Anda di halaman 1dari 21

MATE 202: Introduction to Materials Engineering

Mechanical Failure
Chapter 3

Canada Dept. Mines and Resources/Library and Archives Canada/PA-020614

The Qubec Bridge collapse of 1907


As it stands, the center section of the Qubec
Bridge remains the longest cantilever bridge span
in the world (549m), and a major success of engineering. However, the history of this bridge is
fraught with failure and loss of human life. The
bridge collapsed not once (as shown at left in
1907 killing 75 of 86 workers) but twice, a second time, even after redesign, in 1916 (this time
killing 13). You may ask: why did the bridge
collapse at all? Was it inadequate design, poor
material, poor construction, poor management,
lack of engineering standards? All are fair questions when we consider failure, and in many cases the root cause is not a single element, but the
sum of many small factors.
A host of factors led to the first bridge
failure of 1907: observations of bent beams were
reported but ignored, calculations of stresses
were incorrect and criticism thereof was not acted upon, poor and inadequate supervision was
present, and emphasis from the company in
charge was on cost over safety. All factors were
responsible.
The trouble in 1907 began when workers noticed sheared rivets from the stresses being
imposed on the structure, and bent beams then
buckled; the bridge fell into the river claiming
many lives. The failures associated with the Qubec Bridge motivated engineers all over the world to
create self-regulatory bodies with better standards and management protocols to protect human safety.
The Canadian engineering iron ring has been associated with the Qubec Bridge and rumours
claim that early engineering rings were made from the steel of the collapsed bridge (rings today are made
of stainless steel or wrought iron). Deeper meaning is taken from the bridge failures and the iron ring: the
ring is worn to offer a constant reminder that the professional engineers first, and most important, duty
is to society and its protection. Although tragic, the failures transformed the engineering profession.

Source: http://www.thecanadianencyclopedia.com/index.cfm?PgNm=ArchivedFeatures&Params=A2136

30

Chapter 3: Mechanical Failure

Sadly, there are no shortages of mechanical failures in history as it is rather difficult to design
against. This chapter explores how materials fail, what can be learned about design through failure, and what to do in order to minimize failure as much as possible to ensure public safety.
Issues to address:
What is failure?
Metals - some important notes about structure that influence failure
What is fracture? What can we learn from fracture to aid design?
Impact what happens to a metal when the strain rate is very high?
Fatigue why do some metals have better resistance to cycling stresses than others?
Creep why will metals have accelerated failure at high temperature?
VIDEO: The front fell off
31

MATE 202: Introduction to Materials Engineering

3.0 Introduction In general, failure is an unwanted event. We need to understand how materials
fail, and how to determine component/design lifetimes in order to avoid failure. Proper design often accounts for the observation of failure before it occurs (e.g., a metal will deform
plastically before it fractures or a pressure vessel may leak before it fractures). Failure analysis covers many factors surrounding the situation, but the component which fails ultimately
does so by a physical failure. However, there are often other latent reasons such as politics.
Physical failures are caused by (in order of importance)
1. improper material selection or processing
2. inadequate design
3. misuse (failure to understand/account for mechanical failure modes (fracture,
fatigue, creep) and environmental failure modes (wear & corrosion!)
Examples of failure from history
Titanic
Cold water (ductile to brittle transition of the steel), poor materials processing
and testing, overconfidence, poor safety procedures, and no backup plan, too
few lifeboats etc
Exxon Valdez
Change in route and poor navigational information resulted in puncture of hull,
and huge amount of oil spilled into the ocean. Hull design was of single hull
construction, even though better double wall hull designs were available at the
time (but would cost more!).
Challenger
O-ring failure due to operating conditions (too cold ductile to brittle transition
of the rubber), political pressure, poor communication.
3.1 Metals: Crystals, grain, and grain boundaries (this is important for understanding failure)
Metals are crystalline and composed of grains, which are crystals of the same type but of different orientations (these grains are elongated during plastic deformation below). The narrow
region (2-3 atoms) between grains in which the atoms dont fit into either grain perfectly has
higher energy and is the grain boundary. Grains are typically about 10 to 50 m in diameter, but
their size can be varied for special applications. A 3D image in Figure 3.0 depicts grains of different size in different colours within a titanium alloy bar. Grains and grain deformation are
shown in Fig. 3.1.
Figure 3.0 Reconstruction of 4380 grains
from 200 serial sections of which 2098 grains
were analyzed. Some of the outside grains
are removed to reveal the internal structure.
The color of the grain indicates the spherical
equivalent radius of the grain, with the grain
size distribution shown in the color bar.
(Source: D.J. Rowenhorst and A.C. Lewis,
Image Processing and Analysis of 3-D Microscopy Data, JOM 63 [3] 2011, p53-57).

