Nuclear Materials and Energy: Hao Yu, Sosuke Kondo, Ryuta Kasada, Naoko Oono, Shigenari Hayashi, Shigeharu Ukai

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Nuclear Materials and Energy 25 (2020) 100798

Contents lists available at ScienceDirect

Nuclear Materials and Energy


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

Development of nano-oxide particles dispersed alumina scale formed on


Zr-added FeCrAl ODS ferritic alloys
Hao Yu a, *, Sosuke Kondo a, Ryuta Kasada a, Naoko Oono b, Shigenari Hayashi b, Shigeharu Ukai b
a
Institute of Materials Research, Tohoku University, Sendai 980-8577, Japan
b
Graduate School of Engineering, Hokkaido University, Sapporo 060-8628, Japan

A R T I C L E I N F O A B S T R A C T

Keywords: In order to develop a robust alumina scale dispersed with nano-oxide particles for FeCrAl oxide dispersion
FeCrAl ferrite strengthened (ODS) ferritic alloys, the oxidation behavior of FeCrAl ODS ferritic alloys with zirconium and
ODS excessive oxygen (Ex. O) additions was investigated at 900 ◦ C in air. Zirconia incorporation into alumina scale
Oxidation
was not observed in the Ex. O-added FeCrAl ODS alloys at 900 ◦ C, instead of this, a distribution with dense Y-Zr
Alumina
oxide particles inside the alumina scale was found, which is expected to enhance the stability of the alumina scale
during service at elevated temperatures in nuclear applications. Preexisting large Zr-enriched precipitates formed
during the alloy fabrication process are incorporated into the alumina scales during oxidation, which would
increase the oxidation rate via providing a short-circuit paths for oxygen inward diffusion by themselves and
associated porosity. In addition to a traditional spherical shape, a unique tadpole-like shape of the fine Y-Zr oxide
particles in alumina was also observed, which is considered to be related to those incorporated coarse Zr-Y oxide
inclusions in alumina scale.

1. Introduction limited, it was claimed that the oxygen can be forcibly dissolved into
matrix by the unique fabrication process of ODS alloys via mechanical
Alumina-forming FeCrAl ODS ferritic alloys have been considered alloying (MA), which makes it possible to adjust the oxygen content in
key structural materials for fission and fusion reactors of the next gen­ FeCrAl ODS alloys [8]. The addition of Fe2O3 powder during MA process
eration due to their excellent performance against high-temperature is one of most used methods to introduce the oxygen into the matrix of
creep, irradiation and oxidation [1–5]. The tolerance against creep FeCrAl ODS alloys [9]. Besides, it is inevitable that the oxygen
deformation and neutron irradiation can be further improved by the contamination may happen during the whole fabrication process of ODS
addition of small amounts of Zr, which is attributed to the formation of alloys, such as MA and hot-extrusion, which would also introduce a
fine Y-Zr oxide particles with a high number density [6]. However, at certain amount of oxygen into the alloys. In order to accurately quantify
high temperatures, the Zr also could play a detrimental role via forming the introduced oxygen content in the matrix of FeCrAl ODS alloys, the
a ZrO2 inside the alumina scale. The formed ZrO2 will act as a rapid concept of excessive oxygen (Ex. O) is put forward, which represents the
oxygen inward diffusion pathway, thus weakening the oxidation resis­ oxygen in the FeCrAl ODS alloys caused by the Fe2O3 powder addition
tance of the alumina scale [7]. Such an unbeneficial effect of Zr is and the oxygen contamination. The Ex. O will consume the Zr in the
associated with the Zr distribution type in matrix. Since the solubility of alloy by forming Y-Zr oxides during the alloy fabrication process, which
Zr in the bcc Fe is negligibly small, the Zr atoms may segregate at grain may prevent the formation of ZrO2 inside external alumina scale after­
boundary or particle/matrix interface. If Zr was preferentially consumed ward. The authors investigated the oxidation behavior of the Zr-added
by an internal oxidation of the matrix during the fabrication process, the FeCrAl ferritic alloys at ultra-high temperatures from 1200 ◦ C to
activity of Zr to form ZrO2 precipitation in the alumina scale during the 1500 ◦ C and demonstrated that the Ex. O addition higher than 0.2 wt%
subsequent environmental oxidation exposure will be significantly can effectively suppress the ZrO2 formation inside the alumina scale
reduced. The preferential oxidation is directly related on the oxygen [10].
content in matrix. Even the oxygen solubility in ferritic Fe is very In addition, an interesting point is that the alumina scale formed on

* Corresponding author.
E-mail address: qqyu1111@163.com (H. Yu).

