You are on page 1of 7

Scripta METALLURGICA Vol. 17, pp. 3 1 - 3 7 , 1 9 8 3 Pergamon Press Ltd.

P r i n t e d in the U . S . A .

VIEWPOINT SET No. 5

CONTINUOUS CAVITY NUCLEATION At© CREEP FRACTURE

B.F. Dyson
Division of Materials Applications
National Physical Laboratory
Queens Road, Teddington
Middlesex, TWql OLW, UK

(Received October 14, 1982)

Introduction

Intergranular creep fracture under the stress and temperature conditions experienced by compo-
nents in service is generally acknowledged as resulting from nucleation and growth of inter-
granular cavities. During the last twenty or more years, there have been periods of debate
regarding the relative importance of nucleation and growth in controlling the overall kinetics of
fracture, a point which is still unresolved today. The early views were expressed necessarily
without much data on cavity nucleation and growth kinetics and ideas centred mainly around two
points. Firstly, that there always seemed to be an approximately linear reciprocal relationship
between time-to-fracture, tf , and minimum rate of creep deformation, ~min, the so-called Monkman-
Grant equation (I). Second~y, that cavity growth by stress-airected flow of atoms, first proposed
by Balluffi and Seigle (2), was shown by Hull and Rimmer (3) to lead to lifetime being inversely
proportional to the maximum principal stress, oi, providing that all cavities initiated at the
start of the test. To overcome the obvious inadequacy of the Hull-Rimmer prediction viz, that
tf ~ o I rather than tf ~ I/@min, two points of view emerged in the 1960's. The first rejected
the Hull-Rimmer analysis and proposed that cavities grew at a rate proportional to strain rate
(~,5), thus implying that fracture was still controlled by cavity growth rates. A possible
mechanism was given by Ishida and McLean (6). The other, due independently to Greenwood (7) and
Woodford (8), simply replaced the assumption of a constant density of cavities by one in which
cavitation increased continuously throughout the test, such that nucleation rate was proportional
to strain rate with the implication that fracture was controlled mainly by the cavity nucleation
rate. Recently, it has been demonstrated that the boundary conditions imposed by Hull and Rimmer
were unnecessarily restrictive and that more realistic ones result in cavity growth rates being
proportional to strain rates (9-13). Thus the first viewpoint has gained a measure of credibi-
lity, albeit indirectly, and has resulted in the development of cavity-growth mechanism maps for
some metals and alloys (14,15) which, at least in principle, provide a convenient means of asses-
sing the validity of extrapolation techniques. Nevertheless, there is an increasing body of
experimental evidence on engineering alloys (16-22) and also iron (23) which indicates that cavi-
ties are not all nucleated in a short time interval, as assumed above, but are nucleated conti-
nuously in the manner proposed by Greenwood and Woodford. This paper collates the evidence for
the continuous nucleation of cavities, places this within the framework of creep fracture model-
ling and introduces a quantitative assessment of the likely errors of measurement. A model of
cavity nucleation based on the stochastic nature of transgranular creep deformation is also
discussed.
Measurements of Cavity Nucleation Rate
Two main methods have been used to measure nucleation kinetics. Cane and Greenwood (23) and
Cane (19) took advantage of the inherent low temperature brittleness of iron and a low alloy
ferritic steel respectively to determine directly the number of cavities per unit area of grain
boundary facet, Na, as a function of strain and time. The number of cavities per unit area of
polished microsection, Nm, was measured by Dyson and McLean (16,18), Tipler and Hopkins (17),
Lonsdale and Flewitt (20) and Needham and Gladman (21), while Chen and Argon (22) chose a

31
0036-9748/83/010031-07503.00/0
32 CONTINUOUS CAVITY NUCLEATION Vol. 17, No. 1

similar parameter, Nb, the number of cavities per unit length of cavitated boundary. The three
parameters can be related quantitatively using standard stereological relationships (24,25):-

