You are on page 1of 12

Acta Materialia 234 (2022) 117991

Contents lists available at ScienceDirect

Acta Materialia
journal homepage: www.elsevier.com/locate/actamat

Insights into hardening, plastically deformed zone and geometrically


necessary dislocations of two ion-irradiated FeCrAl(Zr)-ODS ferritic
steels: A combined experimental and simulation study
Peng Song a,b,∗, Kiyohiro Yabuuchi b, Philippe Spätig a
a
Department of Nuclear Energy and Safety, Paul Scherrer Institute, 5232 Villigen PSI, Switzerland
b
Institute of Advanced Energy, Kyoto University, Gokasho, Uji, Kyoto 611-0011, Japan

a r t i c l e i n f o a b s t r a c t

Article history: As one of promising candidate materials for fuel claddings and structural components in the Gen-IV
Received 30 August 2021 fission reactors, FeCrAl(Zr)-ODS ferritic steels were studied to well understand the radiation hardening
Revised 15 March 2022
behavior. Nanoindentation (NI) hardness and plastically deformed zone (PDZ), geometrically necessary
Accepted 1 May 2022
dislocations (GNDs) as well as microstructures were investigated for two FeCrAl(Zr)-ODS ferritic steels
Available online 6 May 2022
irradiated with 6.4 MeV Fe3+ at room temperature (RT) up to the nominal damages of 2, 10 and 50
Keywords: dpa. Irradiation-induced hardening, which was estimated by using the Nix-Gao model regardless of the
Radiation hardening damage gradient effect (DGE), increased continuously with increasing the nominal displacement dam-
Plastically deformed zone age. When taking the DGE into account, the dependence of the irradiation hardening (in MPa) on a local
Geometrically necessary dislocations damage level (dpa) obtained by finite element method (FEM) simulations was 153.65 × (dpa)0.26 and
ODS ferritic steel 158.35 × (dpa)0.25 for the non-Zr steel and the Zr-added one, respectively. The hardening caused by ion-
Nanoindentation
irradiation was discussed in terms of the loss of oxide particles, the formation of dislocation loops and
the solid solution hardening mainly by dissolved oxygen. The addition of Zr reduced the ion-irradiation
hardening of steel to some extent mainly by suppressing the formation of dislocation loops. Meantime, as
the FEM simulations revealed, the PDZ size underneath a conical indenter at the equivalent plastic strain
ε eq > 1.9% can reasonably estimate the density of GNDs. According to the strain gradient plasticity (SGP)
theory, the densities of GNDs at an indentation depth of 200 – 300 nm were slightly higher in the case
of the 50 dpa/nominal than the unirradiated, in spite of similarity for the two FeCrAl(Zr)-ODS steels.
© 2022 The Author(s). Published by Elsevier Ltd on behalf of Acta Materialia Inc.
This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/)

1. Introduction As previous studies revealed, the corrosion resistance [10], the


mechanical properties [1,3] and the radiation tolerance [5,11–13]
High-Cr oxide dispersion strengthened (ODS) ferritic steels with of ODS ferritic steels strongly depend on the dispersion morphol-
Al addition have been developed in Japan to replace Zircaloy for ogy of nano-sized oxide particles inside the matrix, which not only
the accident tolerance fuel (ATF) of current light water reactors effectively hinder the motion of dislocations and grain boundaries
mainly because of their good oxidation resistance arising from a (GBs) resulting in enhanced strength, but also provide point-defect
protective Al2 O3 scale especially at high temperatures in steam traps through particle-matrix (P-M) interfaces. Thus, they promote
[1,2], their superior structural stability and mechanical properties the annihilation or recombination of radiation defects.
(e.g. tensile, creep strength) even at elevated temperatures [1,3,4], As far as the radiation hardening of ODS steels is concerned,
as well as their excellent neutron radiation tolerance [5–8]. They numerous investigations [11–14] focused on the formation of de-
are also considered as promising fuel claddings and structural fect clusters, especially the dislocation loops during ion irradia-
components in the future Gen-IV fast reactors [1,9], where neutron tions, which can be suppressed by trapping sites provided by the
irradiation damage is expected up to 160 dpa [9]. initial microstructures (GBs, P-M interfaces and dislocations). In
our previous work [11], after dual-beam (Fe3+ plus He+ ) irradia-
tion at 550 °C to the nominal damage of 15 dpa (∼20 0 0 appm

Corresponding author at: Department of Nuclear Energy and Safety, Paul Scher-
He), no hardening was found in the three ODS ferritic steels, al-
rer Institute, 5232 Villigen, Switzerland. though cavities (3–7 nm in mean diameter) of high number densi-
E-mail address: peng.song@psi.ch (P. Song). ties up to ∼1 × 1023 m−3 were formed in the irradiated steels and

https://doi.org/10.1016/j.actamat.2022.117991
1359-6454/© 2022 The Author(s). Published by Elsevier Ltd on behalf of Acta Materialia Inc. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/)
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

approximately 80% cavities were adjacent to oxide particles. Duan coherency of particle-matrix interface and dispersion morphology
et al. [12] reported that, the Fe-ion irradiation at 300 or 500 °C up to some extent [33,34]. Therefore, this work was launched to in-
to 10 dpa caused a significant hardness increase in a set of ODS vestigate by means of nanoindentation and FEM simulation: 1)
alloys of different grain sizes, dislocation densities and morpholo- dpa-dependence hardening behaviors of two FeCrAl(Zr)-ODS fer-
gies of particle dispersion, while the hardening was significantly ritic steels under Fe-ion irradiation at RT; 2) the effects of ion-
reduced with increasing the total sink strength (i.e. trapping sites) irradiation on the PDZ size and the associated GNDs.
of the initial microstructures. Similar dependence of hardening on
the sink strength of oxide particles was obtained by Zhang et al. 2. Materials and methods
[13] for three ODS ferritic steels irradiated with high-energy Ni
ions at ∼0.8 dpa and −50 °C. In contrast, Heintze et al. [14] found 2.1. Materials
that, under Fe-ion irradiation at room temperature (RT) to 10 dpa,
the hardness increase in the two ODS ferritic steels depended only The materials used in this study were two FeCrAl(Zr)-ODS fer-
on grain size rather than particle number density, which is ex- ritic steels produced by mechanical alloy processing. The corre-
plained by the fact that coherent interfaces between (Y, Ti) ox- sponding chemical compositions (wt.%) were Fe-15.42Cr-1.85W-
ide particles and matrix cannot efficiently trap self-interstitials. So 0.1Ti-3.8Al-0.36Y2 O3 (excess oxygen: 0.084) and Fe-14.59Cr-1.84W-
both the dispersion morphology and the interface coherency of ox- 0.14Ti-3.46Al-0.27Zr-0.33Y2 O3 (excess oxygen: 0.11), hereinafter,
ide particles could play a vital role in suppressing radiation hard- named as (Y, Al) ODS and (Y, Zr) ODS, respectively. After hot ex-
ening of ODS steels. trusion and forging at 1150 °C, the materials were finally annealed
Collision cascades produced by energetic particles (like neu- at 1150 °C for 1 h followed by air-cooling. More details regarding
tron, proton and heavy ions) modify the oxide particles dispersed the material processing are given in Ref. [3].
in the steel matrix in terms of their chemical composition [15– Prior to ion-irradiation, the specimens were prepared so that
18], the crystal structure [19], the particle-matrix (P-M) interface the specimen surfaces were perpendicular to the extrusion direc-
[19,20], and the dispersion morphology [15,19–23]. Our recent tion, in other words, the ion-irradiation direction is parallel to the
work [19] showed that, in three 15Cr-ODS ferritic steels irradiated extrusion direction. The specimen surfaces were mechanically pol-
with Fe ions at 200 °C up to 10 dpa/nominal, the phase stability of ished with SiC papers, and subsequently with successive grades
the dominant oxide particles was in the order of Y-Zr-O > Y-Ti-O > of diamond spray down to 0.25 μm. Finally, electrolytic polishing
Y-Al-O, according to the corresponding changes in volume fraction was conducted in the solution of 10% (vol.) HClO4 and 90% (vol.)
and amorphization behavior. Therefore, it is necessary to investi- CH3 COOH at room temperature to remove any residual mechanical
gate the irradiation-induced hardening of ODS steels with different damage.
oxide forming elements, where nano-scale oxide particles may re-
spond differently to radiation damages.
Although nanoindentation has been long used to study ion- 2.2. Irradiation conditions
irradiation effects on the mechanical properties of nuclear ma-
terials [24–29], the interpretation of nanoindentation results is For each damage level, specimens of (Y, Al) ODS and (Y, Zr) ODS
still difficult because indentation probes a shallow irradiated layer were irradiated simultaneously with 6.4 MeV Fe3+ ions by using
(commonly < 3 μm) of a steep damage gradient on a soft unirra- the DuET facility at Kyoto University [35]. The beam was raster
diated substrate. At such small indentation depths, numerous ge- scanned at a frequency of 10 0 0 Hz in a horizontal direction and
ometrically necessary dislocations (GNDs) are produced to accom- 300 Hz in a vertical direction under a vacuum of ∼10−6 Pa. The
modate the strain gradient in the plastically deformed zone (PDZ) flux of ion beam was measured to be approximately 2.1 × 1016
underneath the indenter, whose density usually increases with de- ions m−2 s−1 . The specimens during irradiation were set on a cop-
creasing indentation depth. This phenomenon results in a larger per holder cooled by flowing water inside the holder. The irradia-
hardness at a smaller indentation depth, which is known as the in- tion temperature was measured by a thermocouple and controlled
dentation size effect (ISE) [30]. However, for an indentation depth to be 30 °C within the fluctuation of ±5 °C.
below ∼100 nm, dislocation activities under indentation may be- The displacement damages (displacements per atom, dpa) of a
come discrete in nature, which can limit the multiplication of Fe-15Cr alloy were obtained by SRIM [36] simulations, which used
GNDs [31]. According to Johnson’s model [32], the normalized size the Kinchin-Pease method [37] and are shown in Fig. 1. The dam-
Znor of PDZ in elastic-perfectly plastic solids depends on the ratio aged layer was located in the depth range of 0 – 2220 nm. The
of Young’s modulus E to yield stress σ Y by the following equation: nominal displacement damages were defined as the ones at the
ion-penetration depth of 600 nm, which were about 2, 10 and 50
 2 E 1 / 3 dpa, and the damages were ranging roughly 1.2 – 6 dpa, 6 – 30 dpa
Znor = (1) and 30 – 150 dpa in each damaged area, respectively (see Fig. 1).
3 σY The nominal displacement rate was around 6.4 × 10−4 dpa/s. The
Based on the above discussion, the normalized PDZ size and injected Fe concentration for 50 dpa/nominal is shown with the
the associated GNDs density of an ion-irradiated specimen are ex- dashed profile in Fig. 1.
pected to vary with indentation depth within a damage-gradient
layer, which is the source of difficulties for assessing ion-irradiation 2.3. Transmission electron microscopy (TEM)
hardening. By using FEM simulation, Dolph et al. [24] found the
PDZs in an Fe9Cr-ODS steel produced by a conical indenter were In order to observe depth dependent damage structures, cross
7.84, 7.25 and 5.49 times the indentation contact depth for the as- sectional TEM specimens were prepared by the lift-out technique
received, ion-irradiated (50 dpa, 400 °C) and neutron-irradiated (3 in a Focused Ion Beam (FIB, Hitachi, FB2200) instrument, where
dpa, 500 °C), respectively, without taking the damage gradient ef- the final beam used was 20 keV Ga+ . Prior to TEM observations,
fect (DGE) into account. Hence, based upon a good understanding flash polishing was done by using the solution consisting of 5%
of the radiation hardening in ODS steels, the PDZ size and GNDs (vol.) HClO4 and 95% (vol.) CH3 OH at around −35 °C to remove
density can be considered. the damage layer produced by FIB. About five TEM lamellas were
The Zr addition into FeCrAl-ODS ferritic steels changed the crys- finally polished with 1 keV Ar+ ions instead of flash polishing. Dis-
tal structure of dominant oxide particles, and also improved their location loops and oxide particles were observed under bright field

