You are on page 1of 9

Effect of crack length-to-width ratio on crack resistance of high Cr-ODS

steels at high temperature for fuel cladding application


R. Chaouadi

, M. Ramesh, S. Gavrilov
SCKCEN, Boeretang 200, 2400 Mol, Belgium
a r t i c l e i n f o
Article history:
Available online 21 February 2013
a b s t r a c t
Oxide dispersion strengthened (ODS) steels with high Cr-content are extensively investigated in Europe,
Japan and United States by the nuclear materials community for application to both advanced ssion
reactors and fusion systems. In comparison to standard high Cr-steels, the expected operation tempera-
ture range can be extended to 650 C or more because of their improved creep resistance. However, their
crack resistance behavior in the high temperature range was less investigated.
The aim of the present paper is to provide some insight on their fracture behavior at high temperature
and different crack congurations, in particular shallow crack. Crack resistance measurements were
performed on a 12%Cr-ODS steel using compact tension specimens at 650 C considering both shallow
and deep crack congurations. Finite element calculations were performed on a typical fuel cladding tube
geometry to assess the performances in terms of crack resistance. It is found that the temperature
gradient across the wall should be maintained low enough to avoid cracking. After irradiation in corrosive
environment, the boundary conditions might be further affected limiting therefore the lifetime of ODS
cladding.
2013 Elsevier B.V. All rights reserved.
1. Introduction
Among the candidate structural materials for fusion and ssion
advanced nuclear systems, oxide dispersion strengthened (ODS)
steels are receiving considerable attention in Europe, Japan and
in the US [16]. In particular, high Cr-ODS steels offer a number
of properties that are attractive for application at high temperature
under intense irradiation [710]. However, the advantages of such
steels such as their high strength at high temperature, their high
creep resistance and their relatively better irradiation resistance
are counterbalanced by their low fracture toughness leading to
high ductile-to-brittle transition temperature (DBTT) [6,1116]
and their low crack resistance at high temperature [17,18].
Recently, signicant improvement of the DBTT was reached by
rening the oxide dispersion and better control of impurities and
the fabrication process [1922]. Actually, the low fracture tough-
ness at high temperature is more critical as these steels are
supposed to operate in the high temperature range where other
available non-ODS steels cannot be used. As a matter of fact, few
preliminary results were indicating signicantly low toughness at
high temperature [17,18]. It should be mentioned that although
the specimen orientation is not always reported, most of the
published data are taken with the crack propagation occurring
perpendicular to the extrusion direction, namely LT. As it will
be seen later, this is not the weakest orientation and it is expected
that the mechanical properties will be signicantly poor when the
crack is orientated parallel to the extrusion direction.
Within the EURATOM FP7 GETMAT project, three ODS steels
were selected for investigation with 9%Cr, 12%Cr and 14%Cr con-
tent. The fracture toughness of the 12%Cr- and 14%Cr-ODS steels
at high temperature was found signicantly low than commercial
9%Cr steels [23]. The 9%Cr-ODS Eurofer steel reported in [18]
exhibited the same tendency.
From an application point of view, the measured fracture
toughness of high Cr ODS steels exclude them to be used as struc-
tural materials of nuclear systems. This means that the fracture
toughness of high Cr ODS steels is too low to be used in thick com-
ponents. The question remains whether they can be used for thin
walled components such as fuel cladding [2,2431]. However, rep-
resentative cracks in fuel cladding are shallow cracks that can be
found on the inner or outer surface. Indeed, during the fabrication
process, such penny-shape cracks can appear and it is interesting
to assess their impact on the integrity of the fuel cladding during
operation.
In the following, we will experimentally measure the fracture
resistance of the 12%Cr-ODS steel at 650 C in two crack congura-
tions, deep (according to prevailing standards) and shallow crack,
respectively. The tests will be performed in two specimen orienta-
tions, LT and TL. Then, nite element calculations will be
performed on a typical thin wall cladding tube to estimate the
actual loading during normal and accidental conditions to assess
0022-3115/$ - see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jnucmat.2013.02.017

