Sie sind auf Seite 1von 14

1

Strength of materials

Strength of materials, also called mechanics of materials, is a subject which deals with the behavior
of solid objects subject to stresses and strains. The complete theory began with the consideration of the
behavior of one and two dimensional members of structures, whose states of stress can be approximated
as two dimensional, and was then generalized to three dimensions to develop a more complete theory of
the elastic and plastic behavior of materials. An important founding pioneer in mechanics of materials was
Stephen Timoshenko.

The study of strength of materials often refers to various methods of calculating the stresses and strains in
structural members, such as beams, columns, and shafts. The methods employed to predict the response
of a structure under loading and its susceptibility to various failure modes takes into account the
properties of the materials such as its yield strength, ultimate strength, Young's modulus, and Poisson's
ratio; in addition the mechanical element's macroscopic properties (geometric properties), such as its
length, width, thickness, boundary constraints and abrupt changes in geometry such as holes are
considered.

Definition

In materials science, the strength of a material is its ability to withstand an applied load without failure or
plastic deformation. The field of strength of materials deals with forces and deformations that result from
their acting on a material. A load applied to a mechanical member will induce internal forces within the
member called stresses when those forces are expressed on a unit basis. The stresses acting on the
material cause deformation of the material in various manner. Deformation of the material is called strain
when those deformations too are placed on a unit basis. The applied loads may be axial (tensile or
compressive), or [shear strength shear]. The stresses and strains that develop within a mechanical
member must be calculated in order to assess the load capacity of that member. This requires a complete
description of the geometry of the member, its constraints, the loads applied to the member and the
properties of the material of which the member is composed. With a complete description of the loading
and the geometry of the member, the state of stress and of state of strain at any point within the member
can be calculated. Once the state of stress and strain within the member is known, the strength (load
carrying capacity) of that member, its deformations (stiffness qualities), and its stability (ability to
maintain its original configuration) can be calculated. The calculated stresses may then be compared to
some measure of the strength of the member such as its material yield or ultimate strength. The
calculated deflection of the member may be compared to a deflection criteria that is based on the
member's use. The calculated buckling load of the member may be compared to the applied load. The
calculated stiffness and mass distribution of the member may be used to calculate the member's dynamic
response and then compared to the acoustic environment in which it will be used.

Material strength refers to the point on the engineering stressstrain curve (yield stress) beyond which the
material experiences deformations that will not be completely reversed upon removal of the loading and
as a result the member will have a permanent deflection. The ultimate strength refers to the point on the
engineering stressstrain curve corresponding to the stress that produces fracture.

Types of loadings

Transverse loading - Forces applied perpendicular to the longitudinal axis of a member. Transverse
loading causes the member to bend and deflect from its original position, with internal tensile and
compressive strains accompanying the change in curvature of the member. [1] Transverse loading
also induces shear forces that cause shear deformation of the material and increase the transverse
deflection of the member.

Axial loading - The applied forces are collinear with the longitudinal axis of the member. The forces
cause the member to either stretch or shorten.[2]

Torsional loading - Twisting action caused by a pair of externally applied equal and oppositely
directed force couples acting on parallel planes or by a single external couple applied to a member
that has one end fixed against rotation.

Stress terms
2

A material being loaded in a) compression, b) tension, c) shear.

Uniaxial stress is expressed by

where F is the force [N] acting on an area A [m2].[3] The area can be the undeformed area or the deformed
area, depending on whether engineering stress or true stress is of interest.

Compressive stress (or compression) is the stress state caused by an applied load that acts to
reduce the length of the material (compression member) along the axis of the applied load, it is in
other words a stress state that causes a squeezing of the material. A simple case of compression is
the uniaxial compression induced by the action of opposite, pushing forces. Compressive strength
for materials is generally higher than their tensile strength. However, structures loaded in
compression are subject to additional failure modes, such as buckling, that are dependent on the
member's geometry.

Tensile stress is the stress state caused by an applied load that tends to elongate the material
along the axis of the applied load, in other words the stress caused by pulling the material. The
strength of structures of equal cross sectional area loaded in tension is independent of shape of the
cross section. Materials loaded in tension are susceptible to stress concentrations such as material
defects or abrupt changes in geometry. However, materials exhibiting ductile behavior (most
metals for example) can tolerate some defects while brittle materials (such as ceramics) can fail
well below their ultimate material strength.

