Unit-2
Unit-2
When an elastic (not elastomeric, or long range elastic) material is stressed, there is an
immediate and corresponding strain response. Figure 43 illustrates this by showing
schematically the strain response to a particular stress history. Note that when the stress is
removed the strain also returns to zero. So in a perfectly elastic material all the deformation is
returned to the forcing agency. If this energy had not been stored elastically then it would
have been dissipated as either heat or sound. Tyre squeal and the heat build-up in the
sidewalls of car tyres are good examples of such dissipation.
Figure 43 When stressed, a perfectly elastic material deforms in proportion to the applied
stress and returns to its original state when the stress is released
If the material is linear and elastic then the applied stress σ directly proportional to the strain
ε. Then, for simple tension,
So some plastics like HIPS yield, but others fracture in a brittle manner like polystyrene.
Rubbers do not yield, but at high strains some of them crystallise and hence stiffen. When the
stresses are removed from a polymeric material before fracture, the strain recovery path is not
necessarily identical to that of the loading part of the deformation cycle. So energy must have
been dissipated during the deformation of such materials – another indication of deviation
from perfect elasticity. Both the deformation and the subsequent recovery are time-
dependent, suggesting that some part of their behaviour is viscous. In fact solid polymers
show a combination of elastic and viscous behaviour known as viscoelasticity. The degree of
viscoelasticity is strongly dependent upon the temperature of test and the rate at which the
polymer is deformed, as well as such structural variables as degree of crystallinity,
crosslinking, and molecular mass.
5.2.1 Elastic and viscoelastic behaviour When an elastic (not elastomeric, or long range
elastic) material is stressed, there is an immediate and corresponding strain response. Figure
4389 illustrates this by showing schematically the strain response to a particular stress
history. Note that when the stress is removed the strain also returns to zero. So in a perfectly
elastic material all the deformation is returned to the forcing agency. If this energy had not
been stored elastically then it would have been dissipated as either heat or sound. Tyre squeal
and the heat build-up in the sidewalls of car tyres are good examples of such dissipation.
Figure 43 When stressed, a perfectly elastic material deforms in proportion to the applied
stress and returns to its original state when the stress is released If the material is linear and
elastic then the applied stress σ directly proportional to the strain ε. Then, for simple tension,
where E is a constant known as Young's modulus, and is considered to be a property of the
material. For polymers, due to time-dependence and nonlinearity, E is not a constant and the
term tensile modulus is used to reflect this. Various values of tensile modulus are tabulated in
the Data Book for a range of polymers together with those of Young's modulus for other
materials. The stress-strain curves for PS, HIPS and two types of rubber are shown in Figure
4490. While polystyrene apparently obeys Hooke's law (Equation (27)), HIPS yields and
necks before failing. By contrast, rubbers exhibit long-range elasticity and fail only at many
hundred per cent strain. Figure 44 Tensile stress-strain curves for some rubbers and plastics
So some plastics like HIPS yield, but others fracture in a brittle manner like polystyrene.
Rubbers do not yield, but at high strains some of them crystallise and hence stiffen. When the
stresses are removed from a polymeric material before fracture, the strain recovery path is not
necessarily identical to that of the loading part of the deformation cycle. So energy must have
been dissipated during the deformation of such materials – another indication of deviation
from perfect elasticity. Both the deformation and the subsequent recovery are time-
dependent, suggesting that some part of their behaviour is viscous. In fact solid polymers
show a combination of elastic and viscous behaviour known as viscoelasticity. The degree of
viscoelasticity is strongly dependent upon the temperature of test and the rate at which the
polymer is deformed, as well as such structural variables as degree of crystallinity,
crosslinking, and molecular mass.
Viscoelasticity is the property of materials that exhibit both viscous and elastic
characteristics when undergoing deformation. Viscous materials, like honey, resist shear flow
and strain linearly with time when a stress is applied. Elastic materials strain when stretched
and quickly return to their original state once the stress is removed. Viscoelastic materials
have elements of both of these properties and, as such, exhibit time-dependent strain.
Whereas elasticity is usually the result of bond stretching along crystallographic planes in an
ordered solid, viscosity is the result of the diffusion of atoms or molecules inside an
amorphous material.[1]
Stress–strain curves for a purely elastic material (a) and a viscoelastic material (b). The red
area is a hysteresis loop and shows the amount of energy lost (as heat) in a loading and
Unlike purely elastic substances, a viscoelastic substance has an elastic component and a
viscous component. The viscosity of a viscoelastic substance gives the substance a strain rate
dependence on time.[1] Purely elastic materials do not dissipate energy (heat) when a load is
applied, then removed.[1] However, a viscoelastic substance loses energy when a load is
applied, then removed. Hysteresis is observed in the stress–strain curve, with the area of the
loop being equal to the energy lost during the loading cycle.[1] Since viscosity is the resistance
to thermally activated plastic deformation, a viscous material will lose energy through a
loading cycle. Plastic deformation results in lost energy, which is uncharacteristic of a purely
elastic material's reaction to a loading cycle.[1]
Types of viscoelasticity
Linear viscoelasticity is when the function is separable in both creep response and load. All
linear viscoelastic models can be represented by a Volterra equation connecting stress and
strain:
or
where
t is time
is stress
is strain
and are instantaneous elastic moduli for creep and relaxation
K(t) is the creep function
F(t) is the relaxation function
Nonlinear viscoelasticity is when the function is not separable. It usually happens when the
deformations are large or if the material changes its properties under deformations.