Corresponding author.
E-mail address: rachid.chaouadi@sckcen.be (R. Chaouadi).
Journal of Nuclear Materials 442 (2013) 425433
Contents lists available at SciVerse ScienceDirect
Journal of Nuclear Materials
j our nal homepage: www. el sevi er. com/ l ocat e/ j nucmat
the integrity of the fuel cladding using the measured fracture
toughness data.
2. Materials and experimental
Within the GETMAT project, several 12%Cr-ODS steel plates
(605 33 6.5 mm) fabricated by KOBELCO Research Institute in
Japan were distributed among the participants for multi-disciplin-
ary investigations. The chemical composition is given in Table 1.
The nal heat treatment consisted of annealing at 1150 C/1 h.
The plate used is this work is labeled K11. Because of the size
limitations of the available plate, the small compact tension (CT)
geometry was selected. CT specimens were taken in two orienta-
tions, LT and TL. The specimen size corresponds to
1
=
4
in.-CT
specimens with a thickness B = 6.25 mm.
The
1
=
4
in.-CT specimens were precracked at room temperature
up to a crack length-to-width ratio of about 0.5 and further 20%-
side grooved to ensure a uniform crack front propagation and to
promote plane strain conditions. The stress intensity factor during
fatigue precracking was kept below about 18 MPa
p
m at any time.
For the deep crack conguration, the specimens in their stan-
dard conguration, for example E1820 [32], were used. For the
shallow crack conguration, the loading holes were located such
as a/W is between 0.1 and 0.2 (see Fig. 1). Depending on the used
holes for loading, either deep or shallow crack conguration can be
experienced. For illustration, in Fig. 2, the loading corresponds to
the shallow crack conguration. Tests were performed at 650 C
under quasi-static loading rate of about 1 kJ/m
2
s and the J
R
curve
is determined using the energy normalization technique [33,34].
At this high temperature, crack length monitoring with a clip gage
and using the unloading compliance method cannot be used. The
details of the procedure can be found in [33,34]. The reasons of
selecting the energy normalization method rather than a standard-
ized method such as the normalization data reduction (NDR)
technique [32] is its better accuracy and independence on experi-
mental bias. Indeed, the method relies essentially on the specimen
dimensions, the loaddisplacement record and the measured
initial and nal crack lengths. An example is given in Fig. 3 where
the energy normalization method is compared to three standards
methods, namely the unloading compliance, the potential drop
and the normalization data reduction.
It should be noted that at 650 C, the tensile properties of the
12%Cr-ODS steel exhibit a nearly perfectly plastic material with
signicant ductility anisotropy (see Fig. 4). As a matter of fact, this
anisotropy was reported by several authors including both tensile
as well as fracture properties [3540]. All latter authors reported a
typical extruded microstructure with the grain elongated along the
extrusion direction.
3. Testing and analysis
For the deep crack conguration, the tests were duplicated to
check material homogeneity and no signicant difference was
found [23]. Here, only a single test is reported as the aim is to
assess the shallow versus deep conguration. The specimen
dimensions and testing conditions are given in Table 2.
The J-integral is calculated according the ASTM E1820
formulation [32]:
J J
el
J
pl

K
2
1 V
2

E

g U
pl
B
n
W a
0

where K is the stress intensity factor, E is the Youngs modulus, v


is the Poissons ratio, W is the specimen width, B
n
is the specimen
net thickness, a
0
is the initial crack length and U
pl
is the area
under the loadplastic displacement curve.Moreover, g 2:0
0:5221 a
0
=W.
For crack resistance curve, the J-integral values are corrected for
crack growth according to:
J
i

