MO 201 Mechanical Behaviour
MO 201 Mechanical Behaviour
Dr Bratindranath Mukherjee
Dept. of Metallurgical Engineering
bratindra.met@itbhu.ac.in
Why study mechanical properties?
• Deformation
• Fracture or failure
• Deformation
• Tension
• Compression
• Concept of strain
• Change in dimensions
• Angular distortion
Concepts of stress/strain
• Simple definitions
• Stiffness
• Strength
• Hardness
• Ductility
• Toughness
• Hardness test
• Tension test
• Compression test
• Fatigue test
• Impact test
Elastic Anelastic And Viscoelastic Behaviour
Brittle materials, such as concrete, cast iron and silicate glasses, under tensile stress show elastic deformation right up to
the point of fracture. Ductile materials such as copper and aluminium are elastic up to a certain stress called the elastic
limit. Thereafter, they plastically deform. In both these groups of materials, within the elastic region, the strain is
proportional to the stress applied, as given by Hooke’s law. This behaviour can be called ordinarily elastic (or truly elastic)
behaviour.
Materials which undergo recoverable deformation of a few hundred per cent are called elastomers and exhibit rubber-like
elasticity. The stress is not proportional to strain in these materials, in contrast to ordinary elastic materials.
In most engineering materials, there is a time-dependent elastic strain component (elastic deformation will
continue after the stress application, and upon load release, some finite time is required for complete recovery)
This time-dependent elastic behavior is known as anelasticity, and it is due to time-dependent microscopic and
atomistic processes that are attendant to the deformation
• For metals, the anelastic component is normally small and is often neglected; however, for some polymeric
materials, its magnitude is significant – in this case it is termed viscoelastic behavior (rubber-like)
A tangent to the force-distance curve drawn at r0 practically
coincides with the curve over a small range of displacement on
either side of r0,
Strains in the elastic region for both brittle and ductile
materials lie in the range 0.001 to 0.005 and are, therefore,
within this small range of displacement. Then, the negative of
the slope of the force-distance curve at r0 is proportional to
the Young’s modulus Y of a material. Also, the curvature of the
potential energy curve at r0 is proportional to the elastic
modulus
In a crystal, the interatomic distance varies with crystal direction, with a corresponding variation in the bond
strength. This gives rise to elastic anisotropy, that is, the elastic properties become a function of the crystal
direction.
This anisotropy is particularly evident in materials which have two kinds of bonds. For example, the Young’s
modulus of graphite in the a direction parallel to the sheets is 950 GN m –2, which is much larger than that
averaged over all directions, which is only 8 GN m –2.
The transition metals have elastic moduli much higher than those of the alkali metals, as a result of the
partial covalent character of their bonds. Metals of the first transition series have Young’s moduli in the range
200 GN m–2. Metals of the second and the third transition series have higher moduli, reaching up to 600 GN
m–2 . As is the case with some other physical properties, the modulus reaches a peak value for the electronic
configuration of d5
The application of a tensile stress causes an elongation lalong the tensile axis and a transverse contraction t. The ratio of
these two strains defines the Poisson ratio :
The Poisson ratio for metals is around 0.3, for polymers and rubber it is between 0.4 and 0.5; for ionic solids, it is around 0.2.
The shear modulus of a material is defined as the ratio of the shear stress µ applied to the shear strain. The shear modulus
is related to the Young’s modulus Y through the Poisson ratio:
Similarly, the bulk modulus K, which is the ratio of the hydrostatic stress to the relative volume change, is related to Y:
At higher temperatures, under the influence of thermal energy, the atoms vibrate about their mean positions,
the amplitude of the vibrations increasing with increasing temperature. With more thermal energy, we can
visualize the bonds to besomewhat loosened up. This reflects in a decrease in the elastic modulus with
increasing temperature, n a majority of cases, on heating from 0 Kto the melting point, the decrease in elastic
modulus is in the range 10–20%.
Key mechanical design properties
• Hardness For identical shapes, it is proportional to the elastic modulus. Therefore, the elastic
modulus is an important material parameter in mechanical design. Materials with
high stiffness and hence high modulus are often required.
