Sunteți pe pagina 1din 31

4.

08 Oxide Dispersion Strengthened Steels


S. Ukai
Hokkaido University, Sapporo, Japan

 2012 Elsevier Ltd. All rights reserved.

4.08.1 Introduction 242


4.08.2 Nanoscale Oxide Particle Control 242
4.08.2.1 Dissociation and Precipitation 242
4.08.2.2 Structure and Coherency 243
4.08.3 Martensitic 9Cr-ODS Steels 244
4.08.3.1 Chemical Composition and Microstructure 244
4.08.3.2 Residual Ferrite Formation and Strength Characterization 245
4.08.3.2.1 Mechanically alloyed powder characterization 245
4.08.3.2.2 Pinning of a–g interface by oxide particles 245
4.08.3.2.3 Strength characterization 247
4.08.3.3 Cladding Manufacturing 248
4.08.3.3.1 Continuous cooling transformation diagram 248
4.08.3.3.2 Manufacturing process 249
4.08.3.4 Creep and tensile properties 250
4.08.4 Ferritic 12Cr-ODS Steels 250
4.08.4.1 Strength Anisotropy 250
4.08.4.2 Recrystallization Tests 252
4.08.4.3 Cold-Rolling Cladding Manufacturing 252
4.08.4.4 Internal Creep Rupture Property 253
4.08.5 Al-Added 16Cr-ODS Steels 255
4.08.5.1 Application and Technical Issues 255
4.08.5.2 Thermal Aging Embrittlement Due to High Cr Content 256
4.08.5.3 Mechanical Properties 256
4.08.5.4 Cladding Manufacturing 257
4.08.6 Existing ODS Steel Cladding 258
4.08.7 Corrosion and Oxidation 259
4.08.7.1 Sodium Compatibility 259
4.08.7.2 LBE Compatibility 259
4.08.7.3 SCPW Compatibility 262
4.08.7.4 Oxidation 262
4.08.8 Irradiation 263
4.08.8.1 Simulated Irradiation 263
4.08.8.2 Neutron Irradiation of Materials 264
4.08.8.3 Fuel Pin Irradiation 268
4.08.8.3.1 9Cr- and 12Cr-ODS steel cladding in BOR-60 268
4.08.8.3.2 12Cr-ODS steel cladding in EBR-II 269
4.08.8.3.3 DT2203Y05 in Phénix 269
4.08.9 Summary 269
References 270

CVN Charpy V-notch


Abbreviations
EFTEM Energy-filtered transmission electron
CTT Continuous cooling transformation
microscopy
CEN-SCK Centre d’Etude de l’énergie Nucleaire –
EPMA Electron probe microanalysis
Studiecentrum voor Kernenergie

241
242 Oxide Dispersion Strengthened Steels

FFT Fast Fourier transformation example, tensile and creep rupture strengths in the
HIP Hot isostatic pressing hoop directions, and irradiation performance, is
HRTEM High-resolution transmission electron reviewed. The development of Al-added high Cr-
microscopy ODS steel cladding is also addressed, with a focus on
INCO International Nickel Company superior resistance to oxidation and corrosion in a lead–
JAEA Japan Atomic Energy Agency bismuth eutectic (LBE), and supercritical pressurized
LBE Lead–bismuth eutectic water (SCPW) in the international Generation IV
LFR Lead fast reactor advanced nuclear power system. Nanocluster ODS
LMP Larson–Miller parameter steels,1 for example, 14YWT, etc., for fusion blanket
MA Mechanical alloying structure materials, are not addressed in this chapter.
MOX Mixed oxide
ODS Oxide dispersion strengthened
PMW Pulse magnetic Welding 4.08.2 Nanoscale Oxide Particle
PRW Pressurized resistance welding Control
SCPW Super critical pressurized water 4.08.2.1 Dissociation and Precipitation
SEM Secondary electron microscopy
SFR Sodium fast reactor The fine distribution of Y2O3 particles, which is
TIG Tungsten inert gas welding essential to improving the high temperature strength
UTS Ultimate tensile strength of ODS steels, is attained by the dissociation of oxide
particles during MA processing.2 The thermodynam-
ically stale Y2O3 particles are forcedly decomposed
into the ferritic steel matrix during the MA process.
4.08.1 Introduction Subsequent annealing induces oxide particles to pre-
cipitate finely at elevated temperature of around
Recent progress in oxide dispersion strengthened 1000 C. The co-addition of Ti during MA proces-
(ODS) steels produced by mechanical alloying (MA) sing promotes the decomposition of Y2O3 and then
techniques allows them to be used as fuel cladding in the precipitation of Y–Ti complex oxide particles
sodium-cooled fast reactors (SFR). The thermally through an annealing heat treatment.3,4 A field emis-
stable oxide particles dispersed in the ferritic matrix sion ion micro-probe (FIM) analysis confirmed that
improve the radiation resistance and creep resistance this type of complex oxide is constituted of several
at high temperature. As a result, ODS steels have a nanometer-sized Y–Ti–O compounds.5–7
strong potential for high burnup (long-life) and high- The precipitation process of the decomposed Y2O3
temperature applications typical for SFR fuels. The was investigated by means of a small angle neutron
attractiveness of ODS steels is due not only to the scattering (SANS) experiment.8 The neutron-scattering
nanosize oxide particles composed of Y–Ti–O atoms cross-section (dS/dO) versus scattering vector (q2)
but also to their controlled micron-size grain mor- plots for the milled U14YWT(Fe–14Cr–0.4Ti–3W–
phology. We review existing knowledge on the crys- 0.25Y2O3) are shown in Figure 1(a). They indicate
talline structure and lattice coherency of these that the hot isostatic pressing (HIP) of U14YWT at
nanosize particles with their surrounding matrix, 850 C leads to the precipitation of a high number
since these factors dominate the dispersion and density of nanoclusters, as designated by Odette.
strength-determining mechanism through dislocation Figure 1(b) shows the effects of HIP (filled symbols)
interaction. The development of manufacturing pro- and powder annealing (open symbols) at tempera-
cesses is a principal issue for hardened ODS steels to tures of 700, 850, 1000, and 1150 C. The increase
realize long, thin-walled ODS steel cladding on pro- in magnitude and decrease in slope of the dS/dO
duction scales. There was the long-standing problem in versus q2 curves indicate that the radius of nanoclus-
low hoop strength due to the extremely elongated fine ters decreases and their number density increases
grains parallel to the rolling direction. To soften hard- with decreasing temperature at HIP and powder
ened cold-rolled products and modify their grain annealing. Annealing at 700 C produces the highest
morphology, martensitic 9Cr-ODS steels and ferritic scattering and lowest sloping, which indicates that the
12Cr-ODS steels have been developed. Current prog- smallest-sized nanoclusters precipitate with the high-
ress in the development of these ODS steel claddings, est number density at lower temperatures. In terms of
including their relevant mechanical properties, for an X-ray diffraction experiment using Super Photon
Oxide Dispersion Strengthened Steels 243

10 10
HIPed materials Powder anneals
850 C HIP (No Y2O3)

850 C HIP (U14YWT) 1150 C


dS/dW (cm srad)−1

dΣ/dΩ (cm srad)−1


1 1
As-MA (U14YWT) Decreasing <r>
Increasing Nd 1000 C
As-MA (No Y2O3) 850 C
0.1 0.1 700 C

1150 1000 850 C Controls (No Y2O3)


0.01 0.01
0 2 4 6 8 10 0 2 4 6 8 10
(a) q2 (nm−2) (b) q2 (nm−2)
Figure 1 Results of a SANS experiments for as-mechanically alloyed powders and after HIP and annealing in U14YWT
(Fe–14Cr–0.4Ti–3W–0.25Y2O3). (a) As-MA, 850 C HIP and annealing and (b) HIP (filled) and annealing (annealing) at 1150, 1000,
850, and 700 C . Reproduced from Alinger, M. J.; Odette, G. R.; Hoelzer, D. T. J. Nucl. Mater. 2004, 382, 329–333.

ring-8 eV (Spring-8) constructed in Japan, Kim et al.


recently reported that nanoclusters could be in a −
(0 1 1)
noncrystalline state and can be transformed to nano- −
3.06 Å − (2 2 2) −
crystalline oxide particles at around 1000 C.9 (1 0 1) (1 1 0)
70.5 −−
(2 2 2 )

(2 2 2)
4.08.2.2 Structure and Coherency −
3.06 Å

(1 1 0) − − (1 0 1 )
(2 2 2 )
With regard to ODS steels without Ti, high resolu- −
tion (HR) TEM investigations were performed by (0 1 1 )
(a) (b)
Klimiankou to investigate the structure of Y2O3.10
The crystallographic lattice of the metal matrix cor- Figure 2 HRTEM micrograph of the Y2O3 particle with
responds to a-Fe with a bcc structure and a lattice surrounding matrix (a) and FFT image of micrograph
(b). The diffraction spots from Y2O3 particle of {2 2 2} type
constant of a0 ¼ 0.287 nm.11 The Y2O3 has a crystal-
form the rectangle, whereas diffraction spots from the
line bcc structure with a 1.06 nm lattice constant.11 matrix of {1 1 0} type form the hexagon at the [1 1 0] zone
Figure 2 shows an HRTEM image taken from an axis and [1 1 1] of matrix. Reproduced from Kliniankou, M.;
Y2O3 particle that is surrounded by the matrix (M) Lindau, R.; Moslang, A. J. Nucl. Mater. 2004, 329–333,
lattice. This image was taken from the grain, oriented 347–351.
with [1 1 1]M zone axis to the electron beam. A fast
Fourier transformation (FFT) of the image shows the
matrix lattice as a hexagonal pattern with diffraction The interfacial coherency between the Y2O3 particle
spots of the {1 1 0} type and dM(110) ¼ 0.203 nm dis- and the ferritic matrix can be estimated by Klimiankou
tance. In the FFT image, the Y2O3 (YO) lattice as follows:
is rectangular, with diffraction spots of the {2 2 2}
3dMð1 1 0Þ & 2dYOð2 2 2Þ
type and a corresponding atomic planes distance ' 0:5%: ½2%
of dYO(2 2 2) ¼ 0.306 nm. The angle of 70.5 between 2dYOð2 2 2Þ
diffraction spots of the {2 2 2}YO type marked in This result suggests that a coherency could be satisfied
Figure 2(a) confirms that the Y2O3 particle is oriented for the Y2O3 particles surrounded by a ferritic matrix.
with the [1 1 0] zone axis, and consequently [1 1 0]YO// Concerning the Y–Ti–O complex oxide particles
[1 1 1]M. The orientation correlation of both lattices formed in Ti-added ODS steels, Figure 3(a) shows
is (1 1 1)YO//(1 1 0)M. Therefore, the following an energy-filtered (EF) TEM micrograph from a
Kurdjumov–Sachs orientation relationship12 is satisfied: Y–Ti–O particle in which two atomic planes are simul-
taneously visible.10 Figure 3(b) shows the (0 0 4) and
ð11 1 ÞYO ==ð11 0ÞM ; (2 2 2) atomic planes of Y2Ti2O7 cubic (a0 ¼ 1.01 nm)
½110%YO ==½111%M : ½1% phases with the [1 1 0] zone axis. In fact, the measured
244 Oxide Dispersion Strengthened Steels

(a) (b)

−− 222 Residual
222 ferrite
004

5 nm

Figure 3 EFTEM images of Y2Ti2O7 particles. Tempered martensite


Reproduced from Kliniankou, M.; Lindau, R.; Moslang, A.
J. Nucl. Mater. 2004, 329–333, 347–351.

data are equal to the following data calculated from the


Y2Ti2O7 structure: d2 2 2 ¼ 0.29 nm and d0 0 4 ¼ 0.25, 20 µm
and an angle between the (0 0 4) and (2 2 2) atomic
planes of 54.7 . The analysis of EFTEM results defini-
tively shows that Y–Ti–O particles have a Y2Ti2O7 Figure 4 Microstructure of 9Cr-ODS steel showing
composition. residual ferrite and tempered martensite.
These findings suggest that nano-oxide particles
precipitate from the ferritic matrix, maintaining crys-
talline coherency or partial-coherency with a ferritic cold rolling and cold workability. To achieve a bal-
matrix. In general, the nucleation and growth of pre- ance between strength and workability, a value of
cipitates proceeds, as both interfacial and strain ener- 0.2 wt% was selected. Tungsten of 2 wt% is also
gies become minimal. In the case of ODS steels, added in order to improve high-temperature strength
interfacial coherency could be maintained between by means of solid solution hardening.
thermodynamically stable nanoparticle precipitates The microstructure of 9Cr-ODS steels13–18 can be
and the ferritic matrix in order to decrease the free easily controlled by a reversible a–g transformation
energy in the system from the extremely high energy with a remarkably high driving force of a few hundred
state induced by MA. Elucidation of the details of the MJ m&3, as compared with a driving force of irrevers-
nanoscale precipitation is important not only as basic ible recrystallization with a few MJ m&3 for 12Cr-ODS
materials science research but also as the develop- steels. By inducing reversible a–g transformations,
ment of high-strength engineering materials. 9Cr-ODS steel cladding for fast-reactor fuel elements
is currently being manufactured at the JAEA.
The microstructure of 9Cr-ODS steel cladding is
4.08.3 Martensitic 9Cr-ODS Steels basically tempered martensite. However, it has been
recognized that 9Cr-ODS steel cladding manufactured
4.08.3.1 Chemical Composition and
in an engineering process possesses a dual-phase struc-
Microstructure
ture that comprises both tempered martensite and
9Cr-ODS steels are being developed by the JAEA ferrite phases. An example of their microstructure is
( Japan Atomic Energy Agency) for application to shown in Figure 4. The ferrite phase appears white,
SFR fuel cladding. Their standard chemical compo- and the elongated phase is indicated by arrows. Their
sition is 9Cr–0.13C–0.2Ti–2W–0.35Y2O3 (wt%). size is about 30–60 mm in length (3–10 mm in width.
The chromium concentration was determined to be The formation of a ferrite phase in 9Cr-ODS steel is
9 wt% in terms of ductility, fracture toughness, and somewhat unusual, because only the full martensite
corrosion resistance based on a series of irradiation phase can be expected in 9Cr-ferritic steel without
data of ferrite steels. The addition of titanium pro- yttria under normalizing and air-cooling conditions.
duces the nanoscale dispersion of oxide particles, Moreover, the high-temperature strength of manu-
which leads to a markedly improved high-temperature factured 9Cr-ODS steel is significantly improved by
strength. If titanium is added to excess, however, it the presence of the ferrite phase.19–21 This is obvious
creates too much strength, which negatively impacts from the creep rupture data shown in Figure 5.20
Oxide Dispersion Strengthened Steels 245