The secondary bonds of a polymer constantly break and reform due to thermal motion.
Application of a stress favors some conformations over others, so the molecules of the
polymer will gradually "flow" into the favored conformations over time.[13] Because thermal
motion is one factor contributing to the deformation of polymers, viscoelastic properties
change with increasing or decreasing temperature. In most cases, the creep modulus, defined
as the ratio of applied stress to the time-dependent strain, decreases with increasing
temperature. Generally speaking, an increase in temperature correlates to a logarithmic
decrease in the time required to impart equal strain under a constant stress. In other words, it
takes less work to stretch a viscoelastic material an equal distance at a higher temperature
than it does at a lower temperature.
Viscoelastic creep
Main article: Creep (deformation)
a) Applied stress and b) induced strain (b) as functions of time over a short period for a
viscoelastic material.
At a time , a viscoelastic material is loaded with a constant stress that is maintained for a
sufficiently long time period. The material responds to the stress with a strain that increases
until the material ultimately fails, if it is a viscoelastic liquid. If, on the other hand, it is a
viscoelastic solid, it may or may not fail depending on the applied stress versus the material's
ultimate resistance. When the stress is maintained for a shorter time period, the material
undergoes an initial strain until a time , after which the strain immediately decreases
(discontinuity) then gradually decreases at times to a residual strain.
Viscoelastic creep data can be presented by plotting the creep modulus (constant applied
stress divided by total strain at a particular time) as a function of time.[14] Below its critical
stress, the viscoelastic creep modulus is independent of stress applied. A family of curves
describing strain versus time response to various applied stress may be represented by a
single viscoelastic creep modulus versus time curve if the applied stresses are below the
material's critical stress value.
Viscoelastic creep is important when considering long-term structural design. Given loading
and temperature conditions, designers can choose materials that best suit component
lifetimes.
SLIP (MATERIALS SCIENCE)
In materials science, a slip system describes the set of symmetrically identical slip planes
and associated family of slip directions for which dislocation motion can easily occur and
lead to plastic deformation. An external force makes parts of the crystal lattice glide along
each other, changing the material's geometry. Depending on the type of lattice, different slip
systems are present in the material. More specifically, slip occurs on close-packed planes
(those containing the greatest number of atoms per area), and in close-packed directions
(most atoms per length). The magnitude and direction of slip are represented by the Burgers
vector. The picture on the right shows a schematic view of the slip mechanism. The slip
planes and slip directions in a crystal have specific crystallographic forms. The slip planes are
normally the planes with the highest density of atoms, i.e., those most widely spaced, and the
direction of the slip is the direction in the slip plane that corresponds to one of the shortest
lattice translation vectors. Often, this is the direction in which atoms are most closely spaced.
A slip plane and a slip direction constitute a slip system. A critical resolved shear stress is
required to initiate a slip.[1] Slip is an important mode of deformation mechanism in crystals.
For metals and technically used metallic alloys it is by far the most important deformation
mechanism and subject to current research in materials science.
Slip systems
Face centered cubic crystals
Lattice configuration of the close packed slip plane in an fcc material. The arrow represents
the burgers vector in this dislocation glide system.
Slip in face centered cubic (fcc) crystals occurs along the close packed plane. Specifically,
the slip plane is of type {111}, and the direction is of type <110>. In the diagram, the specific
plane and direction are (111) and [110], respectively. Given the permutations of the slip plane
types and direction types, fcc crystals have 12 slip systems. In the fcc lattice, the norm of the
Burgers vector, b, can be calculated using the following equation:[2]
[2]
Lattice configuration of the slip plane in a bcc material. The arrow represents the burgers
vector in this dislocation glide system.
Slip in body-centered cubic (bcc) crystals occurs along the plane of shortest Burgers vector as
well; however, unlike fcc, there are no truly close-packed planes in the bcc crystal structure.