32

Chapter 3: Mechanical Failure


Figure 3.1 Alteration of the grain structure of a
polycrystalline metal as a result of tensile plastic
deformation. a) before deformation the grains
are equiaxed. b) The deformation elongates the
grains in the direction of the tensile strain (top to
bottom). The average grain in a) is about 30m.
From Moffatt, Pearsall and Wulff 1964.

WHATS IN THE BOX?


You can produce this same elongated structure
if you create a structure of balls of Silly PuttyTM
and then stretch them with a tensile force. See
what then happens in compression or shear
Figure 3.2 a) Highly ductile fracture in which
the specimen necks down to a point. This typically occurs in pure gold for example. b)
Moderately ductile fracture after some necking, such as occurs in a mild steel c) Brittle
fracture without any plastic deformation. This
strictly only occurs in ceramics like glass, but
steel at low temperature can be fairly brittle.
(Source: W.D. Callister)

WHATS IN THE BOX?


Can you produce the same structures using a
cylinder of Silly PuttyTM ? How do you do it?

3.2 Fracture
3.2.0 Ductile vs. Brittle Fracture (see Figure 3.2)
We can distinguish ductile fracture and brittle fracture.
Failure and fracture are not necessarily synonymous. Fracture is a type of failure.
Ductile fracture involves extensive plastic deformation like in a tensile test, lots of energy
is absorbed, time is required, and usually the load must increase for failure
Atoms are often rearranged permanently through damage processes. Accumulation of
damage results in final failure. Plastic deformation can more or less be considered damage
in this case.
Brittle fracture is fast and involves little deformation beyond the fracture - like a rock shattering a window (broken but not deformed), little energy is absorbed and the load does
not have to increase, fracture surface tends to be flat
Atomic bonds are completely broken over the section. Atoms are minimally
permanently rearranged (if at all).
There are degrees of brittleness, brittle metals (well see this for steel in
section 3.3) still have considerable plastic energy in fracture; whereas, a
ceramic like glass exhibits no plasticity in fracture and the only energy involved is that to create the new surfaces.

33

MATE 202: Introduction to Materials Engineering

A component may fail in a ductile or brittle manner. In some cases there may be features that
indicate brittle and ductile fracture (as discussed later for fatigue). Ductile and brittle are relative
terms, and a more quantitative approach is to measure the amount of sheared metal by area
percentage of the fracture section to gain an idea of how much ductile fracture occurred.
3.2.1 Stress Concentration Around any discontinuity (e.g. flaw (crack, pore) or change in section size) the stresses tend to be magnified (See Fig. 3.3 below). The stress increase is a
maximum close to the discontinuity and decreases as we move away as shown.
Figure 3.3 a) the geometry of surface and internal cracks b) schematic
stress profile along the X
-X plane in a) showing
the stresses in the material close to the tip.
Note for subsurface
flaws 2a is the width and
for surface flaws a is the
width. (Adapted from
W.D. Callister).