K
2
i
1 V
2

E
J
pli1

g
i1
W a
i1

U
pli
U
pli1
B
n

1 c
i1
a
i
a
i1
W a
i1

where g
i1
2:0 0:5221 a
i1
=W and c
i1
1:0 0:76
1 a
i1
=W.
However, the g and c factors given above are only valid for deep
crack conguration, namely, CT specimens with a crack-to-width
ratio, a/W, close to 0.5. In the case of shallow cracks, these formulas
might not be valid. Therefore, 3D nite element calculations were
conducted by Ramesh [41] on a 1/2T-CT specimen with a crack-to-
with ratio varying between 0.1 and 0.8 to derive the g-factor. Note
that for crack extension levels under consideration, the crack
growth correction accounted by the c-factor remains close to 1.
Therefore, only the g-factor was determined as a function of crack
length-to-width ratio.
It is important to mention that, for these nite element simula-
tions, the compact tension geometry was adapted to avoid large
deformation at the pin holes. Therefore, for shallow cracks, typi-
cally a/W < 0.35, the geometry was modied by increasing the
material thickness around the loading pin holes and consequently
avoiding excessive deformation of the pin holes. Note that modi-
cation holds only for the nite element simulation because a
tougher material was used in those simulations. In the case of
ODS steels, the crack resistance is so low that no excessive defor-
mation occurred at the loading pin holes.
The nite element results can be summarized as:
g 2:72 0:52a=W 3:751 a=W
6
It should be mentioned that by moving the pin hole position to re-
duce the crack length-to-width ratio, the loading changes from pre-
dominantly bending for deep crack to predominantly tensile for
short crack. Therefore, it is interesting to compare the nite element
results obtained by Ramesh [41] to results obtained on single edge
crack tension, SE(T), specimens obtained by [42,43]. The compari-
son is shown in Fig. 5 which indicates a consistent trend, supporting
the use of the above derived equation for shallow crack congura-
tion while the ASTM formula is applied for the deep crack
conguration.
4. Test results
Four
1
=
4
in.-CT specimens were tested at 650 C in two crack con-
gurations (shallow crack versus deep crack) in two orientations
(LT versus TL). The specimen dimensions and testing conditions
are given in Table 2.
As it can be seen, a crack extension of 1 mm was reached by
applying a load equivalent to a J-integral level of only 3.5 kJ/m
2
for the deep crack CT in the TL orientation. This is signicantly
low and lower than the value in the LT orientation in which
55 kJ/m
2
were required to extend the crack by 0.77 mm. Similarly,
for the shallow crack conguration, almost 3 mm crack extension
were reached with an equivalent load of about 35 kJ/m
2
in the
Table 1
Chemical composition (in wt%).
Cr W Ti Si Ni Mn Y
2
O
3
11.59 1.87 0.22 0.10 <0.01 <0.01 0.23
426 R. Chaouadi et al. / Journal of Nuclear Materials 442 (2013) 425433
TL orientation while only 1 mm was achieved with a load corre-
sponding to 184 kJ/m
2
in the LT orientation.
For each specimen, the crack resistance curve is determined
from which the engineering crack initiation toughness established
at 0.2 mm crack extension, J
0.2mm
, the tearing modulus dened as
T
dJ
da

