Sie sind auf Seite 1von 29

CREEP

Review of plastic deformation and failure


Creep Mechanisms (and Maps)
Creep Resistant Materials
MATERIALS SCIENCE
Creep in Nanomaterials Part of & A Learners Guide
ENGINEERING
Superplasticity AN INTRODUCTORY E-BOOK

Superplascity in Nanomaterials Anandh Subramaniam & Kantesh Balani


Materials Science and Engineering (MSE)
Indian Institute of Technology, Kanpur- 208016
Email: anandh@iitk.ac.in, URL: home.iitk.ac.in/~anandh
http://home.iitk.ac.in/~anandh/E-book.htm

Mechanical Metallurgy
George E Dieter
McGraw-Hill Book Company, London (1988)
Review

If failure is considered as change in desired performance*- which could involve changes in


properties and/or shape; then failure can occur by many mechanisms as below.

Mechanisms / Methods by which a can Material can FAIL

Elastic deformation

Creep Chemical / Physical


Fatigue Electro-chemical degradation
Plastic Fracture degradation
deformation
Microstructural
Twinning changes
Wear
Slip Twinning
Corrosion Erosion
Phase transformations
Oxidation
Grain growth

Particle coarsening

* Beyond a certain limit


Review
Though plasticity by slip is the most important mechanism of plastic deformation, there are
other mechanisms as well (plastic deformation here means permanent deformation in the
absence of external constraints):

Plastic Deformation in Crystalline Materials

Slip Twinning Phase Transformation Creep Mechanisms


(Dislocation
motion) Grain boundary sliding
+ Other Mechanisms
Vacancy diffusion
Grain rotation
Dislocation climb

Note: Plastic deformation in amorphous materials occur by other mechanisms including flow (~viscous fluid) and shear
banding
High-temperature behaviour of materials
Designing materials for high temperature applications is one of the most challenging tasks
for a material scientist.
Various thermodynamic and kinetic factors tend to deteriorate the desirable microstructure.
This is because kinetics of underlying processes (like diffusion) are an exponential function
of temperature.
Hence, a small increase in temperature can prove to be catastrophic.
Strength decreases at high temperature and material damage (e.g. void formation) tends to
accumulate.
Phenomena like creep and accelerated oxidation kick-in.
Cycling between high and low temperature will cause thermal fatigue.
High temperature effects (many of the effects described below are coupled)
Increased vacancy concentration at high temperatures more vacancies are
thermodynamically stabilized (this will further increase the diffusion rate).
Thermal expansion material will expand and in multiphase materials/hybrids thermal
stresses will develop due to differential thermal expansion of the components.
High diffusion rate diffusion controlled processes become important.
Phase transformations can occur this not only can give rise to undesirable
microstructure, but lead to generation of internal stresses.
Precipitates may dissolve.
Grain related:
Grain boundary weakening may lead to grain boundary sliding and wedge cracking.
Grain boundary migration
Recrystallization / grain growth decrease in strength.
Dislocation related these factors will lead to decrease in strength
Climb
New slip systems can become active
Change of slip system
Decrease in dislocation density.
Overaging of precipitates and precipitate coarsening decrease in strength.
The material may creep (time dependent elongation at constant load/stress).
Enhanced oxidation and intergranular penetration of oxygen.
Etc.
Creep Creep is phenomenological term, which is responsible for plastic deformation.

In some sense creep and superplasticity are related phenomena: in creep we can think of
damage accumulation leading to failure of sample; while in superplasticity extended plastic
deformation may be achieved (i.e. damage accumulation leading to failure is delayed).
Creep is permanent deformation (plastic deformation) of a material under constant load (or
constant stress) as a function of time. (Usually at high temperatures lead creeps at RT).

Normally, increased plastic deformation takes place with increasing load (or stress)
In creep plastic strain increases at constant load (or stress)
Usually appreciable only at T > 0.4 Tm High temperature phenomenon.
Mechanisms of creep in crystalline materials is different from that in amorphous materials.
Amorphous materials can creep by flow.
At temperatures where creep is appreciable various other material processes may also
active (e.g. recrystallization, precipitate coarsening, oxidation etc.- as considered before).

Creep experiments are done either at constant load or constant stress and can be classified
based on Phenomenology or underlying Mechanism.