Thus, a slip system in bcc requires heat to activate. Some bcc materials (e.g. α-Fe) can
contain up to 48 slip systems. There are six slip planes of type {110}, each with two <111>
directions (12 systems). There are 24 {123} and 12 {112} planes each with one <111>
direction (36 systems, for a total of 48). While the {123} and {112} planes are not exactly
identical in activation energy to {110}, they are so close in energy that for all intents and
purposes they can be treated as identical. In the diagram on the right the specific slip plane
and direction are (110) and [111], respectively.[2]
[2]
Hexagonal close packed crystals
Slip in hexagonal close packed (hcp) metals is much more limited than in bcc and fcc crystal
structures. Usually, hcp crystal structures allow slip on the densely packed basal {0001}
planes along the <1120> directions. The activation of other slip planes depends on various
parameters, e.g. the c/a ratio. Since there are only 3 independent slip systems on the basal
planes, for arbitrary plastic deformation additional slip or twin systems needs to be activated.
This typically requires are much higher resolved shear stress and results in the brittle
behavior of hcp polycrystals.
Cadmium, zinc, magnesium, titanium, and beryllium have a slip plane at {0001} and a slip
direction of <1120>. This creates a total of three slip systems, depending on orientation.
(Remember that a slip system is a combination of a slip plane and a slip direction) Other
combinations are also possible.[3]
There are two types of dislocations in crystals that can induce slip-edge dislocations and
screw dislocations. Edge dislocations have the direction of the burgers vector perpendicular
to the dislocation line, while screw dislocations have the direction of the burgers vector
parallel to the dislocation line. The type of dislocations generated largely depends on the
direction of the applied stress, temperature and other factors. Screw dislocations can easily
cross slip from one plane to another if the other slip plane contains the direction of the
burgers vector.[1]
Plastic Deformation
When a metal is stressed below its elastic limit, the resulting deformation or strain produced
in the metal is temporary. This strain or deformation vanishes after the removal of stress and
the metal goes back to the original dimensions. When it is stressed above the elastic limit,
permanent deformation takes place and the metal does not returns to its original shape after
the removal of stress. The ability of a metal to undergo plastic deformation is one of the
important property which is utilised for shaping of metals by various fabrication process such
as rolling, forging, drawing, extrusion etc.,
Mechanism Of Plastic Deformation
Slip is a permanent displacement of one part of crystal relative to the other part. slip involves
sliding of one plane of atoms over the other. The plane on which the slip occurs are called
slip planes and the direction in which this occurs are called slip direction. Slip occurs when
shear stress applied exceeds a critical value.During slip each atom usually moves same
integral number of atomic distances along the slip plane producing a step,but the orientation
of the crystal remains the same. Slip planes are usually the most close packed planes i.e., the
planes of maximum atomic density. Such planes obviously will be widely spaced i.e., the
interplanar distance between such planes is more. Slip results from the motion of dislocations
from one place to the other place. There are two basic types of dislocations movements called
as glide and climb. In glide, the dislocation moves in a surface defined by its line and
Burger’s vector (glide is conservative motion of dislocations). In climb, the dislocation
moves out of the glide surface and therefore, climb becomes a non conservation motion of
dislocation. Slip is the most common manifestation of glide.
Twinning:
Twinning is a process in which the atoms in a part of the crystal subjected to stress rearrange
themselves so that the orientation of the part changes in such a way that the distorted part
becomes a mirror image of the other part. The plane across which the two part are mirror
images is called twinning plane or composition plane.Like slip, twinning also occurs along
the certain crystallographic planes and directions. These planes and directions are called as
twin planes and twin directions.The important role of twinning in plastic deformation is that
it causes changes in plane orientation so that further slip can occur.
Resolved shear stress is the component of shear stress, resolved in the direction of slip,
necessary to initiate slip in a grain. It is a constant for a given crystal.
Tests have been conducted on single crystals of metals to measure the shear stress required to
initiate plastic deformation, or cause atomic planes to slip. Since this is a threshold value, it is
referred to as critical; and since it is a component of the applied force or stress, it is said to be
resolved; that is, the critical resolved shear stress.
Resolved shear stress is given by τ = σ cos Φ cos λ[1] where σ is the magnitude of the applied
tensile stress, Φ is the angle between the normal of the slip plane and the direction of the
applied force and λ is the angle between the slip direction and the direction of the applied
force. whereas, critical resolved shear stress value is given by τ =σ (cosΦ cosλ)max
Crystalline materials tend to deform or fail by the relative motion of planes of atoms under
the action of stress. This motion is induced by the component of stresses acting across the slip
planes. The deformation process is a collective motion of adjacent slip planes. But all the
planes do not start deforming simultaneously. Rather slip starts from a single plane and then
other planes follow. The first slip occurs when the shear stress across the plane exceeds a
certain value. This threshold value is called Critical Resolved Shear Stress.
Temperature and crystal geometry influence the minimum stress required to cause the shear.
But when the temperature and crystal geometry too is held constant and the crystal is loaded,
a new scenario emerges.