The problem of determining the maximum stress at a discontinuity was solved for elastic deformations nearly 100 year ago by J. Inglis who showed that any flaw can be represented by an
equivalent ellipse with 2a being the width of the flaw perpendicular to the tensile stress and t
being the geometric curvature at the tip of the flaw (radius of the circle that just fits the curvature at the tip). One modern easily understood reference for this development which also summarizes the development of fracture mechanics is J.E. Gordon, The New Science of Strong Materials, re-issued by the Princeton Science Library in 2006.
The maximum stress, m ,near the tip of the flaw (under the nominal applied stress, 0) is given
by:

m = o + 2 o

We see that as a (crack length) increases and t (sharpness of crack tip) decreases, the maximum stress increases.
The stress intensity (concentration) factor Kt, is another way to understand the magnification of
stress at flaw tips, and is given by:
Kt =

m
a
=1+ 2
o
t

For a circular hole in metal however (a = t) so the ratio of maximum stress to applied stress is
3, and is an effective stress magnification (the stress at the edge of the hole is three times the
applied stress:
34

Chapter 3: Mechanical Failure

Kt =

m
a
=1+ 2
=1+ 2 1 = 3
o
t

This factor of 3 sometimes arises in engineering codes for metals and this equation gives us a
hint as to where this factor comes from. The idea of stress concentration is an important concept
in understanding material behaviour the stress magnification due to cracks and flaws in materials can cause fracture at an applied stress which should seemingly not cause failure.
Eexamples of material failures in the real world: the Liberty Bell, popping a balloon with a
pin vs. your finger.
For brittle fracture in ceramics, if the critical stress is reached then catastrophic fracture will
occur (enough energy is present to cause the entire part to fracture in a brittle manner that is
spontaneous and rapid):

2 E S
C =

Different materials have different critical flaw sizes. Ceramics can have critical flaws on the
order of microns to 10s of microns, whereas metals can have critical flaws on the order of cm.
Ceramics are far more likely to contain critical flaws due to the nature of their processing and
the small allowable size, so ceramics tend to have poor fracture resistance. Stress concentrations
are more critical in ceramics, as they do not generally have any mechanisms to blunt the crack
tips; in metals stress concentrations can cause high enough local stresses to cause plastic deformation at the crack tip and effectively blunt the crack. Despite lower moduli of metals than ceramics, metals have higher specific surface energy, and this fact, combined with plasticity account for the higher energy absorption in fracture (higher fracture toughness).
3.2.2 Plane Strain Fracture Toughness is the critical stress intensity factor in plane strain.
Plane strain means there is no strain component along the axis of the crack (e.g., plane
strain in Fig 3.3 would mean there is no strain along the direction in or out of the page,
or there is only strain in one plane). Plane strain tends to occur in thick specimens. The
plane strain fracture toughness in mode one (I) fracture is:

K Ic = Y f ac
Where K-IC (K-one-c) is the critical (c) stress intensity (concentration) factor in mode one fracture (crack opening). ac is the critical flaw size. Y is a dimensionless geometrical crack, specimen, and loading factor (Y would have different values for edge cracks compared to center
cracks). f is the fracture stress, or stress at which failure by fracture occurs. Units are [MPam].
The above equation (KIC) can be applied to brittle materials, or materials that behave in a brittle
manner when loaded appropriately.
35

MATE 202: Introduction to Materials Engineering

Un-notched

Notched

Before

After

**IMPORTANT CONCEPT**
Figure 3.4 Experiments performed on silly putty to show the importance of notching toughness: the susceptibility of a materials to fracture in the presence of a flaw.

Stress concentrations can be responsible for changing fracture behaviour from ductile to brittle;
ductile materials can fail in a more brittle manner when stress concentrations are present.
A prime example of this concept can be demonstrated with your demo kit: Whats in the Box?
Pull on a cylindrical specimen of Silly PuttyTM to put it in simple tension to see if you can get it
to break. Notice that after pulling the sample quickly in simple tension the notched specimen
fractures in a brittle manner with practically no plastic deformation! Your can observe the crack
patterns on the surface of the specimen which resemble hackle patterns similar to those seen
in ceramics and in some metals. You may also see pores or hairs or dirt depending on how many
times you have dropped your putty!
36

Chapter 3: Mechanical Failure

Example in class demo of a much larger specimen of silly putty and the effect of stress concentration due to an edge crack.
3.3 Impact Failure by impact usually takes the form of fracture (ductile or brittle). The amount
of energy absorbed in fracture depends on the material, testing, and environmental conditions.
3.3.1 Impact Testing
Current awareness was developed as a result of the failure of Libertyt ships in WWII
that failed on cold days sometimes while loaded with troops (Fig. 3.5). These ships
were built with stress raisers and lower quality steels to replace the ships lost to the enemy submarines during the conflict.
BCC (body centered cubic) metals like steel (Fe-C alloy) become brittle under the influence of the following factors: 1. T (temperature), lowering T leads to brittleness, 2. At
notches which magnify applied stresses (sharper notch, more brittle). 3. When the load

Figure 3.5 An oil


tanker that failed by
brittle fracture of
the steel structure
by crack propagation around its
girth. Photograph
by Neal Boenzi
from the New York
Times

is applied very quickly (faster application results in more brittle nature).