200 1500
700 C L
1400 δ
1300 γ+δ
1200
Uni-axial stress (MPa)

1100

Temperature ( C)
γ+δ γ + TiC
1000 +TiC
100 A3
900 γ + TiC + M23C6
α
90 800 +TiC A1
80 Solid: with residual ferrite α + M23C6 + TiC
Open: without residual ferrite 700 α
70 600 +TiC
+Laves
60 500 α + M23C6 + laves + TiC
Mechanically milled without Y2O3
400
50 300
10 100 1000 10 000
0 0.1 0.2 0.3
Time to rupture (h)
C content (wt%)
Figure 5 Uni-axial creep rupture strength of 9Cr-ODS
steels at 700 C after the normalizing-and-tempering Figure 6 Computed phase diagram with respect to
(1050 C ( 1 h, Ar-gas cooling (AC) = > 780 C ( 1 h, AC) carbon content for 9Cr–xC–0.2Ti–2W system without Y2O3.
with and without residual ferrite. Reproduced from
Ohtsuka, S.; Ukai, S.; Fujiwara, M.; Kaito, T.; Narita, T.
Mater. Trans. 2005, 46, 487.
ferrite phase as well. The austenite g-phase transforms
to the martensite phase, but the ferrite phase remains
Therefore, the control of ferrite phase formation is a unchanged by quenching. Considering that the ferrite
key to the realization of high-temperature strength in phase is formed only in the specimens containing 0.35
9Cr-ODS steel cladding. and 0.7 wt% Y2O3, and that four types of ODS steels
have an identical chemical composition except for
Y2O3 content, the Y2O3 particles could suppress the
4.08.3.2 Residual Ferrite Formation and a–g reverse transformation.
Strength Characterization Figure 722 shows the results of dilatometric mea-
4.08.3.2.1 Mechanically alloyed powder surement when 9Cr–0.13C–2W–0.2Ti is heated
characterization without and with 0.35 wt% Y2O3. In the case of the
The computed phase diagram of the Fe–0.13C–2W– specimen without Y2O3, the linear thermal expansion
0.2Ti system without Y2O3 is shown in Figure 6 with begins to decrease from an AC1 point of 850 C to an AC3
respect to carbon content. For a carbon content point of 880 C, due to the reverse transformation of
of 0.13 wt%, a single austenite g-phase containing a–g-phase, which corresponds reasonably well with the
TiC carbide exists at a normalizing temperature of computed phase diagram. The addition of 0.35 wt%
1050 C. The equilibrium g/g + d-phase boundary Y2O3 induces an increase up to an AC3 point of
at this temperature corresponds to a carbon content 935 C. By comparing both curves, it was found that
of 0.08 wt%, beyond which d-ferrite is not stable. The the specimen with 0.35 wt% Y2O3 exhibits a smaller
specimens without and with 0.1 wt% Y2O3 exhibit the degree of reduction in linear thermal expansion during
full martensite structure, whereas the specimens with the reverse transformation of the a–g-phase; this obser-
0.35 and 0.7 wt% Y2O3 exhibit a dual phase compris- vation indicates that the entire a-phase could not be
ing both martensite and ferrite phases. Digital image transformed to a g-phase. This untransformed ferrite
analyses show that the area fraction of the ferrite phase phase was designated as a residual ferrite.
is )0.2 for specimens with 0.35 and 0.7 wt% Y2O3.
High-temperature X-ray diffraction measurement at 4.08.3.2.2 Pinning of a–g interface by oxide
950 C showed a considerable difference; the specimen particles
without Y2O3 shows diffraction peaks that correspond Alinger’s results indicate that the mechanically
only to the austenite g-phase, whereas specimens alloyed powder annealed at 700 C shows the smallest
with 0.35 and 0.7 wt% Y2O3 show diffraction peaks radius and highest density in Y–Ti complex oxide
corresponding not only to an austenite g-phase but to a particles,8 as shown in Figure 1. Considering that
246 Oxide Dispersion Strengthened Steels

1.4 12
1.3 F(0.7 mass % Y2O3)
1.2 Without Y2O3 10
AC1
Linear thermal expansion (∆L/L, %)

1.1
1 ∆G(0.2 mass % C)

Driving force (MJ m–3)


0.9 8
F(0.35 mass % Y2O3)
0.8
AC3
0.7 6
0.6
700 750 800 850 900 950 1000 1050 1100
4
F(0.1 mass % Y2O3)
1.4
1.3 0.35 mass % Y O 2
2 3
1.2
1.1 ∆G(0.13 mass % C)
1 AC1 0
0.9
0.8
0.7 AC3 –2
800 900 1000 1100 1200 1300
0.6
700 750 800 850 900 950 1000 1050 1100 Temperature (°C)
Temperature (ºC)
Figure 8 Comparison of the driving force (DG) for a to g
Figure 7 Results of linear thermal expansion reverse transformation derived by using Thermo-Calc
measurement between 700 and 1100 C at temperature code and pinning force (F ) due to oxide particles for 0.1
rising of 0.33 C s&1 for 0 mass % and 0.35 mass % Y2O3 in mass %, 0.35 mass %, and 0.7 mass % Y2O3 in
9Cr–0.13C–2W–0.2Ti specimens. Reproduced from Fe–0.13C–2W–0.2Ti specimens. Driving force (DG) for
Yamamoto, M.; Ukai, S.; Hayashi, S.; Kaito, T.; Ohtsuka, S. 0.13 mass % C and 0.2 mass % C is shown. Reproduced
Mater. Sci. Eng. A 2010, 527, 4418–4423. from Yamamoto, M.; Ukai, S.; Hayashi, S.; Kaito, T.;
Ohtsuka, S. Mater. Sci. Eng. A 2010, 527, 4418–4423.

Y2O3 particles are decomposed during MA, subsequent


annealing results in the formation and precipitation of
3sfp2=3
Y–Ti complex oxide particles at elevated temperatures F¼ ; ½3%
of 700 C or higher. Since the reverse transformation of 8r
a–g-phase takes place at a temperature over 850 C, where, s( Jm&2) is the interfacial energy between
which is higher than the precipitation temperature of a- and g-phases, and its value was selected to be
Y–Ti complex oxide particles, it is possible that the 0.56 J m&2.25 The character r represents the radius of
retention of the residual a-ferrite can be attributed the oxide particles (m) in the a-phase; its value was
to the presence of Y–Ti complex oxide particles in determined as 1.5 nm by using TEM observation. The
9Cr-ODS steels. These particles could block the character fp represents the volume fraction of dispersed
motion of the a–g interface, thereby partly suppres- oxide particles (&), and was derived on the basis of the
sing the reverse transformation from a- to g-phase. experimental evidence that oxide particles consist of
This section presents a quantitative evaluation of Y2Ti2O7. By substituting these values into the afore-
this process. mentioned equation, the value of pinning force F was
The chemical driving force (DG) for the reverse determined for 0.1, 0.35, and 0.7 wt% Y2O3, which are
transformation from a- to g-phase in the Fe–0.13C– also shown in Figure 8.22,23 The value of F increases
2W–0.2Ti system without Y2O3, can be evaluated in with the amount of Y2O3 added according to the
terms of Gibbs energy versus carbon content curves at relation of f 2=3 .
each temperature. These curves were derived using The velocity of the a–g interface motion (v) is
the Thermo-Calc code and the TCFE6 database. The proportional to the difference between F and DG, as
result of the calculation is presented in Figure 8.22,23 shown in the following equation:
The peak value of the driving force for the reverse
v ¼ MðDG & F Þ: ½4%
transformation from a- to g-phase reaches 4 MJ m&3
at 1000 C in the case of 0.13 wt% C. M is the mobility of the interface. DG and F are
The pinning force (F ) against the motion of competitive, and DG > F indicates a positive velocity
the a–g interface can be expressed as the following for the interface motion, that is, the reverse trans-
equation, which was derived from the modified Zener formation from a- to g-phase. On the other hand,
equation of Mishizawa et al.24 DG < F indicates that the a–g interface can be
Oxide Dispersion Strengthened Steels 247

Oxide particle 6.0


AC1 AC3

Hardness (GPa)
5.0

・ ・ ・ ・

・ ・
・ ・ ・ 4.0
γ ・ ・ ・

γ α 3.0
・ ・ ・ ・
・ ・ ・
α ・ ・ ・ 2.0
NT 550 ºC 750 ºC 800 ºC 800 ºC FC
1h 1h 7h 58 h
Carbide
Residual ferrite
Figure 9 Formation process of residual ferrite in Tempered martensite
Average covering residual
9Cr-ODS steel (Fe–0.13C–2W–0.2Ti–0.35Y2O3).
ferrite and tempered
Reproduced from Yamamoto, M.; Ukai, S.; Hayashi, S.; martensite
Kaito, T.; Ohtsuka, S. Mater. Sci. Eng. A 2010, 527,
4418–4423. Figure 10 Hardness change at room temperature as
a function of tempering conditions for the residual
ferrite and tempered martensite. NT: normalizing
pinned by oxide particles so that the a-phase is, and tempering; FC: furnace cooling. Ukai, S.; Ohtsuka, S.;
Kaito, T.; Sakasegawa, H.; Chikata, N.; Hayashi, S.; Ohnuki,
thus, retained. The results of the calculation shown S. Mater. Sci. Eng. A 2009, 510–511, 115–120.
in Figure 822 reveal that in the case of Y2O3 con-
tents of 0.35 and 0.7 wt%, the pinning force is larger
than the driving force for 0.13 wt% C. These results parameter of the tempering conditions.26 The decrease
are reasonably consistent with our observation of in hardness is significantly restricted in the residual
the retainment of residual ferrite during a–g reverse ferrite as compared to that of the martensite phase
transformation. in terms of increasing the tempering conditions.
On the basis of the aforementioned discussion, the The overall hardness measured by the micro-Vickers
formation process of the residual ferrite in Fe–0.13C– tester is also shown by the broken line which covers
2W–0.2Ti–0.35Y2O3 is schematically illustrated in both the residual ferrite and martensite, therefore,
Figure 9. At the AC1 point, the carbide begins to representing the average hardness of both phases.
decompose, and a–g inverse transformation takes Hardness Hv is correlated with yield stress sy using
place in the area of higher carbon content around the relationship provided by Tabor.27 For tempering
the decomposed carbide, where the driving force of conditions at 800 C for 58 h, which is equivalent to
the reverse transformation exceeds the pinning force tempering at 700 C for 10 000 h based on the LMP
because the carbon content may be >0.2 wt% (see (Larson–Miller parameter), hardness can be converted
Figure 8). The g-phase could be enlarged by these to yield stress at room temperature for the individual
processes. Approaching the AC3 point, the matrix phases: 1360 MPa for the residual ferrite and 930 MPa
carbon content achieves equilibrium at 0.13 wt%, for the tempered martensite. The yield strength of the
where the pinning force (0.35Y2O3) exceeds the residual ferrite is 1.5 times higher than that of mar-
driving force (0.13C), and the velocity of the a–g tensite at tempering at 700 C for 10 000 h.
interface motion is markedly reduced due to dragging A full ferrite ODS steel and full martensite ODS
by the oxide particles. Thus, the a-ferrite could be steel were manufactured, and the oxide particle dis-
retained even beyond the AC3 point. tribution in both ODS steels was measured by TEM.
The results are shown in Figure 11.28 It is obvious
4.08.3.2.3 Strength characterization that a few nanometer-sized oxide particles are finely
Nanoindentation measurements were conducted in distributed in the full ferrite ODS steel, whereas their
order to evaluate the mechanical properties of the size is coarsened in the bi-modal distribution in the
residual ferrite itself. The trace of a Berkovich tip can martensite ODS steel. Considering that the residual
be placed within the interiors of the residual ferrite ferrite phase belongs to full ferrite ODS steel, resid-
regions, while conventional micro-Vickers diamond ual ferrite contains fine (nanosized) oxide particles
tips using 100-mN loads cover 7 ( 7 mm. Figure 10 which are responsible for higher strength in residual
shows the hardness change in the individual phases ferrite containing ODS steels. In regard to the bi-
measured by this nanoindentation technique as a modal distribution of oxide particles in martensite
248 Oxide Dispersion Strengthened Steels

ODS steels, the a–g-phase transformation could control of 9Cr-ODS steels. Figure 12 exhibits a CCT
induce the coarsening of oxide particles by disturbing diagram that was experimentally constructed for
the interface coherency between these particles and 9Cr-ODS steel.21 The minimum cooling rate for the
the g-phase matrix. matrix phase in order to fully transform to martensite
is extremely higher in 9Cr-ODS steel (solid circular
symbol) than in mechanically milled EM10 (open
4.08.3.3 Cladding Manufacturing
diamond symbol) that does not contain added
4.08.3.3.1 Continuous cooling transformation Y2O3.29 Residual ferrite plays an important role in
diagram the process of continuous cooling transformation.
The preparation of a CCT (continuous cooling trans- The minimum cooling rate is known to increase
formation) diagram is essential to the microstructure with a decrease in the size of prior austenite (g) grains.
This smaller size of prior g grains provides more
nucleation sites (grain boundaries) for a g–a-phase
transformation, so that a higher cooling rate is
required to enable steel with small prior g grains to
fully transform to a. The presence of residual ferrite
restricts the growth of g grains; the prior grain size of
residual ferrite-containing steel is roughly 5 mm, thus
increasing the minimum cooling rate to produce a full
martensite matrix.
In steel that does not contain residual ferrite and
(a) 20 nm (b) 20 nm the mechanically milled EM10, the size of the prior
g grains is roughly 10 mm and 35 mm, respectively.
Figure 11 TEM photograph of the oxide particles: The results shown in Figure 12 can be explained
(a) finely distributed oxide particles in full ferrite ODS steel by the relationship between the size of prior g grains
and (b) bi-modal distribution of oxide particles with larger
size in the full martensite ODS steel. Yamamoto, M.; Ukai, S.; and the minimum cooling rate.21 As for the normal-
Hayashi, S.; Kaito, T.; Ohtsuka, S. J. Nucl. Mater. 2011, izing heat treatment used in commercial furnaces, the
417, 237–240. cooling rate would be roughly 3000 C h&1, so that