https://doi.org/10.1016/j.nme.2020.100798
Received 1 February 2020; Received in revised form 21 August 2020; Accepted 11 September 2020
Available online 21 September 2020
2352-1791/© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
H. Yu et al. Nuclear Materials and Energy 25 (2020) 100798

the FeCrAl ODS ferrites is possible to be dispersed with Y-Zr oxide with Cu-Kα). In order to analyze cross sections of the oxidized samples,
particles based on other studies on alumina-forming ODS superalloys the oxidized samples were mounted in resins and then kept in grinding
[11], which is expected to mitigate irradiation defects formed in the until using 1 μm-diamond paste. Backscattered electron (BSE) pattern of
alumina scale and keep the alumina scale stable during nuclear irradi­ an electron probe micro-analyzer (EPMA, JEOL JXA-8530F) was utilized
ation [12,13]. Conventional ODS steels are generally pre-oxidized to to observe the polished cross sections. Corresponding elemental distri­
form a passive chromium oxide film and a small amount of iron oxide to bution was also detected with a wavelength dispersive X-ray spectros­
protect the steels during practical application as a tritium breeder copy (WDS) elemental mapping. Grain size of substrates in the cross
blanket material in fusion reactor [14,15]. In contrast, the alumina- sections were observed by electron back-scattered diffraction (EBSD)
forming FeCrAl ferritic ODS alloys is more competitive to act as the jointed in the EPMA. In order to investigate the oxide particles distri­
materials and it would be very exciting if the formed single alumina bution in the formed oxide scales, transmission electron microscopy
scale could be dispersed with dense nano-oxide particles during the pre- (TEM, JEOL ARM-200F) observation was carried out to further inves­
oxidation. However, B.A. Pint [16] reported that the dynamic- tigate the oxide scale by using scanning transmission electron micro­
segregation of reactive-elements (RE) such as Y, Zr and Ti would scopy (STEM) imaging and X-ray energy dispersive spectroscopy (EDS)
happen with a RE outward diffusion from the metal substrate to the gas elemental mapping. Focused ion beam (FIB, HITACHI, FB2200) was
interface of alumina scale during high temperature oxidation. The used to prepare the TEM samples. Before FIB milling, carbon and
dynamic-segregation is sensitively dependent on the temperatures, i.e., tungsten coatings were successively overlaid on the surface of oxidized
high temperatures above 1000 ◦ C accelerate the RE outward diffusion samples to maintain the integrity of the alumina scale.
and induce RE-rich particles ripening and coarsening near the alumina
scale surface. Thus, a relatively low temperature below 900 ◦ C is more 3. Results
suitable for developing the Y-Zr-O particles with a small size and high
number density inside the alumina scale for the Zr-added FeCrAl ODS 3.1. Weight gain by isothermal oxidation
alloys. The fine Y-Zr oxide particles with high number density inside
alumina scale are expected to improve irradiation resistance of the Fig. 1 shows the weight gain per unit surface area as a function of
alumina scale by capturing defects and reducing irradiation swelling exposure time for all the FeCrAl ODS ferritic alloys at 900 ◦ C. It can be
[12]. confirmed that all specimens showed an approximate mass gain during
In the present work, aimed at developing a superior alumina scale in the initial stage of exposure (− 81 h). However, with increasing exposure
which the ZrO2 formation is eliminated and dense Y-Zr oxide particles time, 0Zr specimens showed the smallest increase in weight in the later
are dispersed, relatively high Ex. O contents of 0.2 wt% and 0.3 wt% stage (81–225 h), implying that Zr addition accelerated the oxidation
were added to suppress the ZrO2 formation inside the alumina scale rate at longer exposure times even though with high Ex. O additions.
based on previous study [10]. Besides, 900 ◦ C in air was chosen to Besides, it seems that two Ex. O added samples exhibited similar
conduct the oxidation since the relatively lower temperature is consid­ oxidation kinetics at 900 ◦ C, in spite of different Ex. O contents with 0.2
ered to be beneficial for the fine Y-Zr oxide particles formation in wt% and 0.3 wt%. Assuming that all oxidation rates follow classical
alumina scale. In order to characterize the effect of Zr addition on the parabolic scale-growth kinetics [17], the oxidation rate constant kp
oxidation properties, the FeCrAl ODS ferritic alloys without Zr addition could be derived as follows,
were prepared to act as comparison tests.
ΔW2
kp = (1)
2. Experimental t
in which ΔW is the specific weight change of oxidized samples and t
Table 1 lists the actual chemical compositions of all the used samples is the oxidation time during exposure. Fig. 2 is the re-arranged oxidation
in this study, in which corresponding abbreviations are used to name the kinetics figure derived from mass gains as a function of square root of
samples. All alloys were subjected to MA process and then consolidated time. In order to figure out the kp value, several straight lines were tried
with a hot-extrusion at 1150 ◦ C. The introduced Ex. O content into the to make to link all the dots in the Fig. 2. Interestingly, in contrast to a
Zr-added ODS alloys was controlled via adjusting the Fe2O3 powder straight line running through the results of 0Zr, the oxidation process of
amount during MA process. Specimens at as-extruded condition were both the two Zr-added specimens seems can be separated into two parts.
cut into a dimension of 10 × 6 × 2 mm for the oxidation test, and all the i.e., initial stage (− 81 h) and later stage (81–225 h). Table 2 lists the kp
surfaces were wet ground using SiC paper up to 4000-grit, and then values of each specimens, which are calculated from the squares of the
polished with 3 μm-diamond paste. After confirming that no residual corresponding gradients in Fig. 2.
scratches remain on the surfaces by an optical microscopy, the samples
were ultrasonically cleaned in alcohol, and dried in hot air. Isothermal 3.2. Microstructure of the oxide scales
oxidation test at 900 ◦ C was carried out inside a muffle furnace in air and
then subsequently air cooled to room temperature. In order to determine Fig. 3 shows the XRD patterns for all the FeCrAl ferritic ODS alloys
the oxidation kinetics, the oxidation weight gains in different exposure oxidized at 900 ◦ C with various exposure times of 4 h, 9 h, 25 h, 49 h,
durations (1, 4, 9, 16, 25, 49, 64, 81, 100, 196, 225 h) were measured at 100 h and 225 h, in which (a), (b) and (c) represents the results of 0Zr,
room temperature via a Sartorius precision electronic balance with a Zr0.2Ex. O and Zr0.3Ex. O, respectively. In accordance with noticeable
resolution of 1 μg. characteristic peaks of Al2O3, it is confirmed that alumina formation
The crystalline structure of oxidized samples was detected using X- dominates all the oxide scales during overall exposure at 900 ◦ C in 225
ray diffraction (XRD, RIGAKU Ultima IV/SG operated at 40 kV/40 mA h.
In order to investigate the distribution of elements in the oxidized
scales, cross sections of all samples oxidized within 225 h were prepared
Table 1
and analyzed by EPMA. Based on BSE images, a single-layered oxide
Actual compositions of the investigated alloys (in wt.%) and utilized
scale with dark contrast could be found on the surfaces of all specimens,
abbreviations.
as shown in Fig. 4, in which (a), (b) and (c) represents the cross section
No. Fe Cr Al Y2O3 Ti Zr Ex. O
obtained from 0Zr, Zr0.2Ex. O, Zr0.3Ex. O specimens, respectively.
0Zr Bal. 14.2 6.4 0.5 0.5 0 0.1 Although a general trend indicated that the thickness of all the scales
Zr0.2Ex. O Bal. 14.8 6.4 0.5 0.5 0.4 0.2 formed in the three alloys is not homogeneous, it is apparent that the Zr-
Zr0.3Ex. O Bal. 15.1 7.0 0.5 0.5 0.4 0.3
free specimen possesses the thinnest oxide layer, which agrees well with

