Anda di halaman 1dari 13

Engineering Fracture Mechanics 71 (2004) 587–599

www.elsevier.com/locate/engfracmech

An experimental and theoretical consideration of the effect


of prior creep damage on the heat affected zone
fracture toughness of CrMoV steel
a,*
R. Moskovic , I.J. Lingham a, A.G. Crocker b, G.E. Smith b,
P.E.J. Flewitt a
a
BNFL Magnox Generation, Berkeley Centre, Berkeley GL13 9PB, UK
b
Department of Physics, University of Surrey, Guildford GU2 7XH, UK
Received 30 October 2002; accepted 5 November 2002

Abstract
Material encompassing the heat affected zone has been removed from a region of a manual metal arc weldment in a
CrMoV steel. This zone contained creep cavities on grain boundaries oriented normal to the direction of the maximum
principal stress arising from the stress relief heat treatment. Fracture toughness tests using reconstituted Charpy ge-
ometry specimens with side grooves were undertaken at temperatures of 150 and 350 °C to assess the effect of the creep
cavitation on the measured fracture toughness. The results are compared and discussed with respect to the predictions
of a theoretical two-dimensional model.
Ó 2003 Elsevier Ltd. All rights reserved.

Keywords: Fracture toughness; Creep cavitation damage; CrMoV steel; Heat affected zone; Grain boundaries

1. Introduction

Low alloy ferritic steels may fracture by either initiation of cracks at second phase particles or precip-
itates in the brittle regime or by the formation and linkage of voids at these sites in the ductile regime [1,2].
If a component is subject to a creep loading this can introduce additional defects in the form of cavities at
grain boundaries which could change the mechanical properties of the steel including fracture toughness.
The overall susceptibility to stress relief cracking in the heat affected zone of weldments is controlled by the
conjoint action of (i) the stresses, (ii) the material microstructure and hence mechanical properties and (iii)
the heat treatment cycle [3]. Irrespective of the detailed mechanisms that describe the accumulation of
strain during this process the overall effect is to lead to intergranular cracking as a consequence of the

*
Corresponding author.
E-mail address: rmoskovic2@magnox.co.uk (R. Moskovic).

0013-7944/$ - see front matter Ó 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/S0013-7944(03)00028-6
588 R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599

accumulation and interlinkage of cavities on those grain boundaries oriented normal to the direction of the
maximum principal stress [4].
There is a need to establish if any residual distribution of discrete grain boundary cavitation would have
an influence on the fracture toughness of the material. Little experimental work has been undertaken to
measure the effect of prior creep damage on the subsequent fracture processes, fracture strength or fracture
toughness of ferritic steels over a range of temperature. In the 1940s, Thum and Richard [5,6] measured the
loss of room temperature impact energy in several steels following a period of prior creep deformation. By
the use of a reheat treatment of specimens it has been interpreted that they separated the effect of impurity
segregation to the grain boundaries and therefore boundary fracture energy from the permanent damage
due to grain boundary cavitation [7]. However, in the intervening years little work has been undertaken to
investigate the specific effect of prior creep cavitation damage on the subsequent fracture toughness of
ferritic steels.
In this paper we describe the results of the influence of prior creep cavitation damage, arising from stress
relief heat treatment, within the heat affected zone of CrMoV, BW87A, ferritic steel weldments removed
from service exposed boilers. In Section 2 we describe the predictions of a theoretical model which considers
the contribution of prior creep cavitation damage on grain boundaries to the upper shelf fracture toughness
of ferritic materials. The experimental procedures adopted are described in Section 3 and results for upper
shelf fracture toughness measurements are given in Section 4. These results are discussed with respect to the
predictions of the theoretical model in Section 5.

