You are on page 1of 14

Corrosion Science 195 (2022) 109984

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

High temperature oxidation behavior and mechanism of Al0.3CuCrFeNi2


high-entropy alloy with a coherent γ/γ’ microstructure
Guoqiang Huang a, *, 1, Jie Wu b, 1, Rui Yuan a, Yingxi Li a, Fanqiang Meng a, *, Penghui Lei c,
Chenyang Lu c, *, Fujun Cao d, Yifu Shen d
a
Sino-French Institute of Nuclear Engineering and Technology, Sun Yat-Sen University, Zhuhai 519082, China
b
College of Electronics and Optical Engineering & College of Microelectronics, Nanjing University of Posts and Telecommunications, Nanjing 210023, China
c
Department of Nuclear Science and Technology, Xi’an Jiaotong University, Xi’an 710049, China
d
College of Materials Science and Technology, Nanjing University of Aeronautics and Astronautics (NUAA), Yudao Street 29, 210016 Nanjing, China

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

Keywords: In this study, the oxidation behavior of the Al0.3CuCrFeNi2 HEA with a typical coherent γ/γ’ microstructure was
High-entropy Alloy systematically investigated by the isothermal oxidation tests at 600, 750 and 900 ℃ in air. The results show that
γ/γ’ microstructure the oxidation kinetics of the Al0.3CuCrFeNi2 HEA follow the parabolic law at all three temperatures. As the
Oxidation
oxidation temperature increases, the surface segregation of Cu near the grain boundary occurs, leading to
Formation mechanism
different oxidation behavior in the intragranular regions and near the grain boundary regions. The associated
formation mechanisms of the oxide scales are proposed and discussed in detail.

1. Introduction strength-ductility synergy.


Precipitation strengthening has been demonstrated as an effective
In recent years, high entropy alloys (HEAs) have gained great at­ strategy for further improving the mechanical properties of the FCC
tentions because of their unique microstructural characteristics and HEAs [6]. Various kinds of precipitates like L12, B2, σ, μ, Laves and
promising properties originated from the novel concept of composition carbides, have been introduced into the FCC HEAs as strengthening units
design [1,2]. Unlike conventional metallic alloys with one or two major by purposefully adding the related formers [7–9]. Among these rein­
constituents, HEAs are originally defined to have at least five principle forcing phases, ordered L12-type coherent nanoprecipitates in the FCC
elements in equiatomic or near-equiatomic ratio, and hence endowed HEAs have attracted a great interest due to their promising strength­
with extreme chemical complexity and high configurational entropies, ening effect toward achieving superior strength-ductility synergy
which is generally perceived to facilitate the formation of simple solid [10–14]. Combining the benefits of a ductile FCC HEA matrix and or­
solutions with face centered-cubic (FCC) or body-centered-cubic (BCC) dered L12-type coherent nanoprecipitates has been experimentally
structures, rather than brittle intermetallic compounds (IMCs) [3]. Early demonstrated to enable the resulting alloys to achieve an excellent
HEA research was primarily focused on seeking single-phase solid-­ strength-ductility synergy over a wide range of temperatures. He et al.
solution alloys by carefully controlling configurational entropy [4]. [13] reported that a minor addition of strong γ’ phase formers can
However, single-phase HEAs has been demonstrated experimentally to introduce coherent nanoscale L12 precipitates in the FCC
generally fail to provide the balanced mechanical properties as required CoCrFeNi-based HEA, leading to a high tensile strength of ~1.3 GPa and
for real industrial applications [5]. Specifically, the early interested a good tensile elongation of 17%. Tong et al. [14] revealed that the
single-FCC HEAs normally exhibit low yield strength but exceptionally L12-strengthened HEAs are able to get excellent strength-ductility
terrific ductility and strain hardening capacity, which naturally enables combinations at both ambient and cryogenic temperatures. In partic­
them to serve as a superior base alloy for further strengthening. There­ ular, the newly developed L12 strengthened FCC HEAs have the typical
fore, much effort has been devoted to introducing additional strength­ γ/γ’ microstructure similar to that of the commercial Ni-base superal­
ening mechanisms into the FCC HEAs to achieve superior loys, while incorporating both sluggish diffusion and lattice distortion