1300
1000
9Cr-ODS ferritic steel
900 Containing residual-α
No residual-α 1100
800 EM10

700 α+γ
Temperature, T ( C)

Temperature, T (K)

γ
900
600
α
500
700
400
γ+m α+γ+m
300
500
200 m
α+ m 30 (K/h)
100 18 000 (K/h) Austenite (γ) => Ferrite (α)
Austenite (γ) => Martensite 3000 (K/h)
300
101 102 103 104 105
Time from 800 C, t (s)
Figure 12 CCT diagram of 9Cr-ODS steel. Reproduced from Ohtsuka, S.; Ukai, S.; Fujiwara, M.; Kaito, T.; Narita, T. J. Nucl.
Mater. 2006, 351, 241.
Oxide Dispersion Strengthened Steels 249

Low carbon steel

Elemental
powders

Yttria
powder MA powder
9Cr–0.13C–2W–0.2Ti–0.35Y2O3 Mechanical alloying (MA)

Cladding tube At intermediate heat Cold rolling Hot extrusion (1423 K)


treatment (pilger mill)
At final heat treatment
Figure 13 Cladding tube manufacturing process developed for 9Cr-ODS steel.

the matrix phase of 9Cr-ODS steel cladding consists


of residual ferrite, martensite, and a small amount of 450 Cold rolling (Rd = 50%)
transformed ferrite from the g-phase. 1st 2nd 3rd 4th
Hardness (Hv)

4.08.3.3.2 Manufacturing process


9Cr-ODS steels are promising materials to enable 400
fast reactor fuel cladding to realize a high burnup
of 200 GWd t&1 at 700 C, since they have superior
350
radiation resistance and high temperature strength.
Mother 1st 2nd 3rd 4th
Figure 13 shows a series of manufacturing processes tube
of fuel cladding that is 8.5 mm in diameter by 0.5 mm Heat treatment
300
in thickness by 2 m in length. The element powders
and yttria powder are mechanically alloyed for 48 h in Figure 14 Hardness change in the process of cold rolling
and intermediate and final heat treatments for cladding tube
an argon gas atmosphere using an attrition type ball
manufacturing of 9Cr-ODS steels.
mill with a capacity of 10 kg batch. The mechanically
alloyed powders are sealed in hollow-shaped cans
and degassed at 400 C in a 0.1 Pa vacuum for 2 h. rate of about 150 K h&1 was applied to the intermedi-
The hollow shape of the bars is consolidated by hot- ate heat treatment in order to induce the ferrite phase
extrusion at an elevated temperature of 1150 C to at room temperature without martensite transforma-
the dimensions of 32 mm in outer diameter, 5.5 mm in tion. This phase has a lower degree of hardness. Hard-
wall-thickness, and 4 m in length. After machining ened cladding due to the accumulation of cold
to the precise dimensions, claddings are produced at deformation can be sufficiently softened by this inter-
their final dimension (8.5 mm in outer diameter, mediate heat treatment, and cold rolling can then be
0.5 mm in thickness, and 2 m in length) by four-pass continued with the softened ferrite structure. Figure 14
rolling with about a 50% reduction ratio on each pass represents the typical hardness change of 9Cr-ODS
by using a pilger mill. steel in the process of cladding manufacturing by
Without heat treatment, it is too difficult to man- repeated cold rolling and intermediate heat treatment.
ufacture cladding for ODS steels by the cold-rolling The elongated grain structure induced by the fourth
process. Using the CCT diagram of 9Cr–0.13C–2W– cold rolling can ultimately be made into equi-axed
0.2Ti–0.35Y2O3, as shown in Figure 12, a cooling grain structure by the final heat treatment, which
250 Oxide Dispersion Strengthened Steels

consists of normalizing at 1050 C for 1 h, followed by The ultimate tensile strength (UTS) of 9Cr-ODS
tempering at 800 C for 1 h. ferritic cladding in the hoop direction as measured
in a temperature range from room temperature to
850 C, is shown in Figure 16, together with the
4.08.3.4 Creep and tensile properties
corresponding data for the ferritic–martensitic stain-
The lifetime of a fast reactor fuel pin is most strongly less steel (PNC-FMS)19 that is conventionally used as
determined by the internal creep rupture strength fast reactor fuel cladding. The strength of 9Cr-ODS
of the cladding induced by the internal pressure of steel is superior to that of conventional PNC-FMS.
the fission gas at a temperature of around 700 C. For The uniform elongation that takes place from room
9Cr-ODS steel cladding, internal creep rupture data temperature to 900 C is also shown in Figure 16. In
at 650, 700, and 750 C are shown in Figure 15.30 the temperature range from 400 to 700 C at which a
Additionally, the best fit lines for hoop stress versus fast reactor is commonly operated, the measured
rupture time at each temperature are shown by solid uniform elongation exhibits adequate ductility. This
lines. These results confirmed that creep rupture advantage of superior elongation in the produced clad-
strengths in the hoop and longitudinal directions of dings can probably be ascribed to the pinning of dis-
cladding are almost the same, due to their equi-axed locations by oxide particles, which retard recovery and
grains. The corresponding creep rupture curves for sustain work-hardening.
HT931 and austenitic PNC31632 are also presented
for comparison. PNC316 is a typical austenitic clad-
ding developed by JAEA in the fast reactor program. 4.08.4 Ferritic 12Cr-ODS Steels
Notably, superior performance in rupture time is
4.08.4.1 Strength Anisotropy
shown in 9Cr-ODS steel cladding. The slope of
PNC316 is steeper, and there is a cross-over before When JAEA started to develop ODS steels in
1000 h at 750 C and before 10 000 h at 700 C. The 1985, the ferritic type of ODS steels was applied.3,33
stress condition of the fast reactor fuel pin gradually These are similar to MA957,34 which is single ferrite
increases due to the accumulation of fission gases and phase and does not include the martensite. Based on
reaches around 120 MPa at its final service milestone the results of R&D conducted for several years, three
of 75 000 h at 700 C. In this stress range, it is obvious kinds of claddings, 63DSA, 1DK, and 1DS, were
that 9Cr-ODS steel cladding is of advantage. manufactured in 1990. Their chemical compositions

500

400 PNC316 (923 K)

300 PNC316 (973 K)


Hoop stress (MPa)

PNC316
200 (1023 K)

9Cr-ODS
(923 K)
100 Stress range for
9Cr-ODS
90 SFR fuel cladding
(973 K)
80
70 HT9 (923 K) 9Cr-ODS
HT9 (973 K) (1023 K)
60
HT9 (1023 K)
50
10 102 103 104 105
Time to rupture (h)
Figure 15 Creep rupture curves of 9Cr-ODS steel claddings in hoop direction by using internally pressurized specimens
at temperatures of 650, 700, and 750 C, compared with those of HT9 and PNC316. Reproduced from Allen, T.; Burlet, H.;
Nanstad, R. K.; Samaras, M.; Ukai, S. Mater. Res. Soc. Bull. 2009, 34(1), 20–27.
Oxide Dispersion Strengthened Steels 251

are 13Cr–0.02C–3W–0.7Ti–0.46Y2O3 (63DSA), 13Cr– diameter, 0.4 mm thickness, and 1 m length. In the case
0.05C–3W–0.5Ti–0.34Y2O3 (1DK), and 11Cr–0.09C– of the 63DSA and IDS claddings, only six warm rolling
3W–0.4Ti–0.66Y2O3 (1DS). The manufacturing process passes at 650–700 C with intermediate annealing at
is almost the same as the process shown in Figure 13, 1100 C were conducted. The temperature of the final
except for the rolling process and intermediate heat heat treatment of 1DK cladding was 1150 C for 60 s,
treatment, because cold-rolling processing can be hardly and 63DSA and 1DS claddings at 1100 for 3.6 ks.
applied to these ODS steels. In the case of the 1DK The uni- and bi-axial creep rupture strengths of
cladding, six warm drawings at 800–850 C, followed by the manufactured claddings at 650 C are shown
four warm rolling passes at 500 C with intermediate in Figure 17, where the uni-axial corresponds to
annealing at 1080 C, were repeated to manufacture the the hot working direction and bi-axial belongs
thin-walled cladding in the dimension of 7.5 mm outer to the internal hoop direction.3 It was found that

1500 15
Tensile strength (MPa)

1000 Uniform elongation (%) 10

500 5
PNC-FMS
PNC-FMS

0 0
200 400 600 800 1000 1200 200 400 600 800 1000 1200
Temperature (K) Temperature (K)
Figure 16 Tensile strength and uniform elongation of 9Cr-ODS steel cladding in hoop direction by the ring specimens.
Reproduced from Ukai, S.; Kaito, T.; Otsuka, S.; Narita, T.; Fujiwara, M.; Kobayashi, T. ISIJ Int. 2003, 43, 2038.

1000
Uni-axial Bi-axial
1 DK
1 DS
600 63 DSA

500

400
Stress (MPa)

Uni-axial

300

200 Bi-axial

MA 957 Mol-ODS
(see figure 29) (DT2203Y05, see figure 28)
100
1 10 100 1000 10 000 100 000
Time to rupture (h)
Figure 17 Creep rupture strength of 1DK, 1DS, and 63DSA claddings in hoop direction by using internally pressurized
specimens at 650 C. Reproduced from Ukai, S.; Harada, M.; Okada, H.; Inoue, M.; Nishid, T.; Fujiwara, M. J. Nucl. Mater.
1993, 204, 65–73.
252 Oxide Dispersion Strengthened Steels

there is strong strength anisotropy, and the bi-axial von Mises’ equivalent stress was estimated for the
creep rupture strength is considerably lower than internal hoop stress. The unrecrystallized specimen
that of the uni-axial direction. Microstructure obser- shows significant strength anisotropy in uni-axial and
vations of these claddings exhibited the elongated internal creep rupture strength, whereas the recrystal-
grains like a bamboo structure in parallel to the lized specimens reveal decrease of anisotropy, where
working direction. The strength degradation in the uni-axial creep rupture strength decreases and inter-
bi-axial/internal hoop direction, which is essential nal strength approaches the uni-axial strength. These
for the fuel elements, should be mainly attributed to results demonstrate that the recrystallization process
the grain boundary sliding and crack propagation due adequately improves the creep rupture strength in
to stress concentration. the internal hoop direction. Furthermore, softening
by recrystallization makes it possible to manufacture
cladding by cold-rolling processing.
4.08.4.2 Recrystallization Tests
Based on the aforementioned finding in ODS steels,
4.08.4.3 Cold-Rolling Cladding
the recrystallization processing was extensively stud-
Manufacturing
ied to change the substantially elongated grain struc-
ture to the equi-axed grain structure. The Y2O3 Based on the aforementioned finding, cladding
content should be <0.25 mass % to attain the recrys- manufacturing tests were conducted using cold-
tallized structure in the ODS ferritic steels. The two rolling pilger mill in 12Cr-ODS ferritic steels with
types of 12Cr-ODS steels in the chemical composi- the limited Y2O3 content <0.25 mass % to induce the
tion of 0.23Y2O3–2W–0.4Ti (A3) and 0.34Y2O3–3W– recrystallized structure, and their internal creep rup-
0.4Ti (A15) were extruded at 1150 C, followed by ture properties were evaluated at 700 C, not at
60% cold rolling and annealed at 1200 C for 1 h. 650 C. The chemical composition of the manufac-
From these heat-treated bars, the internal creep rup- tured cladding is listed in Table 1, where the speci-
ture specimens were machined and tested at 650 C. mens are denoted as F1 to F4. The four levels of
Figure 1835 exhibits a comparison of the creep Y2O3 contents were selected in the range of <0.25
rupture strength of recrystallized (A3) and unrecrys- mass %, and the titanium content ranged from 0.13
tallized (A15) 12Cr-ODS steels at 650 C, where to 0.31 mass %. Cold rolling by a pilger mill was

Solid: uni-axial
600
Open: internal
500

400
Equivalent stress (MPa)

300

200

A3-2(1123 K extrusion)
A3-1(1423 K extrusion)
A15

100
10 100 1000 10 000
Time to rupture (h)
Figure 18 Creep rupture strength of recrystallized and unrecrystallized 12Cr-ODS steel claddings by using pressurized
specimens at 650 C. Reproduced from Ukai, S.; Nishida, T.; Okada, H.; Okuda, T.; Fujiwara, M.; Asabe, K. J. Nucl. Sci.
Technol. 1997, 34(3), 256–263.
Oxide Dispersion Strengthened Steels 253

Table 1 Chemical composition of F1–F4 specimens with different titanium and yttria contents (mass %) in 12Cr-ODS
steels

Specimen no. Chemical composition (mass %)

C Cr W Ni Ti Y2O3 O N Ar

F1 0.065 11.8 1.92 0.03 0.13 0.08 0.08 0.010 0.005


F2 0.054 11.8 1.94 0.03 0.13 0.13 0.05 0.010 0.004
F3 0.065 11.8 1.93 0.03 0.22 0.22 0.09 0.012 0.005
F4 0.056 11.7 1.92 0.03 0.31 0.24 0.04 0.010 0.004

Source: Reproduced from Ukai, S.; Okuda, T.; Fujiwara, M.; Kobayashi, T.; Mizuta, S.; Nakashima, H. J. Nucl. Sci. Technol. 2002,
39(8), 872–879.