2
H. Yu et al. Nuclear Materials and Energy 25 (2020) 100798

Fig. 1. Weight gain curves for the FeCrAl ODS ferritic alloys oxidized at 900 ◦ C in air for 225 h.

Fig. 2. The fitted parabolic curves for the FeCrAl ODS ferritic alloys oxidized at 900 ◦ C in air for 225 h.

the mass gain exhibited in Fig. 1. With regard to the Zr-added specimens, oxygen elements at the outer dark layer and the alumina diffraction
even different Ex. O contents with 0.2 wt% and 0.3 wt% were added, it is peaks obtained from the XRD results in Fig. 3, the oxide scales formed on
difficult to obtain a significant scale thickness difference from the BSE all the tested FeCrAl ODS ferritic alloys at 900 ◦ C in air were confirmed
images in the Fig. 4 (b) and (c). to be single-layered alumina scales. Besides, spot-like enrichments of Zr
Fig. 5 shows corresponding EPMA elemental maps of the cross sec­ and Ti inside the substrate also can be observed, which demonstrates the
tions, in which (a), (b) and (c) belong to the 0Zr, Zr0.2Ex. O and Zr0.3Ex. existence of Zr-rich precipitates in the matrix of Zr-added alloys.
O, respectively. According to the enrichment of only aluminum and The grain of substrates in the cross sections was observed with EBSD.

3
H. Yu et al. Nuclear Materials and Energy 25 (2020) 100798

Table 2
Parabolic rate constants of the FeCrAl ODS ferritic alloys oxidized at 900 ◦ C in
air for 225 h.
Superalloys Parabolic rate constant, kp (g2⋅cm¡4⋅s¡1)
14
0Zr 2.39 × 10−
14
Zr0.2Ex. O 2.89 × 10− (~81 h)
14
6.42 × 10− (81 ~ 225 h)
14
Zr0.3Ex. O 2.86 × 10− (~81 h)
14
6.19 × 10− (81 ~ 225 h)

Fig. 4. The cross-section morphologies of oxidized samples at 900 ◦ C in 225 h,


(a), (b) and (c) represents the 0Zr, Zr0.2Ex. O and Zr0.3Ex. O samples,
respectively.

According to the image quality (IQ) maps in Fig. 6, the grain size ranges
in 0Zr, Zr0.2Ex. O and Zr0.3Ex. O alloys were measured to be 0.53–7.89
μm, 0.27–6.54 μm and 0.28–4.64 μm, respectively. It is noticeable that
the average grain size in the 0Zr specimen is larger than that in the two
Fig. 3. XRD results obtained from the 0Zr (a), Zr0.2Ex. O (b), Zr0.3Ex. O (c) Zr-added samples.
alloys oxidized at 900 ◦ C in 4, 9, 25, 49, 100 and 225 h, respectively, in which In order to obtain a more detailed microstructure information in the
the alloys prior to oxidation also have been tested.
formed single-layered alumina scale, TEM observation on the Zr 0.2Ex.

