Вы находитесь на странице: 1из 67

Creep (deformation)

This article needs additional citations for


verification. Learn more

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
persistent 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 are 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. It works on the
principle of Hook's law ( Stress is directly
proportional to strain)
Temperature dependence
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).[1] As a general guideline, the
effects of creep deformation generally
become noticeable at approximately 35%
of the melting point 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.[3]

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.[4] 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
second 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 decrease 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.[5]

Mechanisms of creep
deformation
The mechanism of creep depends on
temperature and stress. Under the
conditions of different temperature and
applied stress, dislocation glide,
dislocation climb, or diffusional-flow
mechanisms may dominate creep
deformation. Some mechanisms of
creep, especially those involving
dislocations, have not been verified by
direct microstructural examination yet.[4]
However, processes just like the
mechanisms conjectured should happen
during creep deformation.

Various mechanisms are:

Bulk diffusion (Nabarro–Herring creep)


Climb — here the strain is actually
accomplished by a 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, σ is the applied
stress, d is the grain size of the material,
k is Boltzmann's constant, and T is the
absolute temperature.[6]

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 the intrinsic activation
energy, but no grain size dependence.

Some alloys exhibit a very large stress


exponent (m > 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:

where A, Q and n can all be explained by


conventional mechanisms (so 3 ≤ m ≤
10). The creep increases with increasing
applied stress, since the applied stress
tends to drive the dislocation past the
barrier, and make the dislocation get into
a lower energy state after bypassing the
obstacle, which means that the
dislocation is inclined to pass the
obstacle. In other words, part of the work
required to overcome the energy barrier
of passing an obstacle is provided by the
applied stress and the remainder by
thermal energy.

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

Nabarro–Herring (NH) creep is a form of


diffusion creep, while dislocation glide
creep does not involve atomic diffusion.
Nabarro–Herring creep dominates at
high temperatures and low stresses. As
shown in the figure on the right, the
lateral sides of the crystal are subjected
to tensile stress and the horizontal sides
to compressive stress. The atomic
volume is altered by applied stress: it
increases in regions under tension and
decreases in regions under compression.
So the activation energy for vacancy
formation is changed by ± , where
is the atomic volume, the " " sign is for
compressive regions and " " sign is for
tensile regions. Since the fractional
vacancy concentration is proportional to
, where is the

vacancy-formation energy, the vacancy


concentration is higher in tensile regions
than in compressive regions, leading to a
net flow of vacancies from the regions
under tension to the regions under
compression, and this is equivalent to a
net atom diffusion in the opposite
direction, which causes the creep
deformation: the grain elongates in the
tensile stress axis and contracts in the
compressive stress axis.

In Nabarro–Herring creep, 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 the 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

Coble creep is the 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 to creep to have
a stronger grain size dependence than
Nabarro–Herring creep, thus, Coble
creep will be more important in materials
composed of very fine grains. 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.
Generally, the diffusional creep rate
should be the sum of Nabarro–Herring
creep rate and Coble creep rate.
Diffusional creep leads to grain-boundary
separation, that is, voids or cracks form
between the grains. To heal this, grain-
boundary sliding occurs. The diffusional
creep rate and the grain boundary sliding
rate must be balanced if there are no
voids or cracks remaining. When grain-
boundary sliding can not accommodate
the incompatibility, grain-boundary voids
are generated, which is related to the
initiation of creep fracture.

Solute drag creep


Solute drag creep is one kind of
mechanism for power law creep (PLC),
involving both dislocation and diffusional
flow. Solute drag creep is observed in
certain metallic alloys. Their creep rate
increases during the first stage of creep
before a steady-state, which can be
explained by a model associated with
solid-solution strengthening. The size
misfit between solute atoms and edge
dislocations could restrict dislocation
motion. The flow stress required for
dislocations to move is increased at low
temperatures due to immobility of the
solute atoms. But solute atoms are
mobile at higher temperatures, so the
solute atoms could move along with
edge dislocations as a "drag" on their
motion, if the dislocation motion or the
creep rate is not too high. The solute
drag creep rate is:

where C is a constant, is the solute


diffusivity, is the solute concentration,
and is the misfit parameter, is the
applied stress. So it could be seen from
the equation above, m is 3 for solute drag
creep. Solute drag creep shows a special
phenomenon, which is called the
Portevin-Le Chatelier effect. When the
applied stress becomes sufficiently large,
the dislocations will break away from the
solute atoms since dislocation velocity
increases with the stress. After
breakaway, the stress decreases and the
dislocation velocity also decreases,
which allows the solute atoms to
approach and reach the previously
departed dislocations again, leading to a
stress increase. The process repeats
itself when the next local stress
maximum is obtained. So repetitive local
stress maxima and minima could be
detected during solute drag creep.

Dislocation climb-glide creep


Dislocation climb-glide creep is observed
in materials at high temperature. The
initial creep rate is larger than the steady-
state creep rate. Climb-glide creep could
be illustrated as follows: when the
applied stress is not enough to for a
moving dislocation to overcome the
obstacle on its way via dislocation glide
alone, the dislocation could climb to a
parallel slip plane by diffusional
processes, and the dislocation can glide
on the new plane. This process repeats
itself each time when the dislocation
encounters an obstacle. The creep rate
could be written as:
where ACG includes details of the
dislocation loop geometry, DL is the
lattice diffusivity, M is the number of
dislocation sources per unit volume, is
the applied stress, and is the atomic
volume. The exponent m for dislocation
climb-glide creep is 4.5 if M is
independent of stress and this value of m
is consistent with results from
considerable experimental studies.

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.[7]

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:
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.[8][9]
Unlike in other creep mechanisms, the
dislocation density here is constant and
independent of the applied stress.[7]
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.[10]

Sintering

At high temperatures, it is energetically


favorable for voids to shrink in a material.
The application of tensile stress opposes
the reduction in energy gained by void
shrinkage. Thus, a certain magnitude of
applied tensile stress is required to offset
these shrinkage effects and cause void
growth and creep fracture in materials at
high temperature. This stress occurs at
the sintering limit of the system.[11]

The stress tending to shrink voids that


must be overcome is related to the
surface energy and surface area-volume
ratio of the voids. For a general void with
surface energy γ and principle radii of
curvature of r1 and r2, the sintering limit
stress is:[12]
Below this critical stress, voids will tend
to shrink rather than grow. Additional
void shrinkage effects will also result
from the application of a compressive
stress. For typical descriptions of creep,
it is assumed that the applied tensile
stress exceeds the sintering limit.

Creep also explains one of several


contributions to densification during
metal powder sintering by hot pressing.
A main aspect of densification is the
shape change of the powder particles.
Since this change involves permanent
deformation of crystalline solids, it can
be considered a plastic deformation
process and thus sintering can be
described as a high temperature creep
process.[13] The applied compressive
stress during pressing accelerates void
shrinkage rates and allows a relation
between the steady-state creep power
law and densification rate of the
material. This phenomenon is observed
to be one of the main densification
mechanisms in the final stages of
sintering, during which the densification
rate (assuming gas-free pores) can be
explained by:[14][15]
In which is the densification rate, is
the density, is the pressure applied, n
describes the exponent of strain rate
behavior, and A is a mechanism-
dependent constant. A and n are from
the following form of the general steady-
state creep equation:

Where is the strain rate, and is the


tensile stress. For the purposes of this
mechanism, the constant A comes from
the following expression, where A' is a
dimensionless, experimental constant, μ
is the shear modulus, b is the Burgers'
Vector, k is Boltzmann's Constant, T is
absolute temperature, is the
diffusion coefficient, and Q is the
diffusion activation energy:[14]

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:

σ = 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.