F1 F2 F3 F4
Longitudinal

20 mm
Transverse

20 mm

Figure 19 Optical microstructure of the F1, F2, F3, and F4 specimens in the final claddings. Reproduced from Ukai, S.;
Okuda, T.; Fujiwara, M.; Kobayashi, T.; Mizuta, S.; Nakashima, H. J. Nucl. Sci. Technol. 2002, 39(8), 872–879.

repeated twice with a reduction ratio of about 50% 4.08.4.4 Internal Creep Rupture Property
per rolling. The intermediate heat-treatment to
The internal creep rupture properties of the manu-
soften the cold-rolled cladding was performed at
factured 12Cr-ODS steels at 700 C are shown in
1100 C for 30 min, and the final heat-treatment was
Figure 20.36 Increasing the yttria and titanium con-
performed at 1150 C for 0.5 h.
tents improves the internal creep rupture strength
Figure 19 shows the optical microstructures of
(F4 > F3 > F2 > F1). The uni-axial creep rupture
the manufactured claddings in the longitudinal and
strength for F4 is also plotted; there is the strength
transverse directions.36 All of the specimens seem
anisotropy between the uni-axial and internal hoop
to be recrystallized. However, the extent of recrys-
directions. This strength anisotropy can be associated
tallization depends on the yttria and titanium con-
with the slightly elongated grain structure shown in
tents. In the transverse cross-section, the grain size
Figure 19.
becomes slightly finer with increasing yttria and
The stress–strain rate relationship was investi-
titanium contents from F1 to F4. In the F4 speci-
gated for ODS ferritic claddings to evaluate the
men, the elongated grains can be still seen and the
creep deformation mode. The results of the analyses
aspect ratio in the longitudinal (L) and transverse
are given in the log–log plot in Figure 21.36 In
(T) directions is large, whereas the aspect ratio
general, the creep strain rate in the steady-state con-
of specimen F1 appears to be nearly unity. These
dition is expressed using applied stress s as:
findings show that F4 specimen with higher yttria
and titanium contents did not attain the perfectly "_ ¼ Asn ½5%
recrystallized grain structure by the annealing of where n is the stress exponent and A is the
1150 C for 0.5 h. temperature-dependent coefficient.37 In the case of
254 Oxide Dispersion Strengthened Steels

1000 mode of F4, and a higher strain rate is found even


F1 below a stress of 200 MPa. A transverse section of this
F2
F3
F4
specimen shows finely equi-axed grains of 5–10 mm
F4 (uni-axial, this work) (Figure 19). Apart from pinning the gliding dislocations
due to oxide particle-dislocation interaction, the defor-
Hoop stress (MPa)

Uni-axial
mation mechanism associated with grain morphology
may be the dominant factor that induced accelerated
strain in the hoop stress mode of the tubular specimen.
In order to characterize the high temperature
strength of manufactured 12Cr-ODS steel cladding,
100 Internal, bi-axial
direction
its strength mechanism was evaluated from the view-
point of the interaction between Y2O3 particles
and dislocations. This interaction could be formu-
lated by the void-hardening mechanism proposed
10 100 1000 10 000 by Srolovitz,38 in which oxide particles were re-
Time to rupture (h)
placed by voids. The oxide particle-hardening stress
Figure 20 The creep rupture strength in hoop direction sp can be evaluated by the following equation based
for pressurized F1 to F4 specimens at 700 C. Reproduced on Scattergood and Bacon’s equation,39 which takes
from Ukai, S.; Okuda, T.; Fujiwara, M.; Kobayashi, T.;
into account the interaction between the branches of
Mizuta, S.; Nakashima, H. J. Nucl. Sci. Technol. 2002,
39(8), 872–879. the bowed-out dislocation around a Y2O3 particle:
sp =G ¼ AMb=ð2plÞ½lnðD=r0 Þ þ B%; ½6%
for screw dislocation;
10−5
F1
F2 A ¼ ð1 þ v sin2 ’Þcos ’=ð1 & vÞ;
F3
F4 B ¼ 0:6
10−6 F4 (uni-axial)
PNC-FMS
for edge dislocation;
A ¼ 1 & v sin2 ’=ð1 & vÞ cos ’;
!
10−7
Strain rate (s−1)

B ¼ 0:7
10−8 where G is the Shear modulus, v is Poisson’s ratio, M
is the Taylor factor, b is the magnitude of Burgers
vector, and r0 is the inner cut-off radius of the dislo-
10−9 cation core. The value of ’ is the critical angle at
which the dislocation detaches from the particles.
This value was estimated to be ’ ¼ 46 for screw
10−10 dislocations and ’ ¼ 19 for edge dislocations. Fur-
50 60 70 80 90 100 200 300
Stress (MPa) ther, l is the average face-to-face distance between
Figure 21 Stress–strain rate relationship for internal creep particles on a slip plane and is given as a function of
of specimens F1–F4 and PNC-FMS, and for uni-axial creep the average particle radius rs and the average center-
of specimen F4 at 700 C. Reproduced from Ukai, S.; to-center distance ls between the particles by
Okuda, T.; Fujiwara, M.; Kobayashi, T.; Mizuta, S.;
Nakashima, H. J. Nucl. Sci. Technol. 2002, 39(8), 872–879. l ¼ 1:25ls & 2rs ; ½7%
where the averages are calculated by considering the
the uni-axial creep mode, a significantly high stress size distribution of the particles. The factor 1.25 is
sensitivity of n ¼ 43.7 appears. This stress exponent the conversion coefficient from regular square distri-
value is typical for an ODS alloy.37 The applied stress bution to random distribution.40 The characters ls and
that initiates the strain is clearly located around rs represent the results of the measurement of oxide
250 MPa; this stress corresponds to the so-called particles by means of TEM. D is the harmonic mean
threshold stress for deformation. On the other hand, of 2rs and l. The values of l were calculated, and
the stress exponent, n, is 10.4 for the internal creep the oxide particle-hardening stress was estimated by
Oxide Dispersion Strengthened Steels 255

substituting l, M ¼ 3.0,41 n ¼ 0.334, b ¼ 2.48 ( lower bars derived from an estimate of edge and screw
10&10 m, and G ¼ 50 600 MPa, at 700 C. dislocations, and with the uncertainty of r0 ranging
Figure 22 shows the results of analyses in rela- from b to (3 ( b). The measured stress in the
tion to the face-to-face distance between particles.36 uni-axial mode of the F4 specimen is shown by an
The oxide particle-hardening stress levels estimated open circle. These results imply that the higher oxide
by using the aforementioned equations at 700 C particle-hardening stress for specimen F4 is due to its
are represented by vertical bars, with the upper and shortened face-to-face particle distance l of 70 nm.
The lower band represents the stress corresponding
to a strain rate of 10&9 s&1 in the internal hoop direc-
400 tional mode. For the F1 specimen, as a stress level
corresponding to a strain rate of 10&9 s&1 approaches
350 the oxide particle-hardening stress, the strong anisot-
F4 F4 (uni-axially longitudinal)
ropy tends to disappear. However, for the F3 and
300
Oxide particle-hardening F4 specimens with a shortened distance between par-
Hoop stress (MPa)

250
stress (s p) from particle ticles, stress levels for a strain rate of 10&9 s&1 in the
distribution by TEM
hoop direction are degraded from the oxide particle-
F3
200 hardening stress. The strong anisotropy still remains in
the F4 specimen. The accelerated deformation in the
F2
150 internal hoop direction could be the result of grain
F1 boundary sliding, since finely equi-axed grains with a
100
small size of 5–10 mm are formed, and the grain bound-
Stress at strain rate of aries occupy a large fractional area in the transverse
50 10−9 S−1 in the internal
hoop direction cross-section of the F4 specimen (see Figure 19).
0 Based on these results, it seems to be difficult to control
0 50 100 150 200 250 300
internal creep rupture strength by recrystallization
Face-to-face distance between particles, l (nm)
processing in 12Cr-ODS steel cladding.
Figure 22 Comparison of oxide particle-hardening stress
estimated from dispersion parameters of F1, F2, F3, and
F4 specimens, uni-axially longitudinal creep strength of
F4 specimen, and internal creep strength in hoop direction
4.08.5 Al-Added 16Cr-ODS Steels
at a strain rate of 10&9s&1 for F1, F3, and F4 specimens, 4.08.5.1 Application and Technical Issues
as functions of face-to-face distance between particles.
Each stress was obtained at 973 K. Note that internal creep Generation IV advanced nuclear power systems are
strength is located below the oxide particle-hardening proposed; the temperature and dose regimes for their
stress due to the grain boundary sliding in the hoop stress
mode. Reproduced from Ukai, S.; Okuda, T.; Fujiwara, M.; operation are shown in Figure 23.42 Among them, the
Kobayashi, T.; Mizuta, S.; Nakashima, H. J. Nucl. Sci. supercritical water-cooled reactor (SCWR) and the
Technol. 2002, 39(8), 872–879. lead fast reactor (LFR) require a higher neutron dose

1400
1200
Temperature ( C)

1000
VHTR GFR
800 MSR
600 SFR
SCWR LFR
400
200
Generations II–III
0
0 50 100 150 200
Displacement per atom
Figure 23 Temperature and dose regimes for Generation IV advanced nuclear power plants. VHTR: very high temperature
reactor; SCWR: supercritical water-cooled reactor; GFR: gas fast reactor; LFR: lead fast reactor; MSR: molten salt reactor;
SFR: sodium fast reactor. Reproduced from Guerin, Y.; Was, G. S.; Zinkle, S. J. Mater. Res. Soc. Bull. 2009, 34(1), 10–14.
256 Oxide Dispersion Strengthened Steels

at an operating temperature of 600 C. It is known that 0.8


9Cr-ODS steels have superior compatibility with Aged at 500 C
0.7
sodium, but their corrosion resistance is not adequate
for SCPW and LBE at a temperature >600 C. Thus, 0.6 As-received

Absorbed energy (J)


the most critical issue for the application of 9Cr-ODS
steels to SCWR and LFR is to improve their resistance 0.5
to corrosion. 0.4
It has been reported that the addition of chromium
(>13 wt%) and aluminum (4 wt%) to ODS steels quite 0.3
effectively suppresses corrosion in an SCPW and LBE 1000 h 100 h
0.2
environment. In general, however, an increase in the Cr 4300 h
content often results in increased susceptibility to ther- 0.1 10 000 h
mal aging embrittlement. Furthermore, the addition of
Al significantly reduces steel strength at high tempera- 0.0
12 14 16 18 20 22
tures. Recent progress in R&D of high Cr–Al-added
Cr content (mass%)
ODS ferritic steels is summarized in the proceedings
of the International Conference of Advanced Power Figure 24 Aging embrittlement of high Cr-ODS steels
with respect to Cr content. Absorbed fracture energy was
Plants (ICAPP) 2009. The oxidation and corrosion
measured at room temperature with the use of miniaturized
performance of Al-added 16Cr-ODS steels in SCPW Charpy V-notch (CVN) specimen which measures 1.5 mm
and LBE environments is described in Section 4.08.7. square with 20 mm length. Reproduced from Kimura, A.;
Kasada, R.; Iwata, N.; et al. In Proceedings of ICAPP ’09,
Tokyo, Japan, May 10–14, 2009; Paper 9220.
4.08.5.2 Thermal Aging Embrittlement Due
to High Cr Content
High Cr concentration often increases susceptibility tensile strength of high Cr-ODS steels. A decrease in
to aging embrittlement through the formation of UTS versus Al content is obvious in 16Cr-ODS steels;
Cr-rich secondary phases. The trade-off between cor- this dependency becomes weaker at higher tempera-
rosion resistance and aging embrittlement caused ture. The effect of Cr concentration on the UTS is not
by increasing Cr content is one of the critical issues so obvious between 13.7 and 17.3 wt% at 450 and
facing the developers of high-Cr ODS steels. The aging 700 C.
effects of ODS steels with different Cr content were In the case of 9Cr–12Cr-ODS steels, high-
investigated by measuring their impact fracture energy temperature strength is considerably enhanced by the
at RT after aging at 500 C up to 10 kh. The results are uniform dispersion of Y–Ti complex oxide (Y2Ti2O7)
shown in Figure 24.43 The fracture energy decreases particles. In Al-added high Cr-ODS steels, however,
with increasing Cr content before aging. Aging, then, Y–Al complex oxides and/or Al oxides are formed
causes a reduction in the fracture energy. ODS steels rather than Y–Ti complex oxides, which leads to
with a Cr content >18 wt% show a significant reduc- larger oxide particles, causing a degradation of high-
tion in fracture energy after aging for 100 h. In contrast, temperature mechanical strength. Therefore, Hf or Zr,
16Cr–4Al ODS steel showed a small reduction in frac- which form thermodynamically stable oxides, were
ture energy even after aging for 10 kh. Microstructure added to form Y–Hf or Y–Zr complex oxide particles
observation by TEM revealed that fine secondary rather than Y–Al complex oxide. The process of
phases were formed in high density after aging for manufacturing these materials is exactly the same as
1000 h at 500 C. These secondary phases are consid- that used for 9Cr-ODS ferritic steels: MA by an attri-
ered to be Cr-rich phases. In order to reduce suscepti- tion type ball mill and hot extrusion at a nominal
bility to aging embrittlement, the Cr content could temperature of 1150 C. The extruded bars were
be <16 wt%. provided for mechanical tests. The creep rupture prop-
erties of Al-added high Cr-ODS steels are summar-
ized in Figure 26.44 The creep strength of the standard
4.08.5.3 Mechanical Properties
steel is generally lower than that of Al-free steel. On
The addition of Al to ODS ferritic steels sometimes the other hand, the addition of Zr and Hf induces
softens their creep and tensile properties. Figure 2544 an improved creep strength which approaches that of
shows the effects of Al addition and Cr content on the Al-free steel. Furthermore, it was found that their
Oxide Dispersion Strengthened Steels 257

1400 1000
Fe–16Cr–0W–0.1–0.35Y2O3 700 C Fe–4Al–0W–0.1–0.35Y2O3
1200 700 C
450 C 800 450 C
1000
UTSave. (MPa)

UTSave. (MPa)
800 600

600 400
400
200
200
0 0
0 1 2 3 4 12 13 14 15 16 17 18 19
Al concentration (mass %) Cr concentration (mass %)
Figure 25 The UTS with respect of Al and Cr content in Al added high Cr-ODS steels at 450 C and 700 C. Reproduced
from Furukawa, T.; Ohtsuka, S.; Inoue, M.; et al. In Proceedings of ICAPP ’09, Tokyo, Japan, May 10–14, 2009; Paper 9221.