Shear stress is the stress state caused by the combined energy of a pair of opposing forces acting
along parallel lines of action through the material, in other words the stress caused by faces of the
material sliding relative to one another. An example is cutting paper with scissors[4] or stresses due
to torsional loading.

Strength terms

Mechanical properties of materials include the yield strength, tensile strength, fatigue strength, crack
resistance, and other characteristics.[5]

Yield strength is the lowest stress that produces a permanent deformation in a material. In some
materials, like aluminium alloys, the point of yielding is difficult to identify, thus it is usually defined
as the stress required to cause 0.2% plastic strain. This is called a 0.2% proof stress. [6]

Compressive strength is a limit state of compressive stress that leads to failure in a material in the
manner of ductile failure (infinite theoretical yield) or brittle failure (rupture as the result of crack
propagation, or sliding along a weak plane - see shear strength).

Tensile strength or ultimate tensile strength is a limit state of tensile stress that leads to tensile
failure in the manner of ductile failure (yield as the first stage of that failure, some hardening in the
second stage and breakage after a possible "neck" formation) or brittle failure (sudden breaking in
two or more pieces at a low stress state). Tensile strength can be quoted as either true stress or
engineering stress, but engineering stress is the most commonly used.

Fatigue strength is a measure of the strength of a material or a component under cyclic loading, [7]
and is usually more difficult to assess than the static strength measures. Fatigue strength is quoted
as stress amplitude or stress range (), usually at zero mean stress, along with the number of cycles
to failure under that condition of stress.
3

Impact strength, is the capability of the material to withstand a suddenly applied load and is
expressed in terms of energy. Often measured with the Izod impact strength test or Charpy impact
test, both of which measure the impact energy required to fracture a sample. Volume, modulus of
elasticity, distribution of forces, and yield strength affect the impact strength of a material. In order
for a material or object to have a high impact strength the stresses must be distributed evenly
throughout the object. It also must have a large volume with a low modulus of elasticity and a high
material yield strength.[8]

Strain (deformation) terms

Deformation of the material is the change in geometry created when stress is applied (as a result of
applied forces, gravitational fields, accelerations, thermal expansion, etc.). Deformation is
expressed by the displacement field of the material.[9]

Strain or reduced deformation is a mathematical term that expresses the trend of the deformation
change among the material field. Strain is the deformation per unit length. [10] In the case of uniaxial
loading the displacements of a specimen (for example a bar element) lead to a calculation of strain
expressed as the quotient of the displacement and the original length of the specimen. For 3D
displacement fields it is expressed as derivatives of displacement functions in terms of a second
order tensor (with 6 independent elements).

Deflection is a term to describe the magnitude to which a structural element is displaced when
subject to an applied load.[11]

Stressstrain relations

Basic static response of a specimen under tension

Elasticity is the ability of a material to return to its previous shape after stress is released. In many
materials, the relation between applied stress is directly proportional to the resulting strain (up to a
certain limit), and a graph representing those two quantities is a straight line.

The slope of this line is known as Young's modulus, or the "modulus of elasticity." The modulus of elasticity
can be used to determine the stressstrain relationship in the linear-elastic portion of the stressstrain
curve. The linear-elastic region is either below the yield point, or if a yield point is not easily identified on
the stressstrain plot it is defined to be between 0 and 0.2% strain, and is defined as the region of strain in
which no yielding (permanent deformation) occurs. [12]

Plasticity or plastic deformation is the opposite of elastic deformation and is defined as


unrecoverable strain. Plastic deformation is retained after the release of the applied stress. Most
materials in the linear-elastic category are usually capable of plastic deformation. Brittle materials,
like ceramics, do not experience any plastic deformation and will fracture under relatively low
strain, while ductile materials such as metallics, lead, or polymers will plasticly deform much more
before a fracture initiation.

Consider the difference between a carrot and chewed bubble gum. The carrot will stretch very little before
breaking. The chewed bubble gum, on the other hand, will plastically deform enormously before finally
breaking.
4

Design terms

Ultimate strength is an attribute related to a material, rather than just a specific specimen made of the
material, and as such it is quoted as the force per unit of cross section area (N/m 2). The ultimate strength
is the maximum stress that a material can withstand before it breaks or weakens. [13] For example, the
ultimate tensile strength (UTS) of AISI 1018 Steel is 440 MN/m2. In general, the SI unit of stress is the
pascal, where 1 Pa = 1 N/m2. In Imperial units, the unit of stress is given as lbf/in or pounds-force per
square inch. This unit is often abbreviated as psi. One thousand psi is abbreviated ksi.