2. Theory

Although there has been a range of modelling undertaken to consider fracture of materials containing,
for example, distributions of periodic arrays of collinear cracks [8] or indeed distributions of dispersions of
particles in composite materials [9], there has been little consideration of the effect of prior creep damage on
subsequent fracture of ferritic steels or indeed metals and alloys in general. However, recently Smith et al.
[10,11] have developed a two-dimensional model of a polycrystalline material to evaluate the fracture
properties for a range of energies for transgranular and intergranular fracture, and prior creep cavitation
damage. The creep damage in, for example, a low alloy ferritic steel is introduced by identifying appropriate
grain boundaries as decohered. This two-dimensional model has been shown by Smith et al. [12,13] to be a
representative approximation of a three-dimensional polycrystalline material. The computer simulation is
based on two-dimensional models of polycrystalline materials with polygonal grains generated using a
random process. The model generation is performed by a purpose written interactive program in C which
runs under Windows 95.
The model has been used to consider the effect of prior creep cavitation damage on low temperature
fracture and fracture energy in a-iron and on upper shelf ductile crack extensions [14]. In the model, each
grain is assigned randomly an orientation and the parameters such as direction of the applied stress, and the
proportion of prior grain boundary creep damage are addressed interactively. A consideration of ductile
fracture is accommodated by introducing slip ahead of the crack and a ductile fracture criterion. Here the
simulations accommodate the alignment of the grain boundary cavitation damage by increasing the pro-
portion of the fully decohered boundaries oriented within 10° to the normal of the maximum principal
stress axis; in this case a uniaxial stress is applied. It is possible to make an estimate of the changes of the
relative fracture energies as the proportion of decohered, completely cavitated, grain boundaries increases.
The results are presented in Fig. 1. They relate to temperatures where transgranular ductile failure is the
most prevalent fracture mode. However, if the slip planes in a particular grain are badly oriented with
respect to the stress axis, the grain may fail by cleavage fracture or along a grain boundary. The model
offers all three possibilities and chooses the fracture mechanism and path that is easiest. It takes into ac-
R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599 589

Fig. 1. Predictions based upon the results of a two-dimensional model of the relative ductile fracture strengths of heat affected zone
material as a function of pre-existing decohered, cavitated, grain boundaries.

count the stress axis and grain orientation, the relative energies of the three fracture modes and any damage
on contiguous boundaries as the crack propagates from one grain to another. As shown in Fig. 1 the model
indicates that the decoherence and hence the proportion of creep cavitation in the grain boundaries has to
be quite severe before it makes a significant contribution to the fracture energy and thereby the propagation
of ductile fracture in the heat affected zone of weldments. In particular it is only when the material has more
than about 20% creep cavitation damage that cavitated boundaries become linked.
Although it is possible to provide a simple link between the linear elastic fracture toughness and relative
fracture strength for low temperature fracture, the mechanisms involved in ductile crack initiation and
growth at higher temperatures are different [2]. In a ferritic steel containing a significant proportion of non-
metallic inclusions, the micromechanism controlling the process of ductile fracture is damage accumulation
in the form of void nucleation, growth and coalescence. Certainly, for small amounts of prior creep cav-
itation this mechanism of crack formation will dominate. As a consequence, during the operation of this
strain controlled mechanism the amount of plastic deformation will be correlated with the proportion of
pre-existing grain boundary creep cavitation. Therefore, as the fracture energy decreases the fracture
toughness will correspondingly be reduced. This will lead to the change of relative fracture energy with the
proportion of prior grain boundary creep cavitation damage shown in Fig. 1, which provides a simple
indication of the expected trend in the upper shelf fracture toughness for a ferritic steel.

3. Experimental procedure

The BW87A material to be tested was taken from creep cavitated heat affected zones located next to
manual metal arc weld fusion boundaries of a boiler subject to a major repair [15]. The chemical com-
position of the weldment is given in Table 1. Slices approximately 11 mm thick were extracted and re-
constituted by electron beam welding so that each specimen contained a portion of heat affected zone
throughout the thickness. Charpy geometry three point bend fracture toughness specimens of 10  10 mm2

Table 1
Chemical composition (wt%) for BW87A plate steel and weld metal (H1)
C Si Mn S P Ni Cr Mo V Fe
Plate BW87A 0.18 0.35 1.29 0.002 0.01 0.68 0.63 0.29 0.07 Bal
Weld metal (H1) 0.04 0.3 0.88 0.006 0.005 1.6 0.1 0.32 – Bal
590 R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599