Phenomenology
Constant load (easier)
Creep can Harper-Dorn creep
Power Law creep Creep tests can
be classified
be carried out at
based on Constant stress
Mechanism
Constant load creep curve
In a typical creep test the load and temperature are kept constant and the elongation is
monitored with time. The strain (typically engineering strain) computed from the elongation
is plotted as function of time. The loads employed are typically below the elastic limit.
Three stages may be observed in such a plot: (i) decreasing rate with time, (ii)
approximately constant rate, (iii) increasing rate with time. These stages have to be
understood keeping in view underlying mechanisms (& necking in stage-III).
The instantaneous strain seen (0) is the elastic strain, which develops on the application of the load.

Stages of creep Measured as strain rate (note that this strain


Constant load creep curve rate is not the one imposed as in UTT, but
the one which develops in the material)
Stage-I
I II Creep rate decreases with time.
Effect of work hardening more than recovery.
Strain ()

A technical term
III Stage-II
Stage of minimum creep rate ~ constant.
Work hardening is balanced by recovery.
The distinguishability of the three Stage-III
stages strongly depends on T and Absent (/delayed very much) in constant
stress tests (shown later).
0 Initial instantaneous strain
0 Necking of specimen starts in this stage.
t Specimen failure processes set in.
Constant Stress creep curve
In stage-III (due to necking) the engineering stress is no longer a correct measure of the
state of stress. To keep the stress constant, the instantaneous area has to be taken into
account.
If this is done, then the increasing strain rate part is not observed. Note: if load is kept
constant then in stage-III the stress is actually increasing (for the material it is stress which
matters and not load).

I II
Strain ()

III
t
Effect of stress on the creep curve (constant load)
On increasing the load at which the experiment is conducted: (i) the instantaneous strain
(elastic) increases, (ii) for a given time (say t1) the strain is more, (iii) the time to failure (tf)
decreases (i.e. as expected, specimens fail earlier).

Fracture

Strain ()

Elastic strains

Increasing stress

With increasing load there is


increased initial elastic strain


0 increases

0''
0 0' 0''
0'
0 t 'f' t t 'f t 0f
t1
Effect of temperature
On increasing the temperature at which the experiment is conducted:
(i) the instantaneous strain (elastic = 0) increases (slightly),
(ii) for a given time (say t1) the strain is more, (iii) the time to failure (tf) decreases.
The instantaneous strain 0 increases with increasing T because of the slight decrease in the
Youngs modulus (E) of the material.

Strain ()

E as T Increasing T

As decrease in E
with temperature
0 increases

is usually small
the 0 increase is
also small
0
0 '
0 ''
0
t1 t 'f' t t 'f t 0f
Creep Mechanisms of crystalline materials
Stress and temperature are the two important variables, which not only affect the creep rate,
but also the mechanism operative. Three kinds of mechanisms are operative in creep:
1 dislocation related,
2 diffusional,
3 grain boundary sliding.
These and their sub-classes are shown in the next page.
At high temperatures the grain boundary becomes weaker than the grain interior and two
grains can slide past one another due to shear stress. The temperature at which the grain is
as strong as the grain boundary is called the equicohesive temperature.
A combination of these mechanisms could also be responsible for the creep strain.
Depending on the stress and temperature other mechanisms of plastic deformation or
microstructural changes may occur concurrently with creep. These include plastic
deformation by slip and dynamic recrystallization.
Deformation mechanism maps can be drawn with homologous temperature (T/Tm) and
normalized shear stress (/G) as the axis (other combination of variables may also be chosen
for these plots: T/Tm vs shear strain rate, normalized shear stress vs shear strain rate, etc.).
Typically these maps overlay descriptors, which are based both on phenomenology and
mechanism.
Creep Mechanisms of crystalline materials

Cross-slip

Dislocation related Climb

Glide

Coble creep
Grain boundary diffusion controlled

Creep Diffusional Nabarro-Herring creep


Lattice diffusion controlled

Dislocation core diffusion creep


Diffusion rate through core of edge dislocation more

Interface-reaction controlled diffusional flow


Grain boundary sliding

Accompanying mechanisms: creep with dynamic recrystallization


Dislocation related mechanisms
Two roles can be differentiated with respect to of dislocations activity: (i) it is the primary
source of strain, (ii) it plays a secondary role to accommodate local strain (while the major
source of strain is another mechanism (e.g. grain boundary sliding).