3.3.1.1 Charpy Testing
The Charpy test is a standard test for determining a steels tendency to become brittle
and has all of the above embrittling factors present. The specimen and tester are shown
in Fig. 3.6.
Steel has a ductile to brittle transition (DBTT) and such a temperature only occurs in
BCC metals, because plastic deformation is thermally activated in those metals but not
FCC metals. The results of Charpy testing are shown schematically in Fig. 3.7 and with
real data in Fig. 3.8. DBTT can also occur in ceramics, where thermally activated diffusion at high temperature can lead to more ductile fracture. Ceramics have much higher
DBTTs than for metals.
The actual tendency to develop brittleness at low temperature depends on the structure
(grain) and chemistry (composition and impurities) of the steel, and varies from batch to
batch. This is a standard test to check the quality of the steel with respect to this property.
The Charpy test measures energy to fracture a notched specimen with a hammer blow as
a function of specimen T.
37

MATE 202: Introduction to Materials Engineering

Figure 3.7 and 3.8 show that BCC metals have an upper and lower shelf energy (shelves
are the average of the data that form a `shelf` on unchanging energy as a function of T.
The curve through the data can be fit with a mathematical function, or the data points
connected with a smooth line.
At low T, brittle fracture occurs by cleavage along certain crystallographic planes, failure occurs with little or no distortion. The surface is faceted and shiny and the facets
sparkle under direct light. Because there is no plastic deformation, the cross section is
square, and generally much smoother than a ductile fracture surface.
At high T, ductile fracture occurs. The sample deforms plastically and we see large distortion of the cross section and shear lips are present. The surface has a dull, fibrous
appearance since it does not reflect light very well.
The appearance of the fracture surfaces is shown in the photomicrographs of Fig. 3.9.

DBTTs can be reported in three main ways:


1. The temperature at which the fracture energy is the average of the upper and lower shelf
energies.
2. The temperature at a particular fracture energy (specified by a standard,e.g., 20J, 50J).
3. The temperature for a particular percentage of ductile vs. brittle fracture surface (e.g.,
50% ductile (also called fibrous or shear fracture)).
A low transition T (DBTT) is a good thing! But be careful how it was determined and
do not use transition temperature directly to reflect safe usage temperatures as it depends
on the three factors mentioned at the beginning of the section, how it was measured and:

Figure 3.6. Schematic diagrams showing


the specimen and tester for the Charpy test.
The test specimen is shown in a) and the test
configuration for Charpy (metal) and Izod
(polymers) is shown in the middle. b)
shows the test configuration. The hammer
is raised to the start position and fractures
the specimen and then swings to some position at the end of swing which is lower the
more energy is absorbed in the fracture process. A ratchet system holds the hammer in
its maximum position after the test. The
figure is from Hayden Moffat and Wulff.
Structure and Properties of Materials Vol.
III.

38

Chapter 3: Mechanical Failure

Figure 3.7 Schematic representation of energy variation of steel with temperature. The dotted line is
the ductile to brittle transition temperature (DBTT).
The lower the DBTT the higher the quality of the
steel for low temperature applications. The line on
the left is called the lower shelf energy and the line
on the right is the upper shelf energy.

Figure 3.8 Charpy impact energy versus temperature for several engineering alloys showing the variation with yield strength. Note that as the strength of steels goes up the toughness goes down. Aluminium is not nearly as tough as steel. From R. Hertzberg Deformation
and Fracture Mechanics of Engineering Materials, Wiley, adapted from ASTM.