4
H. Yu et al. Nuclear Materials and Energy 25 (2020) 100798

Fig. 5. EPMA elemental maps obtained from 0Zr (a), Zr0.2Ex. O (b) and Zr0.3Ex. O samples oxidized at 900 ◦ C for 225 h.

O specimen after exposure at 900 ◦ C for 225 h was conducted. Based on Fig. 7 (a). Remarkably, in the thickest alumina layer, in addition to small
the low-angle annular dark-field (LAADF) STEM images in Fig. 7, it is grains of the substrate with a mean size of 0.61 μm, a coarse precipitate
apparent that the thickness of the alumina scale is heterogeneous, which with an approximate diameter of 0.43 μm could be observed inside the
is consistent with the BSE image observed in the Fig. 4 (b). Underneath a alumina scale. In the magnified LAADF-STEM image of Fig. 7 (b), pores
thin alumina layer, some large grains with an average size of 1.01 μm around the coarse precipitate were also observed, as marked with yellow
can be observed in substrate, as marked with orange broken lines in arrows. Besides, a large number of fine particles could be observed

5
H. Yu et al. Nuclear Materials and Energy 25 (2020) 100798

Fig. 6. EBSD IQ maps obtained from substrates of 0Zr (a), Zr0.2Ex. O (b), Zr0.3Ex. O (c) alloys.

Fig. 7. TEM LAADF images obtained from the Zr0.2Ex. O specimen oxidized at 900 ◦ C for 225 h.

inside the alumina scale. An interesting point should be noted is that the the consolidation of FeCrAl ODS alloys [10]. Thus, during the oxidation
shape of the particles near the coarse precipitate and / or the pores, i.e., process at 900 ◦ C, it is reasonable to suppose that the reservoir of Zr in
in contrast to a common spherical shape in other areas, tadpole-like the alloy is quite limited since the Zr was considerably consumed by the
particles were found in there, as marked with red arrows in Fig. 7 (b), pre-formed fine Y-Zr compound oxide particles, as demonstrated by
which seem to point towards the direction of the coarse precipitate and / previous TEM works [9].
or the pores with their heads. EDS analysis on the coarse precipitate and
fine particles were conducted. As shown in Fig. 8 (a), the coarse pre­
cipitate possesses a relatively high Zr concentration in addition to being 4.2. Oxidation kinetics
enriched in Y and O, which is named coarse Zr-Y oxide inclusions
hereafter. In respect of the fine particles, based on corresponding EDS In terms of oxidation rates, both Ex. O added samples are slightly
maps in Fig. 8 (b), an enrichment of Y and Zr elements can be observed faster than that of the 0Zr alloys in the initial stage (− 81 h), which may
in it, which are considered to be Y-Zr oxide particles in accordance with be caused by a smaller average grain size in the substrate of Zr-added
the references [18,19]. In addition, some slightly larger Y-Zr oxide samples, as shown in Fig. 6, because a fast cation and anion ions diffu­
particles marked with green arrows also can be observed in the Fig. 8 sion via grain boundary could be expected in the small substrate grain.
(b). Dense distribution of fine Y-Zr oxide particles in the Zr-added FeCrAl
ODS alloys is responsible for the smaller average grain size, which
4. Discussion provides a strong pinning effect to inhibit the substrate grain growth
[21,22]. Besides, the grain size in the substrates of all FeCrAl ODS alloys
4.1. ZrO2 elimination in alumina scale (Fig. 6) is inhomogeneous, which derived from the heterogeneous dis­
tribution of oxide particles, i.e., dense oxide particles were dispersed in
The present results confirm that all the FeCrAl ODS ferritic alloys finer grains, while sparse oxide particles were dispersed in larger grains
with and without Zr addition are capable of forming a stable alumina [23]. The distribution of oxide particles is closely related to the growth
scale at 900 ◦ C. Ex. O additions with 0.2 wt% and 0.3 wt% make the Zr- of oxide scales. Stringer et al. [24] claimed that the dispersed Y2O3
added alloys possess a comparable oxidation resistance with Zr-free particles in the matrix of the post-oxidized materials could act as het­
samples. It is noticeable that no ZrO2 precipitates were observed in erogeneous nucleation sites for Cr2O3 grains on the materials surface
the alumina scale formed on Zr-added samples with 0.2 wt% and 0.3 wt and allow a rapid formation of the Cr2O3 scale with a finer Cr2O3 grain
% Ex. O additions. The incorporation of ZrO2 in alumina is directly size. The dispersed oxide particles in the three FeCrAl ODS specimens
determined by the Zr reservoir in alloys [20]. It was claimed that the Zr are also deemed to have the similar effect on the formation of alumina
activity in the alloy matrix decreased significantly through adding Ex. O scale. Alumina scales grow inward since its growth is governed by the
to promote the formation of stable Y-Zr compound oxide particles during inward diffusion of oxygen [25]. The dense oxide particles in small
substrate grains would provide more nucleation sites for Al2O3 grains for