* Corresponding authors.
E-mail addresses: gqhuang1105@gmail.com (G. Huang), mengfq5@mail.sysu.edu.cn (F. Meng), chenylu@xjtu.edu.cn (C. Lu).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.corsci.2021.109984
Received 17 August 2021; Received in revised form 22 November 2021; Accepted 23 November 2021
Available online 26 November 2021
0010-938X/© 2021 Elsevier Ltd. All rights reserved.
G. Huang et al. Corrosion Science 195 (2022) 109984

strengthening effects, making them potential candidates for high tem­ 2.2. Isothermal oxidation test
perature structural applications.
To develop the L12 strengthened FCC HEAs with excellent mechan­ Prior to oxidation, the aged slices at three conditions were machined
ical properties at high temperatures, it is crucial to obtain a high volume into square coupons with the dimension of 10 × 10 × 2 mm3, and then
fraction of L12 precipitates with high thermal stability. At the outset, mechanically polished up to 1 µm diamond paste, followed by ultrasonic
near-equiatomic CoCrFeNi-based FCC HEA is a commonly used system cleaning and drying. Isothermal oxidation tests were carried out at 600,
for precipitating L12 phases by adding trace amounts of L12 forming 750 and 900 ◦ C for various holding time ranging from 5 to 100 h in a
elements, such as Al, Ti or both [13]. However, near-equiatomic HEAs muffle furnace in laboratory air. After a required exposure time (5, 10,
with a minor addition of L12 formers, generally produce a relatively low 20, 35, 55, 75 and 100 h), the oxidized coupons were air-cooled outside
L12 phase fraction, and tend to form brittle IMCs (e.g. NiAl-type B2 and the furnace and the weight gain was determined by measuring the total
complex σ phases) [6,15]. In general, the favorable formation of the L12 weight of the coupons and the alumina crucible using an analytical
phases in the near-equiatomic HEAs requires a high ratio of Ni to balance with an accuracy of 10− 4 g. The mass gain results were averaged
L12-former, which can only be achieved by reducing the addition of the from three parallel coupons.
L12-former [16]. Nevertheless, less addition of the L12-former is not
conducive to obtaining a high volume fraction of L12 precipitates.
2.3. Microstructural characterization
Recently, a novel non-equiatomic composition design strategy, i.e.
AxB1+3xCDE, was proposed, which was found to be able to form high
The phase constitution of the aged Al0.3CuCrFeNi2 HEAs and the
content Ni3Al-type L12 precipitates (B3A) while creating a
developed oxide scales at different temperatures were determined using
near-equiatomic FCC matrix (BCDE) in situ [17].
X-ray diffraction (XRD) with Cu Kα radiation at 45 kV and 40 mA with
As a representative HEA that meets the above design strategy, non-
the 2θ angle from 20◦ to 100◦ and a scanning speed of 2◦ /min. The
equiatomic Al0.3CuCrFeNi2 alloy is of particular interest, in which
microstructure of the aged Al0.3CuCrFeNi2 HEAs, and microstructural
expensive Co is completely substituted with Cu and Ni. The higher
characteristics and chemical compositions of the developed oxide scales
valence electron concentration (VEC) values of Cu (11) and Ni (9) than
at different temperatures were characterized by scanning electron mi­
Co (8) favor the formation of single FCC matrix [18]. Moreover, it was
croscope (SEM, Mira3) equipped with an energy X-ray dispersive spec­
found that in the Al0.3CuCrFeNi2 HEA, Cu-rich clusters readily form due
troscopy (EDX) detector. Initial grain structures of the aged
to the positive mixing enthalpy of Cu with other elements, which can
Al0.3CuCrFeNi2 HEAs were identified by electron backscattered
serve as the heterogeneous nucleation sites for the L12 precipitate, help
diffraction (EBSD) with a EDAX detector. For further analyze the
stabilize L12 phase and improve its solvus temperature [19]. For the
microstructure and chemical composition of the specific layer of the
HEAs targeted for service in elevated temperatures, the
oxide scale formed at 900 ◦ C, thin foils extracted from the corresponding
high-temperature oxidation resistance is amongst the most important
oxide layer using focused ion beam (FIB) lift-out method were analyzed
criteria. However, compared to other aspects, such as their phase for­
using a scanning transmission electron microscopy (STEM) equipped
mation criteria, selection criteria, microstructure, mechanical properties
with an energy X-ray dispersive spectroscopy (EDX) system.
and optimization of preparation methods, the high-temperature oxida­
tion of the HEAs is still a relatively new and not yet extensively explored
3. Results
topic. The existing studies have shown the complexity in the
high-temperature oxidation behavior of the HEAs, which is greatly
3.1. Initial microstructure
affected by the temperature [20,21], the atmosphere [22,23], the ratio
of precursor elements forming precipitated phase [24,25], the types and
Fig. 1 shows XRD patterns of the aged Al0.3CuCrFeNi2 HEAs at three
contents of main constituent elements [26–28], the trace elements
different conditions. After three kinds of aging treatment, only FCC
addition [29,30] and the processing routes [31,32]. As for the
peaks are observed for the Al0.3CuCrFeNi2 HEA. The inverse pole figure
Al0.3CuCrFeNi2 HEA, to the best of authors’ knowledge, its
maps and BSE images of the Al0.3CuCrFeNi2 HEAs aged at three different
high-temperature oxidation behavior has not been explored yet to date.
conditions are shown in Fig. 2. The three aged HEAs all have the equi­
Consequently, it is imperative to ascertain its oxidation behavior in
axed grains with basically similar grain size. After isothermal aging at
order to provide appropriate insights to control its oxidation process.
various conditions, homogenously dispersed precipitates with a dark
In this study, the oxidation behavior of the Al0.3CuCrFeNi2 HEA in
the temperature range from 600 ℃ to 900 ℃ in dry air was systemati­
cally investigated. The microstructural characteristics and chemical
compositions of the developed oxide films at various temperatures were
characterized in detail to elucidate the temperature effect on their for­
mation mechanism.