E
1v
2
r
2
y
and the elasticplastic stress intensity factor
K
J
c

J
0:2mm
E
1v
2

q
are derived. The results are shown in Table 3 for all
specimens. As expected, the stress intensity factor K
J
c
of the deep
crack conguration is only 21 MPa
p
m in the TL orientation. This
is the reason why such a material cannot be used as a structural
material. For the shallow crack conguration in the TL orienta-
tion, the stress intensity factor is enhanced to 46 MPa
p
m. The
crack resistance curves, J-Da, of the various specimens are shown
in Fig. 6 while the characteristic parameters are depicted in Table 3.
As it can be seen, this ODS steel exhibits a signicant crack
resistance anisotropy and in particular, a signicantly weak crack
resistance in the extrusion direction. In terms of stress intensity
factor, the results are shown in Fig. 7. The low fracture resistance
in the TL orientation is very critical as longitudinal cracks can
be generated during the extrusion process. Before investigating
the consequences of such low crack resistance in a component, it
is interesting to examine the cracking behavior and crack surface
using microscopic techniques.
One of the characteristics of the 12%Cr-ODS steel cracking can
be evidenced by optical microscope observation of the crack
Fig. 1. CT-specimen conguration allowing deep (a/W = 0.5) and shallow crack (a/W = 0.1) congurations.
Fig. 2. Shallow crack CT-specimen testing.
Fig. 3. Typical example of crack resistance curve determination procedures
according to various methods as applied to a 508 Cl.3 steel: unloading compliance,
potential drop, normalization data reduction and energy normalization techniques.
Fig. 4. Tensile properties of the 12%Cr-ODS steel at 650 C in the L- and T-
orientation [23].
R. Chaouadi et al. / Journal of Nuclear Materials 442 (2013) 425433 427
propagation ahead of the fatigue precrack. The energy required to
initiate and propagate the crack is so low that the crack extension
occurs with very small crack tip opening displacement (see Fig. 8).
Actually, the nal crack propagation (Fig. 8c) is very similar to a fa-
tigue precrack (Fig. 8a) for which the applied stress intensity factor
does not exceed 18 MPa
p
m. Similar observations were reported by
Alinger et al. [35]. Moreover, the crack propagation occurs on the
same plane as the fatigue precrack without deviation or zigzagging
as usually observed on more ductile materials. This clearly indi-
cates the very low crack resistance of this ODS steel.
The effect of specimen orientation on the crack resistance was
also examined using scanning electron microscopy (SEM) observa-
tions. For LT specimens, we can notice secondary cracks appear-
ing on the fracture surface but that are perpendicularly oriented
with respect to the crack plane (see Fig. 9a). Actually, while the test
is such as to favor crack extension in the T-direction, theses cracks
propagate in the L-direction (extrusion direction). On the TL spec-
imens in which crack propagates in the L-direction, such secondary
cracks are not observed (see Fig. 9b). This clearly demonstrates the
low crack resistance of the 12%Cr-ODS steel when the crack is
propagating in the extrusion direction.
Large SEM magnication of the fracture surface in the region of
interest shows that in the weakest orientation, namely TL, the
grain structure is elongated along the extrusion direction (see
Fig. 10). Although not unambiguously observed, it is likely that
the fracture is intergranular resulting from grain boundary de-
cohesion. We performed stereo SEM micrographic observations
that show that while in the LT orientation, the fracture surface
exhibits a typical dimple structure with a tortuous fracture
landscape, the TL orientation shows a much more atter surface
with textured structure and intergranular aspect (see Fig. 10). Note
that in the high temperature regime, intergranular cracking was
also reported by some authors [4447].
The small oxide particles and other second phase particles
(inclusions) are denitely playing an important role in offering
preferential sites for void nucleation at the interface particle
matrix. The deformation incompatibility between the hard oxide
particles and the soft matrix favors easy void nucleation and their