103 and recrystallization annealing, which is the same


700 C
as 12Cr-ODS steel described in Section 4.08.4.
From mother tubes of 16Cr–2W–0.1Ti–4Al–
0.35Y2O3 with an 18 mm outer diameter and 3 mm
thickness, cold-rolling was repeated four times using a
Stress (MPa)

pilger mill; the reduction rate for each rolling reached


102 about 45%. Annealing is required to soften the
deformed structure in order to make possible the
0Al (16Cr–0.1Ti–0.35Y2O3)
next round of cold rolling. A recrystallized structure
Standard (15Cr–4Al–2W–0.1Ti–0.35Y2O3)
cannot be reproduced by the final heat treatment once
0.63Zr (15Cr–4Al–2W–0.63Zr–0.1Ti–0.35Y2O3) recrystallization has taken place in the intermediate
0.62Hf (15Cr–4Al–2W–0.62Hf–0.1Ti–0.35Y2O3) annealing.45 A heat treatment test was therefore con-
ducted to determine an appropriate annealing tem-
10
perature to induce softening by dislocation recovery
101 102 103 104 105
Time to rupture (h)
but without recrystallization. The temperature for the
final heat treatment was selected in order to realize
Figure 26 Creep rupture strength of various Al added recrystallized grains.
high Cr-ODS steels in hoop direction by using pressurized
specimens at 700 C. Reproduced from Furukawa, T.;
Figure 27 illustrates an orientation image map
Ohtsuka, S.; Inoue, M.; et al. In Proceedings of ICAPP ’09, (OIM) measured after heat treatment at 900 C for
Tokyo, Japan, May 10–14, 2009; Paper 9221. 1 h, 1000 C for 1 h, and 1150 C for 1 h, from which
an evolution of recrystallization can be clearly identi-
fied in the cladding.46 The primary recrystallization
fracture elongation and reduction of area are slightly
begins with an {111}<112> orientation, which is desig-
higher than those of Al-free steel. Under microstruc-
nated by the color blue at 900 C annealing. When
tural observation by TEM, oxide particles consisting of
the temperature increases to 1000 C, a {110}<100>
Y3A5O12, YAlO3, and Al2O3, were observed in typical
Goss orientation, designated by green, partially
Al-added ODS steel, whereas this Y–Al complex oxide
appears. The primary recrystallization orientation of
can be changed to Y2Hf2O7 in Hf-added ODS steel.
{111}<112> is almost replaced by a {110}<100>
The improved creep rupture strength in Hf-added
Goss orientation when the temperature is elevated to
ODS steel could be attributed to the nanosize disper-
1150 C, which corresponds to a secondary recrystalli-
sion of the Y2Hf2O7 complex oxide.
zation. Thus, the final heat treatment was conducted
at 1150 C for 1 h in order to create a perfectly recrys-
tallized structure. The manufactured cladding has sec-
4.08.5.4 Cladding Manufacturing
ondary recrystallized grains with a {110}<100> Goss
16Cr–4Al-ODS steels exhibit a full ferrite structure orientation, which was formed by annealing from the
without a–g-phase transformation, and the manufacture cold rolled {112}<110> orientation through primary
of their cladding is effected by means of cold-rolling recrystallization of the {111}<112> orientation.47
258 Oxide Dispersion Strengthened Steels

4.08.6 Existing ODS Steel Cladding commercialized or are under development. The first
group includes Incoloy MA956 and PM2000. The
The basic chemical composition of the representative former is produced by what was formerly the Inter-
ODS steels is summarized in Table 2. ODS steels are national Nickel Company (INCO) and is now the
divided into two groups which have either been Special Metals (SM) Company. The latter is a prod-
uct of the Plansee Company of Austria. MA956 and
PM2000 are 20% Cr-ODS steels containing 5% Al,
which exhibit superior resistance to oxidation and
corrosion in hot gases at temperatures >1000 C.
Tubes, sheets, and bars made from these steels are
commercially used in various stationary and high-
temperature components in turbines, combustion
chambers, diesel engines, and burners.
The second group is devoted to the application of
fuel cladding for nuclear fast reactors, anticipating its
superior resistance to radiation resistance, and its
excellent creep strength and dimensional stability at
an elevated temperature of 700 C. As shown in
Table 2, DT2906 contains Ti2O3 dispersoids, and
DT2203Y05 is strengthened by Ti2O3 and Y2O3. Both
steels have been developed by SCK+CEN (Centre
d’Etude de l’énergie Nucleaire – Studiecentrum voor
Kernenergie) Mol (Belgium).48–50 The elementary
metallic powders and Y2O3 or TiO2 powder are
mechanically alloyed by means of a pilot scale ball
mill with a capacity of 9.2 kg per batch. Mechanically
alloyed powders are hot-compacted into billets,
200 µm 200 µm 200 µm which are subsequently hot-extruded into the hol-
(a) (b) (c) lows of 20/17 mm. A plug drawing is applied to
manufacture the cladding tube from the hollows.
Figure 27 OIM for 16Cr–4Al–ODS steels heat-treated at
Intermediate annealing is carried out at 1050 C by
900 C for 1 h (a), 1000 C for 1 h (b), and 1150 C for 1 h (c).
Reproduced from Ukai, S.; Ohnuki, S.; Hayashi, S.; et al. using induction heating after a certain number of
In Proceedings of ICAPP’09, Tokyo, Japan, May 10–14, drawing passes. The entire cold drawing is composed
2009; Paper 9232. of 15–20 passed and three intermediate annealing

Table 2 Basic chemical composition of ODS steels (mass %)

Steels Cr Mo W Ti Al Dispersoid Fe Others Development

Turbine, combustion
Incoloy MA956 20 – – 0.5 4.5 0.5Y2O3 Bal SM/US
PM2000 19 – – 0.5 5.5 0.5Y2O3 Bal Plansee/Austria
Fast reactor fuel
DT2203Y05 13 1.5 – 2.2 – 0.5Y2O3, Bal SCK+ CEN
0.9Ti2O3 Mol/Belgium
DT2906 13 1.5 – 2.9 – 1.8Ti2O3 Bal SCK+ CEN
Mol/Belgium
Incoloy MA957 14 0.3 – 1 – 0.25Y2O3 Bal SM/US
9Cr-ODS steel 9 – 2 0.2 – 0.35Y2O3 Bal 0.13C, martensite + JAEA/Japan
residual ferrite
12Cr-ODS steel 12 – 2 0.3 – 0.23Y2O3 Bal JAEA/Japan
16Cr–4Al-ODS steel 15.5 – 2 0.1 4 0.35Y2O3 Bal 0.6Hf or 0.6Zr KU/Japan

SM: Special metals, former International Nickel Company; JAEA: Japan Atomic Energy Agency; KU: Kyoto University.
SCK+CEN: Centre d’Etude de l’énergie Nucleaire – Studiecentrum voor Kernenergie.
Oxide Dispersion Strengthened Steels 259

steps. The final annealing is performed at 1050 C and transverse direction, but with highly elongated grains
800 C to precipitate an w-phase (70%Fe, 15%Cr, 7% with a bamboo-like structure in the longitudinal or
Ti, and 6%Mo). More than 1000 cladding tubes were working direction. Therefore, it turned out that the
manufactured. For defect control, this cladding is non- creep rupture strength of MA957 cladding is signifi-
destructively tested using eddy currents and ultrason- cantly degraded in the hoop direction, which is essen-
ics which employ specified artificial reference defects tial for fuel pins. Some of the stress rupture data are
which define the rejection level for naturally defective shown in Figure 29.52 The pulsed magnetic welding
cladding. For example, the creep rupture strength of (PMW) method was developed in the United States
DT2203Y05 cladding in the hoop direction is shown for MA957 for the manufacture of fuel elements.
in Figure 28.51 For the fabrication of fuel pins with
DT2203Y05 cladding, a special resistance welding
machine was designed at SCK+CEN, because ODS 4.08.7 Corrosion and Oxidation
steels can hardly be welded by conventional fusion
4.08.7.1 Sodium Compatibility
welding methods such as tungsten inert gas (TIG) or
electron beam welding, since they result in an oxide It is essential to evaluate the environmental effects of
particle-free zone. Fuel and blanket pellets were filled sodium on the mechanical strength properties of ODS
into the cladding, and resistance welding with an end- steels to ensure their structural integrity throughout
plug was performed in a glove box at Belgonucleaire. their design life-time in SFR. ODS-steels basically
The two fuel assemblies were fabricated for Phenix display superior compatibility with sodium. For 9Cr-
irradiation. ODS steel (M93) and 12Cr-ODS steel (F95), which are
Incoloy MA957 was developed by the International potential cladding materials for SFR, their UTS at
Nickel Company (INCO) for application to fast reac- 700 C after exposure to sodium in a stagnant state
tor fuel cladding. It is strengthened by a very fine, is shown in Figure 30.53 Both show almost constant
uniformly distributed yttria dispersoid. Its fabrication strength after exposure to sodium, and it was confirmed
involves a MA process and subsequent extrusion, that there is no degradation up to 10 000 h. For conven-
which ultimately results in a highly elongated grain tional ferritic steel without Y2O3, a clear strength
structure. An extruded bar with a diameter of 25.4 mm reduction occurs above 600 C due to decarburization
was gun-drilled in order to generate a tube hollow phenomena in sodium. ODS steel does not show such a
with a 4.75 mm thick wall. Extensive cladding fabrica- clear strength reduction because the fine Y2O3 oxide
tion tests were conducted on the tube hollow using a particles remain stable in steel, thereby maintaining the
rolling and plug draws in the United States, France, strength of the steel.
and Japan. It can be said that MA957 is too hard to Figure 31 shows the results of creep-rupture tests
perform satisfactorily on a small scale without faults. with internally pressurized specimens in a stagnant
The structure of the fabricated MA957 cladding sodium environment.54 The creep-rupture strength
is highly anisotropic with equi-axed grains in the of 9Cr-ODS steel (M11) in sodium is equal to its
strength in air, and no impact from a sodium environ-
ment was observed. However, under a flowing sodium
1000 condition of 4.5 m s&1, the element nickel penetrates
the surface of ODS steel cladding, where an increase in
Hoop stress (N mm–2)

nickel concentration and decrease in chromium con-


600 C
centration were observed at 700 C. These results
650 C suggest that the effects of a sodium environment can
100 700 C
750 C
be ignored under stagnant conditions; however, as
fuel cladding is utilized in an environment with a
high flow rate of sodium, the effects of the microstruc-
ture change associated with nickel diffusion into the
10 cladding surface need to be considered.53
10 100 1000 10 000
Time to rupture (h)
4.08.7.2 LBE Compatibility
Figure 28 Creep rupture strength of DT2203Y05 cladding
in a hoop direction. Reproduced from Huet, J. J.; Coheur, L.; Molten LBE has a high solubility of nickel, iron, and
De Bremaecker, A.; et al. Nucl. Technol. 1985, 70, 215–219. chromium, which are the most important alloy elements
260 Oxide Dispersion Strengthened Steels

10 000

STC STC
TR PNC ORT
650 C
700 C
760 C

1000
s (MPa)

650 C
704 C
100
760 C →
→ →

→ →

10
10 100 1000 10 000
tr (h) 39011045.6

Figure 29 Creep rupture strength of Incoloy MA957 cladding. Reproduced from Hamilton, M. L.; Gelles, D. S. PNNL-13168,
Feb 2000.

80
M93: 650 C M93: 700 C
60 F95: 650 C F95: 700 C
Sodium flow < 0.001 m s–1
UTS

40

20
As-received

0
0 200 400 600 800 1000 1200
Sodium exposure time (h)
Figure 30 UTS of 9Cr-ODS steel (M93) and 12Cr-ODS steel (F95) in hoop direction after sodium exposure.
Reproduced from Yoshida, E.; Kato, S. J. Nucl. Mater. 2004, 329–333, 1393–1397.

in austenitic stainless steels. Thus, nickel super alloys alloy flow velocity in the loop is 1.2 m s&1, and oxygen
and austenitic stainless steels cannot be used as the sensors were used to measure and maintain an oxygen
structural materials for LBE-cooled systems, especially concentration of about 1 ( 10&6 wt%. Samples were
at temperatures >500 C. Ferritic steels have been con- exposed for 200, 400, and 600 h, in order to study the
sidered more appropriate for LBE application. early stages of oxide formation and growth. A cross-
Exposure of 9Cr-ODS steels to an LBE environ- sectional view and the distribution of elements are
ment at 530 C was carried out in the DELTA Loop shown in Figure 32.54 In a short time, the 9Cr-ODS
of the Los Alamos National Laboratory. The molten steel formed a protective duplex oxide layer consisting
Oxide Dispersion Strengthened Steels 261

1000
Sodium flow rate; < 0.001 m s–1 Material: M11
650 C 700 C : In air/Ar
Hoop stress (MPa)
650 C 700 C : In sodium

650 C

700 C

100
10 100 1000 10 000 100 000
Time to rupture (h)
Figure 31 Creep-rupture strength of 9Cr-ODS steel (M11) in hoop direction under sodium exposure at 650 C and 700 C.
Reproduced from Yoshida, E.; Kato, S. J. Nucl. Mater. 2004, 329–333, 1393–1397.