2
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

true stress-true strain relationship was fitted by using the Ludwik-


Hollomon equation [40]:
σ = σYUn + K εpln (2)

where σ is von Mises stress, σYUn is the value at yield point


(ε pl = 0) for the unirradiated specimens, ε pl is equivalent plastic
strain (also PEEQ), K is known as the ‘strength coefficient’ and n as
the ‘strain-hardening exponent’. The measured yield strength val-
ues of (Y, Al) ODS and (Y, Zr) ODS were 933.6 MPa and 1041.3 MPa,
respectively (see Supplementary material S1). In the case of the ir-
radiated material, the constitutive law of the irradiated layers was
assumed to be given by:
σ = σ Irr + σYUn + K εpln (3a)

σ Irr = C × (dpa )m (3b)


where σ Irr is the increased part of flow stress produced by the
ion-irradiation, assumed as a function of the damage level (dpa) by
Fig. 1. The depth profiles of displacement damage (solid) and concentration of in-
Eq. 3(b). Detailed discussion is present in Sections 3.2 and 3.3. The
jected 6.4 MeV Fe3+ ions (dash) into Fe-15Cr alloy predicted by SRIM simulation.
ion-irradiated region in Fig. 2 was divided into 23 layers according
to the depth-dependent displacement damage profile (broad lay-
(BF) and weak beam dark field (WBDF) imaging conditions by us- ers up to 1200 nm and thinner ones for the larger damage gradi-
ing a conventional 200 kV JEOL JEM 2010 microscope equipped ent), to minimize discontinuity in the calculated plastic strain be-
with a LaB6 electron emitter. The thickness of all observation ar- tween two adjacent discrete layers. The possible effects of the layer
eas in TEM lamellas was measured using electron energy-loss spec- number on the plastic strain and the load-indentation depth rela-
troscopy (EELS) log-ratio technique [38]. tionship have been investigated (see Supplementary material S2).
In the depth range of 1650 – 1950 nm, where the damage level
2.4. Nano-hardness measurements decreased steeply with depth, the thickness of each discrete layer
was set to be 60 nm. The constitutive laws of the unirradiated and
Nanoindentation tests were performed on unirradiated and ir- irradiated specimens are shown in Fig. 3(a). The contact friction of
radiated specimens by using an instrumented nanoindenter (Agi- 0.2 was commonly used for the Coulomb friction between the rigid
lent G200) with a Berkovich tip. The continuous stiffness measure- tip and the top surface of a metal.
ment (CSM) method was selected to obtain continuous profiles of The discrepancy of load-depth curves between FEM simulation
hardness and elastic modulus as a function of indentation depth. and experimental tests can be quantified by the dimensionless pa-
Before testing, the indenter tip geometry was calibrated against rameter Sred, the “reduced sum of squares” [41]:
the Young’s modulus of a fused silica reference sample at indenta- N
i=1 (Pi,
M − Pi,E )2
tion depths larger than ∼50 nm. The testing parameters were de- Sred = 2
(4)
NPav,E
scribed as follows: the maximum indentation depth was 20 0 0 nm,
the strain rate was 0.05 s−1 , the harmonic displacement target and where Pi,M and Pi,E are the ith load values predicted by FEM sim-
frequency target were 1 nm and 45 Hz, respectively, and the Pois- ulations and measured by experimental tesst, respectively, Pav,E is
son’s ratio ν for ODS steels was 0.3. For each specimen, 24 inden- the average value of the experimentally measured loads over the
tations (60 μm between two adjacent points) were carried out and whole indentation depth range (up to 20 0 0 nm), and N was set to
the averaged hardness was obtained by calculating the measured be 100 for a displacement increment of 20 nm. Generally, a model
hardness data of ∼20 tests to take into account the significant scat- that captures the material plasticity response reasonably has the
tering in the data. parameter Sred less than 10−3 [41].

2.5. Finite element method (FEM) simulation 3. Results and discussion

A two-dimensional axisymmetric model was developed by us- 3.1. Microstructures before and after irradiations
ing ABAQUS 6.4–1 standard code for modeling the nanoindenta-
tion process. Considering that the contact area function A ver- Microstructures in the unirradiated (Y, Al) ODS and (Y, Zr) ODS
sus the contact depth hc for a perfect Berkovich indenter is A(hc ) have been fully investigated in our previous work [19]. The av-
= 24.56(hc )2 , an axisymmetric rigid conical indenter with the eraged grain sizes of (Y, Al) ODS and (Y, Zr) ODS were approxi-
apex angle of 70.3° was chosen for modeling the Berkovich in- mately 460 and 370 nm, respectively, while the dislocation den-
denter, which is shown in Fig. 2. The tip radius was typically sities in these two ODS ferritic steels were 1.2 × 1014 m−2 [19].
set to be 100 nm to represent the bluntness of the indenter Most of oxide particles in (Y, Al) ODS and (Y, Zr) ODS were Y-
in practical experiments [39]. The specimen had the dimensions Al-O particles (YAP, YAH in YAlO3 ) [33] and Y-Zr-O particles (δ -
of 40 μm × 40 μm (length × width), which was meshed with phase Y4 Zr3 O12 , fluorite Y2 Zr2 O7 ) [19,34], respectively. The mean
36720 8-node biquadratic axisymmetric quadrilaterals (CAX8R). diameters and number densities of oxide particles in (Y, Al) ODS
The ABAQUS code calculated the multi-axial stress/strain state by and (Y, Zr) ODS were 6.0 ± 1.5 nm/(5.2 ± 1.0) × 1022 m−3
using the von Mises stress potential and associated J2 flow law and 6.4 ± 1.5 nm/(6.9 ± 0.4) × 1022 m−3 , respectively [19]. The
[40]. The elastic modulus E and Poisson’s ratio ν of the two ODS higher density of oxide particles, usually possessing a smaller
ferritic steels were 225 GPa and 0.3, respectively, which were iden- inter-particle distance, can effectively impede the dislocation climb
tical with those during experimental nanoindentation testing. The motion at elevated temperatures, which maintained the higher
specimen was modelled as being locally isotropic. The uniaxial creep strength of (Y, Zr) ODS than (Y, Al) ODS [1,4].

