Term of Luen CIA

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

CHAPTER 20

Creep and Creep Fracture

CONTENTS
20.1 Introduction..................................................................................................................311
20.2 Creep testing and creep curves.......................................................................315
20.3 Creep relaxation ........................................................................................................318
20.4 Creep damage and creep fracture..................................................................319
20.5 Creep-resistant materials ....................................................................................320
Worked example...................................................................................................................321
Examples ..................................................................................................................................323

20.1 INTRODUCTION
So far we have concentrated on mechanical properties at room temperature.
Many structures—particularly those that are associated with energy conversion
(e.g., turbines, reactors, steam, and chemical plant)—operate at much higher
temperatures.
At room temperature, most metals and ceramics deform in a way that depends
on stress but which, for practical purposes, is independent of time:
e ¼ f ðsÞ elastic=plastic solid
As the temperature is raised, loads that give no permanent deformation at room
temperature cause materials to creep. Creep is slow, continuous deformation
with time: the strain, instead of depending only on the stress, now depends
on temperature and time as well:
e ¼ f ðs, t, TÞ creeping solid

Engineering Materials I: An Introduction to Properties, Applications, and Design, Fourth Edition


© 2012, Michael F. Ashby and David R. H. Jones. Published by Elsevier Ltd. All rights reserved. 311
312 CHAPTER 20 Creep and Creep Fracture

FIGURE 20.1
A tungsten lamp filament which has sagged undo its own weight owing to creep.

It is common to refer to the former behavior as “low-temperature” behavior,


and the latter as “high temperature.” But what is a “low” temperature and what
is a “high” temperature? Tungsten, used for lamp filaments, has a very high
melting point—well over 3000$ C. Room temperature, for tungsten, is a very
low temperature. If made hot enough, however, tungsten will creep—that is
the reason that lamps ultimately burn out. Tungsten lamps run at about
2000$ C—this, for tungsten, is a high temperature. If you examine a lamp fila-
ment that has failed, you will see that it has sagged under its own weight until
the turns of the coil have touched—that is, it has deformed by creep. Figure 20.1
shows a typical example of a sagging filament.
Figure 20.2 and Table 20.1 give melting points for metals and ceramics and soft-
ening temperatures for polymers. Most metals and ceramics have high melting
points and, because of this, they start to creep only at temperatures well above room
temperature—this is why creep is a less familiar phenomenon than elastic or plastic
deformation. But the metal lead, for instance, has a melting point of 600 K; room
temperature, 300 K, is exactly half its absolute melting point.Room temperature for
lead is a high temperature, and it creeps—as Figure 20.3 shows (see page 313). And
the ceramic ice melts at 0$ C. Temperate glaciers (those close to 0$ C) are at a tem-
perature at which ice creeps rapidly—that is why glaciers move. Even the thickness
of the Antarctic ice cap, which controls the levels of the earth’s oceans, is determined
by the creep spreading of the ice at about –30$ C.
The point, then, is that the temperature at which materials start to creep de-
pends on their melting point. As a general rule, it is found that creep starts when
T > 0:3 to 0:4 TM for metals
T > 0:4 to 0:5 TM for ceramics
where TM is the melting temperature in degrees kelvin. However, special alloy-
ing procedures can raise the temperature at which creep becomes a problem.
20.1 Introduction 313

Ceramics Metals Polymers Composites


5000

Diamond/
4000
Graphite

Tungsten

Tantalum
SiC
3000 MgO Molybdenum
Niobium
(K)

Al2O3 Chromium
Si3N4 Zirconium
Platinum
2000 Titanium
Iron/Steel
Silicon Cobalt/Nickel
Cermets
Copper
Gold
Silver
Alkali
halides Aluminum
1000
Magnesium
Zinc Polyimides
Lead Melamines
Tin Polyester, PC GFRP, CFRP
PE, Epoxy, PS Woods
Ice Mercury Nylon, PP

0
FIGURE 20.2
Melting or softening temperature.

Polymers, too, creep—many of them do so at room temperature. As we said in


Chapter 5, most common polymers are not crystalline and have no well-defined
melting point. For them, the important temperature is the glass temperature, TG,
at which the Van der Waals bonds solidify. Above this temperature, the polymer
is in a leathery or rubbery state and creeps rapidly under load. Below, it becomes
hard (and sometimes brittle) and, for practical purposes, no longer creeps. TG is
near room temperature for most polymers, so creep is a problem.
In design against creep, we select the material and the shape that will carry the
design loads, without failure, for the design life at the design temperature. The
314 CHAPTER 20 Creep and Creep Fracture