BRITTLE

DUCTILE
Temperature in Celsius

Figure 3.9. Transition in fracture surface appearance as a function of test temperature. Specimens are
A36 steel tested in the transverse direction. The test temperature is above the specimen. Notice the
section shape and the degree of sparkle in each specimen. From R. Hertzberg, Deformation and Fracture Mechanics of Engineering Materials, Wiley.
39

MATE 202: Introduction to Materials Engineering

Orientation effects can be important since defects can be oriented by the plastic working
of the metal during manufacture. For example inclusions in steel are frequently elongated in the direction of hot rolling and the properties perpendicular to this direction
(transverse) are often different than along this direction (longitudinal or axial). The
properties must be measured in the direction appropriate for the loading that the part will
receive.
There are ductile to brittle transitions in ceramics and many polymers as well.
These transitions are typically measured with somewhat different tests, but reflect the
fact that we must be aware of temperature of use when considering material selection for
most materials. It is only FCC metals that do not have a serious drop in toughness
(recall Figure 3.8) at some point when the temperature is lowered. Note that any drop
may be well above or well below room temperature.

3.4 Fatigue
Fatigue is the most common mechanical failure mechanism in machines with moving parts.
Fatigue occurs when you stress a component long enough in a dynamic stress cycle (even if the
applied stress is below the yield strength).
Experiment: Whats in the Box? kit. Break a paper clip by bending it back and forth. Observe the difference in the paper clip with score marks and a smooth paper clip and as function
of the bend angle you impose.

Figure 3.10 Schematic diagram of fatigue-testing apparatus for making rotating-bending tests.
(from Materials Science in Engineering 4th Ed. by Carl A. Keyser)
3.4.1 Dynamic Stress
When the stress in a metal is varied (e.g., in a rotating shaft that is loaded in bending) then the
stresses vary sinusoidally between max and min
When a shaft loaded in bending is rotated any point on the outside surface goes through a
tension-compression cycle. This stress cycling is what causes fatigue. (If you havent
taken a course in mechanics of materials look up neutral axis in bending in an introductory text (e.g. E.G. Popov or Byars and Snyder) or on the internet)
There is a relationship between the stress amplitude and the number of cycles before the metal
fails by fatigue. So called S-N curve.

m =

( max + min )
2

a =

( max min )
2

40

Chapter 3: Mechanical Failure

Figure 3.11 Variation of stress with


time in rotating bending. This variation is responsible for fatigue cracking.
a) Reversed stress cycle with tension
maximum and compression minimum
of equal magnitude. The mean stress
is zero.
b) Stress cycle above with a tensile
maximum greater in magnitude than
the compressive minimum. This could
occur in simple rotating bending with a
residual tensile stress in the metal at
the location. Here the mean stress is
greater than zero.
(Source W.D.. Callister)
3.4.4 The S-N Curve a graphical representation of how the stress amplitude, S, and number of
cycles to failure, N, are related. The cycles to failure is plotted on a logarithmic scale. S-N
curves are generally plotted as stress amplitude, sa versus N. S-N curves are displayed in Fig.
3.12. This figure is one of the most important graphs for fatigue!

Figure 3.12 Stress amplitude versus logarithm of the number of cycles to failure (S
-N curves) for:
a) a material that exhibits a fatigue limit,
such as a BCC metal like steel.
b) a material that does not exhibit a fatigue limit, such as a FCC metal like aluminium.

(Source W.D.. Callister)

41

MATE 202: Introduction to Materials Engineering

The effect of mean stress on fatigue is shown in Fig. 3.13. As mean stress increases, the S-N
curve shifts downwards and to the left. The effect of mean stress is considered less significant
than the effect of stress amplitude, but is significant. One way to diminish fatigue is to reduce
the mean stress, by putting the surface of the metal into compression by shot peening.