OKa, 41 PbMb, 124


Bulk
Oxide

CrKa, 26 FeKa, 64

Diff.
zone
LBE

15 kV X 2000 10 µm 24 36 BEC

Figure 32 Backscatter cross-section secondary electron microscopy (SEM) image and Energy dispersive X-ray
spectrometry (EDS) map of 600 h 9Cr-ODS steel, showing much thinner Cr-rich oxide but a thicker diffusion zone.
Reproduced from Machut, M.; Sridharan, K.; Li, N.; Ukai, S.; Allen, T. J. Nucl. Mater. 2007, 371, 134–144.

of an outer magnetite (Fe3O4) layer and an inner Fe–Cr It has been reported that the addition of aluminum
spinel ((Fe,Cr)3O4) layer, which is sometimes accom- to steel effectively prevents LBE corrosion. Figure 33
panied by an O-enriched and Fe-depleted diffusion shows the appearance of ODS steel specimens after
zone at the oxide–bulk interface. Over time, the outer a corrosion test in LBE for 1 ( 104 h at 650 C.43
magnetite layer is removed and the underlying spinel The 18 wt% Cr-ODS steel without the addition of
layer serves to mitigate more catastrophic corrosion Al dissolved markedly into LBE, while those ODS
degradation such as dissolution and liquid metal attack specimens containing 4 wt% Al almost completely
along the grain boundaries. Very thin oxides are maintained their shape even in Al-added 14Cr- and
not particularly protective in regard to loss of metal, 16Cr-ODS steels, indicating a very high resistance to
as manifested by the thick diffusion zones associated LBE corrosion. It is noteworthy that this corrosion
with them. Furukawa pointed out that at tempera- resistance was independent of Cr concentration from
tures above 600 C, the thickness of the oxide layer 13 to 19 wt% in Al-added ODS steels. From the
diminishes with increasing temperature. This behavior distribution of elements across the cladding surface,
can be ascribed to a change in the stable form of we deduce that LBE corrosion can be prevented
iron oxide from magnetite to wustite at 570 C. Beyond by the formation of an Al enriched film.56 It was
this temperature, dissolution attack was observed demonstrated that Al-added 16Cr-ODS steel (16Cr–
on some portions of 9Cr-ODS steel, and the oxide 2W–4Al–0.1Ti–0.35Y2O3) has superior corrosion resis-
layer’s adhesion to the material began to weaken.55 tance at 650 C for 5000 h.
262 Oxide Dispersion Strengthened Steels

14Cr–4Al 16Cr–4Al 6

With 4 wt% Al

Weight gain (mg dm–2)


1800 h
4

600 h
2
100 h

0
14 16 18
(a) Cr content (wt%)

21
18Cr 19Cr–4Al 1800 h
Figure 33 The appearance of Al added high Cr-ODS steel With 16 wt% Cr

Weight gain (mg dm–2)


specimens after corrosion test in LBE for 1 ( 104 h at
923 K (DO: 1 ( 10&6 wt%). Reproduced from Kimura, A.; 14
Kasada, R.; Iwata, N.; et al. In Proceedings of ICAPP ’09, 600 h
Tokyo, Japan, May 10–14, 2009; Paper 9220.

7
100 h
4.08.7.3 SCPW Compatibility
Figure 3457 shows the effects of Cr and Al content on
0
the weight gain of ODS ferritic steels after exposure to 0 2 4
SCPW at 500 C with 8 ppm of dissolved oxygen. (b) Al content (wt%)
Increasing the Cr content from 14 to 17 wt % does Figure 34 Weight gain of Al added high Cr-ODS steels
not affect corrosion resistance if ODS ferritic steels with Cr content (a) and Al content (b) after exposure to
contain 4 wt % Al. For 16 wt% Cr, the addition of A1 SCPW at 500 C with 8 ppm of dissolved oxygen under a
increases corrosion resistance in 16Cr-ODS steels. As pressure of 25 MPa (10 dm = 1 m). Reproduced from
shown in Figure 35,43 tested at SCPW (510 C, Lee, J. H.; Kimura, A.; Kasada, R.; et al. In Proceedings of
ICAPP ’09, Tokyo, Japan, May 10–14, 2009; Paper 9223.
25 MPa) for 600 h, the addition of 4 wt% Al did not
significantly influence corrosion resistance in 19Cr-
ODS steel, though a rather dense chromia film was
observed on the specimen surface. The 16 wt% Cr is controlled atmosphere of dry air. Weight measure-
not large enough to form homogeneous and stable ment to evaluate the degree of oxidation was per-
chromia on the entire surface of the specimen, whereas formed at intervals of 50, 100, 400, 1000, and 2000 h,
a very thin alumina film covers the entire surface of the at temperatures of 650, 750, and 850 C. The results
specimen with the Al addition of 2 wt%. Thus, the of the measured weight gain due to oxidation at
addition of Al effectively improves corrosion resis- 750 C are shown in Figure 36.58 For 9Cr-ODS
tance in 16Cr-ODS steel. As shown in a comparison and 12Cr-ODS steels, the weight gain due to oxida-
with 9Cr-ODS steel in Figure 35, its weight gain is tion was quite small and comparable to that of
much larger than 16Cr-ODS steel, indicating that PNC316 containing 17 wt% Cr. Their weight gain
9Cr-ODS steel is not adequate for application to is limited to below 0.1 mg mm&2. On the other hand, a
SCWR. The suppression of SCPW corrosion by the quite large oxidation of 0.8 mg mm&2 was observed in
addition of Al to 16Cr-ODS steel is due to the for- PNC-FMS. The measured results on SUS430, which
mation of a very thin alumina film on the surface. show a greater weight gain than that of ODS steels,
show that advanced oxidation resistance is attained
with ODS steels, even when compared to higher
4.08.7.4 Oxidation
17 wt% Cr containing stainless steel.
Oxidation tests for 9Cr-ODS and 12Cr-ODS steels The element distribution obtained by Electron
were performed using pickled specimens in a probe microanalysis (EPMA) showed a scale consisting
Oxide Dispersion Strengthened Steels 263

30.0

Weight gain by oxidation (mg mm–2)


0.10
SCW 923 K 3 50 h
20.0 773 K.25 MPa. 600 h 0.08

SUS430 9Cr-ODS 0.06 Y2O3 effects


10.0 (16Cr)
Weight gain (g m–2)

0.04
1.2 Cr supply through
grain boundary diffusion
0.02
16Cr-ODS
0.00
12Cr–ODS 12Cr–ODS PNC–FMS
0.6
(fine grain) (large grain)
Figure 37 Weight gain of 12Cr-ODS steel and PNC-FMS
oxidized at 650 C for 50 h. Reproduced from Kaito, T.;
19Cr-ODS Narita, T.; Ukai, S.; Matsuda, Y. J. Nucl. Mater. 2004,
0 329–333, 1388–1392.
0 2 4
Al content (mass %)
In oxidation tests, Fe, which is a major constituent in
Figure 35 The dependence of the weight gain on the Cr
and Al contents in 16Cr- and 19Cr-ODS steels. SUS430 is
steel, tends to be easily oxidized at an early stage, but
a ferritic steel containing 16 mass % Cr and 4 mass % Al. further oxidation can be suppressed by the formation of
Reproduced from Kimura, A.; Kasada, R.; Iwata, N.; et al. a protective a-Cr2O3 layer. This a-Cr2O3 formation
In Proceedings of ICAPP ’09, Tokyo, Japan, May 10–14, is generally controlled by the rate at which Cr is
2009; Paper 9220. supplied to the reaction front. It is known that a high
Cr content in steel, as well as an increasing diffusion
flux through the grain boundary, that is, finer grains,
Weight gain by oxidation (mg mm–2)

1.0 accelerates both the Cr supply and the formation of


9Cr-ODS
12Cr-ODS (fine grain)
a-Cr2O3. A short-term oxidation test, whose results are
0.8 PNC316 shown in Figure 37, was conducted to investigate the
PNC-FMS
SUS430
mechanism of suppressing oxidation in ODS steels.58
0.6 The decrease in oxidation in fine grain 12Cr-ODS
ferritic steel can be attributed to the enhanced rate at
0.4 which Cr was supplied throughout the accelerated
grain boundary diffusion. In both cases of fine/large
0.2 grains in 12Cr-ODS steels, Raman spectroscopy
detected protective a-Cr2O3 at the interface between
0.0
0 500 1000 1500 2000 2500 the matrix and scale. Comparing 12Cr-ODS large
Testing time (h) grain and PNC-FMS, the Cr content is similar, and
Figure 36 Weight gain of 9Cr-ODS and 12Cr-ODS steels
the grain size is rather smaller in PNC-FMS. Never-
by the oxidation at 750 C. Reproduced from Kaito, T.; theless, protective a-Cr2O3 cannot be detected by
Narita, T.; Ukai, S.; Matsuda, Y. J. Nucl. Mater. 2004, Raman spectroscopy, and oxidation is enhanced in
329–333, 1388–1392. PNC-FMS, implying that the suppression of oxidation
in 12Cr-ODS with large grains could be due to the
effects of the Y2O3 oxide particles themselves. Chen
of Fe-rich oxide in the outer layers and Cr-rich oxide in
et al. showed some TEM images of Y-rich oxides on
the inner layers. At the interface between ODS steel
grain boundaries that may be part of the explanation.59
and the oxide scale, there was a thin layer (a few
micrometers) of further Cr-enriched oxide. Raman
spectroscopy measurement indicated that the outer 4.08.8 Irradiation
Fe-rich and inner Cr-rich layers correspond to
4.08.8.1 Simulated Irradiation
a-Fe2O3 and spinel type (Fe, Cr)3O4, respectively.
It was also confirmed that a-Cr2O3 is formed at the Testing that involves the simulated irradiation of
matrix–scale interface. 9Cr-ODS steel was conducted by Allen et al. at the
264 Oxide Dispersion Strengthened Steels

Environmental and Molecular Science Laboratory at ejection of atoms alone cannot be responsible for the
Pacific Northwest National Laboratory, using 5 MeV loss of diameter in oxide particles. Free point defects
Ni ions at 500, 600, and 700 C with a damage rate of and their diffusion-based mechanism are therefore
1.4 ! 10"3 dpa s"1. The results regarding measured of major importance and play a dominant role in
particle size distribution as a function of dose are the dissolution of oxide particles.
plotted in Figure 38 for irradiation at 500, 600, and
700 C.60 Due to TEM’s limited resolution of the
4.08.8.2 Neutron Irradiation of Materials
images, particles smaller than 2 nm were not detected.
At all temperatures, the size of the oxide particles The 5 mm-wide ring-tensile specimens with a
decreases as the dose increases. At higher tempera- 1.5 mm-wide gauge section were prepared from the
tures (600–700 C), the average size appears to reach a cladding of 12Cr-ODS steels (F94, F95, and 1DS)
value of #5 nm. At all three temperatures, the density and 9Cr-ODS steels (M93).67 This type of specimen
increases as the radiation dose increases. The decrease makes it possible to test mechanical properties in the
in size takes place faster at 600 and 700 C than at hoop direction of the cladding. These ring-tensile
500 C, indicating that the reduction in size is not samples were irradiated in the experimental fast reac-
strictly a ballistic effect and that a diffusion-based tor JOYO using the material irradiation rig at tem-
mechanism is also involved in the dissolution. peratures between 400 and 534 C to fast neutron
Allen extensively reviewed previous papers that fluences ranging from 5.0 ! 1025 to 3.0 ! 1026 nm"2
presented different approaches to the irradiation of (E > 0.1 MeV). The yield strength of the irradiated
ODS ferritic–martensitic steels that employed various samples as a function of test temperature is shown in
ion beams, electrons, and neutrons; the results are Figure 39, together with that of the unirradiated
summarized in Table 3.61 A great many findings ones.67 After irradiation, the yield strength of irra-
asserted that oxide particles are stable under radiation. diated F94, F95, and M93 cladding, is modestly higher
However, as shown in Table 4, the dissolution of oxide (<10%) than that of the unirradiated ones at all test
particles at higher temperatures and doses has been temperatures, due to irradiation hardening. Figure 40
reported in other studies. Dubuisson62 and Monnet63 plots uniform elongation before and after irradiation as
reported that small oxides dissolved under radiation at a function of test temperature.67 Uniform elongation
higher temperatures and doses, but did not dissolve for unirradiated F94 and F95 cladding is almost the
at a lower irradiation dose. Their data will be dis- same at all test temperatures, and that of M93 is lower
cussed in detail in the following section. In material in relation to strength. Uniform elongation in the hoop
irradiated in the JOYO fast reactor at temperatures direction for all three claddings is more than 3%
450–561 C to doses of 21 dpa, Yamashita found that at these test temperatures, though that of 1DS was
small particles disappear and average particles particularly low (<1%) due to its microstructural
increase slightly in size with increasing temperature anisotropy, as shown in Figure 17. Figure 40 indicates
or dose.64 Monnet supplemented neutron radiation that there is no significant degradation in uniform
studies with the electron irradiation of yttrium oxides elongations for F94, F95, and M93, due to irradiation.
and magnesium oxides in the EM10 alloy at tempera- This indicates that the microstructural improvement
tures between 300 and 550 C, and to doses of 100 dpa. by recrystallization or a–g-phase transformation is
In these studies, the yttrium oxides were stable at quite effective in maintaining well-balanced mechani-
400 C when irradiated with 1.0 MeV electrons, but cal properties for ODS steel cladding, especially those
dissolved under 1.2 MeV electron irradiation. of strength and ductility, not only for as-received con-
Allen59 pointed out that the displacement energy ditions but also following irradiation.
for Y and O in yttrium oxide is 57 eV65,66 while that In-pile creep rupture tests were conducted in
for iron is 40 eV. Assuming similar displacement JOYO using the Material Testing Rig with Tempera-
energies in the Y–Ti–O oxide, the radiation-induced ture Control (MARICO-2) as a new irradiation test
vacancy concentration should be larger in the metal device.68 The test specimens were prepared from the
matrix, providing a driving force for a net vacancy claddings of 9Cr-ODS steel (Mm14) and 12Cr-ODS
flux to the precipitate. This could drive the precipi- steel (F14). Both end-plugs of the specimens were
tate mass loss if vacancy absorption frees a precipitate joined by means of pressurized resistance welding
atom. From a comparison between electron irradia- (PRW). The hoop stress was set by adjusting the pres-
tion (Frenkel pairs) and ion irradiation (displacement sure of the enclosed helium gas. To identify the rup-
cascades), Monnet63 also concluded that the ballistic ture of time and specimens, a unique blend of stable
Oxide Dispersion Strengthened Steels 265

20 40
Fraction (%)

15 Unirradiated condition Unirradiated condition

Fraction (%)
30
10
5 20
0 10
20
Fraction (%)