Table 20.1 Melting or Softening(S) Temperature


Material T(K) Material T(K)

Diamond, graphite 4000 Gold 1336


Tungsten alloys 3500–3683 Silver 1234
Tantalum alloys 2950–3269 Silica glass 1100(S)
Silicon carbide, SiC 3110 Aluminum alloys 750–933
Magnesia, MgO 3073 Magnesium alloys 730–923
Molybdenum alloys 2750–2890 Soda glass 700–900(S)
Niobium alloys 2650–2741 Zinc alloys 620–692
Beryllia, BeO 2700 Polyimides 580–630(S)
Iridium 2682–2684 Lead alloys 450–601
Alumina, Al2O3 2323 Tin alloys 400–504
Silicon nitride, Si3N4 2173 Melamines 400–480(S)
Chromium 2148 Polyesters 450–480(S)
Zirconium alloys 2050–2125 Polycarbonates 400(S)
Platinum 2042 Polyethylene, high-density 300(S)
Titanium alloys 1770–1935 Polyethylene, low-density 360(S)
Iron 1809 Foamed plastics, rigid 300–380(S)
Carbon steels 1570–1800 Epoxy, general purpose 340–380(S)
Cobalt alloys 1650–1768 Polystyrenes 370–380(S)
Nickel alloys 1550–1726 Nylons 340–380(S)
Cermets 1700 Polyurethane 365(S)
Stainless steels 1660–1690 Acrylic 350(S)
Silicon 1683 GFRP 340(S)
Alkali halides 800–1600 CFRP 340(S)
Beryllium alloys 1540–1551 Polypropylene 330(S)
Uranium 1405 Ice 273
Copper alloys 1120–1356 Mercury 235

meaning of “failure” depends on the application. We distinguish four types of


failure, illustrated in Figure 20.4.
1. Displacement-limited applications, in which precise dimensions or small
clearances must be maintained (as in the discs and blades of gas turbines).
2. Rupture-limited applications, in which dimensional tolerance is relatively
unimportant, but fracture must be avoided (as in pressure-piping).
3. Stress relaxation-limited applications in which an initial tension relaxes
with time (as in the pretensioning of bolts).
4. Buckling-limited applications, in which slender columns or panels carry
compressive loads (as in structural steelwork exposed to a fire).
To analyze these we need constitutive equations which relate the strain-rate e_ or
time-to-failure tf to the stress s and temperature T.
20.2 Creep Testing and Creep Curves 315

FIGURE 20.3
Lead pipes often creep noticeably over the years. (Source: M.F. Ashby.)

20.2 CREEP TESTING AND CREEP CURVES


Creep tests require careful temperature control. Typically, a specimen is loaded
in tension, usually at constant load, inside a furnace maintained at a constant
temperature, T. The extension is measured as a function of time. Figure 20.5
shows a typical set of results from such a test. Metals, polymers, and ceramics
all show creep curves of this shape.
Although the initial elastic and the primary creep strain cannot be neglected, they
occur quickly, and they can be treated in much the way that elastic deflection is
allowed for in a structure. But thereafter, the material enters steady state, or sec-
ondary creep, and the strain increases steadily with time. In designing against
creep, it is usually this steady accumulation of strain with time that concerns
us most.
By plotting the log of the steady creep rate, e_ ss , against log s at constant T, as
shown in Figure 20.6, we can establish that
e_ ss ¼ Bsn ð20:1Þ
where n, the creep exponent, usually lies between 3 and 8. This sort of creep is
called “power-law” creep. (At low s, a different régime is entered where n %1;
we shall discuss this low-stress deviation from power-law creep in Chapter 22.)
316 CHAPTER 20 Creep and Creep Fracture

"

(a)

(b)

(c)

"

(d)
FIGURE 20.4
Creep is important in four classes of design: (a) displacement-limited, (b) failure-limited, (c) relaxation-
limited, and (d) buckling-limited.
20.2 Creep Testing and Creep Curves 317

Fracture x
F Tertiary
creep

Furnace
windings Strain e Steady-state
creep e ss ef

Initial Primary creep t1


elastic
F strain
Time, t
FIGURE 20.5
Creep testing and creep curves.