Figure 3.13 Effect of mean stress, sm ,on the SN curve. As mean stress, sm, increases
the S-N line shifts DOWN and to the LEFT

(Source: modified from W.D.. Callister)

Fatigue Limit (Fatigue Strength)


A stress below which fatigue failure is very unlikely to occur.
The fatigue strength (for BCC; failure stress at 107 cycles for FCC like Al, Cu, Au) is between one quarter and one half the tensile strength and this limit is often used in fatigue design. If it is not known from testing then one quarter the tensile strength is used as the fatigue strength. FCC metals (soft and ductile metals) will always fail in fatigue no matter the
stress, whereas steels have a fatigue limit, but it may occur at so many cycles that for practical purposes it is not relevant for design since the lifetime will not be as long as the fatigue
life.
3.4.5 Failure surface

Figure 3.14 Fracture surface of a rotating


steel shaft that experienced fatigue failure.
The origin is the corner of the keyway at
the top which acted as a stress raiser to
magnify the stresses as discussed in section 3.2.1. The visible markings are known
as beachmarks and represent locations
where the crack progress was arrested
long enough for the surface to be oxidized
so that the crack front was distinguished.
There is a ductile overload failure in the
lower right corner when the remaining
cross section would no longer carry the
maximum applied stress. From D.J. Wulpi Understanding How Components Fail.
ASM

42

Chapter 3: Mechanical Failure

We see a crack surface with beach marks on it that ringed the crack initiation point, but only
localized deformation near the fracture (no shear lips). These beach marks are arrest markings between periods when the load is cycling. There is a small region of ductile failure
after the fatigue crack has grown to such an extent that the remaining area cannot support
the load.

Fatigue failures are actually statistical in nature. The more serious the consequences of failure the lower the allowed stress amplitude for a given S-N curve, which may be drawn
through average values. To understand why we would use lower S values for a given number of cycles compare designing an automatic stamping machine to an aircraft.

We observe that an increase in the mean stress with the same stress amplitude (figure) will
decrease the fatigue life of a part, since the mean stress increase (at the same stress amplitude) moves the S-N curve downwards and to the left. (An example of an increased mean
stress would be an axial tensile loading superimposed on the bending stresses in the fatigue
testing machine.)

The final rupture surface may exhibit brittle or ductile fracture (without shear lips).

3.4.6 Crack initiation and propagation for fatigue to occur, a crack must form and propagate
through the material. The crack is normally initiated via stress concentration at applied
stresses less than the yield or tensile strength.
Cracks then grow during each stress cycle. The stress amplification will usually increase as the crack grows so that the stress amplitude at the crack tip increases and the
growth may accelerate.
Mechanism of Fatigue
A fatigue crack propagates by the action of the stress reversal causing the crack tip to move
forward a little bit with each stress cycle, each step is a striation. (see Fig.3.15) these
striations are microscopic and there may be a thousand striations between beach marks

Figure 3. 15 Transmission electron


fractograph showing fatigue striations
in aluminium. Each striation corresponds to one stress cycle and these
striation are very small and invisible to
the eye on the fracture surface. They
are used to confirm that the failure was
from fatigue. From Colangelo and
Heiser, Analysis of Metallurgical Failures. John Wiley.

43

MATE 202: Introduction to Materials Engineering

(clamshell markings). Striations result from plastic flow at the crack tip under the loading.
Factors affecting fatigue lifetime (endurance limit) anything that affects crack initiation or
propagation will influence fatigue.
Generally the cycles to failure (for ductile metals) can be increased by:
a. Lowering the mean stress
b. Altering the surface:
i. Geometrical considerations reducing stress concentrations (e.g., designing in fillets
on changes in section)
ii. Treating the surface to:
1. Put it into compression (e.g., shot peening (intentional strain hardening of the surface by impingement) carburizing, case hardening). This makes it more difficult
for cracks to form because a higher stress would be required.
2. Make it smoother (i.e., polishing). Fatigue is strongly dependent on surface finish,
with rougher surfaces showing lower fatigue strength, hence the need to specify
surface finishes in fatigue design (related to stress magnification at the surface).
Dont over specify; polished surfaces are expensive to machine.
c. Reduce exposure to corrosive environments corrosion exacerbates fatigue (corrosion fatigue causes lower fatigue resitance than plain fatigue under the same loading conditions).
d. Reduce thermal gradients particularly for materials that are poor conductors. Important temperature control is important if and only if the temperature gradient is reduced because such
gradients produce additional strains and stresses in the material.
Figure 3.16 is a summary of different observed fracture surfaces caused by fatigue as a function
of the stress state, amount of stress applied, and amount of stress concentration. The guide is
useful for interpreting failures in order to determine if the part is actually behaving as it is
thought to be behaving!
3.5 Creep
DEMO: Silly putty creep, video.
Failure of metal due to slow continuous permanent deformation induced by a static load
(usually tension) at high temperature (i.e., at T > one third to 0.4 of Tmp (K) (homologous temperatures of 0.33 to 0.4).
Creep is important in turbines, engines, boilers, etc. Total deformation before failure may
be less than 5%! (Somewhat surprising considering that the specimen would elongate more
than 20% in a tensile test.)
Ceramics and especially amorphous polymers will also creep (though the temperature at
which this occurs is not as easily determined as it is for metals).
Failure times at different stresses are determined by a creep test, apply a load in a T chamber
and measure strain over time.
Unlike a tension test at room T where strain hardening prevents continuing deformation under
load, strain in a creep test continues under load i.e. homologous Ts above about 0.3 to 0.5
(0.4).