A Factor of safety is a design criteria that an engineered component or structure must achieve. , where FS:
the factor of safety, R: The applied stress, and UTS: ultimate stress (psi or N/m 2) [14]

Margin of Safety is also sometimes used to as design criteria. It is defined MS = Failure Load/(Factor of
Safety * Predicted Load) - 1

For example, to achieve a factor of safety of 4, the allowable stress in an AISI 1018 steel component can
be calculated to be = 440/4 = 110 MPa, or = 110106 N/m2. Such allowable stresses are also known as
"design stresses" or "working stresses."

Design stresses that have been determined from the ultimate or yield point values of the materials give
safe and reliable results only for the case of static loading. Many machine parts fail when subjected to a
non steady and continuously varying loads even though the developed stresses are below the yield point.
Such failures are called fatigue failure. The failure is by a fracture that appears to be brittle with little or no
visible evidence of yielding. However, when the stress is kept below "fatigue stress" or "endurance limit
stress", the part will endure indefinitely. A purely reversing or cyclic stress is one that alternates between
equal positive and negative peak stresses during each cycle of operation. In a purely cyclic stress, the
average stress is zero. When a part is subjected to a cyclic stress, also known as stress range (Sr), it has
been observed that the failure of the part occurs after a number of stress reversals (N) even if the
magnitude of the stress range is below the materials yield strength. Generally, higher the range stress,
the fewer the number of reversals needed for failure.

Failure theories

There are four failure theories: maximum shear stress theory, maximum normal stress theory, maximum
strain energy theory, and maximum distortion energy theory. Out of these four theories of failure, the
maximum normal stress theory is only applicable for brittle materials, and the remaining three theories are
applicable for ductile materials. Of the latter three, the distortion energy theory provides most accurate
results in majority of the stress conditions. The strain energy theory needs the value of Poissons ratio of
the part material, which is often not readily available. The maximum shear stress theory is conservative.
For simple unidirectional normal stresses all theories are equivalent, which means all theories will give the
same result.

Maximum Shear stress Theory- This theory postulates that failure will occur if the magnitude of the
maximum shear stress in the part exceeds the shear strength of the material determined from
uniaxial testing.

Maximum normal stress theory - This theory postulates that failure will occur if the maximum
normal stress in the part exceeds the ultimate tensile stress of the material as determined from
uniaxial testing. This theory deals with brittle materials only. The maximum tensile stress should be
less than or equal to ultimate tensile stress divided by factor of safety. The magnitude of the
maximum compressive stress should be less than ultimate compressive stress divided by factor of
safety.

Maximum strain energy theory - This theory postulates that failure will occur when the strain
energy per unit volume due to the applied stresses in a part equals the strain energy per unit
volume at the yield point in uniaxial testing.

Maximum distortion energy theory - This theory is also known as shear energy theory or von Mises-
Hencky theory. This theory postulates that failure will occur when the distortion energy per unit
volume due to the applied stresses in a part equals the distortion energy per unit volume at the
yield point in uniaxial testing. The total elastic energy due to strain can be divided into two parts:
5

one part causes change in volume, and the other part causes change in shape. Distortion energy is
the amount of energy that is needed to change the shape.

Fracture mechanics was established by Alan Arnold Griffith and George Rankine Irwin. This
important theory is also known as numeric conversion of toughness of material in the case of crack
existence.

Fractology was proposed by Takeo Yokobori because each fracture laws including creep rupture
criterion must be combined nonlinearly.