d.
N _ ,1 N (I)
a ~ DH m

N 2¢
m = d. Nb (2)
l

where d is the mean linear intercept grain size; ~ , the fraction of boundaries cavitated and
DH, the ~armonic mean of intersected cavity diameters, Di, whose frequency of occurrence is fi,
i.e. OH = I / Z fi(q/oi).
The important common observation was that
10 the cavities were not all nucleated immediately
Cr-lMo (19) upon application of the creep load or even
after an incubation period, as is assumed in
modelling studies, but rather their number was
~e 347 steel (21)
8
found to increase progressively during the
o test. Nucleation was either continuous until
® fracture intervened (16,17,20,21) or the
'-'T'
number reached an approximate saturation value
before fracture (18,fl9,22). Another equally
=. 6 important deduction from some of the
~E researches was that strain was the parameter
q
controlling the number of cavities and not
:,T, time (16,18,19,21). Furthermore, in the early
stages, number increased linearly with strain
and the rate of cavity accumulation varied
greatly from material to material, as Figure I
shows. Here N a has been used as the measure
of number since, from equations I and 2, Nm
and Nb additionally reflect kinetics of growth.
Z 2- Only initial rates have been used and for
Z Nimonic 8OA and Type 347 stainless steel, Na
values were calculated from quoted data on
number per unit volume, Nv which is related to
8OA(18) I N a by:-
I I I I !
0 1 2 3 4 5 d.
I
E, Creep strain x 100 Na - 2@ Nv (3)

FIG. 1 and assuming a value of 0.2 for @ . There


Showing that different materials nucleate were insufficient data to calculate N a values
cavities continuously at very different rates. from the other papers.

Also shown in Figure I are data for Nimonic 80A tested in a torsional stress state where,
for a given yon Mises stress, the maximum principal stress o I is only I/ 3 that in tension. The
effect is to reduce the nucleation rate by two thirds whereas there is a corresponding increase
in rate when ~1 becomes larger than in tension (26). These stress state results are paralleled
by the effects of trace elements in a ferritic steel where the rates of cavity accumulation are
raised by certain trace element additions (17). Similar results were reported by Thomas and
Gibbons (27) for a nickel-base superalloy.
The large variation in dNa/de exhibited by the different materials in Figure I is related to
their creep strains at failure as shown in Figure 2. Not surprisingly, a large value of dNa/de
is associated with a low ductility and vice-versa. In compiling Figure 2, Needham and Gladman's
data have been slightly re-interpreted since they chose to ignore slight but systematic
differences in cavity accumulation rate with stress and these have been included in Figure 2 (but
not Figure I) for completeness. This reciprocal relationship between cavity accumulation rate
and ductility has also been noted in the work on trace elements (17,27) and stress state (18,28).
Vol. 17, No. I CONTINUOUS CAVITY NUCLEATION 33

100 10 4

O Nimonic 80A ~' torsion


O z~ tension
x
Type 347 FI
E
A 2 ~ Cr 1 M o E)
Iron •
l T
E 10 3 .
[] SIo
10 [] E
I_ E] I/I
c-
CL
10
® /
(J .1
>
(~
U
10 2 -

1 I I Z

10 4 10 5 10 6 10 ?
dNa mm-2
_4
dE:
FIG. 2
10
Showing that creep ductility is inversely
related to the increase in cavity number I
per unit strain. 0.1 1.0 10
E, Creep strain x 100
Further, if indirect, evidence of continuous
cavity nucleation during creep of Nimonic 80A is FIG. 3
provided in Figure 3, which shows the variation
Variation of cavity density, Nm, in
of Nm with creep strain for material in the usual
Nimonic 8OA, with strain on logorithmic
solution-treated form and for material having
scales for:
received differing amounts of plastic deforma-
(a) solution treated material - solid circles
tion at room temperature prior to creep. Pre-
(b) prestrained material - squares, 10%,
straining Nimonic 80A has been shown to nucleate
triangles, 3% and open circles, 2%.
grain boundary cavities with a density which is
proportional to the magnitude of the prestrain
(29) and which remains constant throughout the duration of the subsequent creep test. It is
apparent from Figure ] that there is a very significant difference in the accumulation of N m with
strain in the two "materials". For the prestrained material, since N a is constant, then N m is
given from equation I by:-

Nm ~ DH
• l
When growth is constrazned (11), D~ ~ s /3 and therefore a plot of log Nm as a function of log
will have a slope of I/3 as observe~ in Figure 3. For the solution treated material, N a a s giving
~ [ for constrained growth and N~ s 4/3 for unconstrained growth (assuming E ~ t). From the
data presented in Figure 3, it appears that growth is unconstrained up to at least 4% strain.
The comprehensive review of model-based predictions of creep fracture by Cocks and Ashby (15)
concentrates solely on those cases where fracture is controlled by growth of an array of cavities
all nucleated at the same time. When materials exhibit continuous cavity nucleation (which
appears to be the rule rather than the exception), fracture may also still be controlled by
growth, but can additionally be controlled by nucleation rate or by a combination of nucleation
and growth rates. Figure 4(a) illustrates schematically the variation of N a with strain and
suggests an arbitrary definition of the strain to nucleation, s n" The only condition necessary
for g rowth - control ms
" that sf ~> c n, where sf m"s the s tr al"n at fracture, which is likely when
dNa/d~: is large and the stress is high (low growth per unit strain). Figure 4(b) illustrates the
situation under nucleation control which becomes increasingly important at lower stresses (high
growth per unit strain). However, it is Figure 4(c) which experiments suggest is representative
34 CONTINUOUS CAVITY NUCLEATION Vol. 17, No. 1

of the behaviour of many materials. Here sf << e n and Na is no longer a constant at fracture
but depends on cavity growth rate for a given nucleation rate and prediction of failure is much
more complicated (18,20,21,30,31). Regime (c) is most likely to occur in weakly cavitating
materials at low stresses.