6
H. Yu et al. Nuclear Materials and Energy 25 (2020) 100798

Fig. 8. STEM elemental maps obtained from Zr0.2Ex. O samples oxidized at 900 ◦ C for 225 h.

the inward growth of alumina scales. As a result, alumina scales with Even a rapid growth rate of alumina could be expected on the substrate
smaller grain sizes would form near the small substrate grains and then with a small grain size, it is obvious that the thickness grows drastically
exhibit faster growth rates due to an accelerated oxygen grain boundary once the coarse Zr-Y oxide inclusions appeared. The formation of coarse
inward diffusion [26,27], which is corresponding to the heterogeneous Zr-Y oxide inclusions is related to the large Zr-enriched precipitates
thickness of the formed alumina scales in the Fig. 7 (a). originally existed in the ODS alloys, which were formed during the hot-
However, with longer exposure time, the oxidation of Zr-added extrusion consolidation and were incorporated into the alumina scale
samples increased significantly and was apparently faster than the 0Zr during the oxide scale growth. In order to demonstrate the presence of
samples in the later stage (81–225 h), as shown in Fig. 2. The accelerated the Zr-enriched precipitates in the matrix of hot-extruded alloys, EPMA
oxidation rates of Zr0.2Ex. O and Zr0.3Ex. O are attributed to the exis­ observation was conducted for the samples in prior to oxidation testing.
tence of the coarse Zr-Y oxide inclusions in the alumina scale. As shown In contrast to the spot-like enrichment of Ti in the 0Zr specimen in Fig. 9
in the LAADF-STEM images in Fig. 7, a most thick alumina scale can be (a), a significant Zr/Zr-Y enrichment was able to be observed in the
observed in the same vertical line with the coarse Zr-Y oxide inclusions. matrix of Zr 0.2Ex. O specimen in Fig. 9 (b). The oxide particles

7
H. Yu et al. Nuclear Materials and Energy 25 (2020) 100798

Fig. 9. EPMA elemental maps obtained from 0Zr (a) and Zr0.2Ex. O (b) samples in prior to the oxidation testing.

dispersed in ODS alloys are formed by decomposition of Y2O3 during MA rate once it was incorporated in alumina scale. Besides, STEM image of
process and its re-precipitation with the added solutes such as Zr/Ti Fig. 7 (b) proved that pores were formed near the large Zr-Y inclusions.
during subsequent consolidation [19]. A lower formation energy of Zr/ Undoubtedly, the porosity could also act as significant short-circuit
Zr-Y oxides than that of Ti/Ti-Y oxides is responsible for the occurrence paths for oxygen transport in the form of molecular gas diffusion. D.
of the Zr-enriched precipitates in the Zr-added alloys during the Naumenko et al. [20] also observed a porosity which prefers to form
consolidation of hot-extrusion [10,28]. Currently, even the EPMA did near zirconia inclusions in alumina scales of Zr-added FeCrAlY alloys,
not carry out for the Zr0.3Ex. O sample, it is reasonable to assume that and proposed it is attributed to the phase transformations of the zirconia
the large Zr-enriched precipitates also appeared after hot-extrusion incorporation caused. In this study, even zirconia precipitates are not
process in the Zr0.3Ex. O sample. As demonstrated in the EPMA maps found in alumina scale, it is supposed that the coarse Zr-Y oxide in­
of oxidized samples in Fig. 5 (b) and (c), obvious Zr-enriched pre­ clusions can also act as a similar effect for the formation of the porosity,
cipitates with a considerably large size could be found in both the sub­ since both of them possess a rapid oxygen diffusion coefficient. Besides,
strates of 2.0 wt% and 3.0 wt% Ex. O samples. The large Zr-enriched unlike the alumina with a large columnar grain structure in other areas,
precipitates located near the surface at the as-extruded condition the alumina beneath the coarse Zr-Y oxide inclusions shows a smaller
would be embedded in the external alumina scale and then promote the alumina grain size. As mentioned above, the smaller alumina grains are
growth of the oxide in both the Zr0.2Ex. O and Zr0.3Ex. O samples. due to the fact that the coarse Zr-Y oxides inclusions provide more
A detailed stoichiometric ratio of the coarse Zr-Y oxide inclusions has nucleation sites for the formation of new alumina grains. Therefore, the
not yet determined, but it is not difficult to estimate the oxygen diffusion coarse Zr-Y oxide inclusions, including associated porosity and the
coefficient in the coarse Zr-Y oxide inclusions at 900 ◦ C in air is near 7.0 smaller alumina grain size, are responsible to the increase of oxidation
× 10− 13 m2 s− 1 in terms of the studies about the oxygen diffusion co­ rate in the Zr0.2Ex. O sample. The detrimental effect begins with the
efficient of yttria stabilized zirconia (YSZ) [29]. It is known that the incorporation of the coarse Zr-Y oxide inclusions inside alumina scale
oxygen diffusion coefficient in Al2O3 can be calculated to be 5.5 × 10− 31 during long exposure period, which is corresponding to the increased kp
m2 s− 1 at 900 ◦ C [30], which is much smaller than that in the Zr-Y in­ values in later stage (81–225 h) in the Table 2.
clusions. Thus, the coarse Zr-Y oxide inclusions are able to provide short-
circuit paths for oxygen inward diffusion and then increase the oxidation