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.[16]
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.[17]

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.[18]
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

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.[19]

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 in epoxy anchor adhesive was


blamed for the Big Dig tunnel ceiling
collapse in Boston, Massachusetts that
occurred in July 2006.[20]

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 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.[21][22]

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, including the Maxwell,
Kelvin–Voigt, standard linear solid and
Jeffrey models.[23]

Nimonic 75 has been certified by the


European Union as a standard creep
reference material.[24]
The practice of tinning stranded wires to
facilitate the process of connecting the
wire to a screw terminal, though having
been prevalent and considered standard
practice for quite a while, has been
discouraged by professional
electricians,[25] owing to the fact that the
solder is likely to creep under the
pressure excerted on the tinned wire end
by the screw of the terminal, causing the
joint to lose tension and hence create a
loose contact over time. The accepted
practice when connecting stranded wire
to a screw terminal is to use a wire
ferrule on the end of the wire.

Preventing creep
Creep resistance can be influenced by
many factors such as diffusivity,
precipitate and grain size.

Take metal as an example, to improve


creep resistance, it is obvious that
diffusion rate should be reduced. And
due to the fact that diffusion activation
energy is proportional to absolute
melting temperature, for a specific creep
temperature, materials with higher
melting temperature will be preferred.
Diffusivity can also be influenced by
material classes. Body-centered cubic
metals, which are not as close packed as
face-centered cubic metals and has
more frequently vibrating atoms,
consequently has higher diffusion
coefficients. That’s the reason why bcc
materials are less creep resistant in high
temperatures.

What’s more, when the primary creep


mechanism is dislocation, shear
modulus will play an important role. In
this situation, materials with higher shear
modulus, which are harder to deform, will
be more creep resistant. However,
modulus changes of different materials
are much less than diffusivity changes,
indicating that modulus strengthening is
not as efficient as diffusion decrease in
creep resistance.
In addition, microstructures, or in this
case, grain sizes and particles at grain
boundaries, are also correlated with
creep resistance. When diffusion creep
dominates, increasing grain sizes can
significantly reduce creep rate because
the rate is proportional to (diameter)−2 in
NH creep and (diameter)−3 in Coble
creep. The second-phase intergranular
particles, on the other hand, will prevent
the grain boundaries from sliding.[4]

Thus, we can conclude that there are


three general ways to prevent creep in
metal. One way is to use higher melting
point metals. The second way is to use
materials with greater grain size. The
third way is to use alloying.

Creep of superalloys

Materials operating at high temperatures, such as


this nickel superalloy jet engine (RB199) turbine
blade, must be able to withstand the significant
creep present at these temperatures.

Materials operating in high-performance


systems, such as jet engines, often reach
extreme temperatures surpassing
1200 °C, which causes creep to be a
serious issue. Superalloys based on Co,
Ni, and Fe are capable of being
engineered to be highly resistant to
creep, and have thus arisen as an ideal
material in high-temperature
environments. As an example, Ni-base
alloys modeled after the Ni-Al system,
known as γ-γ’ alloys are even resistant to
dislocation creep. The γ is the main fcc
matrix, while the γ’ is the precipitate-
phase of Ni3(Al, Ti), which adds particle
strengthening. Solute elements added,
e.g., Ta, W, Mo, Fe, Cr, and Co, contribute
solid-solution hardening, and are often
reacted with carbon to form carbide
particles that deposit at grain
boundaries, and thus inhibit grain
boundary sliding.[4]

Recently, there are also some


developments in creep of superalloys. In
2017, Weili Ren at el. extended the creep
life of Ni-based single crystal superalloy
significantly by adding a static magnetic
field in solidification process. By using
this method, the composition distribution
becomes more homogenous in several
scales, making the lattice in γ’ phase
more coherent with the lattice in matrix.
As a result, dislocation quantity is
reduced during the creep. What’s more,
the size of γ’ as well as other
precipitation is also decreased, which
can be proved by observation of
decrease in crack numbers and raft
thickness near fractures. Although the
magnetic field will damage the integrity
of single crystal, this method shows a
new routine to control microstructure
and mechanical properties in superalloys
and still benefits the creep lifetime.[26]