3
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

Fig. 2. Sketch of the axisymmetric cone tip-specimen model, consisting of 23 irradiated layers within the displacement damage profile.

Fig. 3. (a) scheme of the plastic properties of the unirradiated and the irradiated ODS steels, and the curves of load versus indentation depth for (b) (Y, Al) ODS and (c) (Y,
Zr) ODS, obtained by NI tests and FEM modeling.

After the ion irradiations to 2 and 10 dpa/nominal, TEM obser- dpa/nominal) [22]. This short diffusion distance indicated the neg-
vations have been done for the local microstructures in the depth ligible particle recovery through the recoiled solutes (particularly
range of ∼400 – 1300 nm from the specimen surface to avoid the Y) diffusing to the nanoparticles and the re-precipitation. Oxygen
effects of surface sink and injected Fe interstitials (see Fig. 1). The as an interstitial atom in bcc Fe has an exceptional high affinity for
dpa dependence of microstructures can be obtained from each lo- vacancies [42] and the diffusion coefficient of oxygen in vacancy-
cal depth corresponding to a well-defined local damage level (local included bcc Fe at RT was estimated to be around 9.53 × 10−7
dpa). Fig. 4 shows the averaged diameter and number density of nm2 /s based on Arrhenius equation [42]. Furthermore, the diffu-
oxide particles as a function of local dpa. It is obvious that, both di- sion of oxygen can be further reduced because both oxygen and
ameters and number densities of oxide particles in (Y, Al) ODS and vacancy structures would form strong chemical bonds with the
(Y, Zr) ODS decreased continuously with increasing damage level, solute Ti and Y atoms [43]. So in this work, the dissolved oxide
and the number densities at ∼30 dpa/local were closely one order forming atoms were expected to be at the vicinity of the parent
of magnitude lower than those at 0 dpa. Importantly, it is difficult oxide particles, which was confirmed by the satellite precipitates
to find significant differences between the dispersion morpholo- produced in the two ion-irradiated FeCrAl(Zr)-ODS steels at 200 °C
gies of oxide particles in the two FeCrAl(Zr)-ODS steels under RT [19]. The evolution rate of the particle radius r can be solely con-
irradiations, whereas oxide particles in (Y, Zr) ODS remained little trolled by the ballistic dissolution and depicted by the following
bigger volume fractions than (Y, Al) ODS after the Fe-ion irradia- equation [44]:
tions at 200 °C [19]. The oxide particles in (Y, Ti) ODS had much
higher number densities than those of (Y, Al) ODS and (Y, Zr) ODS dr/dt = −ζ K  (5)
below ∼20 dpa/local, but it became inverse above ∼20 dpa/local, where ζ K represents the flux of atoms from the particle surface, K
which is shown in Fig. 4(b). is the displacement rate (in dpa/s) in the matrix,  is the atomic
According to the vacancy mechanism, the radiation enhanced volume of ejected atoms. Supposing close fluxes of ejected atoms
diffusion coefficient of Y atom, DY at RT was estimated to be from Y-X-O (X = Ti, Al, Zr) particles for a given displacement rate,
5.3 × 10−9 nm2 /s and the corresponding diffusion distance was small oxide particles tended to be invisible during TEM observa-
around 0.02 nm during the irradiation period (∼5 h for 10 tions after a critical period of ion irradiation, which explained the

4
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

Fig. 4. (a) Averaged diameter and (b) number density of oxide particles as a function of local dpa in (Y, Ti) ODS [22], (Y, Al) ODS and (Y, Zr) ODS irradiated to nominal
displacement damages of 2 and 10 dpa. Each local damage corresponds to the calculated value based on SRIM simulations at the corresponding depth.

lower number densities of oxide particles in (Y, Ti) ODS than those dpa/local in Fig. 4(b)), indicating that most of oxide particles in (Y,
of (Y, Al) ODS and (Y, Zr) ODS above ∼20 dpa/local. However, the Al) ODS and (Y, Zr) ODS were dissolved into ferrite matrix.
threshold displacement energies of oxide forming atoms like Y, Ti, Previous studies [49,50] demonstrated that, after ion irradia-
Al, Zr in nano-sized particles require to be known for well un- tions at RT, the primary dislocation loops formed in Fe-Cr al-
derstanding the distinct fluxes ζ K of ejected atoms produced by loys consisted of the 1/2<111> and 100 interstitial types but no
knock-on atoms. Monnet et al. [45] investigated the particle-size 1/2<110> type. According to the g•b = 0 invisibility criteria, a half
evolution of different oxides Y2 O3 , MgAl2 O4 , MgO and Al2 O3 inside of the 1/2<111> loops and two thirds of the 100 loops are visi-
ferritic matrix by using 1 MeV electrons at 400 °C. It is found that ble with g = [11̄0], while all of the 1/2<111> loops and one third
Y2 O3 showed the negligible diameter shrinkage compared with the of the 100 loops are visible with g = [002]. Fig. 5 shows the
other oxides because of the high threshold displacement energy of microstructures of (Y, Al) ODS (Figs. 5(a), 5(b)) and (Y, Zr) ODS
Y [45]. (Figs. 5(d), 5(e)) in an identical grain for each ODS steel after the
In the previous studies [46–48], nearly complete dissolution of irradiation to ∼72 dpa/local at the depth of about 900 nm, which
oxide particles without the formation of new clusters or precipi- were observed by using g = [11̄0] and g = [002] under weak
tates was observed by atom probe tomography (APT) after heavy beam dark field (WBDF, g3g) conditions. In Figs. 5(a), 5(b), 5(d)
ion irradiations at RT [46], at 100 °C and -75 °C [47] where the and 5(e), there are many tiny “white dots”, marked with dashed
damage level ranged from about 50 to 150 dpa. This phenomenon circles. For each ODS steel, the density of “white dots” imaged by
was also confirmed by glancing-incident angle X-ray diffraction g = [002] was evidently bigger than that imaged by g = [11̄0],
(GIXRD) for the ion-irradiation (-123 °C, ∼50 – 100 dpa) [48]. whereas the visibility of oxide particles via strain contrast does
Moreover, the dissolved oxide forming atoms like Y and O were not depend on the chosen g vector. So the “white dots” circled can
present as solute atoms in ferrite confirmed by APT [46,47]. After be considered as dislocation loops and the other weak contrasts
the irradiation up to 50 dpa/nominal (around 30 – 150 dpa/local), without being circled might be some surface contaminations. Ox-
the particle density would be lower than 4.6 × 1021 m−3 (see ∼30 ide particles in the unirradiated grains of (Y, Al) ODS and (Y, Zr)

5
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

Fig. 5. Microstructures in the irradiated grains of (Y, Al) ODS (a, b) and (Y, Zr) ODS (d, e) observed under weak beam dark field (WBDF, g3g) imaging conditions by using g =
[11̄0] and g = [002] close to pole [110] in an identical region at ∼900 nm (∼72 dpa/local) at a nominal damage of 50 dpa. High density of oxide particles in the unirradiated
(Y, Al) ODS (c) and (Y, Zr) ODS (f) imaged by using g = [11̄0] shown here for comparison with dislocation loops. Dislocation loops circled were counted for estimating
number density.

ODS imaged by g = [11̄0] are shown in Figs. 5(c) and 5(f), re- about 330 nm, the PDZ expanded to the unirradiated substrate to
spectively, in contrast with the distinct dispersion morphologies cause the decrease of the relative hardening. It was found by the
of dislocation loops. As a result, the total number densities of FEM simulations that the PDZ at the indentation depth less than
both 1/2<111> and 100 loops in (Y, Al) ODS and (Y, Zr) ODS 330 nm was limited within the irradiated region (0 – 2220 nm).
at ∼72 dpa/local were (1.3 ± 0.2) × 1022 and (6.7 ± 0.7) × 1021 In Nix-Gao model [30] based upon strain gradient plasticity
m−3 , respectively, with the corresponding averaged loop sizes of (SGP) theory, the ISE is modelled by using the concept of GNDs for
(3.9 ± 0.9) and (4.2 ± 1.0) nm. Moreover, the fractions of 1/2<111> a crystalline material, as shown schematically in Fig. 7. All GNDs
loop were 79.9 % and 92.5 % for (Y, Al) ODS and (Y, Zr) ODS, re- are assumed to be circular loops within a hemispherical volume V
spectively. In the (Y, Ti) ODS at the same irradiation condition (RT, given by the contact radius ac , where ac = h/tanθ , (90°-θ ) is the
∼72 dpa/local), the averaged number density and size of disloca- apex angle of a conical indenter (see Fig. 7). The depth-dependent
tion loops were (4.1 ± 0.7) × 1022 m−3 and 2.8 ± 0.7 nm, respec- density ρ G of GNDs is defined as the total line length λ of disloca-
tively [22]. Because of the similar sizes of dislocation loops, the tion loops divided by the hemispherical volume V [30]:
higher density of dislocation loops demonstrated more interstitial π hac
λ 3 tan2 θ
defects formed in the ion-irradiated ODS steels. ρG = = b
= (6)
V 2
3
π a3c 2 bh
3.2. Experimental measurements of nanoindentation (NI) hardness where b is the Burgers vector (0.248 nm for 1/2<111> dislocations
in ferrite). Moreover, the nanoindentation (NI) hardness H at a cer-
Figs. 6(a) and 6(d) show the depth profiles of the averaged tain indentation depth h is given by the following formula [30]:
nanoindentation hardness (HNI ) of (Y, Al) ODS and (Y Zr) ODS,
 