8
H. Yu et al. Nuclear Materials and Energy 25 (2020) 100798

4.3. Nano-oxide particles distribution in alumina scale 5. Conclusions

The oxygen partial pressure in the coarse Zr-Y oxide inclusions / In order to develop nano-oxide particles dispersed alumina scale for
pores regions would be increased [20,31], which contributes the dis­ Zr-added FeCrAl ODS ferritic alloys, oxidation behavior of FeCrAl ODS
tribution of Y-Zr oxide particles in alumina scale, especially for the ferritic alloys with zirconium and excessive oxygen additions at 900 ◦ C
formation of the unique tadpole-like Y-Zr-O particles. It is known that in air has been investigated. Corresponding results could be summarized
the alumina has a very limited solubility for most cations, such as Y as follows:
solubility in the alumina is less than 10 ppm even near to the melting
point of alumina. Thus, RE elements would be forced to disperse mainly 1. With the additions of 0.2 wt% and 0.3 wt% Ex. O, ZrO2 precipitates
at alumina grain boundaries or become precipitates [32–35]. In the case inside the alumina scales of the Zr-added FeCrAl ODS ferritic alloys
of Zr-added FeCrAl ODS alloys, fine Y-Zr oxide particles are dispersed were eliminated at 900 ◦ C, which is supposed to be resulted from a
inside matrix prior to oxidation, which would be incorporated along consumed reservoir of Zr by the pre-formed Zr-Y compound oxide
grain boundaries as well as in the interior of alumina grains as the grains particles in alloys via adding Ex. O. However, some of the pre-formed
continue to grow [33]. The Y-Zr oxide particle/alumina interface in the Zr-Y precipitates with large size may deteriorate the stability of
alumina grains is also expected to act as a preferential site for the alumina scale for long-term exposure at elevated temperatures.
segregation of RE elements, similar to the role of grain boundaries. Thus, 2. Oxidation rate of Zr-added samples would be accelerated at long
a coarsening and/or coalescence of those Y-Zr oxide particles inside exposure with incorporation of coarse Zr-Y oxide inclusions into
alumina scale must happened, which is quite consistent with the slightly alumina scale, resulting from the short-circuit diffusion paths for
larger size of the Y-Zr oxide particles in the STEM image of Fig. 8 (b). In oxygen provided by themselves and associated porosity. Originally
the case of Zr-Y inclusions incorporated, a tadpole-like shape of the Y-Zr formed Zr-enriched precipitates with relatively large size in alloys
oxide particles would be formed. B. A. Pint [16] proposed that RE ions during fabrication process are responsible to the formation of the
tend to diffuse from the metal substrate to the gas interface of the scale, coarse Zr-Y oxide inclusions. In order to avoid the formation of large
in which the driving force for this outward diffusion is the oxygen po­ Zr-Y precipitates, which would deteriorate the oxidation resistance
tential gradient across the scale. Therefore, the locally increased oxygen at 900 ◦ C, a relatively small amounts of Zr should be added to the
partial pressure in the large Zr-Y inclusions and associated pores should alloy.
induce the onset of a similar dynamic-segregation. Even segregation of 3. Dense Y-Zr oxide particles dispersion inside alumina scale was ob­
Y/Zr elements on the Y-Zr particle/alumina interface also happened in tained in the oxidation of 0.2 wt% Ex. O added FeCrAl alloys at
this case, it would be endowed with a certain direction towards to the 900 ◦ C. In addition to a spherical shape of the Y-Zr-O particles, a kind
high oxygen potential, namely, towards the location of the coarse Zr-Y of unique tadpole-like shape was found in the alumina scale, which is
oxide inclusions and associated pores. Besides, tails of the tadpole-like related to an accelerated dynamic-segregation of RE elements driven
Y-Zr particles seem longer near the substrate, which is attributed to by the oxygen potential gradient across the coarse Zr-Y oxide in­
relatively abundant Zr/Y reservoirs near substrate. Some Y-Zr oxide clusions and associated pores in alumina scale.
particles seem to be aligned inside alumina-scale, as shown in the STEM
image in Fig. 7 (b), which may be attributed to the same level of oxygen CRediT authorship contribution statement
partial pressure, since they maintain almost the same distance from the
coarse Zr-Y oxide inclusions and associated pores. The above two models Hao Yu: Investigation, Writing - original draft. Sosuke Kondo:
for the Y-Zr particles distribution in the alumina incorporated without Writing - review & editing. Ryuta Kasada: Writing - review & editing.
and with the large Zr-Y inclusions are shown in Fig. 10 (a) and (b). Naoko Oono: Writing - review & editing. Shigenari Hayashi: Writing -
Fig. 11 illustrates schematically the overall progress of alumina scale review & editing. Shigehara Ukai: Conceptualization.
formation in the Zr-0.2 Ex. O specimens oxidized at 900 ◦ C, in which the
relationship between the large Zr-Y inclusions and the distribution of Y- Declaration of Competing Interest
Zr oxide particles inside alumina scale is exhibited.
The authors declare that they have no known competing financial

Fig. 10. Proposed models for the Y-Zr oxide particles distribution in the alumina scale incorporated without (a) and with (b) the coarse Zr-Y oxide inclusions during
oxidation at 900 ◦ C in air.