Another research done by Te-Kang Tsao


et al. in 2017 is about creep behaviors of
a novel high entropy superalloy(HESA).
The results of creep at 982 °C show that
HESA, which is dominated by tertiary
creep, has comparable creep behaviors
as some NI-based superalloys. The high
creep resistance is primarily due to low
stacking-fault energy in matrix, high anti-
phase boundary energy in precipitates as
well as thermally stable microstructure.
This research is the first to discuss
tensile creep on full scale specimens of
high entropy alloys, which shows the
potential of HESA at high temperature
application.[27]

See also
Biomaterial
Biomechanics
Ductile–brittle transition temperature
in materials science
Deformation mechanism
Downhill creep
Hysteresis
Larson–Miller parameter
Stress relaxation
Viscoelasticity
Viscoplasticity

References
1. "Rheology of Ice" . Archived from the
original on 2007-06-17. Retrieved
2008-10-16.
2. Ashby, Michael (2014). Materials.
Oxford: Elsevier. p. 336. ISBN 978-0-08-
097773-7.
3. "Deformation & Flow | Mechanics" .
Encyclopedia Britannica. Retrieved
2017-03-29.
4. Courtney, Thomas H. (2005).
Mechanical behavior of materials.
Waveland Press. ISBN 9781577664253.
OCLC 894800884 .
5. "Thermal Creep" . NADCA Design.
Retrieved 2017-03-29.
6. "Creep and Stress Rupture" (PDF). NC
State University. 2017-03-29.
7. Mohamed, F. A.; Murty, K. L.; Morris, J.
W. (April 1973). "Harper–dorn creep in al,
pb, and sn". Metallurgical Transactions. 4
(4): 935–940. doi:10.1007/BF02645593 .
8. Kassner, M.E; Pérez-Prado, M.-T
(January 2000). "Five-power-law creep in
single phase metals and alloys". Progress
in Materials Science. 45 (1): 1–102.
doi:10.1016/S0079-6425(99)00006-7 .
9. Paufler, P. (October 1986). J.-P. Poirier.
Creep of crystals. High-temperature
deformation processes in metals,
ceramics and minerals. Cambridge
University Press. Cambridge – London –
New York – New Rochelle – Melbourne –
Sydney 1985. 145 figs., XII + 260 p., price
£ 10.95 (paperback). Crystal Research and
Technology. 21. p. 1338.
doi:10.1002/crat.2170211021 . ISBN 978-
0-521-27851-5.
10. Mohamed, Farghalli A.; Ginter, Timothy
J. (October 1982). "On the nature and
origin of Harper–Dorn creep". Acta
Metallurgica. 30 (10): 1869–1881.
doi:10.1016/0001-6160(82)90027-X .
11. Courtney, Thomas H. (2000).
Mechanical behavior of materials (2nd
ed.). Boston: McGraw Hill. ISBN 978-
0070285941. OCLC 41932585 .
12. Hull, D.; Rimmer, D.E. (2010). "The
growth of grain-boundary voids under
stress". The Philosophical Magazine: A
Journal of Theoretical Experimental and
Applied Physics. 4:42 (42): 673–687.
doi:10.1080/14786435908243264 .
13. Lenel, F. V.; Ansell, G. S. (1966).
Modern Developments in Powder
Metallurgy. Springer, Boston, MA.
pp. 281–296. doi:10.1007/978-1-4684-
7706-1_15 . ISBN 9781468477085.
14. Wilkinson, D.S.; Ashby, M.F. (November
1975). "Pressure sintering by power law
creep" . Acta Metallurgica. 23 (11): 1277–
1285. doi:10.1016/0001-6160(75)90136-
4 . ISSN 0001-6160 .
15. Ratzker, Barak; Sokol, Maxim;
Kalabukhov, Sergey; Frage, Nachum (2016-
06-20). "Creep of Polycrystalline
Magnesium Aluminate Spinel Studied by
an SPS Apparatus" . Materials. 9 (6): 493.
doi:10.3390/ma9060493 . PMC 5456765 .
PMID 28773615 .