Fig. 2. Schematic diagram showing (a) sample of material extracted from the 6/7 circumferential weld and location of samples for
testing and (b) attachment of extensions pieces to 10  15 mm blanks containing cavitated coarse grained heat affected zone.

cross section were manufactured by initially machining 10  15 mm2 blocks from these slices and then
attaching extension pieces by electron beam welding as shown in Fig. 2. Care was taken to ensure that heat
affected zone containing creep cavitation damage was correctly aligned within the finally machined Charpy
geometry test specimens. The electro-discharge machined notch was positioned so that growth of the fa-
tigue pre-crack was located in the cavitated region of the heat affected zone; this had a ratio of pre-crack
length to specimen width of 0.5. Pre-cracked specimens were side grooved by 10% on each side using a
Charpy pre-notch cutter. Fracture toughness tests were carried out in accordance with the ESIS procedure
[16] using the Mayes 100 kN servo-mechanical testing machine. The specimens were enclosed in an envi-
ronmental chamber and heated to a temperature of either 150 or 350 °C and held for a minimum of 30 min.
The temperature was maintained to within 2 °C. Single specimen fracture toughness tests were performed
using the unloading compliance technique to monitor crack lengths. For these Charpy geometry specimens,
displacement was controlled using linear variable differential transducers mounted on the load train. Values
of displacement were corrected for the extraneous compliance of the machine in the post-test analysis. The
elastic displacement rate was set to satisfy the ESIS recommendation of 0.55–2.75 MPa m1=2 per second
[16]. The tests were stopped at, or just after exceeding, the maximum load. After unloading, the specimens
were heat tinted at 300 °C for about 30 min to mark the extent of ductile crack growth, cooled in liquid
nitrogen and broken open in three point bending.
Values of J were evaluated using the relationship:
gU
J¼ ð1Þ
Bn ðW  a0 Þ
R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599 591

where W is the specimen width in mm, a0 is the length of the pre-crack in mm, U is the area under the force
deflection curve in kN mm, corrected for extraneous deflection, Bn the net specimen thickness in mm be-
tween the side grooves and g equals 2 for single edge notch bend specimens. No corrections were made for
crack growth. The value of the fracture toughness parameter KJ , was calculated from J using the rela-
tionship:
 1=2
JE
KJ ¼ ð2Þ
1  m2
where m is the PoissonÕs ratio taken to be 0.3 and E is the YoungÕs modulus of elasticity in GPa given by:
E ¼ 210  0:05T ð3Þ
where T is the temperature in °C. The predicted and measured values of both J and Da, in mm, were
compared with the validity criteria:
ðW  a0 ÞRf BRf
Jmax ¼ and Jmax ¼ ð4Þ
20 20
and
0:2 6 Da 6 0:1ðW  a0 Þ
where Rf is the flow stress (799 MPa).
The fracture surfaces were examined in a JEOL, JSM80, electron microscope operating at 15 kV in the
secondary electron imaging mode. Within the region of brittle cleavage fracture remote from the ductile
fracture, the pre-existing grain boundary creep cavitation was readily identified, Fig. 3. However, to
measure the proportion of creep cavitation within the region of ductile crack extension a method was
developed which has been described in detail elsewhere [17]. In summary, 10 equally spaced areas are se-
lected across the fracture face, adjacent to the fatigue pre-crack, and examined at 200 magnification. A
schematic representation of the fracture surface and the selected analysis positions is given in Fig. 4.
Differentiation of the pre-test creep cavitation damage from ductile dimpling was made by examining the
fracture faces at the magnification of 200. This clearly shows the ductile dimpling, typically greater than 3
lm in diameter, whereas creep cavitation damage, typically less than 1 lm in diameter, is not resolved so

Fig. 3. A secondary electron image of a specimen fractured in liquid nitrogen showing cavities on a grain boundary together with
associated cleavage fracture (120-2).
592 R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599

Fig. 4. Schematic diagram showing end view of a fracture face for charpy geometry fracture toughness specimen and the sampling
positions used to measure proportion of prior cavitation.