Cross-slip
This kind of creep is observed at relatively low temperatures. Herein screw dislocations
cross-slip by thermal activation and give rise to plastic strain as a function of time.

Dislocation climb
Edge dislocations piled up against an obstacle can climb to another slip plane and cause
plastic deformation. In response to stress this gives rise to strain as a function of time. It is
to be noted that at low temperatures these dislocations (being pinned) are sessile and
become glissile only at high temperatures.
Rate controlling step is the diffusion of vacancies.
Diffusional creep Nabarro-Herring creep high T lattice diffusion

Coble creep low T Due to GB diffusion


In response to the applied stress vacancies preferentially move from
surfaces/interfaces (GB) of specimen transverse to the stress axis to
surfaces/interfaces parallel to the stress axis thus causing
elongation. Flow of vacancies
Diffusion of vacancies in one direction can be thought of as flow of
matter in the opposite direction.
This process like dislocation creep (involving climb) is controlled by
the diffusion of vacancies (but diffusional creep does not require
dislocations to operate).
The diffusion could occur predominantly via the lattice (at high
temperatures) or via grain boundaries (at low temperatures). The
former is known as Nabarro-Herring creep, while the later is known
as Coble creep.
Diffusion through edge dislocation cores (pipe diffusion) could play
an important role in creep.
Grain boundary sliding
At low temperatures the grain boundaries are stronger than the crystal interior and impede
the motion of dislocations.
Being a higher energy region, the grain boundaries may pre-melt before the crystal interior.
Above the equicohesive temperature, due to shear stress at the local scale, grain
boundaries slide past one another to cause plastic deformation.
The relative motion of grain boundaries can lead to wedge cracks at triple lines (junction of
three grains). If these wedge cracks are not healed by diffusion (or slip), microstructural
damage will accumulate and will lead to failure of the specimen.

Grains

Wedge crack due to


grain boundary sliding
Phenomenological descriptions of creep
One of the important descriptions of creep is using the power-law formula. The shear strain
rate is a power function of the shear stress. Clearly this formula is not based on a
mechanism operative, but a fit of data. .
Shear strain rate

n
.
G Shear modulus


G Shear stress
n An exponent having a value between ~ 3 - 10

Power-law behaviour can arise from:


Only glide at low temperatures (~0.3TM). Here the exponent n ~ 3.
Glide + climb (referred to as climb controlled creep) occurs at higher temperatures.
Above ~0.6TM climb is lattice-diffusion controlled.
At lower temperatures than this pipe diffusion may play an important role in creep.
At high stresses (> 103G) the power law breaks down. At high stresses the mechanism
changes from climb controlled (creep) to glide controlled (slip). This is bordering on normal
plastic deformation.
Deformation Mechanism Maps
Time and temperature are coupled when it comes to processes like diffusion.
At large values of stresses and at low T, the time available is less (as material immediately
begins to deform plastically) and creep mechanisms do not have time (/activation) to operate.
Usually contours of constant strain rate are superimposed on these diagrams (not shown here).
Stress or strain rate can be used as axes (variable). In components (e.g. truss in a structure,
pressure vessel, etc.) stress is prescribed, while in processing (e.g. extrusion, forging, etc.),
strain rate is prescribed.

At high stresses plastic


flow will take place

The dominant
mechanism is shown in
the diagram Dynamic recrystallization
gives rise to strain-free
grains.

At high temperature and


low stress Diffusional creep
dominates

From Deformation Mechanism Maps: The plasticity and creep of Metals and Ceramics by H.J. Frost and M.F.Ashby, Pergamon Press, Oxford, 1982.
From Deformation Mechanism Maps: The plasticity and creep of Metals and Ceramics by H.J. Frost and M.F.Ashby, Pergamon Press, Oxford, 1982.
Creep Resistant Materials
The is a growing need for materials to operate at high temperatures (and in some
applications for long times). For example, higher operating temperatures gives better
efficiency for a heat engine. Hence, there is a need to design materials which can withstand
high temperatures.
It is to be noted that material should also be good in other properties for high temperature
applications (like it should possess good oxidation resistance). Factors like cost, ease of
fabrication, density, etc. play an important role in determining the final choice of a material.
Some of the material design strategies, which work at low temperature are not useful at
high temperatures (e.g. work hardening, precipitation hardening with precipitates which
coarsen, grain size reduction, etc.).
Some strategies which work are: (i) having grain boundaries aligned along the primary
loading axis, (ii) produce single crystal components (like turbine blades), (iii) use
precipitates with low interfacial energy for strengthen (which will not coarsen easily), (iv)
use dispersoids for strengthening.

