Sie sind auf Seite 1von 6

Journal of Nuclear Materials xxx (2013) xxxxxx

Contents lists available at SciVerse ScienceDirect

Journal of Nuclear Materials


journal homepage: www.elsevier.com/locate/jnucmat

Effects of titanium concentration and tungsten addition


on the nano-mesoscopic structure of high-Cr oxide dispersion
strengthened (ODS) ferritic steels
Peng Dou a,, Akihiko Kimura a, Ryuta Kasada a, Takanari Okuda b, Masaki Inoue c, Shigeharu Ukai d,
Somei Ohnuki d, Toshiharu Fujisawa e, Fujio Abe f
a

Institute of Advanced Energy, Kyoto University, Gokasho, Uji, Kyoto 611-0011, Japan
Kobelco Research Institute, 1-5-5 Takatsukadai, Nishi-ku, Kobe, Hyogo 651-2271, Japan
c
Advanced Nuclear System R&D Directorate, Japan Atomic Energy Agency, 4002 Narita, O-arai, Ibaraki 311-1393, Japan
d
Graduate School of Engineering, Hokkaido University, N13, W8, Kita-ku, Sapporo 060-8628, Japan
e
Graduate School of Engineering, Nagoya University, Furo-cho, Chikusa, Nagoya 464-8603, Japan
f
Structural Metals Center, National Institute for Materials Science, 1-2-1 Sengen, Tsukuba, Ibaraki 305-0047, Japan
b

a r t i c l e

i n f o

Article history:
Available online xxxx

a b s t r a c t
To study the effects of titanium concentration and tungsten addition on the nano-mesoscopic structure of
high-Cr oxide dispersion strengthened (ODS) ferritic steels, the spatial and size distributions, shape, and
coherency of the oxide particles in SOC-5 (Fe15.95Cr0.09Ti0.34Y2O3) and SOC-P3 (Fe13.32Cr1.9W
0.16Ti0.33Y2O3) were studied by diffraction contrast techniques, including weak beam electron microscopy. When the titanium concentration is increased from 0.09 to 0.16 wt.% with 1.9 wt.% W added, the
grain size decreases considerably, while the number density and mean size of the oxide particles significantly increases and decreases, respectively. 63.5% and 96% of the oxide particles in SOC-5 and SOCP3 (diameter <4.5 nm), respectively, are coherent with the bcc steel matrix. In SOC-5, 36% of the oxide
particles (4.510 nm in diameter) are semi-coherent; mist moir fringes are seen across only 12% of
the oxide particles with mist moir fringe spacing of 1.12 nm, indicating the mist strain is 0.11. However, only 4% of the oxide particles in SOC-P3 (4.510 nm in diameter) are semi-coherent; mist moir
fringes are seen across only 2% of the oxide particles with mist moir fringe spacing of 0.98 nm, indicating the mist strain is 0.126. The mean mist strain of coherent oxide particles in SOC-5 is 0.017.
2013 Elsevier B.V. All rights reserved.

1. Introduction
Future fusion energy will require new high-performance
structural alloys with outstanding properties sustainable under
long-term service in ultra-severe environments, including high
operating temperatures (7731273 K), large time-varying stresses, intense neutron radiation eld leading to high level of neutron
damage producing up to 200 dpa and 2000 appm of helium, and
chemically-reactive environments (e.g., highly corrosive coolants)
[1,2]. The stringent requirements of structural materials for fusion
reactors include superior high-temperature strength; excellent
creep resistance; high resistance to irradiation damage and
swelling; compatibility with cooling media (high resistance to corrosion in coolants such as lithium, lithium-lead, lithium-tin, Flibe
(LiFBeF2 or Li2BeF4), Flinak (LiFNaFKF), supercritical pressurized water (SCPW) and He); a low ductile-to-brittle transition
Corresponding author. Address: City University of Hong Kong, Tat Chee Avenue,
Kowloon, Hong Kong. Tel.: +852 3442 7468; fax: +852 3442 0295.
E-mail address: doup@tsinghua.edu.cn (P. Dou).