that the intergranular fracture surfaces appear featureless. The identification of different fracture mor-
phologies was confirmed by a further examination of the grain boundary features in the scanning electron
microscope, but at a higher magnification of 2000 to resolve creep cavitation.
Here, the cavitation corresponds to individual cavities of 61 lm diameter which have been formed at
the prior austenite grain boundaries by creep deformation [4]. They are aligned on those grain boundaries
that are approximately normal to the direction of the maximum principal stress. Their progressive nu-
cleation and growth leads to decoherence of a given grain boundary and therefore in this case the definition
of cavitation damage is the proportion of boundaries containing cavitation compared with the total
fracture surface area. The areas were recorded for subsequent image analysis. Each area of potential creep
cavitation damage recorded from the fractography was assessed at the higher magnification, and the areas
containing creep cavitation marked. The extent of creep cavitation damage was therefore measured as the
projected area fraction of the total fracture surface. The error associated with these measurements is 10%
of the value. This process was repeated for the 10 selected positions and the average percentage damage for
the sample calculated. An example of an area selected for analysis, is shown in Fig. 5, which gave 32%
cavitation damage.

Fig. 5. Typical area used for quantitative measurement of creep cavitation annotated to identify the cavitated regions (1182-2).
R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599 593

4. Results

4.1. Microstructure and hardness

The microstructure of the BW87A parent plates consists of bands of polygonal ferrite and tempered
bainite. The hardness of the parent plates, in the as-received condition, is approximately 213 Hv10 . The weld
metal consists of beads comprising coarse columnar and grain refined microstructures described previously
[18]. The coarse grains in the weld metal are approximately 100 lm wide and 400 lm long as measured by
mean linear intercept. The hardness of H1 weld metal is approximately 236 Hv10 for the as-received ma-
terial. The heat affected zone of the weld was approximately 2–3 mm wide. The microstructure varied with
distance from the weld fusion boundary and, due to the intersection of weld cusps, periodically throughout
the depth of the weld. The heat affected zone immediately adjacent to the fusion boundary contains an
unaltered grain-coarsened microstructure, supercritically reheated grain refined microstructure and inter-
critically reheated grain coarsened regions. The microstructure of the unaltered grain coarsened regions
comprises tempered martensite and bainite with a prior austenite grain size of approximately 100 lm; the
hardness is typically 323 Hv10 . The supercritically reheated heat affected zone comprises bainite with a prior
austenite grain size of about 20 lm; the hardness is about 248 Hv10 . The microstructure of the intercritically
reheated region is a mixture of coarse grained martensite and bainite, decorated with fine grained ferrite
and carbides. The grain size is in the range from 10 to 100 lm; the hardness is typically 244 Hv10 . In
addition to these features, with increasing distance from the weld fusion boundaries, the prior austenite
grain size reduces to about 30 lm with an approximate hardness of 256 Hv10 ; this region is referred to as
fine-grained heat affected zone.

4.2. Fracture toughness

The results from the fracture toughness testing and assessment of creep cavitation damage are presented
in Table 2. The table includes the final values of the crack extension measured on the fracture surface of
each specimen and also the values predicted by the unloading compliance. The variability in measured
crack extension across the specimen thickness was, in most cases, outside the ESIS limits for uniform crack
extension [16].

Table 2
Fracture toughness test results obtained on cavitated specimens (values predicted are from unloading compliance data)
Temp Test code Crack extension Da/mm % Cavitation J/N mm1 Fracture toughness MPa m1=2
(°C) Measured Predicted damage Initiation J0:2 Final Jend Initiation KJ 0:2 Final KJ end
150 1201-1 0.00 – 47 – 18 – 63
1201-2 0.57 0.57 14 46 84 101 136
1201-3 0.40 0.36 20 41 52 95 107
1201-4 0.27 0.26 3 174 188 196 204
1201-6 0.42 0.61 10 85 214 137 218
1201-9 0.42 0.54 1 70 166 125 192
1201-10 0.30 0.40 14 111 177 157 198
350 1182-2 0.69 1.00 32 20 63 65 115
1182-3 0.28 0.28 6 77 104 127 148
1182-4 0.31 – 13 45 59 97 111
1182-5 0.40 0.35 10 54 75 106 125
594 R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599

Fig. 6. Examples of force–displacement plots obtained from the fracture toughness tests conducted at: (a) 150 °C and (b) 350 °C. The
appearance of these curves depends upon the proportion of prior creep cavitation present in the test specimen. K values are those
obtained at termination of the tests.