2. Materials and methods

2.1. Materials

A 2-kg ingot with a nominal composition of Al5.6Cu18.8Cr18.8


Fe18.8Ni37.7 (at%), namely Al0.3CuCrFeNi2, was prepared using a high-
frequency vacuum-induction furnace using pure Fe, Cr, Ni, Cu, and Al
metals. The as-cast Al0.3CuCrFeNi2 HEA ingot was then homogenized at
1100 ◦ C for 2 h. Several slices were cut from the inner part of the ho­
mogenized ingot, which were isothermally aged at 600 ℃ for 60 h, 750
℃ for 20 h and 900 ℃ for 20 h, respectively, to obtain the desired γ/γ’
microstructure. The aging temperatures were selected with aim to match
the target oxidation temperatures used in the present study.
Fig. 1. XRD patterns of the Al0.3CuCrFeNi2 HEAs aged at three
different conditions.

2
­
G. Huang et al. Corrosion Science 195 (2022) 109984

Fig. 2. Inverse pole figure maps and BSE images of the three Al0.3CuCrFeNi2 HEAs aged at three different conditions: (a), (d) 600 ℃/60 h; (b), (e) 750 ℃/20 h; (c),
(f) 900 ℃/20 h.

contrast, can be observed throughout the whole matrix for all three aged to 900 ℃ in air. Clearly, as the oxidation temperature increases, the
HEAs, but they exhibit different size and morphology depending on the oxidation rate accelerates, resulting in a faster mass gain, especially at
aging conditions. After aging at 600 ℃ for 60 h and at 750 ℃ for 20 h, a 900 ◦ C. At each specific temperature, oxidation proceeds rapidly in the
mixing of equiaxed-like and needle-shaped precipitates, with virtually first 5 h and then gradually slows down with the oxidation time
indistinguishable difference in size, is clearly visible. With the aging extending. After oxidation for 100 h at 600, 750 and 1000 ◦ C, the cor­
temperature further increasing to 900 ℃, only long needle-shaped responding mass gains per unit area of the Al0.3CuCrFeNi2 HEA are 0.11,
precipitates are formed. The variation in the size and morphology of 0.23 and 1.06 mg/cm2, respectively. Given that the oxidation kinetics of
the precipitates with aging condition in the present study is consistent the Al0.3CuCrFeNi2 HEA at the three temperatures obey well the para­
with previous report [33]. These precipitates with two typical mor­ bolic law, the square of the corresponding mass gains (Δm/A)2 was
phologies can potentially be γ’ phase. It should be mentioned that the plotted as a function of oxidation time, as shown in Fig. 3b. Thus, based
high-temperature oxidation resistance is greatly affected by grain size on the parabolic law expression [36].
and the oxidation behavior usually varies with grain size [34,35]. In the (Δm)
present study, despite the different aging conditions, the three aged 2
= kp t + C (1)
A
samples are in the similar grain size. Thus, the effect of grain size on the
oxidation behavior of the Al0.3CuCrFeNi2 HEA will not be discussed in where Δm is the mass gain, A is the area, kp is the parabolic rate constant,
this study. t is the oxidation time, and C is the integration constant. kp can be
determined from the slope of (Δm/A)2 vs. t plot. The kp values for the
Al0.3CuCrFeNi2 HEA at 600, 750 and 900 ◦ C are 1.285 × 10− 4,
3.2. Isothermal oxidation kinetics
4.438 × 10− 4 and 1.09 × 10− 2 (mg2⋅cm− 4⋅h− 1), respectively.

Fig. 3a shows the mass gain per unit area versus oxidation time of the
Al0.3CuCrFeNi2 HEA oxidized for 100 h at temperature range from 600

Fig. 3. The kinetic curves of the Al0.3CuCrFeNi2 HEA oxidized for 100 h at 600, 750 and 900 ◦ C in air: (a) Mass gain curves, (b) fitted lines of squared mass gain.

3
G. Huang et al. Corrosion Science 195 (2022) 109984

3.3. Surface morphologies of the oxide scale be two kinds of phases, i.e. one is the raised particles in white contrast
(Point 1 in Fig. 7a), and the other is the flat and dense particles beneath
The surface morphologies and the corresponding elemental distri­ these raised particles in dark contrast (Point 2 in Fig. 7a). The compo­
butions of the Al0.3CuCrFeNi2 HEA after oxidization for 100 h at 600, sition of the while one is 56.28 at% O, 4.99 at% Al, 6.31 at% Cr, 6.08 at