• Ductility
• Toughness
Covalently bonded elements such as diamond have a very high modulus (1140 GN m-2 ). However, they are
not suitable for use in engineering practice, due to high cost, non availability and brittleness. Brittle materials
cannot withstand accidental overloading during service and may fail in a catastrophic manner. Hence, they
are not suitable as structural members, even though they may have a high modulus.
Ductile elements such as metals withstand accidental overloading without catastrophic failure and as such
are suitable for structural components. Among the metals, the elements of the first transition series offer a
good compromise of adequate ductility and a moderately high modulus, in the range 200 GN m–2 . The
metals of the second and the third transition series have an even higher modulus but have the disadvantage
of high density. By suitable alloying, the Young’s modulus of metals can be increased.
For producing a high modulus Fe based material, reinforcement with TiB2 is a promising route. With 50 vol%
of TiB2 particles in the Fe matrix, there is an increase of more than 50% in themodulus. The TiB2 particles
are in stable equilibrium with Fe. As the particles are approximately spherical, the modulus is not dependent
on direction as infibre-reinforced materials.
The Young’s moduli of some ionic solids are given below. Even though the modulus values of some of them are quite high,
they also suffer from the lack of ductility like covalent solids
In composite materials, an attempt is made to increase the stiffness, without the disadvantages of brittleness. Boron has a
low density and is suitable for light weight applications and for air borne structures. Its elastic modulus is one of the
highest for elements (Y = 440 GN m–2 ), but it is brittle. It can be used as a reinforcing fibre for a ductile matrix such as
aluminium. In the Al-B composite, the elastic modulus is increased due to the presence of the boron fibres. At the
same time, the disadvantages of the brittleness of boron are countered, by the cushioning effect of the ductile matrix. The
ductile matrix stops a propagating crack if a fibre embedded in it breaks accidentally. If the entire material were to
consist of boron only, a propagating crack would culminate in the fracture of the entire cross-section.
The Young’s modulus Y of a composite in a direction parallel to the fibres can be expressed as a linear function of the
moduli of the fibre and the matrix,
Yf and Ym:
where Vf and Vm are the volume fractions of the fibre and the matrix. Thus, a 40 vol.% of boron in an aluminium matrix
can raise the Young’s modulus from71 GN m–2 for pure aluminium to 219 GN m–2 for the composite. This
composite would then be as stiff as steel but less than one-third its density!
Materials which undergo recoverable deformation of a few hundred per cent are called elastomers and exhibit rubber-
like elasticity. The stress is not proportional to strain in these materials, in contrast to ordinary elastic materials.
Elastomers are long chain molecules with some cross-linking between the chains. This cross-linking is important,
because this feature is what keeps the molecules from slipping past one another permanently during stretching. After
cross-linking, the translational motion of chains gets restricted to segmental mobility between cross-linking points.
When a stress is applied to an elastomer, equilibrium in the molecular configuration is established fairly quickly so
that we can ignore the time dependent aspects of stretching as a first approximation.
n the unstretched state, the chain molecules are randomly coiled. A large number of configurations of equal potential
energy are then possible. This large number of distinguishable arrangements means an appreciable configurational
entropy and a low free energy. On application of an external stress, the coiled molecules respond by stretching out.
The stretching reduces the number of possible configurations and hence lowers the configurational entropy. In the
limit, when the molecules are all fully stretched out, the possible configuration is only one and the configurational
entropy is zero When a stretched rubber is heated, the increase in thermal energy tends to coil back the uncoiled
molecules against the stretching force.
Several relaxation processes take place within a material in response to an externally applied stress. If the time
scale of a relaxation process is too fast or too slow compared to the time interval over which the stress is applied,
the stress-strain relationship is essentially independent of time. If, however, the time scale of the process is
comparable to the time interval of stress application, the stress-strain relationship is dependent on time and
results in anelastic behaviour.
It is important that, during anelastic deformation, the heat generated during each cycle of loading equal to the area of
the hysteresis loop should be dissipated as soon as possible. The fraction of this energy that is dissipated is
dependent on the nature of the material and is called the damping capacity the material. Gray cast iron has a good
damping capacity as compared to mild steel. This can be attributed to the presence of graphite flakes in
themicrostructure of gray cast iron,. It is, therefore, used as a base for erection of machinery to damp out vibrations
efficiently. Polymers including rubber have very good damping capacity compared to other materials and find
uses such as mounting pads for delicate instruments.