During the fracture toughness tests, non-linear force–displacement records, typical of ductile crack
growth in the upper shelf temperature region, were observed at temperatures of 150 and 350 °C, Fig. 6(a)
and (b). The tests were terminated at, or just after, a fall in the force below the maximum. A large rapid
reduction in the force on the force–deflection record at the end of the trace indicated plastic instability. Fig.
7 shows the measured initiation fracture toughness, K0:2 , in undamaged material compared with the scatter
in K0:2 for creep cavitation damaged material. Initiation fracture toughness values measured at a temper-
ature of 150 °C were higher than those measured at 350 °C. Initiation fracture toughness values obtained at
150 °C exceeds 110 MPa m1=2 up to 10% of cavitation damage and falls slightly below this value when
cavitation damage is 20%. At 350 °C, the initiation fracture toughness is slightly below 110 MPa m1=2
when this damage is 13%. As the amount of grain boundary cavitation damage increases further there is a
progressive decrease in the initiation fracture toughness, but even at 40% damage the residual fracture
toughness is approximately 60 MPa m1=2 .
R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599 595

Fig. 7. The initiation fracture toughness, K0:2 , from previously measured BW87A heat affected zone material (19) that is uncavitated
and with increasing proportion of prior creep cavitation damage.

The appearance of the force–deflection curves, Fig. 6(a) and (b), depends on both the amount of creep
cavitation damage and the type of microstructure sampled by the crack tip. For creep cavitation damage up
to about 10–13%, the force approximately levels off during the test. When the creep cavitation damage is in
the range from 14% to 32% the maximum force is reached quite early during the test and there is a no-
ticeable decrease in the force during the test. For the specimen in which the cavitation damage is 47%,
instability was observed at the maximum load and examination of the broken specimen showed that no
stable ductile crack growth had occurred. Tests 1182-4, 1201-3, 1201-1 and 1201-2 include a small pro-
portion, 610%, of crack growth in the weld metal. In previous work [19], it has been shown that crack
fronts can move into adjacent lower strength materials. It is argued, therefore, that whilst the tests may not
have been conducted in wholly creep cavitated heat affected zone material the results will be conservative
because resistance to ductile crack growth is lower in the manual metal arc weld metal than in the heat
affected zone.
Fig. 8 shows secondary electron images of the fracture surface of a fracture toughness specimen tested at
350 °C and containing 32% of prior creep cavitation damage. The overall fracture surface, Fig. 8(a), is a
mixture of ductile and intergranular fracture. The latter corresponds to prior austenite grain boundaries
containing a distribution of creep cavitation damage, Fig. 8(c). There are within the regions of trans-
granular ductile fracture Fig. 8(d), small regions 61%, of intergranular ductile fracture, Fig. 8(b). This
mixture of fracture in the initiation region of the test specimen was characteristic of tests conducted at both
150 and 350 °C; only the proportion of intergranular creep cavitated boundaries changed.

5. Discussion

The blocks of material extracted from the CrMoV steel weldment were reconstituted into three point
bend specimens which were used to measure the fracture toughness of service exposed heat affected zone
material for various amounts of prior creep cavitation damage at grain boundaries. Due to the overall
narrow width of the real weldment heat affected zone, 3 mm, three point bend Charpy size geometry
specimens are, in general, the largest that can be produced for fracture toughness testing. It is recognised
596 R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599

Fig. 8. A secondary electron image of the fracture surface for a specimen tested at 350 °C with 32% cavitation in the region of im-
mediate crack extension showing (a) a mixture of prior cavitated intergranular facets and ductile fracture, (b) intergranular ductile
facets, (c) creep cavitated facets and (d) transgranular ductile fracture (1182-2).