respectively, before and after irradiations, where all averaged HNI H ρG h∗
continuously decrease with increasing the indentation depth (h) = 1+ = 1+ (7)
H0 ρS h
to 20 0 0 nm, due to the indentation size effect (ISE). The relative
hardening (HIrr /HUnirr - 1) was plotted in Figs. 6(b) and 6(e), where where H0 is the depth-independent bulk equivalent hardness only
HIrr and HUnirr are the average hardness values at the indentation related to the density ρ S of the statistically stored dislocations
depth of the unirradiated and the irradiated specimens, respec- (SSDs); h∗ represents a characteristic length that depends on the
tively. The measured hardness at the indentation depth less than indenter geometry (tanθ ), ρ S , and the shear modulus μ of a given
150 nm, which is strongly affected by the specimen surface condi- material. In spite of the damage gradient effect (DGE), Nix-Gao
tions such as oxidation, roughness [51], was not considered. From model is still widely adopted to evaluate a bulk equivalent hard-
150 to approximately 330 nm the relative hardening increased fast, ness for a thin ion-irradiated layer [25,27,39,52].
and then decreased slowly from approximately 330 to 20 0 0 nm. Figs. 6(c) and 6(f) show the part plots of the square of the
The increasing relative hardening (150 – 330 nm) is mainly as- measured hardness (HNI )2 against the reciprocal of the indenta-
cribed to the increasing displacement damage with depth. Beyond tion depth 1/h, where the linear fitting in the indentation depth

6
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

Fig. 6. The average hardness profiles of (a) (Y, Al) ODS and (d) (Y, Zr) ODS, the corresponding profiles of relative hardening (HIrr /HUnirr −1) (b, e), and each plot of (HNI )2
versus inverse depth 1/h for (c) (Y, Al) ODS and (f) (Y, Zr) ODS, and (g) equivalent bulk hardness H0 measured for the indentation depth (150 nm < h <330 nm) of the ODS
ferritic steels including (Y, Ti) ODS [22] estimated by using Nix-Gao model, as well as (f) the corresponding hardening H0 as a function of nominal dpa.

range of 150 – 330 nm was used to evaluate the bulk equiva- tion depth (∼150 nm). The averaged H0 values of the unirradiated
lent hardness of the surface thin layers for the unirradiated and (Y, Al) ODS and (Y, Zr) ODS were (4.42 ± 0.21) and (4.82 ± 0.20)
the irradiated specimen (see the dashed lines), and thus the aver- GPa, respectively, which can be converted to the corresponding
aged H0 values were deduced. Although the CSM during the ex- yield strength σ Y values of 933.6 and 1041.3 MPa of each ODS ma-
perimental nanoindentation testing should cause the fluctuation of terial with the relationship:
the measured hardness, all the values of R-squared factor (R2 ) (se-  
lectively shown in Figs. 6(c) and 6(f)) were larger than 0.88, which
σY (MPa ) ≈ 214 H0 × 103 MPa (8)
means the good linear fitting of H2 versus 1/h. The applicability of Large scatter in the measured H0 in Fig. 6(g) may be due to inho-
Eq. (7) in the case of irradiated specimen is further discussed in mogeneous microstructures or/and variable crystal orientations of
the Section 3.3. For the unirradiated ODS specimens, each plot of indented grains of ODS ferritic steels [28]. The amount of irradi-
(HNI )2 versus 1/h yielded a nearly straight line rather than a bi- ation hardening, H0 were obtained by subtracting the averaged
linear curve (see black lines in Figs. 6(c) and 6(f)). In other words, H0 of the unirradiated from the irradiated counterpart. As shown
the initial dislocations barriers (e.g. oxide particles) in ODS steels in Fig. 6(h), it is apparent that the amount of hardening increased
ensured a continuous dislocation activity even at a small indenta- considerably with dpa. Despite large scatters, (Y, Zr) ODS exhibited

7
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

Table 1
Maximum depth Z of plastically deformed zone (PDZ) at the indentation depth of
330 nm for equivalent plastic strain ε eq of 0.002 and 0.019.

FeCrAl(Zr)-ODS steels Y-Al-ODS Y-Zr-ODS

ε eq > 0.002 > 0.019 > 0.002 > 0.019


Z, nm of the unirradiated 2732.0 1936.8 2646.9 1904.0
Z, nm of the 50 dpa/nominal 2974.1 2119.8 2815.3 1990.0

nanoindentation experiments. The corresponding FEM load-depth


curves are also shown in Figs. 3(b) and 3(c) for the case of the 50
dpa/nominal. We recognized here that if the strain hardening rate
n would be affected by irradiation, then that the fitted parameter
Fig. 7. Geometrically necessary dislocations produced by a rigid conical indentation, C with Eqs. (3a) and (3b) would be in turn affected.
which are idealized as a circular dislocation loops in Nix-Gao model [30]. Another important issue is the relation between the PDZ size
and the density of GNDs. Figs. 8(a) and 8(b) show the contour plots
of equivalent plastic strain (PEEQ) larger than 0.2 % for (Y, Al) ODS
the lowest hardening at all nominal damage levels compared with before and after the ion-irradiation at 50 dpa/nominal at the in-
(Y, Al) ODS and (Y, Ti) ODS in our previous study [22]. The hard- dentation depth h of 330 nm, which corresponds to the measured
enings H0 (in GPa) of (Y, Al) ODS and (Y, Zr) ODS were fitted to relative hardness peak (see Figs. 6(b) and 6(e)). The PDZ shown
be H0 = 0.62 × (dpa)0.26 and H0 = 0.44 × (dpa)0.25 , respec- in Figs. 8(a) and 8(b), expanded much deeper than the indenta-
tively. The exponents of those two ODS ferritic steels (also the m tion depth and also along the specimen surface beyond the con-
in Eq. 3(b)) are in a good agreement with 0.25 – 0.3 of the Fe-(9 tact radius, which was much larger than the PDZ defined in Nix-
– 10)Cr alloys irradiated with Fe ions at RT [26,52], although the Gao model (see Fig. 7) for which the contact radius corresponds to
exponent of (Y, Ti) ODS is 0.16 much smaller than the other ODS the plastic zone radius. For the equivalent plastic strain ε eq > 0.2
ferritic steels. The possible reason of this difference among the ODS % or 1.9 %, the values of the maximum depth Z below indenter are
steels will be discussed in the Section 3.4. displayed in Table 1. The PDZ size (ε eq > 0.2%) turned out to be
8 – 9 times the indentation depth, irrespective of ion irradiations.
3.3. FEM simulation for hardening, plastically deformed zone and In fact, we can regard Z as the radius of a hemispherical plastically
GNDs deformation zone and the scaling factor f was defined as the ratio
of Z to the contact radius ac :
Based on the elastic-plastic properties of the unirradiated spec-
imens for (Y, Al) ODS and (Y, Zr) ODS, the curves of load versus f = Z / ac (9)
indentation depth were obtained through FEM simulations, which In the case of 50 dpa/nominal at h = 330 nm, the PDZ radius Z for
are plotted in red dashes in Figs. 3(b) and 3(c). For the unirradi- εeq > 0.2 % and > 1.9 % were 2815.3 – 2974.1 nm and 1990.0 –
ated cases, the parameter Sred values of (Y, Al) ODS and (Y, Zr) ODS 2119.8 nm, respectively. The latter is similar with the depth range
were (6.0 ± 1.5) × 10−4 and (8.0 ± 1.8) × 10−4 , respectively. These (0 – 2220 nm) of the ion-irradiated layer. However, the larger PDZ
discrepancies were possibly caused by the harmonic displacement (for ε eq > 0.2 %) clearly extends into the soft unirradiated sub-
oscillation of the CSM or the residual surface roughness [53]. strate. So the PDZ size at ε eq > 1.9 % was considered to estimate
Previous works [6,54–58] revealed that, the neutron irradia- the density of GNDs. Durst et al. [59] found that, compared with
tions modified the elastoplastic and plastic instability behaviors the strain limit ε eq of 0.2 %, the PDZ at ε eq > 1.5% gave rise to a
of bcc and hcp metals or alloys to some extent, which depended reasonable estimation of the density of GNDs and the NI hardness
much on the neutron fluence, irradiation temperature and initial for various metallic materials. Thus, after replacing the contact ra-
microstructures (e.g. grain size). Furthermore, the true stress-true dius ac with Z, the density of GNDs in Eq. (6) is rewritten as:
strain curves after the yield point showed nearly same continu-
ous strain hardening tendency with those of the unirradiated steels π hZ
λ 3 h 3 tan2 θ
[54,56]. Maloy et al. [54] ascribed this phenomena to the fine grain ρG = = b
= = (10)
size on the order of 1 μm or less, which can produce high strain
V 2
3
π z3 2 b( f ac )2 2 f 2 bh
hardening rates and limit the length and propagation of nascent
Accordingly, the relationship of H versus h in the Nix-Gao model
localized shear bands. In this work, the two FeCrAl(Zr)-ODS fer-
becomes
ritic steels have ultrafine grain sizes less than 1 μm [19]. Thus
 