For a given relaxation process, the energy loss during a cycle is obviously a maximum at some frequency of loading. In
automobile tyres, the frequency of loading increases with the speed of the vehicle. In the common type of tyres,
theenergy loss becomes a large value at high speeds such as 100 km per hr. This would increase the heating effect in
the tyres, reducing their lifetime as well as increasing the risk of a tyre burst at high speeds. Recent improvements such
as fibre-glass reinforced polyester tyres yield a much larger relaxation time than the unreinforced tyres for the same
temperature, so that tyre life and the safety factor are increased.
Tensile Test
Tensile test specimen
• Load
• Elongation/Displacement
• Stress
• Strain
Tensile Test
• Toughness
Tensile Test
Yield Stress and Tensile Strength
0.2% offset Yield Stress
Toughness
Resilience
Area under Area under
curve up to curve till
yield fracture
Modulus of resilience
Tensile Test
Tensile curves for polymers
• Brittle
• Plastic
• Elastomeric
(highly elastic)
K is called the strength coefficient and n is the work hardening exponent.
Materials which have a high work hardening exponent such as copper and brass (n ~ 0.5) can
be given a large plastic strain more easily than those which have a smaller n, such as heat
treated steel (n ~ 0.15).
A is a constant and m is the index of strain rate sensitivity. If m = 0, the stress is independent of the strain
rate and the stress-strain curve would be the same for all strain rates.
If m = 0.4–0.9,
the material may exhibit superplastic behaviour, that is, deform by several hundred per cent of strain
without necking. The reason for this is that, as soon as necking starts in some region, the strain rate
increases locally, resulting in a rapid increase of the stress required to cause further deformation in that
region. The deformation then shifts to another region of the material, where there is no necking. Here,
the strain rate and hence the stress to cause deformation are smaller. Some stainless steels and
aluminium alloys with a very fine grain size exhibit superplastic behaviour. The glass blower is able to pull
his working material to very long rods without necking, because the exponent m for glass approaches
one. If m = 1, the material behaves like a viscous liquid and exhibits
Newtonian flow
Plastic Deformation
slip planes are the closest packed planes in the crystal.
the directions along which slip occurs are the closest
packed directions.
➢ In typical metallic crystals, such as copper, the bonds are nondirectional and not so strong as in a covalent crystal.
So, the dislocations are wide and the Peierls– Nabarro stress is low. These crystals exhibit a considerable amount of
plastic deformation and are said to be ductile. A copper wire can be cold drawn to hundred times its original length
without breaking, in spite of the work hardening that occurs during drawing. In metals of the transition groups,
such as iron, some covalent character persists due to d orbital bonding which is directional and, correspondingly,
the transition metal crystals are harder than copper. They cannot be cold worked to the same extent as copper.
➢ In ionic crystals, the bonds are of moderate strength and nondirectional. From this, one would have expected simple
ionic crystals to have some ability to undergo plastic deformation. However, plastic deformation occurs here only
under special circumstances when the crystal surface is free of cracks that can cause brittle failure. The main reason
for this behaviour seems to lie in the fact that the Burgers vector of dislocations in ionic crystals tends to be large.
➢ Intermetallic compounds and other complex crystal structures such as Fe3C and CuAl2 do not have favourable
crystal planes and directions for easy slip andtherefore tend to be brittle. Ordered compounds such as CuZn require
dislocations to move in pairs in order to preserve the order during slip. Hence,they possess very limited ductility.
The Effect of Temperature on the Stress to Move a Dislocation
This indicates that there are sources within the crystal, which
generate new dislocations during plastic deformation. One such
source is called the Frank–Read source.
➢ Strain Hardening
➢ Grain Refinement
➢ Precipitation Strengthening
Strain Hardening
Single crystal experiments show that after a certain stress level is reached, the rate of strain hardening decreases with
further strain. This phenomenon is called dynamic recovery. This stress level is sufficient to activate screw dislocations to
cross-slip. As the Burgers vector of the screw dislocations is parallel to the dislocation line, they are free to move on any out
of several equivalent slip planes. When deformation is initiated, they move on those planes where the resolved shear stress
is a maximum. If they happen to pile up against an immobile dislocation, they can cross-slip provided the applied shear
stress is sufficiently large to give the necessary resolved shear stress on the new slip plane.