that the presence of grain boundary creep cavities reduces the tensile load bearing cross section area of the
specimens [20]. This imposes a limit on cavitation damage that can be evaluated for a particular specimen
size. Large amounts of cavitation damage may introduce plastic instability and this would preclude J
controlled fracture processes at the crack tip. The question that needs to be addressed is whether small
amounts of creep cavitation damage on prior austenite grain boundaries, arising from the stress relief heat
treatment, bring about noticeable changes in fracture toughness properties.
Initiation fracture toughness values, K0:2 , obtained on cavitated specimens for two different test tem-
peratures, presented in Fig. 7 as a function of cavitation damage, are highly scattered. The data associated
with the highest percentages of cavitation damage tend to fall towards the lower limit of the scatter range.
R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599 597

Fig. 9. Normalised initiation fracture toughness (K0:2 cavitated/K0:2 uncavitated) as a function of the proportion of prior cavitation
damage. The solid line is the theoretical prediction.

The fracture toughness data base for cavitated material is relatively small, hence it is difficult to judge the
degree of cavitation damage which may lead to a true reduction in the initiation fracture toughness. Al-
though the control values of the initiation fracture toughness for this particular material in the uncavitated
condition are not known for the individual specimens, it is helpful to compare the data measured for the
cavitated condition with a much larger data base obtained on uncavitated material. The data are compared
in Fig. 9 in which the ratios of the measured to the mean, for uncavitated material, initiation fracture
toughness are plotted against the percentage of cavitation on the fracture surface. Here the spread of
fracture toughness for undamaged material is between approximately 0.74 and 1.26. It is recognised that
there is inherent scatter in the measured fracture toughness data of materials in general due to (i) variability
from material to material of the same nominal specification, (ii) variability in local microstructure and (iii)
that arising from the testing method. In this case (ii) is particularly important due to the inherent variability
of the microstructure within the heat affected zone of the weldment and (iii) because of the need to locate
the pre-crack within the narrow heat affected zone. However, up to approximately 14% of cavitation
damage the results obtained on cavitated material lie within this spread.
Typical examples of force–displacement curves for different levels of cavitation damage, Fig. 6(a) and
(b), show that at low levels up to 13%, there is no noticeable reduction in the force during the test. There is
no evidence of plastic instability brought about by a reduction in the cross section area of the specimen
jointly by the presence of creep cavities and ductile crack growth during the test. Despite this, the fracture
toughness values measured at this level of cavitation are reasonable in relation to the reference lower bound
value of 110 MPa m1=2 for this material. At higher levels of cavitation, there is a noticeable reduction in
both initiation fracture toughness and force during the test, indicating that plastic instability is becoming
the dominant mode of failure and that larger test specimens would be required to provide confidence that
values representative of crack tip controlled cracking have been measured. In the specimen associated with
47% cavitation damage, a rapid plastic instability was observed without any evidence of ductile crack
growth at the crack tip. As the proportion of creep cavitation damage increases then, as indicated in Fig. 9,
there is a progressive change in the microstructure from discrete cavitated boundaries, to discrete plus
interlinked boundaries. This commences at about 15% damage with the interlinkage developing until mi-
crocracks and finally macrocracks are formed. It is the latter interlinkage that limits the relative fracture
598 R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599

strength in Fig. 1 to about 50% in the simple 2D-model. Moreover, the trend in the relative fracture
strength, Fig. 1 and hence fracture toughness is consistent with the measured values given in Fig. 9. This
lends support to the use of lower bound fracture toughness values, obtained from the undamaged material,
for material containing up to about 15% cavitation damage when undertaking an assessment of plant.
Overall similar trends have been predicted for the role of prior creep cavitation damage to the lower shelf
fracture toughness of ferritic steels [2].
With regard to the more general issue of fracture toughness values for use in the assessment of plant
weldment heat affected zones which may contain creep cavities, the overall concern is that fracture
toughness could be impaired either because cavitation reduces the load bearing cross section or because
cavitated material offers a lower resistance to ductile crack growth than material free of cavities. Since the
wall thickness for the shell of structures and components are usually large and for example in the case of a
butt weld [15,18] the depth of sound material is unlikely to be less than half of the wall thickness, the
fracture toughness data obtained on specimens that exhibited plastic collapse can be considered to be ir-
relevant for structural integrity assessments of macrocracks. In some circumstances it could be difficult to
translate cavitation damage observed metallographically on plant into an equivalent percentage of cavi-
tation density observed on fracture surfaces of laboratory specimens. Nevertheless, the present work
demonstrates that for material containing 615% cavitation damage, upper shelf values measured on un-
damaged material provide an adequate measure of fracture toughness properties certainly for steel BW87A
heat affected zones and indeed, for ferritic steels in general to provide conservative assessments.