it was assumed that the strain hardening behavior of the stud- H ρG h∗
= 1+ = 1+ (11)
ied ODS ferritic steels was the same before and after irradiation, H0 ρS f 2h
which is shown in Fig. 3(a). This assumption is further supported
by the analysis of various metals (bcc, fcc, hcp) in terms of true For ε eq > 1.9%, the ratio f and the density of GNDs are shown
stress-true strain curves, where all strain-hardening rates were Fig. 9 as a function of indentation depth. In the unirradiated cases,
nearly not affected by the neutron irradiations at low temperatures the ratio f was almost depth-independent and around 1.95, which
(< 200 °C), even in the case of localized deformation through dis- means that the constant value of the slope h∗ /f2 ensured the va-
location channeling induced by irradiation [58]. After taking the lidity of the linear fitting on the experimental plots of H2 versus
depth dependence of displacement damage into account, the load- 1/h. However, in the cases of 50 dpa/nominal, the ratio f decreased
depth curves for the 50 dpa/nominal could be obtained through from 1.95 to 1.8 at the indentation depth h < 330 nm, and then
the FEM simulations (see Fig. 2) for a given C in Eq. (3b). With went up quickly to approximately 2.2 at h > 330 nm. Although
the parameter C adjusted to be 153.6 and 158.4 for (Y, Al) ODS it is recommended to consider the depth dependence of homoge-
and (Y, Zr) ODS, respectively, the corresponding values of the pa- nously distributed microstructures (e.g. radiation defects or clus-
rameter Sred were (6.2 ± 2.4) × 10−4 and (5.3 ± 1.5) × 10−4 , ters) when using the Nix-Gao model to obtain a bulk equivalent
which explained good consistency between FEM simulation and nano-hardness, the Nix-Gao model could be valid in the case of

8
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

Fig. 8. Contour plots of different equivalent plastic strains in (Y, Al) ODS before irradiation (a) and after the irradiation up to 50 dpa/nominal.

in Fig. 9(b), which display close tendencies for the two FeCrAl(Zr)-
ODS ferritic steels, decreased from ∼1.1 × 1015 m−2 to ∼7.0 × 1014
m−2 for the 50 dpa/nominal when the indentation depth increased
from 200 nm to 300 nm, slightly higher than those of the unir-
radiated (ranging from ∼9.8 × 1014 m−2 to ∼6.7 × 1014 m−2 ).
For comparison we mention that, in the pure Ni irradiated with
6 MeV protons to 0.1 dpa at 120 °C, at the indentation depth of
∼200 – 300 nm, the density of GNDs varied from ∼2.0 × 1016
m−2 to ∼1.5 × 1016 m−2 , much higher than those of the unirra-
diated (∼7.0 × 1015 m−2 ) [60]. Such discrepancy can be associated
with the high density of statistically stored dislocations (SSDs) in-
side the as-received ODS steels, the motion and multiplication of
which can contribute much to the primary plastic deformation un-
derneath nanoindenter. Therefore, it is better to account for the
density of GNDs when estimating the irradiation-induced harden-
ing at a shallow reference depth.
The FEM method in this work, consisting of divided layers of
the irradiation region having a dose dependent hardening, could
certainly be considered for other materials (e.g. Fe-Cr alloy [26])
besides ODS steels. A sufficient number of discrete layers have to
be used so that the numerical solution does not depend on the
number of layers. Additionally, a large number of layers can reduce
the discontinuity in the deformation between adjacent layers.

3.4. Hardening mechanism

According to the results of FEM simulations, where the


dpa-dependence of yield strength σ Y is σ Y (MPa) = 933.6 +
153.6 × (dpa)0.26 and σ Y (MPa) = 1041.3 + 158.4 × (dpa)0.25 for
(Y, Al) ODS and (Y, Zr) ODS, respectively, the depth dependence
of yield strength is shown in Fig. 10. The stress increment σ
for (Y, Al) ODS and (Y, Zr) ODS at ∼72 dpa/local (50 dpa/nominal,
∼900 nm towards surface) was 467.1 and 461.3 MPa, respectively,
Fig. 9. (a) the ratio f of Z to contact radius ac and (b) the density of geometrically
which are just larger than the counterparts of (353.1 ± 84.5) and
necessary dislocations (GNDs) as a function of indentation depth h. (240.3 ± 89.4) MPa, obtained by the Nix-Gao model on the basis
of Eq. (8). Accompanied by the dissolution of oxide particles, the
dissolved solute atoms (e.g. Y, O, Al, Zr) and radiation-induced
ion-irradiated damages based on the good linear fitting on H2 ver- dislocation loops can act as short-range obstacles for dislocation
sus 1/h in this work and also in others [27,39,52]. The reason can motion, hence, causing changes in the flow stress of ODS steels.
be that, the slope h∗ /f2 is balanced by the decreased ratio f and the Actually, the square superposition law was valid just when obsta-
increased ρ S with increasing depth. The densities of GNDs shown cle strengths are weak and nearly the same [61], the combined

9
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

Table 2
Hardness changes of two ODS steels contributed by oxide particles, dislocation loops and dissolved O interstitials.

Oxide particles at 0 dpa Dislocation loops at 50 dpa/nominal O interstitials at 50 dpa/nominal Final stress

Do , nm No , × 1021 m−2 σ o , MPa 1)


Dl , nm Nl , × 1021 m−2 σ l , MPa 2)
CatO , at. % σ so , MPa σ f , MPa
(Y, Al) ODS 6.0 52 428.9 3.9 13 324.2 0.56 561.6 456.9
(Y, Zr) ODS 6.4 69 510.3 4.2 6.7 241.5 0.63 630.3 361.5

1) Strength factor of oxide particles α o : 0.4; 2) Strength factor of dislocation loops α l : 0.75.

dislocation loops cannot be fully imaged even under WBDF due to


their weak strain contrast. Another possibility might be an extra
contributor like oxygen-vacancy complexes. As numerous vacancies
were caused by ion irradiation at RT, oxygen-vacancy complexes
were expected to be formed because of the high affinity between
oxygen and vacancy in bcc Fe [42,43]. These oxygen-vacancy com-
plexes may trap screw dislocations via attractive interactions and
induce a considerate hardening in bcc alloys [65]. It was reported
that high density oxygen-vacancy complexes in bcc vanadium (V)
alloy caused the fracture modes from dimple to a fully transgraular
cleavage as the oxygen concentration increasing from 0.1 to 1.6 at.%
[65].
In the irradiated ODS ferritic steels, both oxide particles and ra-
diation defects impeded the motion of dislocation lines. The num-
ber density of the former was continuously decreased by irradia-
tion (see Fig. 4(b)), which is much more significant in (Y, Ti) ODS
than the other two steels. This result indicates that in (Y, Ti) ODS
dose dependence of hardening, namely the exponent, can be lower
(see Fig. 6(h)). Actually, accompanied by the particle dissolution,
Fig. 10. Depth dependence of yield strength of (Y, Al) ODS and (Y, Zr) ODS before more radiation defects particularly interstitial loops would be ac-
and after the ion-irradiation at 50 dpa/nominal. cumulated in (Y, Ti) ODS due to the weakened trapping capacity of
P-M interfaces. In (Y, Al) ODS and (Y, Zr) ODS, the dispersion mor-
phology of oxide particles responded similarly and less sensitive
change in the flow stress in this work was calculated by the
to ion irradiations at RT. So the exponent of ∼0.25 in Fig. 6(h) has
linear-superposition model:
been considered to be adequate to describe fluence dependence re-
σf = σl + σso − σo (12) flecting the accumulation of radiation defects.
The simulation work based on density functional theory re-
where σ l , σ so and σ o represent the strengthening contribu-
vealed that four distinct Y2 O3 -Fe interfaces can trap strongly both
tors by dislocation loops, solid atoms and oxide particles, respec-
interstitial and vacancy defects via long-range attraction [66]. The
tively. Furthermore, σ l and σ o can be expressed by the follow-
trapping effect of particle-matrix (P-M) interfaces on point defects
ing equation based on Orowan type bypassing mechanism [44]:
can be measured by the sink strength of oxide particles Zo , be-
σ = Mαμb(Nd )0.5 (13) ing commonly expressed by 4π No do , where No and do are num-
ber density and diameter of oxide particles, respectively [5]. Prior
where M is the Taylor factor (3.06); α is the barrier strength factor to irradiation, the initial sink strengths of oxide particles for (Y,
of the contributor, being 0 – 1; μ is the shear modulus (80 GPa Ti) ODS, (Y, Al) ODS and (Y, Zr) ODS were 20.7 × 1014 , 7.5 × 1014
for alpha iron); N and d are the number density and mean diam- and 13.6 × 1014 m−2 , respectively [11]. This is consistent with the
eter of the contributor, respectively. The strengthening produced lower hardening of (Y, Zr) ODS compared to (Y, Al) ODS, which was
by dissolved solute atoms could be formulated linearly on the mainly ascribed to the dislocation loops formed in the latter (see
concentration of solute atom CatX in bcc alloys. Oxygen atoms as Table 2). It is expected that, at a low displacement damage, the
interstitials in bcc ferrite, having the strengthening coefficient of minimum amount of radiation defects would be produced inside
around 10 0 0 MPa per at.% [62], contributed to 561.6 and 630.3 MPa (Y, Ti) ODS. However, this is in contrast with the fact that the hard-
for (Y, Al) ODS and (Y, Zr) ODS MPa, respectively, which are shown ening was slightly higher in (Y, Ti) ODS than the other two at 2
in Table 2. However, the strengthening coefficients of substitu- dpa/nominal (see Fig. 6(h)). The reason is possibly that, tiny oxide
tional solute atoms like Y (20 MPa per at.%) [62] and Al (9 MPa particles in (Y, Ti) ODS cannot trap self-interstitials in an efficient
per at.%) [63] are far smaller than that of O, so the corresponding way because of their coherent particle-matrix (P-M) interfaces of
strengthening components (∼5 MPa for Y or Al) can be negligible. a low interfacial energy [20,33]. Likewise, the response of grain
The stress increments by O interstitials compensated the lost parts boundaries (GBs) to point defects depended much on GB crystal-
of oxide particles for the corresponding obstacle strength factor lographic parameters such as coherency and misorientation angle
α o of 0.4 (see Table 2). As the strength factor of dislocation loops [67–70], while low energy GBs presented weaker capacity of trap-
α l to be 0.75, the final calculated stress increments of (Y, Al) ODS ping point defects than high energy GBs [68,70].
and (Y, Zr) ODS were 456.9 and 361.5 MPa, respectively, which are
close to those obtained by the FEM simulation (at ∼72 dpa/local).
However, the strength factor of dislocation loops (0.75) in this 4. Conclusions
work is obviously larger than 0.17 – 0.34 for 1/2 111 loops
(7.6 ± 3.8 nm) in the 12Cr-6Al-ODS ferritic steel after neutron irra- Two FeCrAl(Zr)-ODS ferritic steels, named by (Y, Al) ODS and (Y,
diation at 300 °C to 2.6 dpa [64]. This fact was possibly caused by Zr) ODS, were irradiated with 6.4 MeV Fe ions up to nominal dis-
the underestimation of the loop density in this work, because tiny placement damages of 2, 10 and 50 dpa at RT. Irradiation-induced