750 and 900 ℃ are shown in Figs. 4–6, respectively. For the sample % Fe, 10.29 at% Ni, and 16.05 at% Cu, whereas that of the grey one is
oxidized at 600 ◦ C for 100 h, a uniform and continuous surface oxide 43.13 at% O, 4.88 at% Al, 12.03 at% Cr, 11.55 at% Fe, 19.62 at% Ni,
scale is formed, as illustrated in Fig. 4. At 750 ◦ C, uneven oxidation and 16.05 at% Cu. For the sample oxidized at 750 ℃ for 100 h, the
occurs through the entire surface after oxidation up to 100 h, with two oxides near and far away from grain boundary appear to be very dense,
distinct types of oxides in terms of the morphology formed in the but show completely different morphology. The former is Al-rich oxide
intragranular regions and near the grain boundary regions respectively, (Points 3 and 4 in Fig. 7b), amounted to 63.25 at% O, 15.54 at% Al,
as shown in Fig. 5. The intragranular regions away from grain boundary 5.49 at% Cr, 6.53 at% Fe, 5.22 at% Ni, and 3.97 at% Cu in Point 3 and
are rich in Cr element while the regions near grain boundary are 51.70 at% O, 17.36 at% Al, 8.16 at% Cr, 7.54 at% Fe, 9.67 at% Ni, and
enriched with obvious Al, Cu, Fe and Ni elements. In particular, obvious 5.58 at% Cu in point 4, while the latter is Cr-rich oxide ((Point 5 in
enrichment of the Al, Cu and Ni elements occurs along grain boundaries, Fig. 7c), with 57.12 at% O, 2.65 at% Al, 19.29 at% Cr, 9.37 at% Fe,
but with few Fe and Cr elements. After oxidation at higher temperature 7.99 at% Ni, and 3.58 at% Cu. Despite the oxidation behavior of the
(900 ◦ C) for 100 h, oxidation still occurs unevenly, but is significantly sample near and far away from grain boundary at 900 ℃ is similar to
increased, thus leading to the formation of raised and lowered oxide that at 750 ℃ in view of overall morphologies of the surface oxides, the
scale from the intragranular regions to the regions near the grain oxide evolves more fully at 900 ℃ due to higher oxidation temperature,
boundary across the whole surface, as shown in Fig. 6. Again, abundant mainly representing as the change of oxide morphology and composi­
Al and Cu elements are observed near the grain boundary regions, while tion in the typical regions of the sample surface. The oxides in the
the distribution of Cr, Fe and Ni elements is significantly different from intragranular central region (Fig. 7d) are octahedral oxides containing
that of the sample oxidized at 750 ◦ C for 100 h. Besides Cr element, Fe 57.12 at% O, 2.65 at% Al, 19.29 at% Cr, 9.37 at% Fe, 7.99 at% Ni, and
element is also mainly concentrated in the intragranular regions, but the 3.58 at% Cu, while these at the edge are granular oxides (Fig. 7d), with
distribution of Ni element is relatively uniform across the intergranular 57.12 at% O, 2.65 at% Al, 19.29 at% Cr, 9.37 at% Fe, 7.99 at% Ni, and
and intragranular regions, which is only depleted in intragranular region 3.58 at% Cu. The oxide near the grain boundary at 900 ℃ is also Al-rich
obviously rich in Cr element. Based on the above preliminary observa­ oxide, exhibiting similar morphology and composition, but larger size
tions, it is reasonable to assume that grain boundary and temperature compared with that of the oxidized sample at 750 ℃.
greatly affect the oxidation behavior of the Al0.3CuCrFeNi2 HEA.
Fig. 7 shows the morphological characteristics of the oxides in the
typical regions of the sample surfaces after oxidation at different tem­ 3.4. Cross-sectional morphologies of the oxide scale
peratures for 100 h, as marked by white rectangular box in Figs. 4–6.
Also, EDS point analysis was performed on the representative oxide in The cross-sectional morphologies and the corresponding elemental
these typical regions to determine their chemical composition, and the distributions of the Al0.3CuCrFeNi2 HEA after oxidation at 600, 750 and
corresponding results are listed in Table 1. For the uniform oxide scale 900 ℃ for 100 h are shown in Fig. 8. Noting that the magnification of
on the surface of the sample oxidized at 600 ℃ for 100 h, there seem to the BSE images (Fig. 8b, d, f) is different from each other in order to
show more clearly the differences of the oxide film formed at different