coalescence is facilitated by the high density of oxide particles. As a
result, very limited energy is required to initiate and further prop-
agate a crack. Due to the extrusion process, it is possible that these
second phase particles (oxides, inclusions) are aligned in colonies
oriented in the extrusion direction that signicantly decrease the
crack resistance. According to Alinger [35], the low toughness
when the crack is parallel to the L-direction is due to the presence
of a large volume fraction of alumina stringers resulting from
impurity contamination. Further microstructural investigations
will be required to better identify all the features that are
responsible of the low crack resistance and anisotropy of the ODS
steel. In any case, the fracture properties of ODS steels, in particular
in the weakest direction, need to be improved to a level that do not
jeopardize their usability.
5. Finite element simulation of cracked thin wall cladding tube
One of the potential application of high Cr-ODS steels is for the
fuel cladding. For high burnup fuel cladding, ODS steels are consid-
ered as most promising [30].
In Japan, a number of authors [27,30,47] reported successful
fabrication of high Cr ODS cladding tubes. The creep rupture
strength of a 9%Cr-ODS steel with 0.3%Y
2
O
3
at 700 C was signi-
cantly increased. A number of fuel pins made of high Cr-ODS are
under irradiation in the BOR-60 reactor [30,31] at high tempera-
ture, 650 and 700 C with promising results, i.e., no fuel pin failure
observed, were obtained after irradiation to 21 DPa. Results at
higher neutron dose are not reported yet.
Table 2
Specimen dimensions and testing conditions of the 12%Cr-ODS steel at 650 C.
Specimen Orientation Width
(mm)
Thickness
(mm)
Net thickness
(mm)
Crack length
(mm)
Crack length to width
ratio
Crack extension
(mm)
J at end of test (kJ/
m
2
)
W B B
n
a
0
a
0
/W Da
end
J
end
K11-32 TL 7.66 6.17 4.86 1.08 0.14 2.91 33.0
K11-27 LT 7.52 6.22 4.87 1.00 0.13 0.98 184.2
K11-24 TL 12.51 6.25 4.97 6.66 0.53 1.05 3.5
K11-20 LT 12.50 6.26 5.12 6.61 0.53 0.77 54.8
Fig. 5. Effect of crack conguration (shallow to deep) on the plastic g factor of
specimens submitted to tensile, bending and combined tensile/bending loading. All
symbols are from nite element calculations.
Table 3
Crack resistance test results of the 12%Cr-ODS steel at 650 C.
Specimen Orientation Crack length to
width ratio ()
J at onset of crack
extension (kJ/m
2
)
Tearing resistance
(kJ/m
2
p
mm)
J at 0.2 mm crack
extension (kJ/m
2
)
Tearing
modulus ()
Stress intensity factor
(MPa
p
m)
a0/W J
i
J
t
J0.2mm T
K11-32 TL 0.14 3.6 17.3 11.3 32 46
K11-27 LT 0.13 13.1 172.5 90.2 315 129
K11-24 TL 0.53 1.5 2.0 2.35 3.7 21
K11-20 LT 0.53 9.3 51.9 32.5 94.9 77
428 R. Chaouadi et al. / Journal of Nuclear Materials 442 (2013) 425433
In operation, the fuel cladding tube is submitted, among others,
to stresses generated from the pressure difference and tempera-
ture gradients. Given the low crack resistance of the material in
the extrusion direction, the question that arises is whether small
penny-shape cracks located at the inner or outer surface may jeop-
ardize the integrity of the fuel cladding tube. Therefore, three
dimensional 3D nite element calculations were performed for a
quarter cladding tube using ANSYS 14.0 package with different
load cases (see Fig. 11). The thin wall tube has an external diameter
of 6.50 mm and a thickness of 0.45 mm. The calculations were
performed assuming a cladding tube containing a longitudinal
semi-elliptical crack located at the inner and outer surface of the
tube, respectively. The penny shape crack has a crack length-to-
width ratio a/W of 0.135 while the aspect ratio c/a of 1.2. The
meshing model consists of 30,000 elements of SOLID 90 and SOLID
186 type for respectively thermal and structural calculations, with
high mesh renement (10,000 elements) near the crack region (see
Fig. 12). Combined steady state thermal and structural calculations
were performed on one quarter of the sample geometry due to
inherent symmetry conditions of the tube. The model is solved
with the Sparse direct solver using uniform reduced integration
method.
We considered both an internal and external crack oriented in
the extrusion direction. Such small defects are not hypothetic but