10
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

hardening, plastically deformed zone (PDZ), geometrically neces- [4] S. Ukai, S. Kato, T. Furukawa, S. Ohtsuka, High-temperature creep deformation
sary dislocations (GNDs) and post-irradiation microstructures were in FeCrAl-oxide dispersion strengthened alloy cladding, Mater. Sci. Eng. A. 794
(2020) 139863, doi:10.1016/j.msea.2020.139863.
investigated by experimental nanoindentation (NI) measurements, [5] G.R. Odette, M.J. Alinger, B.D. Wirth, Recent developments in irradiation-
FEM simulation and TEM observations. Main results are as follow- resistant steels, Annu. Rev. Mater. Res. 38 (2008) 471–503, doi:10.1146/
ings: annurev.matsci.38.060407.130315.
[6] Z. Zhang, T.A. Saleh, S.A. Maloy, O. Anderoglu, Microstructure evolution in
1) Experimental NI measurements revealed that, the radiation MA956 neutron irradiated in ATR at 328°C to 4.36 dpa, J. Nucl. Mater. (2020)
533, doi:10.1016/j.jnucmat.2020.152094.
hardening of the two FeCrAl(Zr)-ODS steels increased contin- [7] B. van der Schaaf, C. Petersen, Y. De Carlan, J.W. Rensman, E. Gaganidze,
uously with elevating the nominal displacement damage. Af- X. Averty, High dose, up to 80 dpa, mechanical properties of Eurofer 97, J. Nucl.
ter taking the damage gradient effect (DGE) into account, the Mater. 386–388 (2009) 236–240, doi:10.1016/j.jnucmat.2008.12.329.
[8] E. Aydogan, J.S. Weaver, U. Carvajal-Nunez, M.M. Schneider, J.G. Gigax,
dependence of the irradiation hardening (in MPa) on local D.L. Krumwiede, P. Hosemann, T.A. Saleh, N.A. Mara, D.T. Hoelzer, B. Hilton,
dpa through the FEM simulations was demonstrated to be S.A. Maloy, Response of 14YWT alloys under neutron irradiation: a comple-
153.65 × (dpa)0.26 and 158.35 × (dpa)0.25 for (Y, Al) ODS and mentary study on microstructure and mechanical properties, Acta Mater 167
(2019) 181–196, doi:10.1016/j.actamat.2019.01.041.
(Y, Zr) ODS, respectively. [9] C.R.F. Azevedo, Selection of fuel cladding material for nuclear fission reactors,
2) The similar hardening behaviors of both steels were related Eng. Fail. Anal. 18 (2011) 1943–1962, doi:10.1016/j.engfailanal.2011.06.010.
to the close dispersion morphologies of oxide particles at a [10] H. Zhao, T. Liu, Z. Bai, L. Wang, W. Gao, L. Zhang, Corrosion behavior of
14Cr ODS steel in supercritical water: the influence of substituting Y2 O3 with
given dpa, which affected the accumulation of radiation defects
Y2 Ti2 O7 nanoparticles, Corros. Sci. (2020) 163, doi:10.1016/j.corsci.2019.108272.
particularly interstitial dislocation loops in specimens. More- [11] P. Song, D. Morrall, Z. Zhang, K. Yabuuchi, A. Kimura, Radiation response of
over, the irradiation-induced stress increments were mainly ODS ferritic steels with different oxide particles under ion-irradiation at 550°C,
J. Nucl. Mater. 502 (2018) 76–85, doi:10.1016/j.jnucmat.2018.02.007.
contributed by the dissolved oxygen and the dislocation loops
[12] B. Duan, C. Heintze, F. Bergner, A. Ulbricht, S. Akhmadaliev, E. Oñorbe, Y. de
formed. Carlan, T. Wang, The effect of the initial microstructure in terms of sink
3) The PDZ sizes at the equivalent plastic strain ε eq > 0.2 % for strength on the ion-irradiation-induced hardening of ODS alloys studied by
the ODS steels before and after irradiations were ∼8 – 9 times nanoindentation, J. Nucl. Mater. 495 (2017) 118–127, doi:10.1016/j.jnucmat.
2017.08.014.
of the indentation depth. However, the PDZ sizes at the strain [13] Z.N. Ding, C.H. Zhang, Y.T. Yang, Y. Song, A. Kimura, J. Jang, Hardening of ODS
limit ε eq > 1.9 % agreed well with the depth dependence of the ferritic steels under irradiation with high-energy heavy ions, J. Nucl. Mater.
relative hardening and were adopted to estimate the density of 493 (2017) 53–61, doi:10.1016/j.jnucmat.2017.05.040.
[14] C. Heintze, I. Hilger, F. Bergner, T. Weissgärber, B. Kieback, Nanoindentation
GNDs. of single- (Fe) and dual-beam (Fe and He) ion-irradiated ODS Fe-14Cr-based
4) According to the strain gradient plasticity (SGP) theory, the es- alloys: effect of the initial microstructure on irradiation-induced hardening, J.
timated densities of GNDs in the two FeCrAl(Zr)-ODS under in- Nucl. Mater. 518 (2019) 1–10, doi:10.1016/j.jnucmat.2019.02.037.
[15] M.J. Swenson, J.P. Wharry, Nanocluster irradiation evolution in Fe-9%Cr ODS
dentation were close: they decreased from ∼1.1 × 1015 m−2 to and ferritic-martensitic alloys, J. Nucl. Mater. 496 (2017) 24–40, doi:10.1016/j.
∼7.0 × 1014 m−2 for the 50 dpa/nominal when the indenta- jnucmat.2017.08.045.
tion depth increased from 200 nm to 300 nm, slightly higher [16] C.P. Massey, P.D. Edmondson, K.G. Field, D.T. Hoelzer, S.N. Dryepondt, K.A. Ter-
rani, S.J. Zinkle, Post irradiation examination of nanoprecipitate stability and α 
than those of the unirradiated (ranging from ∼9.8 × 1014 m−2
precipitation in an oxide dispersion strengthened Fe-12Cr-5Al alloy, Scr. Mater.
to ∼6.7 × 1014 m−2 ). 162 (2019) 94–98, doi:10.1016/j.scriptamat.2018.10.047.
[17] S. Yamashita, K. Oka, S. Ohnuki, N. Akasaka, S. Ukai, Phase stability of oxide
dispersion-strengthened ferritic steels in neutron irradiation, J. Nucl. Mater.
Declaration of Competing Interest 307–311 (2002) 283–288, doi:10.1016/S0022-3115(02)01077-2.
[18] A.G. Certain, K.G. Field, T.R. Allen, M.K. Miller, J. Bentley, J.T. Busby, Response
The authors declare that they have no known competing finan- of nanoclusters in a 9Cr ODS steel to 1 dpa, 525°C proton irradiation, J. Nucl.
Mater. 407 (2010) 2–9, doi:10.1016/j.jnucmat.2010.07.002.
cial interests or personal relationships that could have appeared to
[19] P. Song, A. Kimura, K. Yabuuchi, P. Dou, H. Watanabe, J. Gao, Y.J. Huang, As-
influence the work reported in this paper. sessment of phase stability of oxide particles in different types of 15Cr-ODS
ferritic steels under 6.4 MeV Fe ion irradiation at 200°C, J. Nucl. Mater. 529
(2020) 151953, doi:10.1016/j.jnucmat.2019.151953.
Acknowledgments
[20] T. Chen, J.G. Gigax, L. Price, D. Chen, S. Ukai, E. Aydogan, S.A. Maloy, F.A. Gar-
ner, L. Shao, Temperature dependent dispersoid stability in ion-irradiated
The authors are also grateful to the technical support by mem- ferritic-martensitic dual-phase oxide-dispersion-strengthened alloy: coherent
interfaces vs. incoherent interfaces, Acta Mater 116 (2016) 29–42, doi:10.1016/
bers of Application of Duet and Muster for Industrial Research and
j.actamat.2016.05.042.
Engineering (ADMIRE), Kyoto University. The first author wishes [21] M.L. Lescoat, J. Ribis, Y. Chen, E.A. Marquis, E. Bordas, P. Trocellier, Y. Ser-
to thank Prof. em. Akihiko Kimura at Kyoto University for fruit- ruys, A. Gentils, O. Kaïtasov, Y. De Carlan, A. Legris, Radiation-induced Ostwald
ful discussion on hardening mechanisms of the ion-irradiated ODS ripening in oxide dispersion strengthened ferritic steels irradiated at high ion
dose, Acta Mater 78 (2014) 328–340, doi:10.1016/j.actamat.2014.06.060.
ferritic steels. The financial support of the Swiss National Science [22] P. Song, J. Gao, K. Yabuuchi, A. Kimura, Ion-irradiation hardening accompanied
Foundation (Grant No. 20 0 021_184695) is gratefully acknowledged. by irradiation-induced dissolution of oxides in FeCr(Y, Ti)-ODS ferritic steel, J.
Nucl. Mater. 511 (2018) 200–211, doi:10.1016/j.jnucmat.2018.09.007.
[23] T.R. Allen, J. Gan, J.I. Cole, M.K. Miller, J.T. Busby, S. Shutthanandan, S. The-
Supplementary materials vuthasan, Radiation response of a 9 chromium oxide dispersion strengthened
steel to heavy ion irradiation, J. Nucl. Mater. 375 (2008) 26–37, doi:10.1016/j.
Supplementary material associated with this article can be jnucmat.20 07.11.0 01.
[24] C.K. Dolph, D.J. da Silva, M.J. Swenson, J.P. Wharry, Plastic zone size for
found, in the online version, at doi:10.1016/j.actamat.2022.117991. nanoindentation of irradiated Fe—9%Cr ODS, J. Nucl. Mater. 481 (2016) 33–45,
doi:10.1016/j.jnucmat.2016.08.033.
Reference [25] P. Hosemann, D. Kiener, Y. Wang, S.A. Maloy, Issues to consider using nano in-
dentation on shallow ion beam irradiated materials, J. Nucl. Mater. 425 (2012)
[1] A. Kimura, R. Kasada, N. Iwata, H. Kishimoto, C.H. Zhang, J. Isselin, P. Dou, 136–139, doi:10.1016/j.jnucmat.2011.11.070.
J.H. Lee, N. Muthukumar, T. Okuda, M. Inoue, S. Ukai, S. Ohnuki, T. Fujisawa, [26] C. Shin, H. ha Jin, M.W. Kim, Evaluation of the depth-dependent yield strength
T.F. Abe, Development of Al added high-Cr ODS steels for fuel cladding of next of a nanoindented ion-irradiated Fe-Cr model alloy by using a finite element
generation nuclear systems, J. Nucl. Mater. 417 (2011) 176–179, doi:10.1016/j. modeling, J. Nucl. Mater. 392 (2009) 476–481, doi:10.1016/j.jnucmat.2009.04.
jnucmat.2010.12.300. 011.
[2] B.A. Pint, K.A. Unocic, K.A. Terrani, Effect of steam on high temperature oxida- [27] R. Kasada, Y. Takayama, K. Yabuuchi, A. Kimura, A new approach to evaluate
tion behaviour of alumina-forming alloys, Mater. High Temp. 32 (2015) 28–35, irradiation hardening of ion-irradiated ferritic alloys by nano-indentation tech-
doi:10.1179/0960340914Z.0 0 0 0 0 0 0 0 058. niques, Fusion Eng. Des. 86 (2011) 2658–2661, doi:10.1016/j.fusengdes.2011.03.
[3] S. Ukai, T. Nishida, H. Okada, T. Okuda, M. Fujiwara, K. Asabe, Development of 073.
oxide dispersion strengthened ferritic steels for fbr core application, (I), J. Nul. [28] H.L. Yang, S. Kano, J. McGrady, J.J. Shen, Y.F. Li, D.Y. Chen, K. Murakami, H. Abe,
Sci. Technol. 34 (1997) 256–263, doi:10.1080/18811248.1997.9733658. Hardening behavior and deformation microstructure beneath indentation in