Fig. 4. Surface morphology and corresponding elemental distributions of the Al0.3CuCrFeNi2 HEA after oxidization at 600 ℃ for 100 h.

4
G. Huang et al. Corrosion Science 195 (2022) 109984

Fig. 5. Surface morphology and corresponding elemental distributions of the Al0.3CuCrFeNi2 HEA after oxidization at 750 ℃ for 100 h.

Fig. 6. Surface morphology and corresponding elemental distributions of the Al0.3CuCrFeNi2 HEA after oxidization at 900 ℃ for 100 h.

locations. Despite the differences in the microstructure and composition oxidation at 600 ℃ for 100 h, with no obvious change in the elemental
of the surface oxide films formed at three temperatures, their thickness distributions beneath it. When oxidized at 750 ℃ for 100 h, a contin­
increase significantly with the increasing of oxidation temperature, uous oxide scale with non-uniform thickness is formed on the sample
which is in good agreement with the mass gain results. As shown in surface, where the oxide film is thicker in the intragranular region and
Fig. 8a and b, a flat, uniform and continuous oxide scale is obtained after thinner near the grain boundary region, as indicated in Fig. 8c and d. In

5
G. Huang et al. Corrosion Science 195 (2022) 109984

Fig. 7. Morphological characteristics of the oxides in the typical regions (as marked by white rectangular box in Figs. 4–6) of the Al0.3CuCrFeNi2 HEA after oxi­
dization at various temperatures: (a) 600 ℃; (b), (c) 750 ℃; (d), (e), (f) 900 ℃ for 100 h.

