Anda di halaman 1dari 10

BAN401 AIRCRAFT STRUCTURES I

UNIT V FAILURE THEORIES

Viscoelastic Material

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

Elastic versus Viscoelastic Behavior

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. Purely elastic materials
do not dissipate energy (heat) when a load is applied, then removed. 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. 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.

Specifically, viscoelasticity is a molecular rearrangement. When a stress is applied to a viscoelastic material,


such as a polymer, parts of the long polymer chain change positions and this movement or rearrangement is called
creep. Polymers remain a solid material even when these parts of their chains are rearranging in order to accompany
the stress, and as this occurs, it creates a back stress in the material. When the back stress is the same magnitude as the
applied stress, the material no longer creeps. When the original stress is taken away, the accumulated back stresses
will cause the polymer to return to its original form. The material creeps, which gives the prefix visco-, and the
material fully recovers, which gives the suffix -elasticity.

Ductile and Brittle Materials

S.
Ductile Materials Brittle Materials
No.

Ductile materials can be drawn into wires by Brittle materials break, crack, or snap easily and hence
1
stretching cannot be drawn into wires

2 Ductile materials show deformation Brittle materials do not show deformation

3 Ductility is affected by temperature Brittleness is affected by pressure (or stress)


Ductile materials will withstand large strains before
4 Brittle materials fracture at much lower strains
rupture

Ductile materials relatively have small elastic moduli Brittle materials have relatively large elastic moduli
5
and ultimate stresses and ultimate stresses

Ductile materials exhibit large strains and yielding Brittle materials fail suddenly and without much
6
before they fail warning

7 e.g. Steel, Aluminium e.g. Glass, Cast Iron

Ductile Failure

In ductile fracture, extensive plastic deformation (necking) takes place before fracture. The terms rupture or
ductile rupture describes the ultimate failure of ductile materials loaded in tension. Rather than cracking, the material
"pulls apart," generally leaving a rough surface. In this case there is slow propagation and absorption of large a mount
energy before fracture.

Many ductile metals, especially materials with high purity, can sustain very large deformation of 50–100% or
more strain before fracture under favorable loading condition and environmental condition. The strain at which the
fracture happens is controlled by the purity of the materials. At room temperature, pure iron can undergo deformation
up to 100% strain before breaking, while cast iron or high-carbon steels can barely sustain 3% of strain.

The basic steps in ductile fracture are: void formation, void coalescence (also known as crack formation),
crack propagation, and failure, often resulting in a cup-and-cone shaped failure surface.
Schematic representation of the steps in ductile fracture (in pure tension)

Brittle Failure

In brittle fracture, no apparent plastic deformation takes place before fracture. Brittle fracture typically
involves little energy absorption and occurs at high speeds (up to 7000 ft/sec in steel). In most cases brittle fracture
will continue even when loading is discontinued.

In brittle crystalline materials, fracture can occur by cleavage as the result of tensile stress acting normal to
crystallographic planes with low bonding (cleavage planes). In amorphous solids, by contrast, the lack of a crystalline
structure results in a conchoidal fracture, with cracks proceeding normal to the applied tension.
The basic sequence in a typical brittle fracture is: introduction of a flaw either before or after the material is
put in service, slow and stable crack propagation under recurring loading, and sudden rapid failure when the crack
reaches critical crack length based on the conditions defined by fracture mechanics.

Ductile to Brittle Transition

Under certain conditions, ductile materials can exhibit brittle behavior. Rapid loading, low temperature, and
triaxial stress constraint conditions may cause ductile materials to fail without prior deformation. This is known as a
ductile to brittle transition.

The ductile to brittle transition temperature is strongly dependant on the composition of the metal. Steel is the
most commonly used metal that shows this behaviour. For some steels the transition temperature can be around 0°C,
and in winter the temperature in some parts of the world can be below this. As a result, some steel structures are very
likely to fail in winter.

Fracture mechanics in materials ultimately extends to ease of dislocation movement. For samples in
which dislocation motion in uninhibited, the material may accommodate an applied load through plastic deformation.
For FCC metals, slip systems allow dislocations to move very easily along the close-packed planes. However, for
BCC metals, there are no close-packed planes that allow for easy dislocation migration, and thus dislocation
movement in these materials (BCC metals) require a thermal activation in order to slip. Otherwise, the material
accommodates the applied stress through more drastic mechanisms, such as breaking of bonds. This is the reason
behind the ductile to brittle transition.

Variation of impact energy absorbed before fracture with respect to temperature for different materials

The sinking of the titanic was caused primarily by the brittleness of the steel used to construct the hull of the
ship. In the icy water of the Atlantic, the steel was below the ductile to brittle transition temperature. In these
conditions even a small impact could have caused a large amount of damage. The impact of an iceberg on the ship's
hull resulted in brittle fracture of the bolts that were holding the steel plates together. Nowadays engineers know more
about this phenomenon and the composition of the steels used is much more controlled, result ing in a lower
temperature at which the ductile to brittle transition occurs.

Creep

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's 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.

Temperature Dependence of Creep

The temperature range in which creep deformation may occur differs in various materials. Lead can creep at
room temperature and 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). As a general guideline, the effects of creep deformation
generally become noticeable at approximately 35% of the melting point and at 45% of melting point for ceramics.
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, or transient creep, the strain rate is relatively high, but decreases with increasing time
and strain due to a process analogous to work hardening at lower temperatures. For instance, the dislocation density
increases and, in many materials, a dislocation subgrain structure is formed and the cell size decreases with strain. The
strain rate diminishes to a minimum and becomes near constant as the secondary stage begins. This is due to the
balance between work hardening and annealing (thermal softening). The secondary stage, referred to as "steady-state
creep", is the most understood. The microstructure is invariant during this stage, which means that recovery effects are
concurrent with deformation. No material strength is lost during these first two stages of creep.