High melting point E.g. Ceramics


Creep Dispersion hardening ThO2 dispersed Ni (~0.9 Tm)
resistance Solid solution strengthening
Single crystal / aligned (oriented) grains
Creep Resistant Materials, cotd..

Commonly used materials Fe, Ni (including superalloys), Co base alloys.


Precipitation hardening involving usual precipitates* is not a good method as precipitates
coarsen (smaller particles dissolve and larger particles grow interparticle separation
thus lowering the strength)
Ni-base superalloys have Ni3(Ti,Al) precipitates, which form a low energy interface with
the matrix. This reduces the driving force for coarsening. (Note: other phenomena like
rafting may lead to the deterioration of the properties of such materials).
Cold work cannot be used for increasing creep resistance, as recrystallization can occur
which will produced strain free crystals.
Fine grain size is not desirable for creep resistance (this is contrary to what is usually practiced for increasing
the low temperature strength) grain boundary sliding can cause creep elongation/cavitation. Hence,
the following two strategies can be used:
Use single crystals (single crystal Ti turbine blades in gas turbine engine have been used though they are very costly).
Aligned/oriented polycrystals as all the grain boundaries are aligned along the
primary tensile axis, they experience no shear stress and creep is negated.

* Which coarsen at high temperatures due to high interfacial energy.


Creep in Nanomaterials
Due to fine grain size nanostructured materials (grain size in the nanoscale regime) are
expected to: (i) show creep at relatively lower temperatures, (ii) display higher creep rates
for a give temperature, (iii) experience predominance of mechanisms like grain boundary
diffusion and grain boundary sliding. We now see what is actually seen in experiments.
In nanocrystalline Pd (~40 nm) and Cu (~20 nm), there seemed to be no increase in creep
rate as compared to micron grain sized materials (in some temperature regimes even a lower
creep rate was observed for Pd). This is in direct contradiction with the expectation that
nanocrystalline materials will experience a higher creep rate.
Studies on Cu (10-25 nm GS), Pd (35-55 nm GS) (TEM showed porocity in sample) [1]
creep in the low T regime (0.24-0.33 Tm) low creep rate, low grain growth
creep in the medium T regime (0.33-0.48 Tm) creep rate decreasing even after long
testing time, grain growth (25 nm 100s of nm)

Cu creep rates of nc sample was


comparable to micron GS sample
Pd nc sample exhibited lower creep rates

[1] P.G. Sanders, M. Rittner, E. Kiedaisch, J.R. Weertman, H.Kung, Y.C. Lu, Nanostruct. Mater. 9 (1997) 433.
[2] D.L. Wang, Q.P. Kong, J.P. Shui, Scr. Metall. Mater. 31 (1994) 47.
In some cases the creep rate increased with a decrease in grain size in the nanoscale
regime of grain sizes (e.g. in Ni-P nanocrystalline material the creep rate of ~30 nm grain
sized material was higher than that of 250 nm material [2]).
In cases where high creep rate expected for nanocrystalline materials (e.g. Pd, Cu) was not
observed, the reason attributed are:
(i) presence of low angle grain boundaries and twin boundaries (which are not prone to
sliding and have low diffusivity for vacancies),
(ii) reduced dislocation activity in nanocrystalline samples.
Creep of nc-Ni at RT (GS: 6, 20, 40 nm) [1]:
Smaller grain size (6nm) showed faster creep rate.
Behaviour consistent with Grain boundary sliding controlled by grain boundary
diffusion mechanism.
At high stresses and larger GS (20, 40 nm), dislocation creep was observed.