might appear during the cladding tube manufacture process. The
temperature at the crack tip is xed at 650 C. The stressstrain
ow curves that were used were derived from the tensile tests re-
ported in [23].
The calculations were performed under various boundary con-
ditions of pressure and temperature gradients. The differential
pressure was varied between 5 and 50 bars while the temperature
gradient was set to 0 (isothermal), 30, 50, 100 and 150 C, respec-
tively. Note that some of the calculations, such as the external
crack with a temperature gradient of 150 C, did not converge
probably because of extensive deformation.
If we dene the crack parametric angle, h, such as it is zero at
the tube surface and 90 at the deepest point of the crack, it is
found that the maximum stress intensity factor is located close
to the tube surface, namely h 2. This can be seen in Fig. 13 which
shows the stress intensity factor distribution along the penny
shape crack front for two temperature gradients of 50 and 100 C
and three differential pressure values of 5, 25 and 50 bar, the pres-
sure being higher inside the tube. Otherwise, it can also be noted
that the stress intensity factor remains relatively constant along
the crack front.
In the case of the internal crack, the maximum stress intensity
factor is not located close to the tube surface but at an angle vary-
ing between about 10 and 25 with increasing the temperature
gradient.
In both cases, namely external and internal crack, it is the tem-
perature gradient that dominates the stress intensity factor, the
pressure gradient effect is relatively small and quasi-independent
of the temperature gradient.
If we represent the obtained stress intensity factor as a function
of temperature gradient, we obtain for the three differential pres-
sures of 5, 25 and 50 bar the results plotted in Figs. 14 and 15 for
the externally- and internally-cracked cladding tube, respectively.
A quasi-linear dependence between the stress intensity factor
and the temperature gradient is observed. Note that in the case
of internal crack (see Fig. 15), two observations can be made. First,
the trend is reversed in comparison to the case of external crack.
Fig. 6. Crack resistance (J-Da) behavior of 12%Cr-ODS steel for shallow and deep
crack in the LT versus TL orientation.
Fig. 7. Elasticplastic stress intensity factor versus crack extension of the 12%Cr-
ODS steel for shallow and deep crack in the LT versus TL orientation.
(a) (b) (c)
Fig. 8. Observation of the crack tip region. Fatigue precrack before testing (a); after crack resistance testing (b); magnication of the crack extension part (c).
R. Chaouadi et al. / Journal of Nuclear Materials 442 (2013) 425433 429
(a) (b)
Fig. 9. SEM observations revealing secondary cracks in the extrusion direction perpendicular to the cracking plane (a) and elongated structure in the extrusion direction (b).
(a) (b)
Fig. 10. SEM fractography of the fracture surface: LT (a) versus TL (b) orientation.
Fig. 11. Finite element meshing of the cladding tube. Because of symmetry, only a quarter of the tube is simulated (symmetry in the Y and Z axes).
430 R. Chaouadi et al. / Journal of Nuclear Materials 442 (2013) 425433
The lower the differential pressure, the higher the stress intensity
factor. Second, at low temperature gradients and high differential
pressure, the stress intensity factor decreases when the tempera-
ture gradient. All these observations can be explained by the crack
closure that occurs when increasing the differential pressure as a
result of the induced compressive stresses.
At this stage, it is interesting to assess the possibility of using
the 12%Cr-ODS steel characterized in the previous sections as a
cladding material and evaluate its integrity.
Fig. 12. Detailed nite element meshing at the penny shape crack.
Fig. 13. Distribution of the stress intensity factor along the crack front in the case of
an external crack for DT = 50 and 100 C and DP = 5, 25 and 50 bar.
Fig. 14. Dependence of the maximum stress intensity factor as a function of DP and
DT for the externally-cracked cladding tube.
Fig. 15. Dependence of the maximum stress intensity factor as a function of DP and
DT for the internally-cracked cladding tube.
Fig. 16. Dependence of the maximum stress intensity factor as a function of DP and