Slope
n ≈ 3 to 8
Power-law
creep
Log e ss

Low ! tail, n ≈ 1

log !
FIGURE 20.6
Variation of creep rate with stress.
By plotting the natural logarithm (ln) of e_ ss against the reciprocal of the absolute
temperature (1/T) at constant stress, as shown in Figure 20.7, we find that:
!
e_ ss ¼ Ce&ðQ=RTÞ ð20:2Þ
Here R! is the Universal Gas Constant (8.31 J mol&1 K&1) and Q is called the
Activation Energy for Creep—it has units of J mol&1. Note that the creep rate in-
creases exponentially with temperature (Figure 20.7). An increase in tempera-
ture of 20$ C can double the creep rate.
Combining these two dependences of e_ ss gives
!
e_ ss ¼ Asn e&ðQ=RTÞ ð20:3Þ
where A is the creep constant. The values of the three constants A, n, and Q char-
actize the creep of a material; if you know these, you can calculate the strain rate
318 CHAPTER 20 Creep and Creep Fracture

e!ss
Slope

–Q/R T
In e!ss

1/T
FIGURE 20.7
Variation of creep rate with temperature.

at any temperature and stress by using the last equation. They vary from mate-
rial to material, and have to be found experimentally.

20.3 CREEP RELAXATION


At constant displacement, creep causes stresses to relax with time. Bolts in hot
turbine casings must be regularly tightened. Plastic paper clips are not, in the
long-term, as good as steel ones because, even at room temperature, they slowly
lose their grip.
The relaxation time (arbitrarily defined as the time taken for the stress to relax to
half its original value) can be calculated from the power-law creep data as fol-
lows. Consider a bolt that is tightened onto a rigid component so that the initial
stress in its shank is si. In this geometry (refer to Figure 20.4(c)) the length of
the shank must remain constant—that is, the total strain in the shank etot must
remain constant. But creep strain ecr can replace elastic strain eel, causing the
stress to relax. At any time t
etot ¼ eel þ ecr ð20:4Þ
But
eel ¼ s=E ð20:5Þ
and (at constant temperature)
e_ cr ¼ Bsn
20.4 Creep Damage and Creep Fracture 319

!/E

!iE

Creep
strain
Total
strain

Elastic
strain

t
FIGURE 20.8
Replacement of elastic strain by creep strain with time at high temperature.

Since etot is constant, we can differentiate Equation (20.4) with respect to time
and substitute the other two equations into it to give
1 ds
¼ &Bsn ð20:6Þ
E dt
Integrating from s ¼ si at t ¼ 0 to s ¼ s at t ¼ t gives
1 1
& ¼ ðn & 1ÞBEt ð20:7Þ
sn&1 sn&1
i

Figure 20.8 shows how the initial elastic strain si/E is slowly replaced by creep
strain, and the stress in the bolt relaxes. If, as an example, it is a casing bolt in a
large turbogenerator, it will have to be retightened at intervals to prevent steam
leaking from the turbine. The time interval between retightening, tr, can be cal-
culated by evaluating the time it takes for s to fall to (say) one-half of its initial
value. Setting s ¼ si/2 and rearranging gives
ð2n&1 & 1Þ
tr ¼ ð20:8Þ
ðn & 1ÞBEsn&1
i

Experimental values for n, A, and Q for the material of the bolt thus enable us to
decide how often the bolt will need retightening. Note that overtightening the
bolt does not help because tr decreases rapidly as si increases.

20.4 CREEP DAMAGE AND CREEP FRACTURE


During creep, damage accumulates in the form of internal cavities. The damage
first appears at the start of the tertiary stage of the creep curve and grows at an
increasing rate thereafter. The shape of the tertiary stage of the creep curve
320 CHAPTER 20 Creep and Creep Fracture

Final
fracture x
e Damage
accumulates

Creep damage
starts
ef

Voids appear on tf
grain boundaries

t
FIGURE 20.9
Creep damage.