temperature with extremely low susceptibility to thermal aging


embrittlement; high resistance to hydrogen embrittlement
and/or helium embrittlement; high resistance to stress corrosion
cracking (SCC) in hot pressurized water; high thermal conductivity; low thermal expansion coefcient; low residual activation;
excellent workability; and good weldability [15].
High-Cr ODS ferritic steels are promising structural materials
for nuclear reactors because they exhibit high resistance to
corrosion in SCPW [4], lead-bismuth eutectic (LBE) [4], and Flinak
(LiFNaFKF) [5]; superior mechanical properties (e.g., tensile and
creep strength at high temperature) [6,7]; high irradiation
resistance to hardening, swelling, and embrittlement due to
self-healing mechanism [810]; high resistance to hydrogen
and/or helium embrittlement [11,12]; high resistance to SCC in
hot pressurized water [13]; and good weldability [14]. To improve
the corrosion resistance to SCPW, two high-Cr ODS steels (SOC-5
(Fe15.95Cr0.09Ti0.34Y2O3) and SOC-P3 (Fe13.32Cr1.9W
0.16Ti0.33Y2O3)) have been newly developed for fusion reactor
application. It was found that SOC-P3 has excellent strength and
optimal creep resistance, which are much better than that of SOC-5.

0022-3115/$ - see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jnucmat.2013.04.090

Please cite this article in press as: P. Dou et al., J. Nucl. Mater. (2013), http://dx.doi.org/10.1016/j.jnucmat.2013.04.090

P. Dou et al. / Journal of Nuclear Materials xxx (2013) xxxxxx

To clarify the underlying mechanism for the superior mechanical strength of SOC-P3 owing to the highly stabilized oxide particles and understand the physics process for the formation of
nanometer-scale YTiO features with different Ti contents and
W addition so as to achieve optimum chemical composition and
the improvement of ODS steels, the spatial and size distributions,
shape, and coherency of the oxide particles in SOC-5 and SOC-P3
were studied by diffraction contrast techniques, including weak
beam electron microscopy.
2. Experimental
Two of the newly developed high-Cr ODS ferritic steels, i.e.,
SOC-5 and SOC-P3, which have chemical compositions of
Fe0.04C0.01Si0.01Mn15.95Cr0.09Ti0.34Y2O3 and Fe0.046C
0.03Si0.03Mn13.32Cr0.16Ti1.9W0.33Y2O3, respectively, were
used in the present research. The fabrication procedure of the
ODS steels is detailed in Ref. [15].
Disc type transmission electron microscopy (TEM) specimens
with 3 mm diameter were punched from the sheets parallel and
perpendicular to the extrusion axis, mechanically thinned to
100 lm and then electro-polished in a TENUPOL device using
HClO4 + 95%CH3COOH as electrolyte at around 293 K. The nanomesoscopic structures of the ODS steels, especially the morphology
and coherency of oxide particles, were characterized by TEM using
a JEOL JEM 2010 transmission electron microscope (accelerating
voltage 200 kV) equipped with a double tilt specimen holder. Foil
thickness was determined by the convergent beam electron diffraction method.
Specically, the metal/oxide interface structures of the nanoparticles in SOC-5 and SOC-P3 were studied by diffraction contrast
techniques, including weak beam electron microscopy. The details
of the principle and process for the judgment of particle coherency
from Ashby and Brown contrast [16] and McIntyre and Brown
contrast [17], and the methods of measuring the lattice mist
and mist strain in the matrix surrounding the coherent and
semi-coherent particles from diffraction contrast images are
described in Section 2.2 and Section 2.3 of Ref. [2], respectively.
The shear moduli of SOC-5 and SOC-P3 were all assumed to be
equal to that of PM 2000 (65 GPa at room temperature) [18]. The
Youngs moduli, E, of the matrix of SOC-5 and SOC-P3, were then
estimated to be 169 GPa, because t is 0.3. The bulk modulus of
Y2TiO5 particle, K, is 134.6 GPa [19].
3. Results and discussion
3.1. Effect of titanium concentration and tungsten addition on grain
morphology
The grain morphologies of SOC-5 and SOC-P3 are shown in
Fig. 1ad, respectively. The mean intercept grain diameters of
SOC-5 (Fig. 1b) and SOC-P3 (Fig. 1d) measured in a plane perpendicular to the extrusion axes are 1.18 lm (uncertainty: +0.3 lm
and 0.2 lm) and 0.78 lm (uncertainty: +0.1 lm and 0.3 lm),
respectively. For SOC-5 (Fig. 1a), the average grain length, average
grain width, and grain aspect ratios (GARs) measured in a plane
parallel to the extrusion axes are 3.96 lm (uncertainty: +0.45 lm
and 0.4 lm), 1.16 lm (uncertainty: +0.35 lm and 0.25 lm),
and 3.4 (uncertainty: +0.45 and 0.4), respectively. However, for
SOC-P3 (Fig. 1c), the average grain length, average grain width,
and GAR are 3.28 lm (uncertainty: +0.4 lm and 0.35 lm),
0.75 lm (uncertainty: +0.15 lm and 0.25 lm), and 4.3 (uncertainty: +0.35 and 0.4), respectively. An average grain size decreases signicantly when the Ti concentration is increased from
0.09 to 0.16 wt.% and 1.9 wt.% W is added.