6. Conclusions

Up to 15% prior creep cavitation damage arising from a stress relief heat treatment on a CrMoV steel
weldment, the initiation fracture toughness values are within the scatter of data obtained for uncavitated
material. At higher levels of cavitation, initiation fracture toughness values are reduced due to the inter-
vention of the prior cavitation damage leading to plastic instability during the test. However, even at 40%
cavitation damage the material retains a measurable fracture toughness.

Acknowledgements

This paper is published with the permission of the Head of Reactor Services Organisation, BNFL
Magnox Generation. Professor A.G. Crocker and Dr. G. Smith acknowledge the support of EPSRC via an
ERCOS award.

References

[1] Ritchie RO, Knott JF, Rice JR. J Mech Phys Solids 1977;21:395.
[2] Ritchie RO, Thompson AW. Met Trans 1985;16A:233.
[3] Dhooge A, Vinckier V. Inst Press Vess Pip 1987;27:239.
[4] Lonsdale D, Flewitt PEJ. Proc Roy Soc Lond A 1981;373:491.
[5] Thum A, Richard K. Arch Eisenhutt 1941;15:33 [ASM English Translation No. 1131, Brutcher H].
[6] Thum A, Richard K. Arch Eisenhutt 1949;20:229 [ASM English Translation No. 2450].
[7] Wilkinson DS, Abiko K, Thyagenajon N, Pope D. Met Trans 1980;11A:1827.
[8] Galybin AN. Engng Fract Mech 1998;59:281.
[9] Kim B, Watanabe M, Ednoki E, Kishi T. Engng Fract Mech 1998;59:289.
R. Moskovic et al. / Engineering Fracture Mechanics 71 (2004) 587–599 599

[10] Crocker AG, Smith GE, Flewitt PEJ, Moskovic R. In: Lejcek P, Paidar V, editors. Intergranular and interphase boundaries in
materials. Switzerland: Trans Tech Pub; 1999. p. 673.
[11] Smith GE, Flewitt PEJ, Crocker AG, Moskovic R. The lifetime of plant structures and components; evaluation, design, extension
and management. In: Edwards JH, editor. Proceedings of Engineering Structural Integrity Assessment. UK-EMAS; 2001. p. 257.
[12] Smith GE, Crocker AG, Flewitt PEJ, Moskovic R. In: Rossmanith PR, editor. Damage and failure at interfaces. Rotterdam:
Balkema; 1997. p. 229.
[13] Smith GE, Flewitt PEJ, Crocker AG, Moskovic R. In: Zielinski A, Desjardins D, Labanowski J, Cwiek J, editors. Environmental
degradation of engineering materials, EDEM99. Gdansk: Gdansk Scientific Society; 1999. p. 146.
[14] Smith GE, Crocker AG, Moskovic R, Flewitt PEJ. Phil Mag, A 82:3443–53.
[15] Flewitt PEJ, Taylor EG, Mitchell LA. Ingenia 2000:(August);6.
[16] ESIS procedure for determining the fracture toughness of materials. European Structural Integrity Society, Document ESIS, 2,
1992.
[17] Lingham IJ. A procedure for measuring the proportion of cavitation damage on fracture surfaces produced by fracture toughness
testing. BNFL Magnox Generation Report M/TE/SXA/EAN/0057/00, 2000.
[18] Exworthy LF, Ellis BJC, Flewitt PEJ. Boiler shell weld repair: Sizewell A Nuclear Power Station, I. Mech Eng, Seminar Pub.
London: Professional Eng Publishing 1999. p. 9.
[19] Lingham IJ, Moskovic R, Priest M. BNFL Magnox Report M/TE/SXA/REP/0099/99, 2000.
[20] Perry AJ. J Mater Sci 1974;9:1016.

Anda mungkin juga menyukai