T = 500 C, dose = 5 dpa 0


15
40
10
T = 600 C, dose = 5 dpa

Fraction (%)
5 30
0 20
20
Fraction (%)

15 T = 500 C, dose = 50 dpa 10


10 0
5 40
0 T = 600 C, dose = 50 dpa

Fraction (%)
30
20
Fraction (%)

15 T = 500 C, dose = 150 dpa 20


10 10
5
0
0 0 2 4 6 8 10 12 14 16 18 20 22
0 2 4 6 8 10 12 14 16 18 20 22
Particle size (nm)
Particle size (nm)
40
Unirradiated condition
Fraction (%)

30
20
10
0
40
T = 700 C, dose = 5 dpa
Fraction (%)

30
20
10
0
40
T = 700 C, dose = 50 dpa
Fraction (%)

30
20
10
0
40
T = 700 C, dose = 150 dpa
Fraction (%)

30
20
10
0
0 2 4 6 8 10 12 14 16 18 20 22
Particle size (nm)
Figure 38 Particle size (diameter) distribution for 9Cr-ODS steel irradiated at 500, 600, and 700 C to doses of 0, 5, 50,
and 150 dpa. Reproduced from Allen, T. R.; Gan, J.; Cole, J. I.; et al. J. Nucl. Mater. 2008, 375, 26–37.
266

Table 3 Historical survey of yttrium–titanium-oxides reported to be stable under radiation

Author Material Irradiation particle (dpa) Temperature ( C) Dose (dpa) Dose rate (dpa s"1) Result
74 "4
Pareige et al. 12YWT 150 keV Fe 300 0.7 1.9 ! 10 Stable dispersions
Asano et al.75 MA957 1 MeV He (þ) 450 150 2 ! 10"3 Stable oxides
4 MeV Ni 650
Hide et al.76 MA957 42 keV He 475 200 1.0–1.4 ! 10"2 Stable oxides
at 25 C (þ) 525
200 keV C" 575
625
Hide et al.76 MA957 220 keV He at 25 C 475 150 3.0 ! 10"3 Stable oxides
(+)3 MeV Niþ 525
Little77 DT2203YO5 52 MeV Cr6 þ (þ) 475 50 3.0 ! 10"4 Stable oxides
Oxide Dispersion Strengthened Steels

4 MeV He
Saito et al.61 13Cr–0.5TiO2 –0.2Y2O3 1 MeV electron 400 12 2.2 ! 10"3 Stable oxides
500
Kinoshita et al.78 13Cr-ODS (+) Nb, V, Zr 1 MeV electron 350 15 2 ! 10"3 Stable oxides
450
Akasaka et al.79 9Cr and 12Cr-ODS JOYO 330 7.0 Not reporteda Stable oxides
400 2.5
450 14.0
500 15.0
Mathon et al.80 MA957 Thermal neutrons 325 0.8, 2.0 1 ! 1014 n cm"2 Stable dispersions
(OSIRIS) 3.5, 5.5 (E > 1 MeV)
Monnet et al.63 DY EM10þY2O3EM10 þ MgO 1 MeV Helium 400 0.05 Not reported No change in oxide particles

Kimura et al.81 (13–19)Cr–4Al-ODS 300 20 1 ! 10"4 No reported change in oxide size


–500

a
Typical fast reactor displacement rates in the driver fuel portion of the core are 1 ! 10"6 dpa s"1.
Source: Reproduced from Allen, T. R.; Gan, J.; Cole, J. I.; et al. J. Nucl. Mater. 2008, 375, 26–37.
Oxide Dispersion Strengthened Steels 267

Table 4 Historical survey of yttrium–titanium-oxides reported change size under radiation

Author Material Irradiation Temperature Dose Dose rate Result


( C) (dpa) (dpa s"1)

Yamashita IDS (11Cr) JOYO 450–561 21 Not reporteda Small particles disappear.
et al.64 Average particles
increase slightly with
increasing temperature
or dose.
IDK (13Cr)
Dubuisson DT2203YO5 Phenix 400–580 81 Not reporteda Oxide particles are totally
et al.62 dissolved (small oxides)
or reduced in size and
were surrounded by a
halo of smaller oxides
(large oxides).
Monnet DT2203YO5 Phenix 400–580 81 Not reporteda Disappearance of small
et al.63 oxides and significant
halo of smaller oxides at
higher temperatures and
doses.
Monnet DY EM10 + 1 MeV and 300–550 100 3–6 ! 10"3 Oxides stable at 400 C
et al.63 Y2O3 1.2 MeV under 1.0 MeV electrons
EM10 + Electron but dissolve under
MgO 1.2 MeV.

a
Typical fast reactor displacement rates in the driver fuel portion of the core are about 1 ! 10"6 dpa s"1.
Source: Reproduced from Allen, T. R.; Gan, J.; Cole, J. I.; et al. J. Nucl. Mater. 2008, 375, 26–37.

F94 10 F94
1200 F94 unirrad. F94 unirrad.
F95
(Fluence; 3 1026 n m–2(E > 0.1MeV) F95
(Fluence; 3 1026 n m–2 (E > 0.1 MeV)
F95 unirrad. (0.5) (2.8) (3.0) F95 unirrad.
(1.35) 8
Uniform elongation (%)

M93 M93
1000 M93 unirrad. (1.4, 2.5) M93 unirrad.
Yield strength (MPa)

1DS
(3.56) 1DS unirrad. 1DS unirrad.
6
800

(0.45) 4
600
Y2O3(wt%)
2 Fluence
F94 0.24 (3.56)
400 F95 0.24 (1.35) (0.45)
M93 0.35
Fluence (0.5) (2.8) (3.0) (1.4, 2.5) 1DS 0.40 0
200 600 650 700 750 800 850 900
600 650 700 750 800 850 900 Test temperature (K)
Test temperature (K)
Figure 40 Uniform elongation of 9Cr-ODS steel (M93)
Figure 39 Yield strength of 9Cr-ODS steel (M93) and and 12Cr-ODS steels (F94, F95, 1DS) in hoop direction
12Cr-ODS steels (F94, F95, 1DS) in hoop direction by ring by ring specimens before and after irradiation.
specimens before and after irradiation. Reproduced from Reproduced from Yoshitake, T.; Abe, Y.; Akasaka, N.;
Yoshitake, T.; Abe, Y.; Akasaka, N.; Ohtsuka, S.; Ukai, S.; Ohtsuka, S.; Ukai, S.; Kimura, A. J. Nucl. Mater. 2004,
Kimura, A. J. Nucl. Mater. 2004, 329–333, 342–346. 329–333, 342–346.

xenon and krypton tag gases was enclosed. The irradi- MA957 and MA956 were irradiated in Fast Flux
ation temperatures were 700, 725, and 750 C, and the Test Facility (FFTF)-Materials Open Test Assembly
hoop stress ranged from 45 to 155 MPa. The maximum (MOTA) at 420 C up to 200 dpa.69 No voids were seen
neutron dose reached 20 dpa. It was confirmed that in- in this area, but precipitates did appear, which were
pile creep rupture time is located within the out-of- expected to be a0 . The results regarding the radiation
pile data band, and there is no degradation in creep damage resistance of ODS steels were highly encourag-
strength due to irradiation.68 ing. Evidence was apparent in both MA956 and MA957
268 Oxide Dispersion Strengthened Steels

0.2 mm

Figure 41 Longitudinal cross-sectional structure in the vicinity of welded section by PRW (9Cr-ODS steel cladding
and endplug). Reproduced from Ukai, S.; Kaito, T.; Seki, M.; Mayorshin, A. A.; Shishalov, O. V. J. Nucl. Sci. Technol. 2005,
42(1), 109–122.

of a0 precipitation, and in regions where recrystalliza-


tion occurred before irradiation in MA957, a few voids
were slightly observed. Gelles69 pointed out that these
could be overcome by employing suitable alloy design
and that ODS steel microstructures, when properly
manufactured to provide a uniform oxide dispersoid
in a structure, appear to be completely resistant to
radiation damage at doses as high as 200 dpa.

4.08.8.3 Fuel Pin Irradiation


4.08.8.3.1 9Cr- and 12Cr-ODS steel cladding
in BOR-60
In order to weld 9Cr- and 12Cr-ODS steel claddings
with end-plugs for the manufacture of fuel pins, the
Figure 42 Optical micrograph of 9Cr-ODS fuel pin after
PRW method was developed in JAEA, which makes
irradiation at 700 C, 5 at.% burnup and 25 dpa in BOR-60.
joining possible in the solid state condition.70 This Reproduced from Kaito, T.; Ukai, S.; Povstyanko, A. V.;
method is based on the electrical resistance heating Efimov, V. N. J. Nucl. Sci. Technol. 2009, 46(6), 529–533.
of the components, while maintaining a continuous
force sufficient to forge-weld without melting. The the lower end-plug ensured that its integrity would be
appropriate conditions, for example, electric current, maintained at a lower temperature of 400 C. The
voltage, and contact force, were selected. For the PRW- inspection and quality control of the fabricated ODS
welded specimens, tensile, internal burst, and creep fuel pins were done through X-ray analysis, gamma
rupture tests, were conducted and their integrity was scanning, and leak testing, etc., which confirmed that
confirmed. In addition, a nondestructive ultrasonic the fuel pins satisfied BOR-60 requirements. The fuel
inspection method was developed to assure the integ- pins were loaded into two dismountable experimental
rity of the weld between the cladding and end-plug. assemblies to satisfy the cladding middle wall tempera-
Using this PRW method, upper end-plugs were ture within 700 C and 650 C, and irradiation was
welded for two types of 9Cr-ODS steel cladding conducted in the BOR-60 up to 5 at.% burnup and
(Mm13) and 12Cr-ODS steel cladding (F13) at JAEA. 25 dpa as the collaborative work between JAEA in
Figure 4171 shows a cross-section of the welded part Japan and RIAR of Research.71
between the 9Cr-ODS steel cladding and end-plug. The results of the postirradiation examination
The ODS steel cladding welded to the upper end- are shown in Figure 42 in the optical micrographs
plug was shipped to the fuel production facility of the of the upper part of the fuel column of 9Cr-ODS
Institute of Atomic Reactor (RIAR) in Russia where the steel fuel; no obvious corrosion inside the cladding
MOX and UO2 granulated fuels, as well as uranium was observed.72 The maximum depth of corrosion
metal getter particles, were vibro-packed into the ODS of 25 mm is partially confirmed in the upper part of
steel cladding, and the lower end-plug was welded by the fuel column. The inner corrosion of the ODS
the TIG end-face method. The TIG-welded part at cladding can be reduced by using a lower O/M
Oxide Dispersion Strengthened Steels 269

(a) (b) (c)

100 nm 1 µm 1 µm

Figure 43 Precipitation occurring during the in-pile service (a) a0 -phases at 400 C, &0 dpa, (b) w-phases at 523 C to
78.8 dpa, and (c) Laves phases at 580 C to 30.5 dpa. Reproduced from Dubuisson, P.; Schill, R.; Higon, M.-P.; Grislin, I.;
Seran, J.-L. In Effects of Radiation on Materials: 18th International Symposium; Nanstad, R. K., Hamilton, M. L., Garner, F. A.,
Kumar, A. S., Eds.; American Society for Testing and Materials: Philadelphia, PA; p 882, ASTM STP 1325.

ratio fuel, even in lower Cr content cladding such as density, was observed in the lower part of the fuel pin
9Cr-ODS steel. at temperatures <500 C. These precipitates were
found to be a0 -phase, as shown in Figure 43(a).63
4.08.8.3.2 12Cr-ODS steel cladding in EBR-II At irradiation temperatures above 500 C, precipita-
JAEA manufactured 12Cr-ODS steel cladding (1DK tion of w-phase was uniformly distributed throughout
and 1DS) and Argonne National Laboratory in the all grains, as shown in Figure 43(b). Their chemical
United States qualified a welding process that employs composition was slightly different from that of the
PRW. Fuel pins composed of 12Cr-ODS steel cladding intergranular w-phase that was present before irradi-
and MOX fuel pellets were successfully fabricated and ation. At a high temperature and low dose, w-phase is
qualified, and irradiated up to 35 dpa at EBR-II.73 The replaced by the thermal precipitation of Laves phase,
ODS cladding with high smear density solid pellet as shown in Figure 43(c).
MOX fuel did induce some diametral strain, demon- From the results of tensile tests at levels corres-
strating some in-core ductility. This program demon- ponding to the fissile column, rupture occurred with-
strated the viability of ODS steel as a potential out striction, and uniform and total elongations were
cladding material for long-life advanced FRs. equal. The elongation values reached 0.2% close
to the maximum dose. These results indicate that
4.08.8.3.3 DT2203Y05 in Phénix DT2203Y05 cladding was highly embrittled by irra-
Fuel pins with DT2203Y05 cladding were irradiated diation. At the bottom of the fuel pin, where the
in an experimental capsule placed in a special subas- temperature is below 500 C, a0 precipitation, oxide
sembly in Phénix. The process by which they were redistribution and dislocation loops are the main
manufactured was described in Section 4.08.6. The features of the microstructure. Dubuisson63 pointed
dose reached at midplane was 81 dpa and the temper- out that dislocation glides on the dislocation denuded
ature along the fuel pin ranged from 400 to 580 C. bands in the hardened materials, that the deformation
It was observed by TEM that the uniform distribu- is all localized in these bands, and that this heteroge-
tion of fine oxides totally disappeared, and a few large neous shear could nucleate cracks that then propa-
oxides were also fragmented into smaller ones. The gate along these channels. At higher temperatures,
recoil resolution of particles is a process where the w precipitation induces a loss of ductility.
atoms that compose particles are ballistically ejected
by an impinging neutron. Dubuisson63 pointed out
that the atoms ejected from oxides by ballistic dissolu- 4.08.9 Summary
tion depend on radiation-enhanced solute diffusivity
and enhanced solubility under irradiation. The formation process of nanosized oxide particles
A uniform distribution of tiny particles <10 nm in through decomposition by MA and subsequent pre-
size and with a density higher than the original oxide cipitation by annealing was reviewed. Based on
270 Oxide Dispersion Strengthened Steels