Grain Refinement
The normal range of grain sizes in metals varies in ASTM specification
from ASTM 1 to ASTM 8. In recent years, ultra fine grain sizes have
been produced in microalloyed steels, which contain small quantities
of strong carbide forming elements such as Nb, V or Ti. These
elements go into solution in austenite on reheating the steel billet to
about 1300°C. When the steel is subsequently hot rolled, the
temperature falls gradually an the solubility of the alloying elements
decreases. Very fine particles of alloy carbides precipitate from
austenite. These precipitates effectively pin down the migrating grain
boundaries during repeated recrystallization of the deformed
austenite between passes at successive stands of the rolling mill.
This decreased growth rate due to pinning yields a
fine-grained recrystallized austenite. A fine-grained austenite
provides more potential sites at the grain boundaries for the
nucleation of ferrite as the temperature further drops. the end result
is thus a very fine grain size of ferrite in the steel of about 2–3 m
(ASTM 14–15). The yield strength increases as per the Hall–Petch
Electron micrograph of dislocations piled up against twin
equation. The increase is about 50%.
boundariesin Cu-8%Al alloy. Magnification 15 000 X .
Solid Solution Strengthening
The strengthening effect of solute atoms in a crystal depends on the size difference between the solute and
the solvent atoms and the concentration of the solute. The more is the difference in the atomic sizes, the
more is the strengthening effect. Higher concentrations of the solute also increase the strength
proportionately. It is not possible to dissolve a large concentration of a solute, which differs substantially in
size from the solvent, as this would be contradicting the Hume–Rothery’s rules. However, this difficulty could
be partly overcome by quenching from an elevated temperature and retaining the larger solubility at the
higher temperature in the form of a supersaturated solution at the lower temperature. Quenching austenite
retains all the carbon in solution in the product phase martensite, producing very hard steels.
Martensite in steels is a supersaturated solution of carbon in iron, obtained by quenching the steel and not allowing the
carbon to diffuse out of the iron lattice. For example, an eutectoid steel contains 0.8% carbon in martensite, which is
some 40 times more than the equilibrium solubility in ferrite (~0.02%) at the eutectoid temperature. On quenching, the
carbon gets trapped in the interstitial positions, as the austenite shears over to form the martensitic structure. The stress
field produced by the oversized carbon atoms is so intense that the dislocation motion is very effectively hindered.
Indeed, it is so effective that it becomes necessary to temper the martensite to restore some ductility, at the expense of
some hardness.
In addition to solute strengthening, the martensitic plates may contain either a high dislocation density or very fine
transformation twins, depending on the type of martensite obtained. These features also aid in increasing the strength of
martensite.
Precipitation Strengthening
A microstructure that consists of a very fine, submicroscopic distribution of precipitates in a matrix has a
good strength. The dislocations moving in the matrix are effectively hindered by the closely spaced
precipitate particles, as they have to bend around and bypass the particles. The strengthening effect is
inversely proportional to the particle spacing, the minimum attainable spacing being of the order of 100 Å.
Such a microstructure tends to become unstable at elevated temperatures, as the particles undergo
coarsening which increases the interparticle spacing and decreases the strength correspondingly. This
happens during overageing in aluminium alloys.
CREEP
Creep is the permanent deformation of a material under load as a function of time. It is appreciable only at
temperatures above 0.4Tm . Room temperature for iron is 0.16Tm and for copper it is 0.22Tm . so that creep
at room temperature is negligible in these materials. On the other hand, room temperature is about half
of the melting point (in kelvin) for lead and so it undergoes creep at room temperature. Creep of lead under
its own weight is seen in old time roofs, where a thick rim of lead is found at the edge of the roofs.
In stage I, the creep rate decreases with time; the effect
of work hardening is more than that of recovery in this
stage.
silicon nitride (Si3N4 :) selected parts of a heat engine such as piston rings and cylinder heads can be produced from
ceramic materials.
➢ Metals and alloys can be used under more versatile conditions. Most creep resistant alloys consist of a base metal of a
fairly high melting point.
The creep resistance of these alloys in the temperature range from 0.5Tm up to the melting point is considerably improved
by a special strengthening process called dispersion hardening. In TD (thoria dispersed) nickel, fine particles of thoria
(ThO2) are dispersed in the nickel matrix and the interparticle distance is small enough for effective hindrance of
dislocation motion in the matrix. TD nickel maintains its strength up to 0.9Tm.