9
H. Yu et al. Nuclear Materials and Energy 25 (2020) 100798

Fig. 11. Schematic illustration for the alumina formation of Zr0.2Ex. O and Zr0.3Ex. O samples during oxidation at 900 ◦ C in air.

interests or personal relationships that could have appeared to influence [16] B.A. Pint, Experimental observation in support of the dynamic-segregation theory
to explain the reactive-element effect, Oxid. Met. 45 (1996) 1–37.
the work reported in this paper.
[17] W.J. Quadakkers, D. Naumenko, E. Wessel, V. Kochubey, Growth rates of alumina
scales on Fe-Cr-Al alloys, Oxid. Met. 61 (2004) 17–37.
Acknowledgments [18] P. Dou, A. Kimura, R. Kasada, T. Okuda, M. Inoue, S. Ukai, S. Ohnuki, T. Fujisawa,
F. Abe, TEM and HRTEM study of oxide particles in an Al-alloyed high-Cr oxide
dispersion strengthened steel with Zr addition, J. Nucl. Mater. 444 (2014)
This work is supported by Grant-in-Aid for Research Activity start-up 441–453.
(19K21029) and Cooperative Research and Development Center for [19] N.H. Oono, S. Ukai, K. Tominaga, N. Ebisawa, K. Tomura, Precipitation of various
Advanced Materials, IMR (19G0413). oxides in ODS ferritic steels, J. Mater. Sci. 54 (2019) 8786–8799.
[20] D. Naumenko, V. Kochubey, L. Niewolak, A. Dymiati, J. Mayer, L. Singheiser, W.
J. Quadakkers, Modification of alumina scale formation on FeCrAlY alloys by
References minor additions of group IVa elements, J. Mater. Sci. 43 (2008) 4550–4560.
[21] S.J. Wu, J. Li, W.H. Li, S. Liu, Characterization of oxide dispersoids and mechanical
[1] S. Ukai, S. Ohtsuka, Nano-mesoscopic structure control in 9Cr–ODS ferritic steels, properties of 14Cr-ODS FeCrAl alloys, J. Alloy. Compd. 814 (2020), 152282.
Energy Mater. 2 (1) (2007) 26–35. [22] H.J. Xu, Z. Lu, S. Ukai, N. Oono, C.M. Liu, Effects of annealing temperature on
[2] S. Ukai, T. Kaito, S. Ohtsuka, T. Narita, M. Fujiwara, T. Kobayashi, Production and nanoscale particles in oxide dispersion strengthened Fe-15Cr alloy powders with Ti
properties of nano-scale oxide dispersion strengthened (ODS) 9Cr martensitic steel and Zr additions, J. Alloy. Compd. 693 (2017) 177–187.
claddings, ISIJ Int. 43 (2003) 2038–2045. [23] H.T. Zhang, Y.N. Huang, H.P. Ning, C.A. Williams, A.J. London, K. Dawson, Z.
[3] S. Ukai, M. Fujiwara, Perspective of ODS alloys application in nuclear L. Hong, M.J. Gorley, C.R.M. Grovenor, G.J. Tatlock, S.G. Roberts, M.J. Reece, H.
environments, J. Nucl. Mater. 307–311 (2002) 749–757. X. Yan, P.S. Grant, Processing and microstructure characterisation of oxide
[4] S. Ukai, T. Okuda, M. Fujiwara, T. Kobayashi, S. Mizuta, H. Nakashima, dispersion strengthened Fe-14Cr-0.4Ti-0.25Y2O3 ferritic steels fabricated by spark
Characterization of high temperature creep properties in recrystallized 12Cr-ODS plasma sintering, J. Nucl. Mater. 464 (2015) 61–68.
ferritic steel claddings, J. Nucl. Sci. Technol. 39 (2002) 872–879. [24] J. Stringer, B.A. Wilcox, R.I. Jaffee, The high-temperature oxidation of Nickel-20
[5] T. Kaito, T. Narita, S. Ukai, Y. Matsuda, High temperature oxidation behavior of wt.% Chromium alloys containing dispersed oxide phases, Oxid. Met. 5 (1972)
ODS steels, J. Nucl. Mater. 329–333 (2004) 1388–1392. 11–47.
[6] J. Ren, L.M. Yu, Y.C. Liu, C.X. Liu, H.J. Li, J.F. Wu, Effects of Zr addition on the [25] K. Hellstrom, N. Israelsson, M. Halvarsson, S. Canovic, J.E. Svensson, L.
strengthening mechanisms of Al-alloyed high-Cr ODS steels, Materials 11 (2018) G. Johansson, The oxide scales formed on a dispersion-strengthened powder