11
P. Song, K. Yabuuchi and P. Spätig Acta Materialia 234 (2022) 117991

heavy ion irradiated 12Cr-ODS steel at elevated temperature, J. Nucl. Mater. [50] Z. Yao, M. Hernandez-Mayoral, M.L. Jenkins, M.A. Kirk, Heavy-ion irradiations
(2021) 543, doi:10.1016/j.jnucmat.2020.152606. of Fe and Fe-Cr model alloys Part 1: damage evolution in thin-foils at lower
[29] C.D. Hardie, S.G. Roberts, A.J. Bushby, Understanding the effects of ion irradi- doses, 2008. https://doi.org/10.1080/14786430802380469.
ation using nanoindentation techniques, J. Nucl. Mater. 462 (2015) 391–401, [51] G.M. Pharr, E.G. Herbert, Y. Gao, The indentation size effect: a critical examina-
doi:10.1016/j.jnucmat.2014.11.066. tion of experimental observations and mechanistic interpretations, Annu. Rev.
[30] W.D. Nix, H. Gao, Indentation size effects in crystalline materials: a law for Mater. Res. 40 (2010) 271–292, doi:10.1146/annurev- matsci- 070909- 104456.
strain gradient plasticity, J. Mech. Phys. Solids. 46 (1998) 411–425, doi:10.1016/ [52] A. Kareer, A. Prasitthipayong, D. Krumwiede, D.M. Collins, P. Hosemann,
S0 022-5096(97)0 0 086-0. S.G. Roberts, An analytical method to extract irradiation hardening from
[31] Z. Zong, J. Lou, O.O. Adewoye, A.A. Elmustafa, F. Hammad, W.O. Soboyejo, In- nanoindentation hardness-depth curves, J. Nucl. Mater. 498 (2018) 274–281,
dentation size effects in the nano- and micro-hardness of fcc single crystal doi:10.1016/j.jnucmat.2017.10.049.
metals, Mater. Sci. Eng. A. 434 (2006) 178–187, doi:10.1016/j.msea.2006.06.137. [53] G.M. Pharr, J.H. Strader, W.C. Oliver, Critical issues in making small-depth me-
[32] M. Mata, O. Casals, J. Alcalá, The plastic zone size in indentation experiments: chanical proeprty measurements by nanoindentation with contiuous stiffness
the analogy with the expansion of a spherical cavity, Int. J. Solids Struct. 43 measurements, J. Mater. Res. 24 (2009) 653–666, doi:10.1557/jmr.2009.0096.
(2006) 5994–6013, doi:10.1016/j.ijsolstr.20 05.07.0 02. [54] S.A. Maloy, T.A. Saleh, O. Anderoglu, T.J. Romero, G.R. Odette, T. Yamamoto,
[33] P. Dou, A. Kimura, T. Okuda, M. Inoue, S. Ukai, S. Ohnuki, T. Fujisawa, F. Abe, S. Li, J.I. Cole, R. Fielding, Characterization and comparative analysis of the
Polymorphic and coherency transition of Y-Al complex oxide particles with ex- tensile properties of five tempered martensitic steels and an oxide dispersion
trusion temperature in an Al-alloyed high-Cr oxide dispersion strengthened strengthtned ferritic alloy irradiated at ≈ 295°C to ≈ 6.5 dpa, J. Nucl. Mater.
ferritic steel, Acta Mater 59 (2011) 992–1002, doi:10.1016/j.actamat.2010.10. 468 (2016) 232–239, doi:10.1016/j.jnucmat.2015.07.039.
026. [55] H.S. Cho, R. Kasada, A. Kimura, Effects of neutron irradiation on the ten-
[34] P. Dou, W. Sang, A. Kimura, Morphology, crystal and metal/oxide inter- sile properties of high-Cr oxide disperison strengthened ferritic steels, J. Nucl.
face structures of nanoparticles in Fe–15Cr–2W–0.5Ti–7Al–0.4Zr–0.5Y2 O3 ODS Mater. 367–370 (2007) 239–243, doi:10.1016/j.jnucmat.2007.03.140.
steel, J. Nucl. Mater. 523 (2019) 231–247, doi:10.1016/j.jnucmat.2019.05.055. [56] D.L. Krumwiede, T. Yamamoto, T.A. Saleh, S.A. Maloy, G.R. Odette, P. Hosemann,
[35] A. Kohyama, Y. Katoh, M. Ando, K. Jimbo, New Multiple Beams-Material Inter- Direct compariison of nanoindentation and tensile test results on reactor ir-
action Research Facility for radiation damage studies in fusion materials, Fu- radiated materials, J. Nucl. Mater. 504 (2018) 135–143, doi:10.1016/j.jnucmat.
sion Eng. Des. 51–52 (20 0 0) 789–795, doi:10.1016/S0920-3796(0 0)0 0181-2. 2018.03.021.
[36] http://www.srim.org/. [57] T.S. Byun, Dose dependence of true stress parameters in irradiated bcc, fcc, and
[37] R.E. Stoller, M.B. Toloczko, G.S. Was, A.G. Certain, S. Dwaraknath, F.A. Garner, hcp metals, J. Nucl. Mater. 361 (2007) 239–247, doi:10.1016/j.jnucmat.2006.12.
Nuclear Instruments and methods in physics research B On the use of SRIM 014.
for computing radiation damage exposure, Nucl. Inst. Methods Phys. Res. B. [58] T.S. Byun, K. Farrell, Plastic instability in polycrystalline metals after low tem-
310 (2013) 75–80, doi:10.1016/j.nimb.2013.05.008. perature irradiation, Acta Mater. 52 (2004) 1597–1608, doi:10.1016/j.actamat.
[38] T. Mails, S.C. Cheng, R.F. Egerton, N. Labor, EELS log-ratio technique for 2003.12.023.
specimen-thickness measurment in the TEM, J. Electron. Microsc. Tech. 8 [59] K. Durst, B. Backes, M. Göken, Indentation size effect in metallic materials:
(1988) 193–200, doi:10.1002/jemt.1060080206. correcting for the size of the plastic zone, Scr. Mater. 52 (2005) 1093–1097,
[39] T. Miyazawa, T. Nagasaka, R. Kasada, Y. Hishinuma, T. Muroga, H. Watanabe, doi:10.1016/j.scriptamat.20 05.02.0 09.