NQ

Nma x ...... ~ ~I . . . . . ¥~j.~.


I __.~T-~-

I
I
I
I I

En E~f En Ef Ef E n

(a) (b) (c)


FIG. 4
Illustrating schematically the conditions under continuous cavity nucleation for (a) growth
control; (b) nucleation control and (c) nucleation rate and growth rate control.

Errors in Nucleation Rate Measurements


Measurements of cavity nucleation will usually be in error because of the small scale of the
event and the finite resolving power of instruments and we have to be content with an operational
definition which always includes an element of growth (22). Only when growth to an observable
size is constant in time will the measured nucleation rate be accurate. Unfortunately, because
of interactions between cavities, growth rate is a function of number and some errors in the
measurements reported in the previous section are inevitable. These will become negligible only
when the times for growth to an observable size are small compared with the time between measure-
ments. Cavity growth rates during the initial stages (but at sizes much larger than critical)
and at relatively low densities are given by:-
8~DB~ Kc
- kT (4)

where O is the volume growth rate, ~ the atomic volume, D_ the grain boundary diffusivity,
the grain boundary width, s the applied stress, kT the thermal energy and K is a measure of
cavity interaction given by (32):-

K = I

41n2~_ [I. (~)~[3. (~)21 (5)

where k and 2r are the cavity spacing and diameter respectively. K is a slowly varying
function of 2r/k and over the large range of 2r/X which may be covered in experiments viz, 10 -4
to 2 x 10 -I, K changes from 3 x 10 -2 to 3 x 10 -I and may be taken as having a mean value of 0.1.
If D R is the smallest diameter of cavity which can be detected by a microscope, then the
time required before cavities can be resolved, tR, is given from equation 4 by:-

kTD~
(6)
tR = 4.8 DB w o

Using ~ = 1.7 x I0 -29 m 3, w = 5 x IO-I0 m, k = 1.4 x IO -23 J K -q and D B = IO -5 exp - m2s


(33) then:-
Vol. 17, No. 1 CONTINUOUS CAVITY NUCLEATION 35

~.4 x 1019[~m]ex p [~]I)~

tR (7)
~/T m

100

0
o
(n x
=lpm 10
:.=
> 103.
o
u u

> / e= 10"6 s-1


I0 2-
IIJ
P
"- 1

101 ._.g

& DR=
& '~'E = lo-%-'
100 Id
10 -1_
Practical zero ~ -]

0.4 0.5 0.6


I = = r
0.4 O.5 0.6 0.7 0.8 TITm

T/TIn FIG.6
Showing the strain accumulated before
FIG.5 cavities can be observed in an optical
microscope as a function of T/T m and for
Time required before cavities are
two rates of straining
resolved as a function of T/T using
limits of detection of O.2Bm mand l~m

to is plotted in Fi~ure 5 as a function of T/T using two limits of detection viz D_ = 0.2um
and um, whlch are conservative detectlon llmz~s for scannlng and optlcal metallography
respectively. For illustrative simplicity, typical values for high strength engineering alloys
of o = 2OOMPa and T = 16OOK were used.
m
Testing temperatures for ferritic alloys u~ed in power generation are typically at about
0.ST m and thus, from Figure 5, approximately lO~s are needed for cavities to grow to lum diameter.
With tests lasting several thousand hours, measurements will be taken at intervals in excess of
105s and there will be only small errors in the measurement of nucleation rate. For shorter
term tests, a scanning electron microscope (SEM) becomes necessary. Testing temperatures for
nickel-base superalloys used for aero engine applications are in the range O.6-O.8T and accurate
optical measurements should be achievable even with tests lasting only a few tens o~ hours.
The strain accumulated before cavities can be resolved, £R' ie the apparent nucleation
strain can be calculated approximately by multiplying tR by the average strain rate and is
plotted in Figure 6 as a function of T/T m for two values of strain rate and a l u m limit of
detection. A resolution strain of zero is taken operationally as O.1%. The experiments of
Dyson and McLean (16,18), Cane and Greenwood (23), Lonsdale a~d Flewitt (20) and Needham and
Gladman (21) were all performed at strain rates less than I0-6s -1 and Figure 6 indicates that
£D should be near zero. In agreement with this, only Lonsdale and Flewitt detected a value of
6 ~ ( < 1 % ) but which they interpreted as a nucleation strain. Additionally, an early study by
wlngrove and Taplin (34) using iron (temperature not quoted but probably~ O.ST ) found that 6-
was z~ro at a straln • rate of 10 - 7s - 1 4% at 10 - 6s - 1 and 12o
% at 10 - 5 s- 1 , which mIs in qualitative
H
36 CONTINUOUS CAVITY NUCLEATION Vol. 17, No. 1