3.2. Effect of titanium concentration and tungsten addition on oxide


particle morphology
Fig. 2 shows the bright eld images of the oxide particles dispersed in the matrix of SOC-5 (Fig. 2a) and SOC-P3 (Fig. 2b).
63.5% of the oxides in SOC-5 and almost all the particles in
SOC-P3 (96%) are very small with diameter <4.5 nm. The very
small oxide particles keep their shape under different diffraction
conditions. Thus, it can be concluded that they are spherical. In
SOC-5, the relatively larger particles (diameter P4.5 nm), which
constitute 36% of the oxides in the ODS steel, appear to adopt a
cubical shape, Fig. 2a. Weak beam dark eld images of the oxide
particles in SOC-5 and SOC-P3 are shown in Fig. 3a and b, respectively. According the contrast features, the oxide particles were divided into three groups, showing lobelobe contrast, black/white
dot contrast, and oxide particles surrounded by fringes, which
are thought to be mist moir fringes.
When determining the number density and size distribution
of the oxide particles in SOC-5 and SOC-P3, all the three groups
of particles were counted and measured together with no distinction made among them. The length of the no contrast line
was taken as the diameter of the oxide particles exhibiting
lobelobe contrast [16]. The diameter of the oxide particles
showing black/white dot contrast was regarded as the diameter
of the dots [3,20]. The diameter of the semi-coherent oxide particles surrounded by mist moir fringes was dened as the
average length along and perpendicular to the mist moir
fringes [3]. The total number density was estimated based on
the bright eld images shown in Fig. 2 and other bright-eld
down-zone images.
The size distribution histograms of the oxide particles in both
ODS steels are presented in Fig. 4. SOC-5 has bimodal oxide particle
size. There are two local maximum numbers of oxide particles at
12 nm and 45 nm. For the oxide particles in SOC-5 and SOCP3, the mean diameters are 3.85 and 2.8 nm, the total number densities are 7.07  1022 and 1.46  1023 m3, the inter-particle spacings are 60.6 and 49.5 nm, and the volume fractions are 0.211% and
0.168%, respectively. When the Ti concentration is increased from
0.09 to 0.16 wt.% with 1.9 wt.% W added, the mean diameter and
inter-particle spacing of oxide particles decrease considerably
while the number density increases very remarkably, which may
lead to more intensive suppressing of dislocation loop growth
[21] and, consequently, still better irradiation resistance due to
the signicantly enhanced self-healing mechanism resulting from
the presence of much higher proportion of metal/oxide interface
to bulk [9,10].
In general, many factors inuence the spatial and size distributions of YTi complex oxide particles (e.g., the amount of
excess oxygen [2224], the Ti concentration [2225], the
processing temperature history (especially the consolidation
temperature) [26], and the W addition [24]). It was found that
increasing Ti concentration is very effective for producing ner
oxide particles with higher number density [24,25], and moreover, the effect of W on the microstructure is smaller than that
of excess O and Ti, but still produces a favorable effect in reducing the size and increasing the number density of the oxide particles [24].
Note that the amount of excess oxygen and fabrication procedure (including the processing temperature) of SOC-5 are similar
to those of SOC-P3. However, the Ti concentrations of SOC-5 and
SOC-P3 are 0.09 and 0.16 wt.%, respectively, and moreover, the
concentration of W of SOC-P3 is 1.9 wt.%, while there is no W in
SOC-5. Therefore, the experimental fact that SOC-5 has oxide particles with larger size and much lower number density, relative to
SOC-P3, is due to the much lower concentration of Ti and the absence of W.