16. Rosato, D. V. et al. (2001) Plastics
Design Handbook. Kluwer Academic
Publishers. pp. 63–64. ISBN 0792379802.
17. M. A. Meyers; K. K. Chawla (1999).
Mechanical Behavior of Materials.
Cambridge University Press. p. 573.
ISBN 978-0-521-86675-0.
18. McCrum, N.G.; Buckley, C.P.; Bucknall,
C.B. (2003). Principles of Polymer
Engineering. Oxford Science Publications.
ISBN 978-0-19-856526-0.
19. Zdeněk Bažant and Yong Zhu, "Why
Did the World Trade Center Collapse?—
Simple Analysis" , Journal of Engineering
Mechanics, January 2002
20. "Ceiling Collapse in the Interstate 90
Connector Tunnel" . National
Transportation Safety Board. Washington,
D.C.: NTSB. July 10, 2007. Retrieved
2 December 2016.
21. Lakes, Roderic S. (1999). Viscoelastic
Solids. p. 476. ISBN 978-0-8493-9658-8.
22. "Is glass liquid or solid?" . University of
California, Riverside. Retrieved
2008-10-15.
23. Osama Abuzeid; Anas Al-Rabadi;
Hashem Alkhaldi (2011). "Recent
advancements in fractal geometric-based
nonlinear time series solutions to the
micro-quasistatic thermoviscoelastic
creep for rough surfaces in contact" .
Mathematical Problems in Engineering.
2011: 1–29. doi:10.1155/2011/691270 .
691270.
24. Gould, D.; Loveday, M.S. (1990). The
certification of nimonic 75 alloy as a creep
reference material — CRM 425 (PDF).
Luxembourg: Office for Official
Publications of the European
Communities. ISBN 978-92-826-1742-7.
Archived from the original (PDF) on 2015-
04-03.
25. IPC J-STD-001 Rev. E, Requirements
for Soldered Electrical and. Electronic
Assemblies
26. Ren, Weili; Niu, Chunlin; Ding, Biao;
Zhong, Yunbo; Yu, Jianbo; Ren,
Zhongming; Liu, Wenqing; Ren, Liangpu;
Liaw, Peter K. (2018-01-23). "Improvement
in creep life of a nickel-based single-
crystal superalloy via composition
homogeneity on the multiscales by
magnetic-field-assisted directional
solidification" . Scientific Reports. 8 (1):
1452. doi:10.1038/s41598-018-19800-5 .
ISSN 2045-2322 . PMC 5780385 .
PMID 29362394 .
27. Tsao, Te-Kang; Yeh, An-Chou; Kuo,
Chen-Ming; Kakehi, Koji; Murakami,
Hideyuki; Yeh, Jien-Wei; Jian, Sheng-Rui
(2017-10-04). "The High Temperature
Tensile and Creep Behaviors of High
Entropy Superalloy" . Scientific Reports. 7
(1): 12658. doi:10.1038/s41598-017-
13026-7 . ISSN 2045-2322 .
PMC 5627260 . PMID 28978946 .
Further reading
Ashby, Michael F.; Jones, David R. H.
(1980). Engineering Materials 1: An
Introduction to their Properties and
Applications. Pergamon Press.
ISBN 978-0-08-026138-6.
Frost, Harold J.; Ashby, Michael F.
(1982). Deformation-Mechanism Maps:
The Plasticity and Creep of Metals and
Ceramics. Pergamon Press. ISBN 978-
0-08-029337-0.
Turner, S (2001). Creep of Polymeric
Materials. Oxford: Elsevier Science Ltd.
pp. 1813–1817. ISBN 978-0-08-
043152-9.

External links
Creep Analysis Research Group –
Politecnico di Torino
Deformation-Mechanism Maps, The
Plasticity and Creep of Metals and
Ceramics
The National Institute of Standards
and Technology – WTC Briefing
"Introduction to Creep" . Archived from
the original on 2008-06-17. Retrieved
2008-10-16.

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Creep_(deformation)&oldid=894642548"

Last edited 15 hours ago by an an…


Content is available under CC BY-SA 3.0 unless
otherwise noted.

Вам также может понравиться