information concerning the irradiation embrittle- 17. Ukai, S.; Kaito, T.; Otsuka, S.; Narita, T.; Fujiwara, M.;
Kobayashi, T. ISIJ Int. 2003, 43, 2038.
ment of DT2203Y05 cladding produced by 18. Ukai, S.; Narita, T.; Alamo, A.; Pamentier, P. J. Nucl.
CEN-SCK Mol due to the formation of a0 -phase Mater. 2004, 329–333, 356.
below 500 C and w-phase above 500 C, 9Cr-ODS 19. Ohtsuka, S.; Ukai, S.; Fujiwara, M.; Kaito, T.; Narita, T.
J. Nucl. Mater. 2004, 329–333, 372–376.
and 12Cr-ODS steels containing low Cr and low Ti 20. Ohtsuka, S.; Ukai, S.; Fujiwara, M.; Kaito, T.; Narita, T.
were developed. The manufacture of cladding and Mater. Trans. 2005, 46, 487.
improvement of the creep rupture strength in the 21. Ohtsuka, S.; Ukai, S.; Fujiwara, M.; Kaito, T.; Narita, T.
J. Nucl. Mater. 2006, 351, 241.
hoop direction were successfully achieved by introdu- 22. Yamamoto, M.; Ukai, S.; Hayashi, S.; Kaito, T.;
cing a–g reverse transformation or recrystallization. Ohtsuka, S. Mater. Sci. Eng. A 2010, 527, 4418–4423.
9Cr-ODS steel has a unique structure consisting of 23. Ukai, S.; Ohtsuka, S. Energy Mater. 2007, 2(1), 26–35.
24. Nishizawa, T.; Ohnuma, I.; Ishida, K. Mater. Trans. JIM
tempered martensite and residual ferrite that induces 1997, 38(11), 950–956.
superior strength through finely dispersed oxide par- 25. Martion, J. W.; Doherty, R. D. Stability of Microstructure in
ticles, which are promising candidates for advanced Metallic Systems; Cambridge University Press:
Cambridge, 1976; p 173.
SFR fuel cladding. 16Cr–4Al–ODS steels present an 26. Ukai, S.; Ohtsuka, S.; Kaito, T.; Sakasegawa, H.;
advantage due to their superior resistance to corro- Chikata, N.; Hayashi, S.; Ohnuki, S. Mater. Sci. Eng. A
sion and oxidation in LBE and SCPW environments. 2009, 510–511, 115–120.
27. Tabor, D. The Hardness of Metals; Oxford university press:
There is still uncertainty concerning the irradiation Oxford, 1951.
performance of ODS steels, such as the oxide particle 28. Yamamoto, M.; Ukai, S.; Hayashi, S.; Kaito, T.;
dissolution due to their diffusion-based mechanism. In Ohtsuka, S. J. Nucl. Mater. 2011, 237–240, 417.
29. Lambard, V. Development of ODS ferritic-martensitic
order to substantiate the use of cladding materials as steels for application to high temperature and
advanced fast reactor fuels, abundant ODS cladding irradiation environment; Rapport CEA-R-5918;
fuel pins should be irradiated, and their results should France 2000.
30. Allen, T.; Burlet, H.; Nanstad, R. K.; Samaras, M.; Ukai, S.
provide feedback contributing to the further improve- Mater. Res. Soc. Bull. 2009, 34(1), 20–27.
ment of ODS steels. 31. Patriarca, P. In Proceedings of the Topical Conference on
Ferritic Alloys for Use in Nuclear Energy Technologies,
Snowbird, Utah, June 19–23 1983; p 107.
References 32. Shibahara, I.; Ukai, S.; Onose, S.; Shikakura, S. J. Nucl.
Mater. 1993, 204, 131.
33. Ukai, S.; Harada, M.; Okada, H.; et al. J. Nucl. Mater. 1993,
1. Odette, G. R.; Alinger, M. J.; Wirth, B. D. Annu. Rev. Mater. 204, 74–80.
Res. 2008, 38, 471–503. 34. Fischer, J. J. US Patent 4075010, 1978.
2. Okuda, T.; Fujiwara, M. J. Mater. Sci. Lett 1995, 35. Ukai, S.; Nishida, T.; Okada, H.; Okuda, T.; Fujiwara, M.;
14, 1600. Asabe, K. J. Nucl. Sci. Technol. 1997, 34(3), 256–263.
3. Ukai, S.; Harada, M.; Okada, H.; Inoue, M.; Nishid, T.; 36. Ukai, S.; Okuda, T.; Fujiwara, M.; Kobayashi, T.; Mizuta, S.;
Fujiwara, M. J. Nucl. Mater. 1993, 204, 65–73. Nakashima, H. J. Nucl. Sci. Technol. 2002, 39(8), 872–879.
4. Ukai, S.; Fujiwara, M. J. Nucl. Mater. 2002, 307–311, 749. 37. Arzt, E. Res. Mech. 1991, 31, 399–453.
5. Okuda, T.; Ukai, S.; Miyahara, K. Materia (Japan) 1999, 38. Srolovitz, D. J.; Perkovic-Luton, R. A.; Luton, M. J. Phil.
39, 954. Mag. A 1983, 48, 795.
6. Larson, D. J.; Maziasz, P. J.; Kim, I.-S.; Miyahara, K. 39. Scattergood, R. O.; Bacon, D. J. Phil. Mag. A 1975, 31, 170.
Scr. Mater. 2001, 44, 359. 40. Foreman, A. J. E.; Makin, M. J. Phil. Mag. A 1966, 14, 911.
7. Miller, M. K.; Kenik, E. A.; Russell, K. F.; Heatherly, L.; 41. Stoller, R. E.; Zinkle, S. J. J. Nucl. Mater. 2000, 283–287,
Hoelzer, D. T.; Maziasz, P. J. J. Nucl. Mater. 2003, A353, 349.
140. 42. Guerin, Y.; Was, G. S.; Zinkle, S. J. Mater. Res. Soc. Bull.
8. Alinger, M. J.; Odette, G. R.; Hoelzer, D. T. J. Nucl. Mater. 2009, 34(1), 10–14.
2004, 382, 329–333. 43. Kimura, A.; Kasada, R.; Iwata, N.; et al. In Proceedings of
9. Kim, S.-W.; Shobu, T.; Ohtsuka, S.; Kaito, T.; Inoue, M.; ICAPP ’09, Tokyo, Japan, May 10–14, 2009; Paper 9220.
Ohnuma, M. Mater. Trans. 2009, 50(4), 917–921. 44. Furukawa, T.; Ohtsuka, S.; Inoue, M.; et al. In Proceedings
10. Kliniankou, M.; Lindau, R.; Moslang, A. J. Nucl. Mater. of ICAPP ’09, Tokyo, Japan, May 10–14, 2009; Paper 9221.
2004, 329–333, 347–351. 45. Narita, T.; Ukai, S.; Kaito, T.; Otsuka, S.; Kobayashi, T.
11. Smith, J. V. American Society for Testing Materials, J. Nucl. Sci. Technol. 2004, 41(10), 1008.
Powder diffraction file, 13. Inorganic; Philadelphia, PA, 46. Ukai, S.; Ohnuki, S.; Hayashi, S.; et al. In Proceedings of
1962–1965. ICAPP’09, Tokyo, Japan, May 10–14, 2009; Paper 9232.
12. Luo, C. P.; Dahmen, U. Acta. Mater. 1998, 46, 2063. 47. Dorner, D.; Zaefferer, S.; Raabe, D. Acta. Mater. 2007, 55,
13. Ukai, S.; Nishida, T.; Okuda, T.; Yoshitake, T. J. Nucl. Sci. 2519.
Technol. 1998, 35, 294. 48. Seran, J. L.; Levy, V.; Dubuisson, P.; et al. In Effects of
14. Ukai, S.; Nishida, T.; Okuda, T.; Yoshitake, T. J. Nucl. Radiation on Materials: 15th International Symposium;
Mater. 1998, 1745, 258–263. Philadelphia, PA 1992; pp 1209–1233, ASTM STP 1125.
15. Ukai, S.; Muzuta, S.; Fujiwara, M.; Okuda, T.; 49. Lippens, M.; Ehrlich, K.; Levy, V.; Brown, C.; Calzabini, A.
Kobayashi, T. J. Nucl. Sci. Technol. 2002, 39, 778. In Proceedings of the International Conference on Ferritic
16. Ukai, S.; Mizuta, S.; Fujiwara, S.; Okuda, T.; Kobayashi, T. Alloys for Use in Nuclear Energy Technologies, Utah, 1983;
J. Nucl. Mater. 2002, 307–311, 758. pp 329–334.
Oxide Dispersion Strengthened Steels 271

50. DeWilde, L.; Gedopt, J.; DeBurbure, S.; Delbrassibe, A.; 67. Yoshitake, T.; Abe, Y.; Akasaka, N.; Ohtsuka, S.; Ukai, S.;
Driesen, C.; Kazimierzak, B. In Proceeding in Materials for Kimura, A. J. Nucl. Mater. 2004, 329–333, 342–346.
Nuclear Reactor Core Applications; BNES: London, 1987; 68. Kaito, T.; Ohtsuka, S.; Inoue, M.; et al. J. Nucl. Mater.
271–276. 2009, 386–388, 294–298.
51. Huet, J. J.; Coheur, L.; De Bremaecker, A.; et al. Nucl. 69. Gelles, D. S. Fusion Materials, Semiannual Progress
Technol. 1985, 70, 215–219. Report for Period Ending March 31, DOE/ER-0313/16
52. Hamilton, M. L.; Gelles, D. S. PNNL-13168, Feb 2000. 1994; pp 146–160.
53. Yoshida, E.; Kato, S. J. Nucl. Mater. 2004, 329–333, 70. Seki, M.; Hirako, K.; Kono, S.; Kihara, Y.; Kaito, T.; Ukai, S.
1393–1397. J. Nucl. Mater. 2004, 329–333, 534–1538.
54. Machut, M.; Sridharan, K.; Li, N.; Ukai, S.; Allen, T. J. Nucl. 71. Ukai, S.; Kaito, T.; Seki, M.; Mayorshin, A. A.; Shishalov, O. V.
Mater. 2007, 371, 134–144. J. Nucl. Sci. Technol. 2005, 42(1), 109–122.
55. Furukawa, T.; Muller, G.; Schumacher, G.; et al. J. Nucl. 72. Kaito, T.; Ukai, S.; Povstyanko, A. V.; Efimov, V. N. J. Nucl.
Sci. Technol. 2004, 41(3), 265–270. Sci. Technol. 2009, 46(6), 529–533.
56. Takaya, S.; furukawa, T.; Aoto, K.; et al. Fall meeting of 73. Bottcher, J.; Ukai, S.; Inoue, M. Nucl. Technol. 2002, 138,
Japan Atomic Energy Society, Sept 2008. 238–245.
57. Lee, J. H.; Kimura, A.; Kasada, R.; et al. In Proceedings of 74. Pareige, P.; Miller, M. K.; Stoller, R. E.; Hoelzer, D. T.;
ICAPP ’09, Tokyo, Japan, May 10–14, 2009; Paper 9223. Cadel, E.; Radiguet, B. J. Nucl. Mater. 2007, 360, 136.
58. Kaito, T.; Narita, T.; Ukai, S.; Matsuda, Y. J. Nucl. Mater. 75. Asano, K.; Kohno, Y.; Kohyama, A.; Suzuki, T.;
2004, 329–333, 1388–1392. Kusanagi, H. J. Nucl. Mater. 1988, 155–157, 928.
59. Chen, Y.; Sridharan, K.; Ukai, S.; Allen, T. R. J. Nucl. 76. Hide, K.; Sekimura, N.; Fukuya, K.; et al. In Effects of
Mater. 2007, 371, 118–128. Radiation on Materials: 14th International Symposium, vol. I.;
60. Allen, T. R.; Gan, J.; Cole, J. I.; et al. J. Nucl. Mater. 2008, Packan, N. H., Stoller, R. E., Kumar, A. S. Eds.; American
375, 26–37. Society for Testing and Materials: Philadelphia, PA, 1988;
61. Saito, J.; Suda, T.; Yamashita, S.; et al. J. Nucl. Mater. p 61, ASTM STP 1046.
1998, 258–263, 1264. 77. Little, E.A. In Effects of Radiation on Materials:
62. Dubuisson, P.; Schill, R.; Higon, M.-P.; Grislin, I.; 17th International Symposium; Gelles, D.S., Nanstad, R.K.,
Seran, J.-L. In Effects of Radiation on Materials: 18th Kumar, A.S., Little, E.A., Eds.; American Society for Testing
International Symposium; Nanstad, R. K., Hamilton, M. L., and Materials: Philadelphia, PA, 1996; p 739, ASTM STP
Garner, F. A., Kumar, A. S., Eds.; American Society for 1270.
Testing and Materials: Philadelphia, PA, 1999; p 882, 78. Kinoshita, H.; Akasaka, N.; Takahashi, H.; Shibahara, I.;
ASTM STP 1325. Onose, S. J. Nucl. Mater. 1992, 191–194, 874.
63. Monnet, I.; Dubuisson, P.; Serruys, Y.; Ruault, M. O.; 79. Akasaka, N.; Yamashita, S.; Yoshitake, T.; Ukai, S.;
Kaitasov, O.; Jouffrey, B. J. Nucl. Mater. 2004, 335, 311. Kimura, A. J. Nucl. Mater. 2004, 329–333, 1053.
64. Yamashita, S.; Oka, K.; Ohnuki, S.; Akasaka, N.; Ukai, S. 80. Mathon, M.-H.; de Carlan, Y.; Averty, X.; Alamo, Ch.-H.;
J. Nucl. Mater. 2002, 307–311, 283. de Novion. J. ASTM Int. 2005, 2(8), Paper ID JAI 12381.
65. Rechtin, M. D.; Wiedersich, H. Radiat. Eff. 1977, 31, 181. 81. Kimura, A.; Cho, H.; Toda, N.; et al. J. Nucl. Sci. Technol.
66. Zinkle, S. J.; Kinoshita, C. J. Nucl. Mater. 1997, 251, 200. 2007, 44(3), 323.

S-ar putea să vă placă și