Please cite this article in press as: P. Dou et al., J. Nucl. Mater. (2013), http://dx.doi.org/10.1016/j.jnucmat.2013.04.090

P. Dou et al. / Journal of Nuclear Materials xxx (2013) xxxxxx

Fig. 1. Grain morphologies of SOC-5 and SOC-P3 in: (a) longitudinal section of SOC-5, (b) transverse section of SOC-5, (c) longitudinal section of SOC-P3, and (d) transverse
section of SOC-P3.

3.3. Effect of titanium concentration and tungsten addition on the


coherency of oxide particles
In Fig. 3, the image contrast of the oxides seems to depend on
the particle size. For the very small oxide particles (diameter
<4.5 nm) in SOC-5 and SOC-P3, some are evidently coherent with
the bcc steel matrix according to Ashby and Brown contrast behavior [16] because they appear as small lobe-lobe contrast with a no
contrast line perpendicular to the g vector. However, others are observed as dot contrast. Based on equation: P S gjejr30 n2
g (where g
1
(g dhkl ), e, r0, and ng denote the modulus of the active diffraction
vector, the mist strain, the constrained particle radius, and the
corresponding extinction distance, respectively) [17], an estimate
of the maximum possible value of the parameter PS (assuming a
particle size of 4.5 nm and a mist strain of 0.1) yields values from
0.006 to 0.02 for all reasonable diffraction conditions (e.g., for
1
Fig. 3, g dhkl is 0.20271 nm1, and ng calculated to be 50.1 nm
[27]). In general, the modulus of the mist strain of coherent precipitates is less than 0.05 (|e| 6 0.05). Hence, 0.1 is a conservative

assumption and, therefore, PS is far below McIntyre and Browns


critical value of PS 6 0.2 [17].
Therefore, the strain eld contrast behavior of the very small
oxide particles in SOC-5 and SOC-P3, which is shown in Fig. 3,
agrees well with the contrast features of tiny precipitates proposed
by McIntyre and Brown [17] and mentioned in Section 2.2 of Ref.
[2]. Evidently, whether the very small oxide particles appear as
small lobes with no contrast line perpendicular to the g vector or
as black/white dots only depends on their depth in TEM foils. It
is reasonable to assume that the very small oxide particles appearing as black/white dots are also coherent with the matrix. The
assumption has already been conrmed by high-resolution transmission electron microscopy (HRTEM) analyses on both ODS steels,
and will be detailed in another paper.
Although 180 and 200 very small oxide particles in SOC-5 and
SOC-P3, respectively, have been analyzed by HRTEM, no semicoherent or incoherent oxide has been identied because all the
very small oxide particles were found to be coherent with the
matrix. HRTEM observations indicate that the relatively larger

Please cite this article in press as: P. Dou et al., J. Nucl. Mater. (2013), http://dx.doi.org/10.1016/j.jnucmat.2013.04.090

P. Dou et al. / Journal of Nuclear Materials xxx (2013) xxxxxx

Fig. 2. Bright eld images of the oxide particles dispersed in the matrix of SOC-5 (a) and SOC-P3 (b) under near dynamical two-beam imaging conditions and in slightly
under-focused regime.

Fig. 4. Size distributions of the oxide particles dispersed in the matrix of SOC-5 and
SOC-P3. Black bars denote SOC-5 and white bars denote SOC-P3.

Fig. 3. Weak beam dark eld images of the oxide particles dispersed in the matrix
of SOC-5 (a) and SOC-P3 (b).

particles (4.510 nm in diameter) in SOC-5 and SOC-P3 tend to be


semi-coherent. According to the statistics results based on Fig. 4,
coherent oxides constitute 63.5% and 96% while semi-coherent
oxides constitute 36% and 4% of the particles in SOC-5 and
SOC-P3, respectively. Mist moir fringes were seen across semicoherent particles constituting only 12% and 2% of the oxides

in SOC-5 (Fig. 3a) and SOC-P3 (Fig. 3b), respectively. According to


the results of HRTEM analyses, the large particles (diameter
>10 nm) in SOC-5 and SOC-P3 tend to be incoherent. Thus, the
coherency of the oxides in both ODS steels is size-dependent.
The size dependence of the particle coherency was analyzed
from the micrographs shown in Fig. 3 and other weak beam dark
eld images, and the statistics results of the proportion of coherent, semi-coherent, and incoherent particles are present in Table
1. The coherency of the oxide particles with the bcc steel matrix
is enhanced very signicantly with the titanium concentration
increased from 0.09 to 0.16 wt.% and 1.9 wt.% W is added.
Incidentally, most of the very small particles in both ODS steels
were identied as hexagonal Y2TiO5 oxide while most of the rela-