A material's strength is dependent on its microstructure. The engineering processes to which a material is
subjected can alter this microstructure. The variety of strengthening mechanisms that alter the strength of
a material includes work hardening, solid solution strengthening, precipitation hardening and grain
boundary strengthening and can be quantitatively and qualitatively explained. Strengthening mechanisms
are accompanied by the caveat that some other mechanical properties of the material may degenerate in
an attempt to make the material stronger. For example, in grain boundary strengthening, although yield
strength is maximized with decreasing grain size, ultimately, very small grain sizes make the material
brittle. In general, the yield strength of a material is an adequate indicator of the material's mechanical
strength. Considered in tandem with the fact that the yield strength is the parameter that predicts plastic
deformation in the material, one can make informed decisions on how to increase the strength of a
material depending its microstructural properties and the desired end effect. Strength is expressed in
terms of the limiting values of the compressive stress, tensile stress, and shear stresses that would cause
failure. The effects of dynamic loading are probably the most important practical consideration of the
strength of materials, especially the problem of fatigue. Repeated loading often initiates brittle cracks,
which grow until failure occurs. The cracks always start at stress concentrations, especially changes in
cross-section of the product, near holes and corners at nominal stress levels far lower than those quoted
for the strength of the material.

Creep (deformation)

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.

Temperature dependence

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 35% of the melting point (as measured on a
thermodynamic temperature scale such as Kelvin or Rankine) for metals, and at 45% of melting point for
ceramics.[2] 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.
6

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 or internal voiding decreases the effective area of the specimen. Fracture always occurs at the
tertiary stage.

Mechanisms of creep

The mechanism of creep depends on temperature and stress. Various mechanisms are:

Bulk diffusion (Nabarro-Herring creep)

Climb here the strain is actually accomplished by climb

Climb-assisted glide here the climb is an enabling mechanism, allowing dislocations to get
around obstacles

Grain boundary diffusion (Coble creep)

Thermally activated glide e.g., via cross-slip

General creep equation


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, 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 = 46, 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," th, below which creep can't be measured. The modified power law
equation then becomes:

where A, Q and n can all be explained by conventional mechanisms (so 3 n 10).

Nabarro-Herring creep
7

A diagram showing the diffusion of atoms and vacancies under Nabarro-Herring Creep.

Nabarro-Herring creep is a form of diffusion creep. In Nabarro-Herring creep, atoms diffuse through the
lattice causing grains to elongate along the stress axis; k is related to the diffusion coefficient of atoms
through the lattice, Q = Q(self diffusion), m = 1, and b = 2. Therefore, Nabarro-Herring creep has a weak
stress dependence and a moderate grain size dependence, with the creep rate decreasing as grain size is
increased.

Nabarro-Herring creep is strongly temperature dependent. For lattice diffusion of atoms to occur in a
material, neighboring lattice sites or interstitial sites in the crystal structure must be free. A given atom
must also overcome the energy barrier to move from its current site (it lies in an energetically favorable
potential well) to the nearby vacant site (another potential well). The general form of the diffusion
equation is D = D0exp(E/KT) where D0 has a dependence on both the attempted jump frequency and the
number of nearest neighbor sites and the probability of the sites being vacant. Thus there is a double
dependence upon temperature. At higher temperatures the diffusivity increases due to the direct
temperature dependence of the equation, the increase in vacancies through Schottky defect formation,
and an increase in the average energy of atoms in the material. Nabarro-Herring creep dominates at very
high temperatures relative to a material's melting temperature.

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.

Harper-Dorn creep

Harper-Dorn creep is a climb-controlled dislocation mechanism at low stresses that has been observed in
aluminum, lead, and tin systems, in addition to nonmetal systems such as ceramics and ice. It is
characterized by two principal phenomena: a linear relationship between the steady-state strain rate and
applied stress at a constant temperature, and an independent relationship between the steady-state strain
rate and grain size for a provided temperature and applied stress. The latter observation implies that
Harper-Dorn creep is controlled by dislocation movement; namely, since creep can occur by vacancy
diffusion (Nabarro-Herring creep, Coble creep), grain boundary sliding, and/or dislocation movement, and
since the first two mechanisms are grain-size dependent, Harper-Dorn creep must therefore be dislocation-
motion dependent.[3]

However, Harper-Dorn creep is typically overwhelmed by other creep mechanisms in most situations, and
is therefore not observed in most systems. The phenomenological equation which describes Harper-Dorn
creep is:
8

where: is dislocation density (constant for Harper-Dorn creep), is the diffusivity through the volume of
the material, is the shear modulus, is the Burger's vector, is the applied stress, is Boltzmann's constant,
and is temperature.