agreement with Figure 6.


Ca vit~ Nucle~tip.n
Cavity nucleation has traditionally been ascribed to stress concentrations produced at
obstacles, such as particles, during relaxation of" shear stresses by ~rairl boundary s]iding.
The n~ed for stress amplification can be appreciated by notin~ that s normal stress in excess of
4x10-]E is required before s cavity becomes-ritical by the ag~reffation of vacancies (35) whereas,
even in engineering alloys, applied stresses are rarely greater tham IO-3E. Currently, cavity
nucleation by boundary sliding i~ being trea'ed within the Framework off classical nucleation
theory (35,36, 37). Two important predictions or these model~ are those or (i] a marked stress
dependence on nucleation rate coupled with (it) all cavities being nucleated within a very short
time interval. These predictions contrast sharply with the experimental results detailed shove
and Argon et al (35) have argued, on the basis of experiments by Chang and Grant (38), that this
is because boundary sliding and matrix creep are stochastic processes with individual grains
undergoing large transients and quies]ent periods thus leading to a diffuse poriod of nucleation.
An alternative view, which has some similarities b~t also some significant differences, extends
an idea of Dyson and McLean (18) and assumes that:-
(a) stress concentrations are produced by localised grain de~'ormation rather than by
sliding since one characteristic feature oF engineering alloys is their high area Fraction oF
boundary particles which leads to low levels of stress intensification by sliding
(b) cavities are not produced in the reversible manner mentioned above, but grow from
athermal decohesions which arise from the stress concentrations.

, cavities i cavities It is envisaged that because creep is a


N(r)
stochastic process, an increment of strain,
"--unstable "~l~" stable
I de, produces decohesions with a range of'
I equivalent spherical radii, r, as illustrated
in Figure 7. Cavities nucleate only from those
decohesions that satisfy the condition r > 2y/o n
where y is the surface energy and O n the local
steady state tensile stress. The maximum value
oF o will be in the direction oF the maximum
a.
princlpal stress, oi, and will usually be
similar in magnitude. Cavities will thus be
more numerous on boundaries perpendicular to
the stress axis in uniaxial tension even though
the decohesions are random. If' grain boundary
particles are the sites for cavity nucleation
2~ then the increase in the number of decohesions
r'¢ = - - r
O" 1 of all sizes per increment of strain will be
given by:

~ic. ? aN D = K(Np - ND)aE (8)


Matrix creep is assumed to produce a where N_ is the particle density and N_ the
F . D
distribution of particle/matrix decohesions of decoheslon denslty. Upon integration:-
effective spherical radii r. Only those with N D = Np(1 - exp -KE) (9)
r> 2y/oi will continue to grow by vacancy
absorption to form cavities
and therefore the cavity density N a is:-
N = fNp(1 - exp -K£) (10)
a
where f is a function of y and 01 and controls the fraction of decohesions which become
stable. N as a function of strain exhibits an initial linear part where N a = fKNp s and then
approache~ N a = fNp asymtotically. Anything that decreases y or increases Np (such as trace
elements) or increases 01 (such as triaxial stresses) will increase the slope of N a as a function
of e , in agreement wit experiments, and also increase the value of the asymtote.