Please cite this article in press as: P. Dou et al., J. Nucl. Mater. (2013), http://dx.doi.org/10.1016/j.jnucmat.2013.04.090

P. Dou et al. / Journal of Nuclear Materials xxx (2013) xxxxxx


Table 1
Coherency of oxide particles dispersed in the matrix of SOC-5 and SOC-P3.
Steels

SOC-5

Uncertainty

SOC-P3

Uncertainty

Coherent

Diameter range (nm)


Number ratio (%)
Mist strain e

<4.5
63.5
0.017

+0.0020.002

<4.5
96

Semi-coherent

Diameter range (nm)


Number ratio (%)
Spacing of mist moir fringes (nm)
Lattice mist d
Mist strain e

4.510
36 (12%)
1.12
0.181
0.11

+0.020.02
+0.0030.003
+0.0020.002

4.510
4 (2%)
0.98
0.207
0.126

+0.020.01
+0.0020.004
+0.0010.003

Incoherent

Diameter range (nm)


Number ratio (%)

>10
0.5

>10
/

tive larger particles in SOC-5 and SOC-P3 (diameter >4.5 nm) were
identied as orthorhombic or hexagonal Y2TiO5 oxide by the
HRTEM analyses.
3.4. Effect of titanium concentration and tungsten addition on the
mist strain of oxide particles
For SOC-5, measurements and calculations of mist strain, e,
were done on 30 coherent particles, appearing as small black lobes
in two beam dynamic bright eld images or nearly strong two
beam bright eld images, which are shown in Fig. 2a and other
images. The mean e of coherent particles in SOC-5 is 0.0145.
However, the e obtained by AshbyBrown technique is always
too low by as much as 15% because of the system error due to
the two-beam over-estimation of extinction distance [16]. Therefore, the real mean e of coherent oxide particles in SOC-5 is 0.017.
However, for coherent particles in SOC-P3, at least three different values of mist strain have been obtained. Because the proportion of the coherent particles exhibiting lobe-lobe contrast with no
contrast line perpendicular to the g vector is extremely low while
almost all the coherent particles appear as black/white dots, Figs.
2b and 3b, it is very difcult to get statistic results of the proportion of coherent oxide particles with different specic value of mist strain from diffraction contrast images alone.
The mist moir fringe spacing (DM) of semi-coherent particles in
SOC-5 is 1.12 nm (Figs. 2a and 3a), according to which the modulus
of the lattice mist, d, was calculated to be 0.181, based on Eq. (4) in
Ref. [2] since d1 10 is 0.2027 nm, and then the modulus of the e was
calculated to be 0.11, based on Eq. (2) in Ref. [2]. The DM of semicoherent particles in SOC-P3 is 0.98 nm (Fig. 3b), according to which
the modulus of the d was calculated to be 0.207, and then the modulus of the e was calculated to be 0.126. The e of coherent particles
in SOC-5 and SOC-P3, and the moduli of the d and e of semi-coherent
particles in both ODS steels are also listed in Table 1.
The yield stresses of SOC-5 and SOC-P3 at room temperature are
910 and 1285 MPa, respectively. That is, the mechanical strength
was successfully improved with the Ti concentration increased
from 0.09 to 0.16 wt.% and 1.9 wt.% W added. This is mainly attributed to the increased grain boundary strengthening due to the signicant reduction in grain size (Section 3.1), the enhanced particle
strengthening due to the remarkably ner oxides with much denser dispersion [28] (Section 3.2), and the improved solid solution
strengthening due to the addition of 1.9 wt.% W. W (ferrite-former
elements) improves the creep strength, thermal stability, and corrosion resistance [5] of ODS alloys.
Increasing the Ti concentration from 0.09 to 0.16 wt.% with
1.9 wt.% W added leads to excellent lattice coherency between
the oxide particles and the bcc steel matrix (Section 3.3 and Table
1 in Section 3.4), which gives rise to very low interface energy. The
low interface energy of coherent oxides, which reduces the Gibbs
Thomson effect at the interface, the very low solubility of O and Y