The volumetric activation energy indicates that the rate of Harper-Dorn creep is controlled by vacancy
diffusion to and from dislocations, resulting in climb-controlled dislocation motion. [4][5] Unlike in other creep
mechanisms, the dislocation density here is constant and independent of the applied stress. [3] Moreover,
the dislocation density must be low for Harper-Dorn creep to dominate. The density has been proposed to
increase as dislocations move via cross-slip from one slip-plane to another, thereby increasing the
dislocation length per unit volume. Cross-slip can also result in jogs along the length of the dislocation,
which, if large enough, can act as single-ended dislocation sources. [6]

Examples

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:

= applied stress

C0 = instantaneous creep compliance

C = creep compliance coefficient

= retardation time

= distribution of retardation times

When subjected to a step constant stress, viscoelastic materials experience a time-dependent increase in
strain. This phenomenon is known as viscoelastic creep.

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.
9

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. [7] 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. [8]

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. [9]

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 more
boxes were placed on top of it. Due to the weight, the portions of the empty box not upheld by the lower
support gradually deflected 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. [10]

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.

The design of tungsten light bulb filaments attempts to reduce creep deformation. Sagging of the filament
coil between its supports increases with time due to 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
10

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.

Creep can cause gradual cut-through of wire insulation, especially when stress is concentrated by pressing
insulated wire against a sharp edge or corner. Special creep-resistant insulations such as Kynar
(polyvinylidene fluoride) are used in wirewrap applications to resist cut-through due to the sharp corners of
wire wrap terminals. Teflon insulation is resistant to elevated temperatures and has other desirable
properties, but is notoriously vulnerable to cold-flow cut-through failures caused by 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-design 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 conditions, 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.[11][12]

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. [13]

Nimonic 75 has been certified by the European Union as a standard creep reference material.[14]

Preventing creep
There are four general ways to prevent creep in metal. One way is to use higher melting temperature
metal. Second way is to use materials with greater grain size. Third way is use alloying. Fourth way use
intelligent design to reduce the possible factors of creep.
Deformation (engineering)

This article is about deformation in engineering. For a more rigorous treatment, see Deformation
(mechanics).

Compressive stress results in deformation which shortens the object but also expands it outwards.
In materials science, deformation refers to any changes in the shape or size of an object due to-

an applied force (the deformation energy in this case is transferred through work) or

a change in temperature (the deformation energy in this case is transferred through heat).
The first case can be a result of tensile (pulling) forces, compressive (pushing) forces, shear, bending or
torsion (twisting).

In the second case, the most significant factor, which is determined by the temperature, is the mobility of
the structural defects such as grain boundaries, point vacancies, line and screw dislocations, stacking
11

faults and twins in both crystalline and non-crystalline solids. The movement or displacement of such
mobile defects is thermally activated, and thus limited by the rate of atomic diffusion. [1][2]

Deformation is often described as strain.

As deformation occurs, internal inter-molecular forces arise that oppose the applied force. If the applied
force is not too great these forces may be sufficient to completely resist the applied force and allow the
object to assume a new equilibrium state and to return to its original state when the load is removed. A
larger applied force may lead to a permanent deformation of the object or even to its structural failure.

In the figure it can be seen that the compressive loading (indicated by the arrow) has caused deformation
in the cylinder so that the original shape (dashed lines) has changed (deformed) into one with bulging
sides. The sides bulge because the material, although strong enough to not crack or otherwise fail, is not
strong enough to support the load without change, thus the material is forced out laterally. Internal forces
(in this case at right angles to the deformation) resist the applied load.

The concept of a rigid body can be applied if the deformation is negligible.

Types of deformation

Depending on the type of material, size and geometry of the object, and the forces applied, various types
of deformation may result. The image to the right shows the engineering stress vs. strain diagram for a
typical ductile material such as steel. Different deformation modes may occur under different conditions,
as can be depicted using a deformation mechanism map.

Typical stress vs. strain diagram indicating the various stages of deformation.
Elastic deformation

For more details on this topic, see Elasticity (physics).


This type of deformation is reversible. Once the forces are no longer applied, the object returns to its
original shape. Elastomers and shape memory metals such as Nitinol exhibit large elastic deformation
ranges, as does rubber. However elasticity is nonlinear in these materials. Normal metals, ceramics and
most crystals show linear elasticity and a smaller elastic range.