elemental distributions of the typical oxide films formed after oxidation


Table 1
at 600, 750 and 900 ℃ for 100 h. As shown in Fig. 9, at 600 ℃, the oxide
Chemical composition (at%) of the representative oxides corresponding to the
scale is continuous and dense with the average thickness of ~0.64 µm,
points marked by red cross shown in the Fig. 7.
and has a typical duplex structure consisting of an outer layer enriched
Position O Al Cr Fe Ni Cu with Fe and O and an inner layer rich in Cr and O, which is usually
1 56.28 4.99 6.31 6.08 10.29 16.05 observed for the chromia-forming alloys (such as Fe-Ni-Cr alloys) after
2 43.13 4.88 12.03 11.55 19.62 8.79 high temperature oxidation [39–41]. Ni, Cu and Al are relatively uni­
3 63.25 15.54 5.49 6.53 5.22 3.97
formly distributed throughout the oxide scale. For the oxide scale
4 51.70 17.36 8.16 7.54 9.67 5.57
5 57.12 2.65 19.29 9.37 7.99 3.58 formed at 750 ℃, the thin oxide scale near the grain boundary region
6 54.80 1.57 23.24 6.51 10.31 3.57 (Fig. 10a) is duplex with an outer layer enriched with Al, Fe, Ni, Cu and
7 67.04 1.74 10.73 10.75 8.06 1.68 O, and an inner layer rich in Al and O, while the thick oxide scale in the
8 64.25 17.05 6.77 4.16 4.96 2.81 intragranular region (Fig. 10b) is mainly composed of an thin outer layer
9 52.78 22.59 9.52 4.48 6.14 4.49
rich in Fe and O, a thick middle layer rich in Cr and O mixed with several
particles enriched with Cu and Ni, and an internal oxidation zone of Al.
the case of 900 ◦ C, an obviously thickened oxide film but still with un­ In particular, the enrichment of Cu element can be clearly observed
even thickness is formed, where the thickest part is still in the intra­ under the thin oxide scale near the grain boundary region. In the case of
granular region and the thinnest at the grain boundary, as shown in 900 ℃, the thin oxide scale near the grain boundary region comprises an
Fig. 8e and f. It is apparent that at 750 and 900 ◦ C, the microstructure outermost (Al, Ni)-rich layer with alloyed Cu, Cr and Fe, a middle (Cr,
and composition of the thick oxide film in the intragranular region is Fe, Al)-rich oxide layer, and an innermost layer of Al oxide, while the
clearly different from that of the thin oxide film near the grain boundary thick oxide scale in the intragranular region predominantly displays a Cr
region. This coincides well with the surface morphologies and element oxide thick layer, with a thin layer of (Fe, Ni)-rich oxide overlying it and
distributions of the oxide scale formed at 750 and 900 ◦ C, as shown in an obvious internal oxidation zone of Al below it. Again, Cu element is
Figs. 5 and 6. In particular, Cu element is found to be enriched in the found to be enriched near the grain boundary region, even at the oxi­
region close to the grain boundary, leading to the formation of numerous de/substrate interface right above the grain boundary where obvious Cu
Cu clusters in the vicinity, which has been previously reported to occur segregation occurs. It should be noted that at 900 ℃, the middle (Cr, Fe,
in the Al0.3CuCrFeNi2 HEA [19,33]. Moreover, Cu segregation is Al)-rich oxide layer formed near the grain boundary region seems to
observed at the oxide/substrate interface right above the grain bound­ have a smooth transition with the Cr oxide thick layer formed in the
ary. Similar phenomenon that Cu segregated at the oxide/alloy interface intragranular regions, suggesting that they are likely to have the same
during high temperature oxidation of Cu containing alloys has also been origin, but subsequently undergo the different evolution.
found by other researchers [37,38]. Considering the significantly
different microstructure and composition of the oxide film formed in the
3.5. Phase structures of the oxide scale
intragranular regions and near the grain boundary regions, it can be
claimed that there is a significant difference in the oxidation behavior
Fig. 12 shows the XRD patterns of the Al0.3CuCrFeNi2 HEA oxidized
near and far from the grain boundary when the Al0.3CuCrFeNi2 HEAs are
in air at 600, 750 and 900 ℃ for 100 h. It is clear that for the
oxidized at 750 ℃ and 900 ℃.
Al0.3CuCrFeNi2 HEA, the FCC matrix phase can be easily detected under
Given the different microstructure and composition of the oxide
all three oxidation conditions. This signifies that the oxide films formed
films formed at different temperatures or in different regions at the same
at the three temperatures are relatively thin and can be completely
temperature, the detailed cross-sectional microstructure and chemical
penetrated by X-rays. At 600 ℃, the oxide scale mainly consist of Cr2O3,
composition of the representative oxide film corresponding to the three
Fe2O3 and/or Fe3O4. With the increasing temperature, in addition to
temperatures was further revealed. Figs. 9–11 show the cross-sectional
Cr2O3, new peaks of Al2O3, NiFe2O4, NiAl2O4 and Fe(Cr, Al)2O4 are
microstructures at higher magnification and the corresponding
detected in the oxide scale after oxidation at 750 ◦ C. When the oxidation