A pure crystalline solid, when pulled along different orientations, requires different amount
of loads in each instance for the very first slip to occur (though the stress required for the
planes to slip remains same). This implies that the shear stress resulting from the loading is
heavily affected by the orientation of the slip plane with the tensile axis.
To find a generalized mathematical expression for the dependence of the load on the
orientation let us assume that a cylindrical specimen of the crystalline solid is being loaded
with tension.
As the slip plane is inclined at an angle(φ) with the tensile axis, the cross sectional area of the
slip plane will be A/cos φ. Slip occurs on a direction which is contained within the slip plane.
The tensile load will have a component along the slip direction which will be responsible for
the shear stress developed on the slip plane. The component of the load will be P · cos λ.
Shear stress on the slip plane along the slip direction is given by
P/A can be substituted for tensile stress (σ), and the equation then becomes:
τ = σ · cos λ · cos φ
In materials science, creep (sometimes called cold flow) is the tendency of a solid material to
move slowly or deform permanently under the influence of mechanical stresses. It can occur
as a result of long-term exposure to high levels of stress that are still below the yield strength
of the material. Creep is more severe in materials that are subjected to heat for long periods,
and generally increases as they near their melting point.
The rate of deformation is a function of the material properties, exposure time, exposure
temperature and the applied structural load. Depending on the magnitude of the applied stress
and its duration, the deformation may become so large that a component can no longer
perform its function — for example creep of a turbine blade will cause the blade to contact
the casing, resulting in the failure of the blade. Creep is usually of concern to engineers and
metallurgists when evaluating components that operate under high stresses or high
temperatures. Creep is a deformation mechanism that may or may not constitute a failure
mode. For example, moderate creep in concrete is sometimes welcomed because it relieves
tensile stresses that might otherwise lead to cracking.
Unlike brittle fracture, creep deformation does not occur suddenly upon the application of
stress. Instead, strain accumulates as a result of long-term stress. Therefore, creep is a "time-
dependent" deformation.
The temperature range in which creep deformation may occur differs in various materials.
For example, tungsten requires a temperature in the thousands of degrees before creep
deformation can occur, while ice will creep at temperatures near 0 °C (32 °F).[1] As a general
guideline, the effects of creep deformation generally become noticeable at approximately
30% of the melting point (as measured on a thermodynamic temperature scale such as Kelvin
or Rankine) for metals, and at 40–50% of melting point for ceramics.[citation needed] Virtually any
material will creep upon approaching its melting temperature. Since the creep minimum
temperature is related to the melting point, creep can be seen at relatively low temperatures
for some materials. Plastics and low-melting-temperature metals, including many solders, can
begin to creep at room temperature, as can be seen markedly in old lead hot-water pipes.
Glacier flow is an example of creep processes in ice.
Stages of creep
Strain as a function of time due to constant stress over an extended period for a viscoelastic
material.
In the initial stage, or primary creep, the strain rate is relatively high, but slows with
increasing time. This is due to work hardening. The strain rate eventually reaches a minimum
and becomes near constant. This is due to the balance between work hardening and annealing
(thermal softening). This stage is known as secondary or steady-state creep. This stage is the
most understood. The characterized "creep strain rate" typically refers to the rate in this
secondary stage. Stress dependence of this rate depends on the creep mechanism. In tertiary
creep, the strain rate exponentially increases with stress because of necking phenomena.
Fracture always occurs at the tertiary stage. Creep is a very important aspect of material
science.
Mechanisms of creep
The mechanism of creep depends on temperature and stress. Various mechanisms are:
where is the creep strain, C is a constant dependent on the material and the particular creep
mechanism, m and b are exponents dependent on the creep mechanism, Q is the activation
energy of the creep mechanism, σ is the applied stress, d is the grain size of the material, k is
Boltzmann's constant, and T is the absolute temperature.
Dislocation creep
Main article: Dislocation creep
At high stresses (relative to the shear modulus), creep is controlled by the movement of
dislocations. For dislocation creep, Q = Q(self diffusion), m = 4–6, and b = 0. Therefore,
dislocation creep has a strong dependence on the applied stress and no grain size dependence.
Some alloys exhibit a very large stress exponent (n > 10), and this has typically been
explained by introducing a "threshold stress," σth, below which creep can't be measured. The
modified power law equation then becomes:
Nabarro-Herring creep
Coble creep
Main article: Coble creep
Coble creep is a second form of diffusion controlled creep. In Coble creep the atoms diffuse
along grain boundaries to elongate the grains along the stress axis. This causes Coble creep to
have a stronger grain size dependence than Nabarro-Herring creep. For Coble creep k is
related to the diffusion coefficient of atoms along the grain boundary, Q = Q(grain boundary
diffusion), m = 1, and b = 3. Because Q(grain boundary diffusion) < Q(self diffusion), Coble
creep occurs at lower temperatures than Nabarro-Herring creep. Coble creep is still
temperature dependent, as the temperature increases so does the grain boundary diffusion.
