress is proportional with strain
3. Material to be linearly elastic
4. Proportional limit
a. The upper limit to linear line
b. The material still respond elastically
c. The curve tend to bend and flatten out
5. Elastic limit
a. Upon reaching this point, if load is remove, the specimen still return to original shape
Yielding
1. A Slight increase in stress above the elastic limit will result in breakdown of the material
and cause it to deform permanently.
2. This behaviour is called yielding
3. The stress that cause = YIELD STRESS@YIELD POINT
4. Plastic deformation
5. Once yield point is reached, the specimen will elongate (Strain) without any increase in
load
6. Material in this state = perfectly plastic
1. STRAIN HARDENING
a. When yielding has ended, further load applied, resulting in a curve that rises continuously
b. Become flat when reached ULTIMATE STRESS
c. The rise in the curve = STRAIN HARDENING
d. While specimen is elongating, its cross sectional will decrease
e. The decrease is fairly uniform
2. NECKING
a. At the ultimate stress, the cross sectional area begins its localised region of specimen
b. it is caused by slip planes formed within material
c. Actual strain produced by shear strain
d. As a result, “neck” tend to form
e. Smaller area can only carry lesser load, hence curve donward
f. Specimen break at FRACTURE STRESS
MEE1005: Materials Engineering and Technology - Dr. Raghukiran 314
STRESS STRAIN DIAGRAM
• Simple fracture is the separation of a body into two or more pieces in response to an
imposed stress that is static (i.e., constant or slowly changing with time) and at
temperatures that are low relative to the melting temperature of the material.
• For metals, two fracture modes are possible: ductile and brittle.
• Ductile metals typically exhibit substantial plastic deformation with high energy
absorption before fracture.
• On the other hand, there is normally little or no plastic deformation with low energy
absorption accompanying a brittle fracture
• The stress applied to the material is intensified at the flaw, which acts as a stress raiser
• For a simple case, the stress intensity factor “K” is
Ø where f is a geometry factor for the specimen and flaw, σ is the applied stress, and a is
the flaw size as defined in figure in previous slide.
Ø If the specimen is assumed to have an “infinite” width, f = 1.0.
Ø For a small single-edge notch [Figure (a)], f = 1.12
By performing a test on a specimen with a known flaw size, we can determine the value of K
that causes the flaw to grow and cause failure. This critical stress intensity factor is defined
as the fracture toughness Kc:
1. Fatigue is the lowering of strength or failure of a material due to repetitive stress which
may be above or below the yield strength.
2. It is a common phenomenon in load-bearing components in cars and airplanes, turbine
blades, springs, crankshafts and other machinery, biomedical implants, and consumer
products, such as shoes, that are subjected constantly to repetitive stresses in the form
of tension, compression, bending, vibration, thermal expansion and contraction, or
other stresses.
3. These stresses are often below the yield strength of the material; however, when the
stress occurs a sufficient number of times, it causes failure by fatigue!
4. Quite a large fraction of components found in an automobile junkyard belongs to those
that failed by fatigue.
5. The possibility of a fatigue failure is the main reason why aircraft components have a
finite life.
1. The fatigue test can tell us how long a part may survive or the maximum allowable loads
that can be applied without causing failure.
2. The endurance limit, which is the stress below which there is a 50% probability that
failure by fatigue will never occur, is our preferred design criterion.
3. Fatigue life tells us how long a component survives at a particular stress.
• If we apply stress to a material at an elevated temperature, the material may stretch and
eventually fail, even though the applied stress is less than the yield strength at that
temperature.
• Time dependent permanent deformation under a constant load or constant stress and
at high temperatures is known as creep.
• A large number of failures occurring in components used at high temperatures can be
attributed to creep or a combination of creep and fatigue.
• Diffusion, dislocation glide or climb, or grain boundary sliding can contribute to the
creep of metallic materials.
• Polymeric materials also show creep.
• In ductile metals and alloys subjected to creep, fracture is accompanied by necking, void
nucleation and coalescence, or grain boundary sliding.
• Possibly the most important parameter from a creep test is the slope of the secondary
portion of the creep curve; this is often called the minimum or steady-state creep rate.
• It is the engineering design parameter that is considered for long-life applications, such
as a nuclear power plant component that is scheduled to operate for several decades,
and when failure or too much strain are not options.
• On the other hand, for many relatively short-life creep situations (e.g., turbine blades in
military aircraft and rocket motor nozzles), time to rupture, or the rupture lifetime tr, is
the dominant design consideration.
(a) concentration, (b) size, (c) shape, (d) distribution, and (e)
orientation