in bcc Fe [29], and the extremely low diffusion coefcient of Y element in a-Fe matrix at various temperatures [30], can effectively
prevent the coarsening of the YTiO particles. Moreover, if a solute atom (e.g., Y, and/or Ti) can attach less easily to a coherent
interface, it could be imagined that this process is even more difcult if the particle has a complex lattice structure composed of
three or two distinct elements, as is the case for the orthorhombic/hexagonal Y2TiO5 complex oxide. This can also lead to the
excellent coarsening resistance of the oxide particles at high temperature. Finally, W improves the thermal stability of ODS steels by
suppressing the coarsening of oxides. Therefore, oxide particles in
SOC-P3 can be expected to be very stable and exhibit still higher
thermal stability than those in SOC-5, leading to even higher capability for keeping attractive properties during long-term exposure
to high temperature for nuclear applications.
Creep threshold stresses observed in precipitation-, or dispersion-strengthened alloys are explained by mechanisms based on
particle shearing [31], bypass by climb [31,32], or detachment
[33]. The former two are applicable to coherent [31,32,34] and
semi-coherent particles [35] with the lattice mist and mist
strain having signicant inuence on the creep threshold stresses
[31,32,34,35] while the third is operative for incoherent particles
[33]. Thus the proportion of coherent, semi-coherent, and incoherent particles in both ODS steels and the magnitude of lattice mist
and mist strain of the oxides have been obtained in this work. In
addition, the magnitude of lattice mist and mist strain of the
oxide particles has been obtained by HRTEM and will be detailed
in another paper with the quantitative analyses of the creep
strengthening in both ODS materials.
4. Conclusions
For the two newly-developed high-Cr ODS ferritic steels, i.e.,
SOC-5 (Fe15.95Cr0.09Ti0.34Y2O3) and SOC-P3 (Fe13.32Cr
1.9W0.16Ti0.33Y2O3), the following was determined:
(1) When the titanium concentration is increased from 0.09 to
0.16 wt.% with 1.9 wt.% W added, the grain size decreases
considerably, and the number density and mean diameter
of the oxide particles signicantly increase and decrease,
respectively.
(2) 63.5% and 96% of the oxide particles in SOC-5 and SOC-P3
(diameter <4.5 nm) are coherent with the bcc steel matrix,
respectively. 36% and only 4% of the oxide particles in
SOC-5 and SOC-P3 are semi-coherent, respectively.
(3) Mist moir fringes are seen across only 12% and 2% of
the oxide particles in SOC-5 and SOC-P3, with the mist
moir fringe spacings of 1.12 and 0.98 nm, indicating the
mist strains are 0.11 and 0.126, respectively.
(4) The mean mist strain of coherent oxide particles in SOC-5 is
0.017.

Please cite this article in press as: P. Dou et al., J. Nucl. Mater. (2013), http://dx.doi.org/10.1016/j.jnucmat.2013.04.090

P. Dou et al. / Journal of Nuclear Materials xxx (2013) xxxxxx

Acknowledgements
Present study includes the result of R&D of corrosion resistant
super ODS steel for highly efcient nuclear systems entrusted to
Kyoto University by the Ministry of Education, Culture, Sports, Science and Technology of Japan (MEXT).
References
[1] G.R. Odette, M.J. Alinger, B.D. Wirth, Annu. Rev. Mater. Res. 38 (2008) 471503.
[2] P. Dou, A. Kimura, T. Okuda, M. Inoue, S. Ukai, S. Ohnuki, T. Fujisawa, F. Abe,
Acta Mater. 59 (2011) 9921002.
[3] P. Dou, A. Kimura, T. Okuda, M. Inoue, S. Ukai, S. Ohnuki, T. Fujisawa, F. Abe, J.
Nucl. Mater. 417 (2011) 166170.
[4] A. Kimura, R. Kasada, N. Iwata, H. Kishimoto, C.H. Zhang, J. Isselin, P. Dou, et al.,
J. Nucl. Mater. 417 (2011) 176179.
[5] B. El-Dasher, J. Farmer, J. Ferreira, M. Serrano de Caro, A. Rubenchik, A. Kimura,
J. Nucl. Mater. 419 (2011) 1523.
[6] R. Kasada, C.H. Zhang, P. Dou, et al., in: Proc. Int. Cong. Adv. Nucl. Power Plants,
(ICAPP-2009), CD-ROM le, 2009, Paper No: 9972.
[7] T. Furukawa, S. Ohtsuka, M. Inoue, T. Okuda, F. Abe, S. Ohnuki, T. Fujiwara, A.
Kimura, in: Proc. Int. Cong. Adv. Nucl. Power Plants, (ICAPP-2009), CD-ROM
le, 2009, Paper No: 9221.
[8] T. Yoshitake, Y. Abe, N. Akasaka, S. Ohtsuka, S. Ukai, A. Kimura, J. Nucl. Mater.
329333 (2004) 342346.
[9] X.M. Bai, A.F. Voter, R.G. Hoagland, M. Nastasi, B.P. Uberuaga, Science 327
(2010) 16311634.
[10] G. Ackland, Science 327 (2010) 15871588.
[11] J.S. Lee, A. Kimura, S. Ukai, M. Fujiwara, J. Nucl. Mater. 329333 (2004) 1122
1126.
[12] K. Yutani, H. Kishimoto, R. Kasada, A. Kimura, J. Nucl. Mater. 367370 (2007)
423427.