44

Chapter 3: Mechanical Failure

Figure 3.16. A schematic of fatigue failure surfaces as a function of stress, stress concentration, and
stress state. After ASM Metals Handbook.

45

MATE 202: Introduction to Materials Engineering

3.5.1 The creep test A tensile stress is induced at high temperature and the strain measured. A
constant load (or stress) is used in creep testing in contrast to the increasing stress in a tensile
test. Specimen geometry is similar to tensile testing (tabs with a reduced section to localize deformation). Schematically the applied load is constant over time, as show below (engineering
stress) Fig. 3.17. You will see a version of creep testers in the
laboratory Fig. 3.18. The applied stress is actually lower

than the yield stress (defined by short time testing)! Most


,
creep tests are run in constant load, but constant stress may also
be used for greater accuracy in material response (much like
true-stress true-strain versus engineering stress-strain). Constant stress tests require compensating for the decreasing cross
section as the sample elongates

Figure 3.17 The creep sample and the engineering stress time curve.

Figure 3.18 A creep test machine for lead, where creep occurs at room temperature because the melting
point is only 300C.

3.5.2 Creep Curves The creep strain as a function of time is plotted until rupture Fig. 3.19.
The three stages of creep
In a tensile creep test (and in creep) there are 3 stages (see Fig. 3.19) plus the instantaneous deformation caused by instantaneously loading the material under stress (instantaneous deformation is mostly elastic).
46

Chapter 3: Mechanical Failure

Primary creep stage (transient), strain rate decreases as strain hardening occurs (structural
defects called dislocations are building up in the structure and interfering with deformation
yield stress is increasing).
Secondary creep stage, the strain rate is constant as the number of dislocations (defects in the
grains which cause plastic deformation and strain hardening) stays constant, the production
of dislocations by continued plastic deformation is balanced by a spontaneous removal of
them at these temperatures by a thermally-activated annealing process that removes excess
energy/defects. Because the annealing cancels out the strain hardening effect, this stage is
often called steady state creep. A material spends most of its time in secondary creep, and
the transition to tertiary creep is generally sudden and rapid.
Tertiary creep stage, an instability occurs usually necking and the strain rate increases leading
to failure by rupture.
Final rupture may be ductile or brittle fracture. Note from above that the amount to plastic
strain to failure may be drastically lower than room temperature tensile ductility. Rupture lifetime (tr) is the total amount of time in creep before rupture (failure by fracture).

Figure 3.19 Typical creep curve of strain versus time at constant elevated temperature
(relative to the melting point of the metal).
The minimum creep rate is the slope of the
linear segment in the region of secondary
creep. Rupture lifetime is the total time to rupture. In design creep is not usually allowed to
proceed past the secondary stage.

Effect of T Creep is thermally activated, so it obeys an Arrhenius equation.

d
Qc
= Aexp

dt
RT

Secondary creep rate is given as:

Here, A is a constant (metal dependent), Qc is the activation energy for creep, R is the gas
constant and T is in K. The equation allows us to predict creep rate at any temperature if it
is known at one temperature.
We say that creep rate increases exponentially with T. A plot of ln of creep rate versus 1/T
gives a straight line of slope Q/R.