T. Yamamoto, S. Nogami, M. Hatakeyama, Evaluation of irradiation harden- [60] M.A. Mattucci, I. Cherubin, P. Changizian, T. Skippon, M.R. Daymond, Inden-
ing of ion-irradiated V-4Cr-4Ti and V-4Cr-4Ti-0.15Y alloys by nanoindentation tation size effect, geometrically necessary dislocations and pile-up effects in
techniques, J. Nucl. Mater. 455 (2014) 440–444, doi:10.1016/j.jnucmat.2014.07. hardness testing of irradiated nickel, Acta Mater 207 (2021) 116702, doi:10.
059. 1016/j.actamat.2021.116702.
[40] A. Bhaduri, Mechanical proeprties and working of metals and alloys, Springer [61] M. Hiratani, V.V. Bulatov, Solid-solution hardening by point-like obsta-
Nature Singapore Pte Ltd, 2018, doi:10.1007/978- 981- 10- 7209- 3. cles of different kinds, Philos. Mag. Lett. 84 (2004) 461–470, doi:10.1080/
[41] J.E. Campbell, R.P. Thompson, J. Dean, T.W. Clyne, Experimental and computa- 0950 0830410 0 01726969.
tional issues for automated extraction of plasticity parameters from spherical [62] M.J. Swenson, C.K. Dolph, J.P. Wharry, The effects of oxide evolution on me-
indentation, Mech. Mater. 124 (2018) 118–131, doi:10.1016/j.mechmat.2018.06. chanical properties in proton- and neutron-irradiated Fe-9%Cr ODS steel, J.
004. Nucl. Mater. 479 (2016) 426–435, doi:10.1016/j.jnucmat.2016.07.022.
[42] C.L. Fu, M. Krčmar, G.S. Painter, X.Q. Chen, Vacancy mechanism of high oxygen [63] Q. Lu, W. Xu, S. Van Der Zwaag, Designing new corrosion resistant ferritic heat
solubility and nucleation of stable oxygen-enriched clusters in Fe, Phys. Rev. resistant steel based on optimal solid solution strengthening and minimisa-
Lett. 99 (2007) 1–4, doi:10.1103/PhysRevLett.99.225502. tion of undesirable microstructural components, Comput. Mater. Sci. 84 (2014)
[43] S.L. Shang, H.Z. Fang, J. Wang, C.P. Guo, Y. Wang, P.D. Jablonski, Y. Du, Z.K. Liu, 198–205, doi:10.1016/j.commatsci.2013.12.009.
Vacancy mechanism of oxygen diffusivity in bcc Fe: a first-principles study, [64] J. Gao, P. Song, Y.J. Huang, K. Yabuuchi, A. Kimura, K. Sakamoto, S. Yamashita,
Corros. Sci. 83 (2014) 94–102, doi:10.1016/j.corsci.2014.02.009. Effects of neutron irradiation on 12Cr–6Al-ODS steel with electron-beam weld
[44] G.S. Was, Fundementals of Radiation Materials Science, Springer, Berlin Heidel- line, J. Nucl. Mater. 524 (2019) 1–8, doi:10.1016/j.jnucmat.2019.06.028.
beg, New York, 2007. [65] J. Zhang, W.Z. Han, Oxygen solutes induced anomalous hardening, toughening
[45] I. Monnet, T. Van Den Berghe, P. Dubuisson, A comparison between different and embrittlement in body-centered cubic vanadium, Acta Mater 196 (2020)
oxide dispersion strengthened ferritic steel ongoing in situ oxide dissolution in 122–132, doi:10.1016/j.actamat.2020.06.023.
high voltage electron microscope, J. Nucl. Mater. 424 (2012) 204–209, doi:10. [66] J. Brodrick, D.J. Hepburn, G.J. Ackland, Mechanism for radiation damage re-
1016/j.jnucmat.2012.03.011. sistance in yttrium oxide dispersion strengthened steels, J. Nucl. Mater. 445
[46] M.L. Lescoat, J. Ribis, Y. Chen, E.A. Marquis, E. Bordas, P. Trocellier, Y. Ser- (2014) 291–297, doi:10.1016/j.jnucmat.2013.10.045.
ruys, A. Gentils, O. Kaïtasov, Y. De Carlan, A. Legris, Radiation-induced Ostwald [67] W.Z. Han, M.J. Demkowicz, E.G. Fu, Y.Q. Wang, A. Misra, Effect of grain bound-
ripening in oxide dispersion strengthened ferritic steels irradiated at high ion ary character on sink efficiency, Acta Mater 60 (2012) 6341–6351, doi:10.1016/
dose, Acta Mater 78 (2014) 328–340, doi:10.1016/j.actamat.2014.06.060. j.actamat.2012.08.009.
[47] A. Certain, S. Kuchibhatla, V. Shutthanandan, D.T. Hoelzer, T.R. Allen, Radia- [68] H. Gleiter, Grain boundaries as point defect sources or sinks—Diffusional creep,
tion stability of nanoclusters in nano-structured oxide dispersion strengthened Acta Metall 27 (1979) 187–192, doi:10.1016/0 0 01-6160(79)90 095-6.
(ODS) steels, J. Nucl. Mater. 434 (2013) 311–321, doi:10.1016/j.jnucmat.2012.11. [69] X.M. Bai, A.F. Voter, R.G. Hoagland, M. Nastasi, B.P. Uberuaga, Efficient anneal-
021. ing of radiation damage near grain boundaries via interstitial emission, Science
[48] A.J. London, B.K. Panigrahi, C.C. Tang, C. Murray, C.R.M. Grovenor, Glancing an- 327 (2010) 1631–1634, doi:10.1126/science.1183723.
gle XRD analysis of particle stability under self-ion irradiation in oxide disper- [70] A.H. King, D.A. Smith, On the mechanisms of point-defect absorption by
sion strengthened alloys, Scr. Mater. 110 (2016) 24–27, doi:10.1016/j.scriptamat. grain and twin boundaries, Philos. Mag. A 42 (1980) 495–512, doi:10.1080/
2015.07.037. 01418618008239372.
[49] R. Schäublin, B. Décamps, A. Prokhodtseva, J.F. Löffler, On the origin of the pri-
mary ½ a0 <111>and a0 <100>loops in irradiated Fe(Cr) alloys, Acta Mater.
133 (2017) 427–439, doi:10.1016/j.actamat.2017.02.041.

12

You might also like