The characterized "creep strain rate" typically refers to the constant 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 or internal cracks or voids decreases the effective area of the specimen. Strength
is quickly lost in this stage while the material's shape is permanently changed. The acceleration of creep deformation
in the tertiary stage eventually leads to material fracture.
Fatigue

Fatigue is the weakening of a material caused by repeatedly applied loads. It is the progressive and localized
structural damage that occurs when a material is subjected to cyclic loading.

Fatigue Life

The number of stress cycles of a specified character that a material withstands before failure is known as
fatigue life.

Endurance

The endurance or fatigue limit is defined as the maximum value of the stress which a standard specimen can
withstand without failure, for infinite number of cycles when subjected to completely reversed load.

Reversing cyclic stress, S – N Curves and endurance strength

A purely reversing or cyclic stress means when the stress alternates between equal positive and negative peak
stresses sinusoidally during each cycle of operation, as shown. In this diagram the stress varies with time between
+250 MPa to -250MPa. This kind of cyclic stress is developed in many rotating machine parts that are carrying a
constant bending load.

When a part is subjected cyclic stress, also known as range or reversing stress (Sr), it has been observed that
the failure of the part occurs after a number of stress reversals (N) even it the magnitude of Sr is below the material’s
yield strength. Generally, higher the value of Sr, lesser N is needed for failure.
For a typical material, the table and the graph above (S-N curve) show the relationship between the
magnitudes Sr and the number of stress reversals (N) before failure of the part. For example, if the part were subjected
to Sr= 81,000 psi, then it would fail after N=1000 stress reversals. If the same part is subjected to Sr = 61,024 psi, then
it can survive up to N=16,000 reversals, and so on.

It has been observed that for most of engineering materials, the rate of reduction of Sr becomes negligible near
the vicinity of N = 106 and the slope of the S-N curve becomes more or less horizontal. For the curve shown above, at
N= 106 , the slope of the curve has become horizontal at Sr=40,000 psi. Because the slope of the above S-N curve is
horizontal at N=106 reversals, that means if we keep the cyclic stress value less than Sr = 40,000 psi, then the part will
not fail no matter how many cycles have been operated, ie survive indefinitely.

The maximum completely reversing cyclic stress that a material can withstand for indefinite (or
infinite) number of stress reversals is known as the fatigue strength or endurance strength (Se) of the part
material. This is essentially the max completely reversing cyclic stress that a material can withstand for N=106 or
more, as the curve is horizontal after this point. For the example S-N curve shown above, Se = 40,000 psi.

Environmental Stress Failure

In materials science, environmental stress fracture or environment assisted fracture is the generic name given
to premature failure under the influence of tensile stresses and harmful environments of materials such as metals and
alloys, composites, plastics and ceramics.

Metals and alloys exhibit phenomena such as stress corrosion cracking, hydrogen embrittlement, liquid metal
embrittlement and corrosion fatigue all coming under this category. Environments such as moist air, sea water and
corrosive liquids and gases cause environmental stress fracture. Metal matrix composites are also susceptible to many
of these processes.
Plastics and plastic-based composites may suffer swelling, debonding and loss of strength when exposed to
organic fluids and other corrosive environments, such as acids and alkalies. Under the influence of stress and
environment, many structural materials, particularly the high-specific strength ones become brittle and lose their
resistance to fracture. While their fracture toughness remains unaltered, their threshold stress intensity factor for crack
propagation may be considerably lowered. Consequently, they become prone to premature fracture because of sub-
critical crack growth.

Stress Potential

Stress function or stress potential is a scalar potential function that can be used to find the stress. This function
satisfies equilibrium in the absence of body forces and is valid only for two-dimensional problems (plane stress/plane
strain).

e.g. Airy’s Stress Potential

Safe life and fail-safe structures

The danger of a catastrophic fatigue failure in the structure of an aircraft may be eliminated completely or
may become extremely remote if the structure is designed to have a safe life or to be fail-safe. In the former approach,
the structure is designed to have a minimum life during which it is known that no catastrophic damage will occur. At
the end of this life the structure must be replaced even though there may be no detectable signs of fatigue. If a
structural component is not economically replaceable when its safe life has been reached the complete structure must
be written off. Alternatively, it is possible for easily replaceable components such as undercarriage legs and
mechanisms to have a safe life less than that of the complete aircraft since it would probably be more economic al to
use, say, two lightweight undercarriage systems during the life of the aircraft rather than carry a heavier undercarriage
which has the same safe life as the aircraft.

The fail-safe approach relies on the fact that the failure of a member in a redundant structure does not
necessarily lead to the collapse of the complete structure, provided that the remaining members are able to carry the
load shed by the failed member and can withstand further repeated loads until the presence of the failed member is
discovered.

Such a structure is called a fail-safe structure or a damage tolerant structure. Generally, it is more economical
to design some parts of the structure to be failsafe rather than to have a long safe life since such components can be
lighter. When failure is detected, either through a routine inspection or by some malfunction, such as fuel leakage
from a wing crack, the particular aircraft may be taken out of service and repaired. However, the structure must be
designed and the inspection intervals arranged such that a failure, for example a crack, too small to be noticed at one
inspection must not increase to a catastrophic size before the next. The determination of crack propagation rates is
discussed later.

Some components must be designed to have a safe life; these include landing gear, major wing joints, wing–
fuselage joints and hinges on all-moving tailplanes or on variable geometry wings. Components which may be
designed to be fail-safe include wing skins which are stiffened by stringers and fuselage skins which are stiffened by
frames and stringers; the stringers and frames prevent skin cracks spreading disastrously for a sufficient period of time
for them to be discovered at a routine inspection.

Anda mungkin juga menyukai