Effect of stress () Creep increases in a power law equation with stress:

d = B n where n 7
dt
47

MATE 202: Introduction to Materials Engineering

Here B (sometimes called K2) is a metal dependent constant. When doubles strain rate goes
up 27 or 128 times! The equation allows us to predict creep rates at different stresses from the
rate at one stress.
Figure 3.20. Tensile creep curves at
a variety of increasing temperatures.
Below homologous temperatures of
about 0.4 there is no creep. As T
increases the curves become shorter
and steeper. From Callister.

The effects of temperature and stress are shown in Figure 3.20. Notice that higher stresses and
higher temperatures result in higher secondary creep strain rates, shorter rupture times, and higher creep strains to failure.
3.5.2.4 Log-log plots
Creep information about materials is often given in concentrated form as a plot of log stress versus log steady-state creep rate (strain rate in the second stage of creep is also called minimum
creep rate). These plots are used to determine creep lifetimes based on an allowable secondary
creep deformation of 2 or 3 per cent .
Increasing temperature and stress cause large increases in creep rate. i.e. even small changes in
stress and T are important (see Figure 3.20 and 3.21).
This information all pertains to dislocation creep, the most common kind, but there are other
kinds of creep at higher Ts and lower stresses, whose description is beyond this course.
See for example R. Hertzberg, Deformation and Fracture Mechanics of Engineering Materials, Wiley

Figure 3.21 Logarithm of stress plotted versus steadystate (second stage) creep rate for a low carbon nickel
alloy at three temperatures. From ASM Metals Handbook. Vol. 3 9th edition.

48

Chapter 3: Mechanical Failure

The combination of stress and temperature are shown in stress versus creep rate plots as shown
above. The equations to describe such lines are logarithmic, because the scales are both logarithmic. See worked examples below.

The base 10 log-log plot can be used to find values on the curve, then the
natural logarithm to solve for activation energy, n, etc.

49

MATE 202: Introduction to Materials Engineering

3.5.2.4.1 Example
A specimen 1015 mm long of a low carbon-nickel alloy (Figure belowW.D. Callister) is to be
exposed to a tensile stress of 70 MPa at 427oC. Determine its total elongation after 10,000 h.
Assume that the total of both instantaneous and primary creep elongation is 1.3 mm

Solution:
Draw line to intersect creep rupture curve of 427C
Drop line down to steady-state creep rate NOTE units of strain rate are %/1000hr this
will not always be the case!
The creep rate for the stress of interest will likely fall between two log divisions.
In this case these divisions are 10-2 and 10-1.
You can eyeball the distance, but the division is not a linear change.
Note that the distance between log divisions is always the same on a log scale!
Use a ruler to measure distance from the nearest log division on the left of the line to the line
(value of creep rate). This distance is A.
Use a ruler to measure the distance from the nearest log division on the left to the nearest log
division on the right. This distance is B.
To find the value of the creep rate we know that it has to be greater than 10-2 but less than 10-1,
so we need to find the value of the exponent by: 10(-2 + A/B)
-2 +33/50 = -1.34 (between -2 and -1? YES, looks okay)
The values 33 and 50 will change based on scale
Creep rate is then 10-1.34%/1000hr
Start converting units to get a value of the strain 10-1.34 %/1,000hr x 10,000hr = 10-1.34 % x 101
Add exponents to give 10-0.34 %
Take antilog (base 10) of exponent: log-1(-0.34) = 0.457%
THIS IS THE STRAIN after 10,000 hrs at 427C under 70 MPa in secondary creep!
We want the total elongation in mm:
By the time the specimen enters secondary creep it has already elongated 1.3mm
(instantaneous and primary creep elongation), so the initial length at the start of secondary
creep is 1015+1.3mm = 1016.3 mm
So, 1016.3x0.457% = 4.64mm
FINAL ANSWER : the total elongation is = 1.3 + 4.6 = 5.9mm
50

Anda mungkin juga menyukai