118. metallurgical FeCrAl alloy at 900 oC in O2 and in O2 + H2O, Oxid. Met. 84 (2015)
[7] D. Naumenko, B.A. Pint, W.J. Quadakkers, Current thoughts on reactive element 1–19.
effects in alumina-forming system: in memory of John Stringer, Oxid. Met. 86 [26] W.J. Quadakkers, H. Holzbrecher, K.G. Briefs, H. Beske, Difference in growth
(2016) 1–43. mechanisms of oxide scales formed on ODS and conventional wrought alloys, Oxid.
[8] S. Ukai, N. Oono, S. Ohtsuka, T. Kaito, K. Sakamoto, T. Torimaru, A. Kimura, S. Met. 32 (1989) 67–88.
Hayashi, Development of FeCrAl-ODS steels for ATF cladding, in: Proc. Light Water [27] W.J. Quadakkers, Oxidation of ODS alloys, J. Phys IV. 3 (1993) 177–186.
Reactor (LWR) Fuel Performance Meeting, TOP FUEL, 2016. [28] P. Dou, S.M. Jiang, L.L. Qiu, A. Kimura, Effects of contents of Al, Zr, and Ti on
[9] S. Ukai, N. Oono, K. Sakamoto, T. Torimaru, T. Kaito, A. Kimura, S. Hayashi, oxide particles in Fe-15Cr-2W-0.35Y2O3 ODS steels, J. Nucl. Mater. 531 (2020),
Development of FeCrAl-ODS steel claddings for accident tolerant fuel of light water 152025.
reactors, Proceedings of 2017 international congress on advances in nuclear power [29] M. Kilo, C. Argirusis, G. Borchardt, R.A. Jackson, Oxygen diffusion in yttria
plants (ICAPP2017), 49 (2017), 2573. stabilized zirconia-experimental results and molecular dynamics calculations,
[10] T. Maeda, S. Ukai, S. Hayashi, N. Oono, Y. Shizukawa, K. Sakamoto, Effects of Phys. Chem. Chem. Phys. 5 (2003) 2219–2224.
zirconium and oxygen on the oxidation of FeCrAl-ODS alloys under air and steam [30] Y. Oishi, K. Ando, N. Suga, W.D. Kingery, Effect of surface condition on oxygen
conditions up to 1500 oC, J. Nucl. Mater. 516 (2019) 317–326. self-diffusion coefficients for single crystal Al2O3, J. Am. Ceram. Soc. 66 (1983)
[11] H. Yu, S. Ukai, S. Hayashi, N.H. Oono, Effect of Cr and Y2O3 on the oxidation 130–131.
behavior of Co-based oxide dispersion strengthened superalloys at 900 oC, Corros. [31] N. Mommer, T. Lee, J.A. Gardner, Stability of monoclinic and tetragonal zirconia at
Sci. 127 (2017) 147–156. low oxygen partial pressure, J. Mater. Res. 15 (2) (2000) 377–381.
[12] N. Oono, S. Ukai, S. Kondo, O. Hashitomi, A. Kimura, Irradiation effects in oxide [32] J.D. Cawley, J.W. Halloran, Dopant distribution in nominally yttrium-doped
dispersion strengthened (ODS) Ni-base alloys for Gen. IV nuclear reactors, J. Nucl. sapphire, J. Am. Ceram. Soc. 69 (1986) C195.
Mater. 465 (2015) 835–839. [33] C.M. Wang, G.S. Cargill III, H.M. Chan, M.P. Harmer, Structure of Y and Zr
[13] A. Konno, N.H. Oono, S. Ukai, S. Kondo, O. Hashitomi, A. Kimura, He-Cavity segregated grain boundaries in alumina, Interface Sci. 8 (2000) 243–255.
accumulation at oxide particle-matrix interface in Ni-base ODS superalloy, Mater. [34] M.A. Gulgun, R. Voytovych, I. Maclaren, M. Ruhle, Cation segregation in an oxide
Trans. 58 (12) (2017) 1636–1639. ceramic with low solubility: yttrium doped α-alumina, Interface Sci. 10 (2002)
[14] M. Wang, M.Q. Xiang, Y.C. Zhang, Z.J. Zhou, Study on the chemical compatibility 99–110.
study between Li2TiO3 pebbles and 14Cr-ODS steel, J. Fusi. Ener. 37 (2018) [35] J. Bruley, J. Cho, H.M. Chan, M.P. Harmer, J.M. Rickman, Scanning transmission
247–254. electron microscopy analysis of grain boundaries in creep-resistant yttrium- and
[15] K.A. Terrani, S.J. Zinkle, L.L. Snead, J. Nucl. Mater. 448 (1–3) (2014) 420–435. lanthanum-doped alumina microstructure, J. Am. Ceram. Soc. 82 (10) (1999)
2865–2870.

10

You might also like