Conclusions
I. Continuous, rather than spontaneous, nucleation of cavities is the common characteristic
feature in those metals and engineering alloys investigated.
Vol, 17, No. 1 CONTINUOUS CAVITY NUCLEATION 37

2. Strain rate is the mportant parameter control]im~ n n c ' l , aLior~ r a t ~ a~:d m a t e r i a l < ,4itfU~r
widel~¢ in the numb~.r produced per unit '~train•
3. The applied stres~ state a::d the level o~ trace imK ur[~ier both ~!'~cl czvJty nu,'leatio~!
rate•
4. It is suggested that nucleation is {:o.-ti!..ucu?~and related to ~ [r¢[~ b.~,:su:-e o ~" the
stochastic na~ture o? matr:ix defo~ation i,J polycry~*a] <-.
Re f e g e r ! c e ,~.
I. F. C. Monkman and N. J. Grant, Proc. ASTPi 56, 593, (~956)
2. ~. W. Belluffi and L. L. Setgle, A(:ta Met~ll., _B, 449, (~95T)
~. D. Hull and D. E. Rimmcr, Phil. M~g., '~, (73 (1959)
4. P. W. Davies and B. Wilshire, Shluctur~] Proces~e; in Creep, 2ton a~i] Sleel Ir!s:~i[t~te,
London, (1961)
9- D. McLean, Rep. Progress ]n Physics, ~, (1966)
6. Y. Ishid: and D. McLean, Met. Sci. J., 1, 171, (1967)
7. G. W. Greenwoo?., Phil. Mag., 19, 423, ('~,969)
8. D. A. Woodford, Metal Sci., 3, 50, (1969)
9. B. F. Dyson, M~t. Sci., 10, 349, (1976)
10. W. Beere and ~. ~. Speight, Met. Sc]., 12, 172, (19"28)
11. B. ~ Dyson, Can. Met. qu~rt., ~8 31, -~979)
~2. G. M. Edward an~ M. F. Ashby, Act~ Meteli., 2__7~, 1505, (1'979)
13. B. ~aj and A. K. Ghosh, Met. Trans., 12A, 1291 , (198 ~)
!~+. L. E. Sv(nsson a~d G. L. Dunlop, Int.' .'~et. Reviews , 26 ~o:,
i ~ ~09 (1981)
!~. A. C. F. Cocks and M. F. Ashby, Prog. M~t. Sc~ ~, (19-~2)
16. B. F. Dyscn and D. McLean, Met. Sci., 6, 220, (1972)
I?. H. R. Tipler and B. E. Ho[:klns, Met. Sci., IO, 47, (fl976)
"18. B. F. Dyson and D. McLean, Met. Sci., 11, 3~7, (1977)
19. B. J. Cane, Metal Sci., 13, 287, (1979~--
20. D. Lonsda]e and P. E. J. F!ewit~, Mst. ;3ci. and Engng., [9, 2~7, (1979)
21. N. Needham and T. G!adman, Met. Sci., 14, 64, (!980)
22. I-W Chen at~d A. S. Argo~, Creep and FracturL~ .O~ En~n~ Mat. and Structures. Edited by
B. Wilshire and O. R. J. Ow~en,Pireridge Press <1991)
23. B. J. Csne and G. W. Gre nwood, Met. Sc~., 9, ~,~ (~995)
24. C. S. Smith and L. Guttma:, Trans A.I.!J.M.E~- 19~, ~I, (19u ])
25. R. L. Fullman, Trans. A.i.M.E. 197, 442, (~953--~
26. M. S Loveday and B. F. Dyson, ICI< 3, P 21Z Ed. by K. J. MJl!er and R. F. Smith,Pergamon
(1979)
22. G. B.Thomas and T. B. G-bbons, Metals Technology, 6, 95, (1979)
28. B. F. Dyson and M. S. Loveday, Creep in Structures,-- Proc . or 3rd L.U.T.A.~[. Co,uference,
Leicester, Spri~ger-Verlag,(1981)
29. B. F. Dyson and M. J. Rodgers, Met. Sc~., _8, 261, (~9,4, ~ ~
30. G. W. Gree~wood, Proc. of third Inl. Conf. on 'Strength of Metals a~,d Alloys' Cambridge
(1973)
~I. A. S. Argon, "Recent Advances in Creep and Fracture o~' Engineering Materials", B. Wilshire
and D. ]9. J. Owen, P:iner[@ge Press (1982)
32- M. V. Speight and W. Beere, Met. Sci., 9, 190, (1975)
]3- P. Shewmon, Diffusion in Solids, McGraw-Hill (1963)
34. A. L. Wingrow~ ~nd D. M. R. Tay]in, University of Waterloo Solid Mech. Div. P,eport No.2A
Oct (1969)
3~• A. S. Argon, I-W Chen, and C. W. Lau, "Creep and Fatigue Fracture" Eds. R. ~I. N. P{~lloux
and N. Sto]off, A.I.M.E. ?'.Y (~950)
36. B. l~aj and M. F. Ashby, Acts, Met., 2_.~Z,6.53 (197~)
3?. R. Raj, Acta M e . , 26, 995 (1978)
38. H. C. Chang and N. J. Gr~nt, Trans A I.M.E. J. Metals, ~ 11~5, (1'953)

You might also like