In nickel base superalloys, the coarsening is prevented by a different mechanism. Here, the precipitate particles of
Ni3(Ti,Al) form an interface with the matrix of a very low energy, about 0.005 J m–2 (5 erg/cm2). As the decrease in the
total surface energy is the driving force for coarsening, very little driving force is available here.
➢ cold working cannot be used for creep resistance. At temperatures above 0.4Tm , recrystallization
will occur quite readily and the cold-workedstrength will be lost on recrystallization.
➢ Solid solution strengthening can be used for better creep resistance, in the same way as in plastic
deformation.
➢ A fine grained material is desirable for better mechanical properties in a low temperature
application, where creep is not important. On the other hand, for high temperature applications,
fine-grained materials are to be avoided, as grain boundary sliding can add to creep deformation.
There are no grain boundaries in a single crystal and, therefore, grain boundary sliding is not a
problem here. Single crystal titanium turbine blades have been tried out, even though the costis
an inhibiting factor here.
➢ Fracture refers to the failure of a material under load by breaking into two or more pieces. Fracture can occur under all
service conditions.
➢ Materials subjected to alternating or cyclic loading (as in machines) fail due to fatigue. The fracture under such
circumstances is called fatigue fracture.
➢ Ductile fracture is the rupture of a material after a considerable amount of plastic deformation.
Materials begin to neck beyond the ultimate tensile strength, which is at the maximum point in the load-elongation curve.
Fully ductile materials will continue to neck down to an infinitesimally thin edge or a point and thus fail, as the cross-section
at the neck becomes so small that it cannot bear the load any longer.
Brittle materials, however, break at a much lower stress, of the order of Y/1000.
Existence of tiny cracks in brittle materials is the reason for their poor tensile strength.
When the critical stress is applied to a brittle material, the pre-
existing crack propagates spontaneously with a decrease in
energy, culminating in fracture
For the atomically sharp crack shown in Fig. 12.4, is
about 1 Å (10–10 m). With c ~ 1 m, this yields
max= 200. This concentrated stress is of the same order
as the ideal strength (Y/6) and, therefore, is sufficient to
cause rupture of the atomic bonds at the tip.
➢ In brittle materials such as silicate glass, plastic deformation is nonexistent. The stress concentration at
the tip of the crack is not relaxed due to plastic deformation and consequently the crack tip remains
sharp.
Gc gives the value of the strain energy released per unit area of the crack surface when unstable crack extension
(leading to fracture) takes place.
Critical Stress Intensity Factor Kic (Or Kc).
For a sharp crack in an infinitely wide plate, when the applied tensile stress is perpendicular to the crack
faces, the critical stress intensity factor is given by
Fracture initiates in a material as soon as K Ic is reached, either through increasing stress or increasing c or both.
Kic is a material property, just as yield strength, UTS, etc., and can be determined from fracture tests.
In fracture safe design, if the size of the most damaging defect is known, the design stress can be computed from KIc
Alternatively, if the design stress is given, the critical crack size calculated from Kic must be sufficiently larger than
the smallest size of the crack detectable by available inspection techniques.
The Ductile-Brittle Transition
Common BCC metals become brittle at low temperatures or at extremely high rates of strain. BCC metals generally
require a high stress to move dislocations and this stress increases rapidly with decreasing temperature. The stress
required to propagate a crack, on the other hand, is not a strong function of temperature. So, at some temperature
called the ductile-brittle transition temperature, the stress to propagate a crack, is equal to the stress to move
dislocations,. At temperatures higher than the transition temperature, the material first yields plastically. At
temperatures lower than the transition temperature, the material is brittle. Here, the actual brittle fracture stress
may be controlled by the yield stress, as some microscopic yielding may be necessary to nucleate a crack. As soon as
the applied stress reaches yield strres , the crack is nucleated at the intersection of slip planes and propagates
rapidly.
Many FCC metals, on the other hand, remain ductile even at very low temperatures.
When the slip systems on the basal plane are the only ones operating, polycrystalline HCP metals are brittle, as there
are not enough slip systems to maintain the grain boundary integrity. For the same reason, polycrystalline ionic
crystals are also brittle.