DT for the internally- and externally-cracked cladding tube.
R. Chaouadi et al. / Journal of Nuclear Materials 442 (2013) 425433 431
The nite element results are summarized in Fig. 16. In Table 3,
we found that for the shallow crack length-to-width ratio, the
stress intensity factor when the crack is parallel to the extrusion
direction is 46 MPa
p
m. Under those conditions, Fig. 16 indicates
that the temperature gradient should not exceed about 75 C.
However, for a deeper crack, the maximum allowable temperature
gradient must be decreased. To assess the effect of crack size on the
applied stress intensity factor, we performed few nite element
simulations with a penny shape crack which is twice larger than
the former, namely, the crack length-to-width ratio is 0.25, and
the results are shown in Fig. 17 which indicates a signicant in-
crease of the stress intensity factor in comparison to a/W = 0.135.
We have addressed here only the case of a fuel cladding with
properties representative of the unirradiated condition. In addi-
tion, we considered typically a small crack of about 61 lm depth
and 0.15 mm length, with a/W = 0.13 and showed that the applied
stress intensity factor is signicantly increased if a/W is increased
to 0.25. It must be emphasized that, during operation, the materi-
als properties will be affected by irradiation: interaction between
the reactor coolant and the external cladding surface on one hand
and interaction of the fuel pellet with the internal surface of the
cladding might signicantly decrease the allowable temperature
gradient, limiting therefore the lifetime of the cladding. It is there-
fore of prime importance to correctly assess both the operating
conditions that are governing the applied loading and the material
crack resistance to guarantee the fuel cladding integrity during
normal and accidental conditions.
6. Conclusions
The crack resistance of the 12%Cr-ODS steel is severely reduced
at high temperature (650 C). Moreover, in the TL orientation,
grain boundary decohesion rather than microvoid coalescence is
observed on the fracture surface. This severe crack resistance
reduction is attributed to the presence of a high density of nanosize
oxide particles used to strengthen the material. Because of defor-
mation incompatibility between the hard oxide particles and the
soft matrix, void nucleation can easily occur and further crack
propagation can proceed by a void coalescence process requiring
a very low energy. This is an inherent characteristics of this kind
of materials due to the presence of a high density of strong nano-
size oxide particles ODS in a soft matrix facilitating crack initiation
and propagation at their interface. When the crack is parallel to the
extrusion direction, the crack resistance becomes even lower, with
an elasticplastic stress intensity factor of about 21 MPa
p
m in the
case of a deep crack. For shallow crack (a/W 0.13), it increases to
about 46 MPa
p
m. With such low values, the 12%Cr-ODS steel
investigated here cannot be considered at this stage for thick com-
ponents except if it can be demonstrated that the loading condi-
tions remain very low. For thin components, such as fuel
cladding, nite element calculations have shown that temperature
gradient across the clad wall thickness should be reduced as much
as possible to avoid fracture. Finally, in this work, we have ad-
dressed only the case of unirradiated material. Upon irradiation,
the degradation of the cladding material (irradiation, corrosion,
fuel pelletcladding interaction) might further decrease the maxi-
mum allowable temperature gradient.
Acknowledgements
GETMAT was co-sponsored by the European Commission 7th
Framework Programme of EURATOM for Nuclear Research and
Training Activities, Topic: Reactor systems, Fission-2007-6.0.02
and the project partners under Grant Agreement Number
FP7-212175. The authors are grateful to the LHMA scientic and
technical staff, in particular R. Mertens, P. Wouters, J. Van Eyken
and W. Vandermeulen.
References
[1] N. Baluc et al., J. Nucl. Mater. 417 (2011) 149153.
[2] S. Ukai, M. Fujiwara, J. Nucl. Mater. 307311 (2002) 749757.
[3] Y. de Carlan et al., J. Nucl. Mater. 386388 (2009) 430432.