However, since the number of nearest neighbors is effectively limited along the interface of
the grains, and thermal generation of vacancies along the boundaries is less prevalent, the
temperature dependence is not as strong as in Nabarro-Herring creep. It also exhibits the
same linear dependence on stress as Nabarro-Herring creep.
Creep of polymers
a) Applied stress and b) induced strain as functions of time over a short period for a
viscoelastic material.
Creep can occur in polymers and metals which are considered viscoelastic materials. When a
polymeric material is subjected to an abrupt force, the response can be modeled using the
Kelvin-Voigt model. In this model, the material is represented by a Hookean spring and a
Newtonian dashpot in parallel. The creep strain is given by the following convolution
integral:
where:
σ = applied stress
C0 = instantaneous creep compliance
C = creep compliance coefficient
= retardation time
= distribution of retardation times
At a time t0, a viscoelastic material is loaded with a constant stress that is maintained for a
sufficiently long time period. The material responds to the stress with a strain that increases
until the material ultimately fails. When the stress is maintained for a shorter time period, the
material undergoes an initial strain until a time t1 at which the stress is relieved, at which time
the strain immediately decreases (discontinuity) then continues decreasing gradually to a
residual strain.
Viscoelastic creep data can be presented in one of two ways. Total strain can be plotted as a
function of time for a given temperature or temperatures. Below a critical value of applied
stress, a material may exhibit linear viscoelasticity. Above this critical stress, the creep rate
grows disproportionately faster. The second way of graphically presenting viscoelastic creep
in a material is by plotting the creep modulus (constant applied stress divided by total strain
at a particular time) as a function of time.[2] Below its critical stress, the viscoelastic creep
modulus is independent of stress applied. A family of curves describing strain versus time
response to various applied stress may be represented by a single viscoelastic creep modulus
versus time curve if the applied stresses are below the material's critical stress value.
Additionally, the molecular weight of the polymer of interest is known to affect its creep
behavior. The effect of increasing molecular weight tends to promote secondary bonding
between polymer chains and thus make the polymer more creep resistant. Similarly, aromatic
polymers are even more creep resistant due to the added stiffness from the rings. Both
molecular weight and aromatic rings add to polymers' thermal stability, increasing the creep
resistance of a polymer.[3]
Both polymers and metals can creep. Polymers experience significant creep at temperatures
above ca. –200°C; however, there are three main differences between polymeric and metallic
creep.[4]
Polymers show creep basically in two different ways. At typical work loads (5 up to 50%)
ultra high molecular weight polyethylene (Spectra, Dyneema) will show time-linear creep,
whereas polyester or aramids (Twaron, Kevlar) will show a time-logarithmic creep.
Creep of concrete
Main article: Creep and shrinkage of concrete
The creep of concrete, which originates from the calcium silicate hydrates (C-S-H) in the
hardened Portland cement paste (which is the binder of mineral aggregates), is fundamentally
different from the creep of metals as well as polymers. Unlike the creep of metals, it occurs at
all stress levels and, within the service stress range, is linearly dependent on the stress if the
pore water content is constant. Unlike the creep of polymers and metals, it exhibits multi-
months aging, caused by chemical hardening due to hydration which stiffens the
microstructure, and multi-year aging, caused by long-term relaxation of self-equilibrated
micro-stresses in the nano-porous microstructure of the C-S-H. If concrete is fully dried it
does not creep, though it is difficult to dry concrete fully without severe cracking.
Applications
Creep on the underside of a cardboard box: a largely empty box was placed on a smaller box,
and fuller boxes were placed on top of it. Due to the weight, the portions of the empty box
not sitting on the lower box gradually crept downward.
Though mostly due to the reduced yield strength at higher temperatures, the collapse of the
World Trade Center was due in part to creep from increased temperature operation.[5]
The creep rate of hot pressure-loaded components in a nuclear reactor at power can be a
significant design constraint, since the creep rate is enhanced by the flux of energetic
particles.
Creep was blamed for the Big Dig tunnel ceiling collapse in Boston, Massachusetts that
occurred in July 2006.
An example of an application involving creep deformation is the design of tungsten light bulb
filaments. Sagging of the filament coil between its supports increases with time due to creep
deformation caused by the weight of the filament itself. If too much deformation occurs, the
adjacent turns of the coil touch one another, causing an electrical short and local overheating,
which quickly leads to failure of the filament. The coil geometry and supports are therefore
designed to limit the stresses caused by the weight of the filament, and a special tungsten
alloy with small amounts of oxygen trapped in the crystallite grain boundaries is used to slow
the rate of Coble creep.