[1] N. Wang, Z. Wang, K.T. Aust, U. Erb, Mater. Sci. Eng., A 237 (1997) 150.
Superplasticity
Superplasticity
The phenomenon of extensive plastic deformation without necking is termed as structural
superplasticity. Superplastic deformation in tension can be >300% (up to even 2000%).
Typically superplastic deformation occurs when:
(i) T > 0.5Tm
(ii) grain size is < 10 m
(iii) grains are equiaxed (which usually remain so after deformation)
(iv) grain boundaries are glissile (with a large fraction of high angle grain boundaries).
Presence of a second phase (of similar strength to the matrix- reduces cavitation during
deformation), which can inhibit grain growth at elevated temperatures helps (e.g. Al-
33%Cu, Zn-22% Al)).
Many superplastic alloys have compositions are close to eutectic or eutectoid points.
Superplastic flow is diffusion controlled (can be grain boundary or lattice diffusion
controlled).
A plot of stress versus strain rate is often sigmoidal and shows three regions:
(i) Region-I- low stress, low strain rate regime ( <105 /s) m (0.2,0.33)

Note: low m in region I and III


Sensitive to the purity of the sample. Lower ductility and grain boundary diffusion.
(ii) Region-II- intermediate stress & strain rate regime [ (105, 102)] m (0.4,0.67)
Extended region covering several orders of magnitude in strain rate. Region of maximum
ductility. Strain rate insensitive to grain size and insensitive to purity.
Often referred to as the superplastic region.
Mechanism predominantly grain boundary sliding accommodated by dislocation
activity (Activation energy (Q) corresponding to grain boundary diffusion (Qgb)).
(iii) Region-III- high stress & strain rate regime ( > 102 /s) m > 0.33
Creep rates sensitive to grain size.
Mechanism intragranular dislocation process (interacting with grain boundaries).

C ,T
m
Superplasticity in Nanomaterials
In most cases the superplasticity has not fulfilled the initial expectations.
In many cases superplasticity is only observed in nanocrystalline samples, where it is already
observed in their microcrystalline counterparts.
Superplasticity was observed in nanocrystalline Ni (20 nm grain size) at 0.36Tm (more than
450C lower than that for the bulk material) [1].
Nanocrystalline Ni3Al (grain size 50 nm) also became superplastic about 450C below its
microcrystalline counterparts.
Ni3Al had a ductility of 350% at 650C (strain rate of 103 /s).
1420-Al alloy showed superplasticity at a high strain rate of 101 /s. High amount work
hardening and higher flow stress for superplastic deformation as compared to micron grain
sized material is observed in these cases.
Superplasticity was observed in ~40 nm grain size Zn-Al alloy at 373 K, tested at a strain
rate of 104 /s [2]. Microcrystalline samples showed no superplasticity!

Ni3Al (cP4, Pm-3m)


[1] S. X. McFadden, R. S. Mishra, R. Z. Valiev, A. P. Zhilyaev and A. K. Mukherjee, Nature 398 (1999) 684.
[2] R.S. Mishra, R.Z. Valiev, A.K. Mukherjee, Nanostruct. Mater. 9 (1997) 4732.
Superplasticity at low temperature (or equivalently Superplasticity at high strain rates (>
102 /s) at a given temperature in the superplastic regime) is caused by:
increased diffusion, grain boundary sliding and dislocation activity.
Grain growth is a serious issue during superplasticity experiments. In the case of nc-Ni it
was seen that the grain size could increase to micron sizes, from the starting grain size of
the order of 20 nm. In other materials the grain growth could be less. Grain growth is
expected to be less is two phase mixtures (2nd phase as a precipitate preferred) and
intermetallic compounds. In two phase mixtures the 2nd phase has a pinning effect on the
grain boundaries; while in intermetallics (like Ni3Al) order (with respect to the sublattices)
has to be maintained during grain growth, which restrains the process.
In cases where grain boundary sliding is the predominant mechanism for superplasticity
(e.g. in some Mg alloys), it is seen that non-equilibrium grain boundaries give lower
elongation as compared to equilibrium grain boundaries (due to the long range stress
fields associated with non-equilibrium grain boundaries, which is expected to hamper
grain boundary sliding).
In Ni3Al the high flow stresses and extensive strain hardening during superplastic
deformation has been attributed to depletion of dislocations and high stresses required for
the nucleation of new ones [1].

[1] R.S. Mishra, R.Z. Valiev, S.X. McFadden, A.K. Mukherjee, Mater. Sci. Eng., A 252 (1998) 174.

Das könnte Ihnen auch gefallen