[4] M.J. Alinger, G.R. Odette, D.T. Hoelzer, J. Nucl. Mater. 329333 (2004) 382386.
[5] R.J. Kurtz et al., J. Nucl. Mater. 386388 (2009) 411417.
[6] Z. Oksiuta, P. Olier, Y. de Carlan, N. Baluc, J. Nucl. Mater. 393 (2009) 114119.
[7] M. Inoue, T. Kaito, S. Ohtsuka, Research and development of oxide dispersion
strengthened ferritic steels for sodium cooled fast breeder reactor fuels, in: V.
Ghetta et al. (Eds.), Materials Issues for Generation IV Systems, 2008, pp. 311
325.
[8] Y. Yano, R. Ogawa, S. Yamashita, S. Ohtsuha, T. Kaito, N. Akasaka, M. Inoue, T.
Yoshitake, K. Tanaka, J. Nucl. Mater. 419 (2011) 305309.
[9] S.J. Zinkle, N.M. Ghoniem, Fusion Eng. Des. 5152 (2000) 5571.
[10] M.K. Miller, D.T. Hoelzer, J. Nucl. Mater. 418 (2011) 307310.
[11] R. Lindau et al., Fusion Eng. Des. 7579 (2005) 989996.
[12] E. Lucon, A. Leenaers, W. Vandermeulen, Fusion Eng. Des. 82 (2007) 2438
2443.
[13] J. Henry, X. Averty, A. Alamo, J. Nucl. Mater. 417 (2011) 99103.
[14] A. Alamo, V. Lambard, X. Averty, M.H. Mathon, J. Nucl. Mater. 329333 (2004)
333337.
[15] A. Alamo, J.L. Bertin, V.K. Shamardin, P. Wident, J. Nucl. Mater. 367370 (2007)
5459.
[16] M.A. Sokolov, D.T. Hoelzer, R.E. Stoller, D.A. McClintock, J. Nucl. Mater. 367
370 (2004) 213216.
[17] T.S. Byun, J.H. Kim, J.H. Yoon, D.T. Hoelzer, J. Nucl. Mater. 407 (2010) 7882.
[18] R. Chaouadi, J. Nucl. Mater. 403 (2010) 1518.
[19] D.T. Hoelzer, J. Bentley, M.A. Sokolov, M.K. Miller, G.R. Odette, M.J. Alinger, J.
Nucl. Mater. 367370 (2007) 166172.
[20] D.A. McClintock, D.T. Hoelzer, M.A. Sokolov, R.K. Nanstad, J. Nucl. Mater. 386
388 (2009) 307311.
[21] D.A. McClintock, M.A. Sokolov, D.T. Hoelzer, R.K. Nanstad, J. Nucl. Mater. 392
(2009) 353359.
[22] R.K. Nanstad, D.A. McClintock, D.T. Hoelzer, L. Tan, T.R. Allen, J. Nucl. Mater.
392 (2009) 331340.
[23] R. Chaouadi, Crack resistance behavior of 12%Cr- and 14%Cr-ODS steels at
elevated temperatures, SCKCEN, report ER-190, 2011.
[24] A. Kimura et al., J. Nucl. Mater. 417 (2011) 176179.
[25] L. Toualbi et al., J. Nucl. Mater. 428 (2012) 4753.
[26] S. Ukai, T. Okuda, M. Fujiwara, T. Kobayashi, S. Mizuta, H. Nakashima, J. Nucl.
Sci. Technol. 39 (2002) 872879.
[27] S. Ukai, S. Mizuta, M. Fujiwara, T. Okuda, T. Kobayashi, J. Nucl. Sci. Technol. 39
(2002) 778788.
[28] P. Dubuisson, Y. de Carlan, V. Garat, M. Blat, J. Nucl. Mater. 428 (2012) 612.
[29] A. Kimura et al., J. Nucl. Sci. Technol. 44 (2007) 323328.
[30] S. Ukai, T. Kaito, M. Seki, A.A. Mayorshin, O.V. Shishalov, J. Nucl. Sci. Technol. 42
(2005) 109122.
[31] T. Kaito, S. Ukai, A.V. Povstyanko, V.N. Emov, J. Nucl. Sci. Technol. 46 (2009)
529533.
[32] ASTM E1820, Standard test method for measurement of fracture toughness,
Annual Book of ASTM Standards, Section 3, vol. 03.01, American Society for
Testing and Materials, 2008.
[33] R. Chaouadi, J.L. Puzzolante, Int. J. Press. Vess. Pip. 85 (2008) 752761.
[34] R. Chaouadi, J. Test. Eval. 32 (2004) 469475.
[35] M.J. Alinger, G.R. Odette, G.E. Lucas, J. Nucl. Mater. 307311 (2002) 103114.
[36] A. Steckmeyer, V.H. Rodrigo, J.M. Gentzbittel, V. Rabeau, B. Fournier, J. Nucl.
Mater. 426 (2012) 182188.
Fig. 17. Effect of crack length on the stress intensity factor assuming DP = 25 bar.
432 R. Chaouadi et al. / Journal of Nuclear Materials 442 (2013) 425433
[37] R. Kasada et al., J. Nucl. Mater. 417 (2011) 180184.
[38] M. Serrano, M. Hernandez-Mayoral, A. Garcia-Junceda, J. Nucl. Mater. 428
(2012) 103109.
[39] R.L. Klueh, J.P. Shingledecker, R.W. Swindeman, D.T. Hoelzer, J. Nucl. Mater.
341 (2005) 103114.
[40] H. Hadraba et al., J. Nucl. Mater. 411 (2011) 112118.
[41] M. Ramesh, private communication, 2012.
[42] K.N. Shivakumar, J.C. Newman Jr., Three-dimensional elasticplastic analysis
of shallow cracks in single-edge-crack-tension specimens, NASA Technical
Memorandum 102636, National Aeronautics and Space Administration, April
1990.
[43] C. Ruggieri, Eng. Fract. Mech. 79 (2012) 245265.
[44] A. Steckmeyer et al., J. Nucl. Mater. 405 (2010) 95100.
[45] J.H. Kim, T.S. Byun, D.T. Hoelzer, J. Nucl. Mater. 407 (2010) 143150.
[46] K. Turba, R.C. Hurst, P. Hhner, J. Nucl. Mater. 428 (2012) 7681.
[47] T. Narita, S. Ukai, T. Kaito, S. Ohtsuka, T. Kobayashi, J. Nucl. Sci. Technol. 41
(2004) 10081012.
R. Chaouadi et al. / Journal of Nuclear Materials 442 (2013) 425433 433

You might also like