In steam turbine power plants, pipes carry steam at high temperatures (566 °C (1,051 °F)) and
pressures (above 24.1 MPa or 3500 psi). In jet engines, temperatures can reach up to
1,400 °C (2,550 °F) and initiate creep deformation in even advanced coated turbine blades.
Hence, it is crucial for correct functionality to understand the creep deformation behavior of
materials.
Creep deformation is important not only in systems where high temperatures are endured
such as nuclear power plants, jet engines and heat exchangers, but also in the design of many
everyday objects. For example, metal paper clips are stronger than plastic ones because
plastics creep at room temperatures. Aging glass windows are often erroneously used as an
example of this phenomenon: measurable creep would only occur at temperatures above the
glass transition temperature around 500 °C (932 °F). While glass does exhibit creep under the
right con
ditions, apparent sagging in old windows may instead be a consequence of obsolete
manufacturing processes, such as that used to create crown glass, which resulted in
inconsistent thickness.[6][7]
Fractal geometry, using a deterministic Cantor structure, is used to model the surface
topography, where recent advancements in thermoviscoelastic creep contact of rough
surfaces are introduced. Various viscoelastic idealizations are used to model the surface
materials, for example, Maxwell, Kelvin-Voigt, Standard Linear Solid and Jeffrey media. [8]
Nimonic 75 has been certified by the European Union as a standard creep reference material.
[9]
DUCTILITY
Tensile test of an AlMgSi alloy. The local necking and the cup and cone fracture surfaces are
typical for ductile metals.
This tensile test of a nodular cast iron demonstrates low ductility.
In materials science, ductility is a solid material's ability to deform under tensile stress; this
is often characterized by the material's ability to be stretched into a wire. Malleability, a
similar property, is a material's ability to deform under compressive stress; this is often
characterized by the material's ability to form a thin sheet by hammering or rolling. Both of
these mechanical properties are aspects of plasticity, the extent to which a solid material can
be plastically deformed without fracture. Also, these material properties are dependent on
temperature and pressure (investigated by Percy Williams Bridgman as part of his Nobel
Prize–winning work on high pressures).
Ductility and malleability are not always coextensive – for instance, while gold has high
ductility and malleability, lead has low ductility but high malleability.[1] The word ductility is
sometimes used to encompass both types of plasticity.[2]
Materials science
Ductility can be quantified by the fracture strain , which is the engineering strain at which
a test specimen fractures during a uniaxial tensile test. Another commonly used measure is
the reduction of area at fracture .[3] The ductility of steel varies depending on the alloying
constituents. Increasing levels of carbon decrease ductility. Many plastics and amorphous
solids, such as Play-Doh, are also malleable. The most ductile metal is platinum and the most
malleable metal is gold.[4][5]
The ductile–brittle transition temperature (DBTT), nil ductility temperature (NDT), or nil
ductility transition temperature of a metal is the temperature at which the fracture energy
passes below a predetermined value (for steels typically 40 J[6] for a standard Charpy impact
test). DBTT is important since, once a material is cooled below the DBTT, it has a much
greater tendency to shatter on impact instead of bending or deforming. For example, zamak 3
exhibits good ductility at room temperature but shatters when impacted at sub-zero
temperatures. DBTT is a very important consideration in materials selection when the
material in question is subject to mechanical stresses. A similar phenomenon, the glass
transition temperature, occurs with glasses and polymers, although the mechanism is different
in these amorphous materials.
In some materials this transition is sharper than others. For example, the transition is
generally sharper in materials with a body-centered cubic (bcc) lattice than those with a face-
centered cubic (fcc) lattice. This can be observed in the brittleness of martensitic steel,
obtained by quenching and used for instance in files (body-centered tetragonal lattice) in
comparison with the toughness of austenitic steel, obtained by tempering (fcc lattice). DBTT
can also be influenced by external factors such as neutron radiation, which leads to an
increase in internal lattice defects and a corresponding decrease in ductility and increase in
DBTT.
The most accurate method of measuring the DBTT of a material is by fracture testing.
Typically, four point bend testing at a range of temperatures is performed on pre-cracked bars
of polished material.
BRITTLENESS
"Brittle" redirects here. For other uses, see Brittle (disambiguation).
A material is brittle if, when subjected to stress, it breaks without significant deformation
(strain). Brittle materials absorb relatively little energy prior to fracture, even those of high
strength. Breaking is often accompanied by a snapping sound. Brittle materials include most
ceramics and glasses (which do not deform plastically) and some polymers, such as PMMA
and polystyrene. Many steels become brittle at low temperatures (see ductile-brittle transition
temperature), depending on their composition and processing.
When used in materials science, it is generally applied to materials that fail when there is
little or no evidence of plastic deformation before failure. One proof is to match the broken
halves, which should fit exactly since no plastic deformation has occurred.