[13] H.S. Cho, H. Ohkubo, N.Y. Iwata, A. Kimura, et al. in: Proc. Int. Cong. Adv. Nucl.
Power Plants, (ICAPP-2005), CD-ROM le, 2005, Paper No: 5457.
[14] S.H. Noh, R. Kasada, A. Kimura, Acta Mater. 59 (2011) 31963204.
[15] S. Ukai, S. Mizuta, T. Yoshitake, T. Okuda, S. Hagi, M. Fujiwara, S. Hagi, T.
Kobayashi, J. Nucl. Mater. 283287 (2000) 702706.
[16] M.F. Ashby, L.M. Brown, Philos. Mag. 8 (1963) 10831103.
[17] K.G. McIntyre, L.M. Brown, J. Phy. (Paris) 27 (1966) C31783180.
[18] C. Capdevila, M.K. Miller, I. Toda, J. Chao, Mater. Sci. Eng. A 527 (2010) 7931
7938.
[19] Y. Jiang, J.R. Smith, G.R. Odette, Acta Mater. 58 (2010) 15361543.
[20] U. Holzwarth, H. Stamm, J Nucl. Mater. 279 (2000) 3145.
[21] C.Z. Yu, H. Oka, N. Hashimoto, S. Ohnuki, J. Nucl. Mater. 417 (2011) 286288.
[22] S. Ohtsuka, S. Ukai, M. Fujiwara, T. Kaito, T. Narita, J. Nucl. Mater. 329333
(2004) 372376.
[23] S. Ohtsuka, S. Ukai, M. Fujiwara, T. Kaito, T. Narita, J. Phys. Chem. Solids 66
(2005) 571575.
[24] M. Ohnuma, J. Suzuki, S. Ohtsuka, S.W. Kim, T. Kaito, M. Inoue, H. Kitazawa,
Acta Mater. 57 (2009) 55715581.
[25] S. Ukai, M. Harada, H. Okada, M. Inoue, S. Nomura, S. Shikakura, K. Asabe, T.
Nishida, M. Fujiwara, J. Nucl. Mater. 204 (1993) 6573.
[26] M.J. Alinger, G.R. Odette, D.T. Hoelzer, Acta Mater. 57 (2009) 392406.
[27] D.B. Williams, C.B. Carter, Transmission Electron Microscopy A Textbook for
Materials Science, second ed., Springer, New York, 2009.
[28] D.J. Bacon, U.F. Kocks, R.O. Scattergood, Philos. Mag. 28 (1973) 12411263.
[29] A. Hirata, T. Fujita, Y.R. Wen, J.H. Schneibel, C.T. Liu, M.W. Chen, Nat. Mater. 10
(2011) 922926.
[30] C. Hin, B.D. Wirth, J. Nucl. Mater. 402 (2010) 3037.
[31] E. Nembach, Particle Strengthening of Metals and Alloys, John Wiley & Sons,
Inc., New York, 1997.
[32] E.A. Marquis, D.C. Dunand, Scripta Mater. 47 (2002) 503508.
[33] J. Rsler, E. Arzt, Acta Metall. Mater. 38 (1990) 671683.
[34] M.E. Krug, D.C. Dunand, Acta Mater. 59 (2011) 51255134.
[35] E.A. Marquis, D.N. Seidman, D.C. Dunand, Acta Mater. 51 (2003) 47514760.

Please cite this article in press as: P. Dou et al., J. Nucl. Mater. (2013), http://dx.doi.org/10.1016/j.jnucmat.2013.04.090

Das könnte Ihnen auch gefallen