Linear elastic deformation is governed by Hooke's law, which states:

Where is the applied stress, E is a material constant called Young's modulus or elastic modulus, and is
the resulting strain. This relationship only applies in the elastic range and indicates that the slope of the
stress vs. strain curve can be used to find Young's modulus (E). Engineers often use this calculation in
tensile tests. The elastic range ends when the material reaches its yield strength. At this point plastic
deformation begins.
12

Note that not all elastic materials undergo linear elastic deformation; some, such as concrete, gray cast
iron, and many polymers, respond in a nonlinear fashion. For these materials Hooke's law is inapplicable. [3]

Plastic deformation

See also: Plasticity (physics)


This type of deformation is irreversible. However, an object in the plastic deformation range will first have
undergone elastic deformation, which is reversible, so the object will return part way to its original shape.
Soft thermoplastics have a rather large plastic deformation range as do ductile metals such as copper,
silver, and gold. Steel does, too, but not cast iron. Hard thermosetting plastics, rubber, crystals, and
ceramics have minimal plastic deformation ranges. One material with a large plastic deformation range is
wet chewing gum, which can be stretched dozens of times its original length.

Under tensile stress, plastic deformation is characterized by a strain hardening region and a necking
region and finally, fracture (also called rupture). During strain hardening the material becomes stronger
through the movement of atomic dislocations. The necking phase is indicated by a reduction in cross-
sectional area of the specimen. Necking begins after the ultimate strength is reached. During necking, the
material can no longer withstand the maximum stress and the strain in the specimen rapidly increases.
Plastic deformation ends with the fracture of the material.

Metal fatigue
Another deformation mechanism is metal fatigue, which occurs primarily in ductile metals. It was originally
thought that a material deformed only within the elastic range returned completely to its original state
once the forces were removed. However, faults are introduced at the molecular level with each
deformation. After many deformations, cracks will begin to appear, followed soon after by a fracture, with
no apparent plastic deformation in between. Depending on the material, shape, and how close to the
elastic limit it is deformed, failure may require thousands, millions, billions, or trillions of deformations.

Metal fatigue has been a major cause of aircraft failure, especially before the process was well understood
(see, for example, the De Havilland Comet accidents). There are two ways to determine when a part is in
danger of metal fatigue; either predict when failure will occur due to the material/force/shape/iteration
combination, and replace the vulnerable materials before this occurs, or perform inspections to detect the
microscopic cracks and perform replacement once they occur. Selection of materials not likely to suffer
from metal fatigue during the life of the product is the best solution, but not always possible. Avoiding
shapes with sharp corners limits metal fatigue by reducing stress concentrations, but does not eliminate it.

Diagram of a stressstrain curve, showing the relationship between stress (force applied) and strain
(deformation) of a ductile metal.

Compressive failure
Usually, compressive stress applied to bars, columns, etc. leads to shortening.

Loading a structural element or specimen will increase the compressive stress until it reaches its
compressive strength. According to the properties of the material, failure modes are yielding for materials
with ductile behavior (most metals, some soils and plastics) or rupturing for brittle behavior (geomaterials,
cast iron, glass, etc.).
13

In long, slender structural elements such as columns or truss bars an increase of compressive force F
leads to structural failure due to buckling at lower stress than the compressive strength.

Fracture

See also: Concrete fracture analysis and Fracture mechanics


This type of deformation is also irreversible. A break occurs after the material has reached the end of the
elastic, and then plastic, deformation ranges. At this point forces accumulate until they are sufficient to
cause a fracture. All materials will eventually fracture, if sufficient forces are applied.

Misconceptions

A popular misconception is that all materials that bend are "weak" and those that don't are "strong." In
reality, many materials that undergo large elastic and plastic deformations, such as steel, are able to
absorb stresses that would cause brittle materials, such as glass, with minimal plastic deformation ranges,
to break.[4]

Berikut adalah beberapa sifat mekanik yang penting untuk diketahui :

Kekuatan (strength), menyatakan kemampuan bahan untuk menerima tegangan tanpa


menyebabkan bahan menjadi patah. Kekuatan ini ada beberapa macam, tergantung pada jenis
beban yang bekerja atau mengenainya. Contoh kekuatan tarik, kekuatan geser, kekuatan tekan,
kekuatan torsi, dan kekuatan lengkung.