When a material has reached the limit of its strength, it usually has the option of either
deformation or fracture. A naturally malleable metal can be made stronger by impeding the
mechanisms of plastic deformation (reducing grain size, precipitation hardening, work
hardening, etc.), but if this is taken to an extreme, fracture becomes the more likely outcome,
and the material can become brittle. Improving material toughness is therefore a balancing
act.
FATIGUE FRACTURE
Metal Fatigue
A phenomenon which results in the sudden fracture of a component after a period of
cyclic loading in the elastic regime. Failure is the end result of a process involving the
initiation and growth of a crack, usually at the site of a stress concentration on the
surface. Occasionally, a crack may initiate at a fault just below the surface. Eventually
the cross sectional area is so reduced that the component ruptures under a normal
service load, but one at a level which has been satisfactorily withstood on many
previous occasions before the crack propagated. The final fracture may occur in a
ductile or brittle mode depending on the characteristics of the material. Fatigue
fractures have a characteristic appearance which reflects the initiation site and the
progressive development of the crack front, culminating in an area of final overload
fracture. Fig. la illustrates fatigue failure in a circular shaft. The initiation site is shown
and the shell-like markings, often referred to as beach markings because of their
resemblance to the ridges left in the sand by retreating waves, are caused by arrests in
the crack front as it propagates through the section. The hatched region on the opposite
side to the initiation site is the final region of ductile fracture. Sometimes there may be
more than one initiation point and two or more cracks propagate. This produces
features as in Fig. 1b with the the final area of ductile fracture being a band across the
middle. This type of fracture is typical of double bending where a component is
cyclically strained in one plane or where a second fatigue crack initiates at the opposite
side to a developing crack in a component subject to reverse bending. Some stress-
induced fatigue failures may show multiple initiation sites from which separate cracks
spread towards a common meeting point within the section.
Fig. 1
Fatigue strength is determined by applying different levels of cyclic stress to individual
test specimens and measuring the number of cycles to failure. Standard laboratory test
use various methods for applying the cyclic load, e.g. rotating bend, cantilever bend,
axial push-pull and torsion. The data are plotted in the form of a stress-number of
cycles to failure (S-N) curve, fig 2. Owing to the statistical nature of the failure, several
specimens have to be tested at each stress level. Some materials, notably low-carbon
steels, exhibit a flattening off at a particular stress level as at (a) in Fig.2 which is
referred to as the fatigue limit. As a rough guide, the fatigue limit is usually about 40%
of the tensile strength. In principle, components designed so that the applied stresses do
not exceed this level should not fail in service. The difficulty is a localised stress
concentration may be present or introduced during service which leads to initiation,
despite the design stress being normally below the 'safe' limit. Most materials, however,
exhibit a continually falling curve as at (b) and the usual indicator of fatigue strength is
to quote the stress below which failure will not be expected in less than a given number
of cycles which is referred to as the endurance limit.
Fig2
Although fatigue data may be determined for different materials it is the shape of a
component and the level of applied stress which dictate whether a fatigue failure is to
be expected under particular service conditions. Surface condition is also important.
Often complete components or assemblies, e.g. railway bogie frames or aircraft
fuselage, will be tested by subjecting them to an accelerated loading spectrum
reproducing what they are likely to experience over their entire service lifetime.
o E = modulus of elasticity
o s= specific surface energy
o a = half the length of an internal crack
Fatigue
In materials science, fatigue is the weakening of a material caused by repeatedly applied
loads. It is the progressive and localised structural damage that occurs when a material is
subjected to cyclic loading. The nominal maximum stress values that cause such damage may
be much less than the strength of the material typically quoted as the ultimate tensile stress
limit, or the yield stress limit.
Fatigue occurs when a material is subjected to repeated loading and unloading. If the loads
are above a certain threshold, microscopic cracks will begin to form at the stress
concentrators such as the surface, persistent slip bands (PSBs), and grain interfaces.[1]
Eventually a crack will reach a critical size, the crack will propagate suddenly, and the
structure will fracture. The shape of the structure will significantly affect the fatigue life;
square holes or sharp corners will lead to elevated local stresses where fatigue cracks can
initiate. Round holes and smooth transitions or fillets will therefore increase the fatigue
strength of the structure.
S-N curve
S-N curves are derived from tests on samples of the material to be characterized (often called
coupons) where a regular sinusoidal stress is applied by a testing machine which also counts
the number of cycles to failure. This process is sometimes known as coupon testing. Each
coupon test generates a point on the plot though in some cases there is a runout where the
time to failure exceeds that available for the test (see censoring). Analysis of fatigue data
requires techniques from statistics, especially survival analysis and linear regression.
The progression of the S-N curve can be influenced by many factors such as corrosion,
temperature, residual stresses, and the presence of notches. The Goodman-Line is a method to
estimate the influence of the mean stress on the fatigue strength.