(refer to Figure 20.5) reflects this: as the cavities grow, the section of the sample
decreases, and (at constant load) the stress goes up. Since e_ 1 sn , the creep rate
goes up even faster than the stress does (Figure 20.9).
It is not surprising—since creep causes creep fracture—that the time-to-failure, tf
is described by a constitutive equation which looks very like that for creep itself:
!
tf ¼ A0 s&m eþðQ=RTÞ
Here A0 , m, and Q are the creep-failure constants, determined in the same way as
those for creep (the exponents have the opposite sign because tf is a time
whereas e_ ss is a rate).
In many high-strength alloys this creep damage appears early in life and leads to
failure after small creep strains (as little as 1%). In high-temperature design it is
important to make sure:
n That the creep strain ecr during the design life is acceptable
n That the creep ductility ef , cr (strain to failure) is adequate to cope with the
acceptable creep strain
n That the time-to-failure, tf, at the design loads and temperatures is longer (by
a suitable safety factor) than the design life
Times-to-failure are normally presented as creep-rupture diagrams (Figure 20.10).
Their application is obvious: if you know the stress and temperature you can
read off the life; if you wish to design for a certain life at a certain temperature,
you can read off the design stress.

20.5 CREEP-RESISTANT MATERIALS


From what we have said so far it should be obvious that the first requirement
that we should look for in choosing materials that are resistant to creep is that
they should have high melting (or softening) temperatures. If the material can
Worked Example 321

T3 > T2 > T1
log !

T1

T2
T3

log t f

FIGURE 20.10
Creep-rupture diagram.

then be used at less than 0.3 of its melting temperature creep will not be a prob-
lem. If it has to be used above this temperature, various alloying procedures can
be used to increase creep resistance. To understand these, we need to know
more about the mechanisms of creep—the subject of the next two chapters.

WORKED EXAMPLE
The photograph on the next page shows an easy experiment you can rig up at
home to demonstrate creep—and even measure the creep exponent, n. Take a
length of wooden curtain pole or broom handle (mine was 28.5 mm diame-
ter—11 8”), and wrap electronic solder wire (mine was 1.60 mm diameter—
=

1 ”) around the pole to make a coil that has no gaps between the turns. I wound
=
16

36 turns of wire on to the pole. Then hook the end of the coil over the end of the
pole as shown, and put the axis of the pole vertical (hold the lower end in a vice
or clamp). Do the experiment indoors in the warm, and within a few minutes
the coil will start to open out. After about 15 minutes it will look like the one in
the photograph. Obviously the solder (an alloy of lead and tin) has experienced
considerable creep deformation. This is exactly what you would expect—elec-
tronic solder melts at 183$ C (456 K), so a room temperature of 20$ C (293 K) is
0.64 TM—well into the régime where creep will occur.
You can see from the photograph that the coils have opened out much more at
the top (where the wire supports the maximum load) than they have toward the
bottom (where the wire supports a much smaller load). This tells you that the
creep rate is increasing with stress, as it should do. But how fast does it increase?
The value of the creep exponent n can easily be found as follows.
322 CHAPTER 20 Creep and Creep Fracture

Starting from the top of the coil, measure the positions of the crests of the wire
(e.g., 1, 18, 34, 49, 63 mm). List the differences between these measurements
which is the pitch of each turn (e.g., 17, 16, 15, 14 mm) and correlate these
pitches with the number of turns of wire they support (e.g., 36, 35, 34, 33).
The opening-out of each turn is found by subtracting the wire diameter
(1.6 mm) from the pitch, giving 15.5, 14.5, 13.5, 12.5 mm, and so on (rounded
to the nearest 0.5 mm).
Then plot log (opening out) versus log (number of turns supported), e.g., log
15.5 versus log 36, log 14.5 versus log 35, etc. The full graph is shown in the
following diagram. Over most of the range, there is an excellent fit to a straight
line with a slope of 2.1, so n ¼ 2.1. Only toward the bottom of the coil do
the data points go off, but this is because it is very difficult to make
accurate measurements of pitch from the photograph when the turns have
opened out so little.
Examples 323

1.0

1.0 – 0
n= = 2.1
1.47 – 1.0
log "

0.5
Slope
n

0
0.8 1.0 1.2 1.47 1.6
log N

EXAMPLES
20.1 A cylindrical tube in a chemical plant is subjected to an excess internal pressure
of 6 MN m&2, which leads to a circumferential stress in the tube wall. The tube
wall is required to withstand this stress at a temperature of 510$ C for 9 years.
A designer has specified tubes of 40 mm bore and 2 mm wall thickness made from
a stainless alloy of iron with 15% by weight of chromium. The manufacturer’s
specification for this alloy gives the following information.

Temperature ($ C) 618 640 660 683 707


Steady-state 1.0 ( 10&7 1.7 ( 10&7 4.3 ( 10&7 7.7 (10&7 2.0 ( 10&6
creep rate e_ (s&1),
for an applied
tensile stress s of
200 MN m&2

You might also like