Kekerasan (hardness), dapat didefenisikan sebagai kemampuan suatu bahan untuk tahan terhadap
penggoresan, pengikisan (abrasi), identasi atau penetrasi. Sifat ini berkaitan dengan sifat tahan
aus (wear resistance). Kekerasan juga mempunya korelasi dengan kekuatan.

Kekenyalan (elasticity), menyatakan kemampuan bahan untuk menerima tegangan tanpa


mengakibatkan terjadinya perubahan bentuk yang permanen setelah tegangan dihilangkan. Bila
suatu benda mengalami tegangan maka akan terjadi perubahan bentuk. Apabila tegangan yang
bekerja besarnya tidak melewati batas tertentu maka perubahan bentuk yang terjadi hanya
bersifat sementara, perubahan bentuk tersebut akan hilang bersama dengan hilangnya tegangan
yang diberikan. Akan tetapi apabila tegangan yang bekerja telah melewati batas kemampuannya,
maka sebagian dari perubahan bentuk tersebut akan tetap ada walaupun tegangan yang diberikan
telah dihilangkan. Kekenyalan juga menyatakan seberapa banyak perubahan bentuk elastis yang
dapat terjadi sebelum perubahan bentuk yang permanen mulai terjadi, atau dapat dikatakan
dengan kata lain adalah kekenyalan menyatakan kemampuan bahan untuk kembali ke bentuk dan
ukuran semula setelah menerima bebang yang menimbulkan deformasi.

Kekakuan (stiffness), menyatakan kemampuan bahan untuk menerima tegangan/beban tanpa


mengakibatkan terjadinya perubahan bentuk (deformasi) atau defleksi. Dalam beberapa hal
kekakuan ini lebih penting daripada kekuatan.

Plastisitas (plasticity) menyatakan kemampuan bahan untuk mengalami sejumlah deformasi plastik
(permanen) tanpa mengakibatkan terjadinya kerusakan. Sifat ini sangat diperlukan bagi bahan
yang akan diproses dengan berbagai macam pembentukan seperti forging, rolling, extruding dan
lain sebagainya. Sifat ini juga sering disebut sebagai keuletan (ductility). Bahan yang mampu
mengalami deformasi plastik cukup besar dikatakan sebagai bahan yang memiliki keuletan tinggi,
bahan yang ulet (ductile). Sebaliknya bahan yang tidak menunjukkan terjadinya deformasi plastik
dikatakan sebagai bahan yang mempunyai keuletan rendah atau getas (brittle).

Ketangguhan (toughness), menyatakan kemampuan bahan untuk menyerap sejumlah energi tanpa
mengakibatkan terjadinya kerusakan. Juga dapat dikatakan sebagai ukuran banyaknya energi yang
diperlukan untuk mematahkan suatu benda kerja, pada suatu kondisi tertentu. Sifat ini dipengaruhi
oleh banyak faktor, sehingga sifat ini sulit diukur.
14

Kelelahan (fatigue), merupakan kecendrungan dari logam untuk patah bila menerima tegangan
berulang ulang (cyclic stress) yang besarnya masih jauh dibawah batas kekuatan elastiknya.
Sebagian besar dari kerusakan yang terjadi pada komponen mesin disebabkan oleh kelelahan ini.
Karenanya kelelahan merupakan sifat yang sangat penting, tetapi sifat ini juga sulit diukur karena
sangat banyak faktor yang mempengaruhinya.

Creep, atau bahasa lainnya merambat atau merangkak, merupakan kecenderungan suatu logam
untuk mengalami deformasi plastik yang besarnya berubah sesuai dengan fungsi waktu, pada saat
bahan atau komponen tersebut tadi menerima beban yang besarnya relatif tetap.

Beberapa sifat mekanik diatas juga dapat dibedakan menurut cara pembebanannya, yaitu

Sifat mekanik statis, yaitu sifat mekanik bahan terhadap beban statis yang besarnya tetap atau
bebannya mengalami perubahan yang lambat.

Sifat mekanik dinamis, yaitu sifat mekanik bahan terhadap beban dinamis yang besar berubah
ubah, atau dapat juga dikatakan mengejut.

Ini perlu dibedakan karena tingkah laku bahan mungkin berbeda terhadap cara pembebanan yang
berbeda.

Das könnte Ihnen auch gefallen