FRACTURE TOUGHNESS
n materials science, fracture toughness is a property which describes the ability of a material
containing a crack to resist fracture, and is one of the most important properties of any
material for many design applications. The linear-elastic fracture toughness of a material is
determined from the stress intensity factor ( ) at which a thin crack in the material begins to
grow. It is denoted KIc and has the units of or . Plastic-elastic fracture
toughness is denoted by JIc, with the unit of J/cm2 or lbf-in/in2, and is a measurement of the
energy required to grow a thin crack.
The subscript Ic denotes mode I crack opening under a normal tensile stress perpendicular to
the crack, since the material can be made deep enough to stand shear (mode II) or tear (mode
III).
Fracture mechanics, which leads to the concept of fracture toughness, was broadly based on
the work of A. A. Griffith who, among other things, studied the behavior of cracks in brittle
materials.
Example values
The following table shows some typical values of fracture toughness for various materials:
Aluminum 14–28
Concrete 0.2–1.4
Consider a body with flaws (cracks) that is subject to some loading; the stability of the crack
can be assessed as follows. We can assume for simplicity that the loading is of constant
displacement or displacement controlled type (such as loading with a screw jack); we can
also simplify the discussion by characterizing the crack by its area, A. If we consider an
adjacent state of the body as being one with a broader crack (area A+dA), we can then assess
strain energy in the two states and evaluate strain energy release rate.
The rate is reckoned with respect to the change in crack area, so if we use U for strain energy,
the strain energy release rate is numerically dU/dA. It may be noted that for a body loaded in
constant displacement mode, the displacement is applied and the force level is dictated by
stiffness (or compliance) of the body. If the crack grows in size, the stiffness decreases, so the
force level will decrease. This decrease in force level under the same displacement (strain)
level indicates that the elastic strain energy stored in the body is decreasing—is being
released. Hence the term strain energy release rate which is usually denoted with symbol G.
The strain energy release rate is higher for higher loads and broader cracks. If the strain
energy so released exceeds a critical value Gc, then the crack will grow spontaneously. For
brittle materials, Gc can be equated to the surface energy of the (two) new crack surfaces; in
other words, in brittle materials, a crack will grow spontaneously if the strain energy released
is equal to or more than the energy required to grow the crack surface(s). The stability
condition can be written as
If the elastic energy released is less than the critical value, then the crack will not grow;
equality signifies neutral stability and if the strain energy release rate exceeds the critical
value, the crack will start growing in an unstable manner. For ductile materials, energy
associated with plastic deformation has to be taken into account. When there is plastic
deformation at the crack tip (as occurs most often in metals) the energy to propagate the crack
may increase by several orders of magnitude as the work related to plastic deformation may
be much larger than the surface energy. In such cases, the stability criterion has to be restated
as
Practically, this means a higher value for the critical value Gc. From the definition of G, we
can deduce that it has dimensions of work (or energy) /area or force/length. For ductile metals
GIc is around 50–200 kJ/m2, for brittle metals it is usually 1–5 and for glasses and brittle
polymers it is almost always less than 0.5.
The problem can also be formulated in terms of stress instead of energy, leading to the terms
stress intensity factor K (or KI for mode I) and critical stress intensity factor Kc (and KIc).
These Kc and KIc (etc.) quantities are commonly referred to as fracture toughness, though it is
equivalent to use Gc. Typical values for KIcare 150 MN/m3/2 for ductile (very tough) metals,
25 for brittle ones and 1–10 for glasses and brittle polymers. Notice the different units used
by GIc and KIc. Engineers tend to use the latter as an indication of toughness.
Conjoint action
There are number of instances where this picture of a critical crack is modified by corrosion.
Thus, fretting corrosion occurs when a corrosive medium is present at the interface between
two rubbing surfaces. Fretting (in the absence of corrosion) results from the disruption of
very small areas that bond and break as the surfaces undergo friction, often under vibrating
conditions. The bonding contact areas deform under the localised pressure and the two
surfaces gradually wear away. Fracture mechanics dictates that each minute localised fracture
has to satisfy the general rule that the elastic energy released as the bond fractures has to
exceed the work done in plastically deforming it and in creating the (very tiny) fracture
surfaces. This process is enhanced when corrosion is present, not least because the corrosion
products act as an abrasive between the rubbing surfaces.
Fatigue is another instance where cyclical stressing, this time of a bulk lump of metal, causes
small flaws to develop. Ultimately one such flaw exceeds the critical condition and fracture
propagates across the whole structure. The fatigue life of a component is the time it takes for
criticality to be reached, for a given regime of cyclical stress. Corrosion fatigue is what
happens when a cyclically stressed structure is subjected to a corrosive environment at the
same time. This not only serves to initiate surface cracks but (see below) actually modifies
the crack growth process. As a result the fatigue life is shortened, often considerably.
Griffith Theory of Brittle Fracture