6
G. Huang et al. Corrosion Science 195 (2022) 109984

Fig. 8. Cross-sectional view and corresponding elemental distributions of the Al0.3CuCrFeNi2 HEA after oxidation at various temperatures: (a) (b) 600 ℃; (c) (d)
750 ℃; (e) (f) 900 ℃ for 100 h.

temperature further rises to 900 ℃, apart from the increase in the exhibiting the lower intrinsic diffusivity for Al than ferritic microstruc­
number of NiAl2O4 peaks, similar diffraction peaks to these obtained at ture [42]. At 900 ℃, a visually continuous Al2O3 inner layer is formed,
750 ◦ C but with increased intensity are observed, thus suggesting similar but significant internal oxidation occurs below it in the intragranular
oxidation mechanism at 750 and 900 ◦ C. region. Therefore, the Al2O3 layers formed in the intragranular region
Combining the XRD results with the previous microstructural and and near the grain boundary region were further examined by
compositional analysis of different oxide films formed at the three TEM-STEM to reveal their detailed microstructure and chemical
temperatures, the phase composition of each sub-layer comprising the composition. The samples were taken from the specified regions by
specific oxide film can be identified. At 600 ℃, the duplex oxide scale focused ion beam (FIB) milling. Figs. 13 and 14 show the STEM images
has an outer Fe2O3& Fe3O4 layer and an inner Cr2O3 layer; At 750 ℃, a and the corresponding EDS maps of the Al2O3 layers formed in the
thin and uniform duplex oxide scale consisting of an outer [NiAl2O4 + Fe intragranular region and near the grain boundary region respectively,
(Cr, Al)2O4] spinel layer and an inner Al2O3 layer is formed near the along with the associated SAED patterns of the corresponding areas. As
grain boundary regions, while a thick and non-uniform duplex oxide shown in Figs. 13 and 14, for two TEM-STEM samples taken from the
scale consisting of an outer NiFe2O4 layer, a middle Cr2O3 layer and an vicinity of the aluminum oxide formed in the intragranular region and
internal Al2O3 zone is developed in the intragranular regions; At 900 ℃, near the grain boundary region, the SAED patterns reveal that the upper
the oxide scale formed near the grain boundary region consists of an oxide layer of Cr-rich is Cr2O3, the middle oxide layer is Al2O3, and the
outer NiAl2O4 spinel layer, a middle Fe(Cr, Al)2O4 layer and an inner lower layer is γ/γ’ matrix, which agrees well with the EDS and XRD
Al2O3 layer, while that formed in the intragranular region is composed results. However, the detailed microstructure of Al2O3 layer formed in
of a very thin NiFe2O4 outer layer, a thick Cr2O3 middle layer and a the two regions is obviously different. The Al2O3 layer formed in the
continuous inner Al2O3 layer with an internal Al2O3 zone below it. intragranular region (Fig. 13) contains a large number of pores holes,
with several long pegs of Al2O3 embedded in the lower alloy matrix,
while that formed near the grain boundary region (Fig. 14) is highly
3.6. TEM-STEM analysis dense, but has detached from the underlying substrate. Such spallation is
likely to occur during the cooling process due to no sign of a thin
For the Al0.3CuCrFeNi2 HEA, it is clear that the higher temperature re-oxidation layer forming beneath it, which to some extent reflects the
facilitates the establishment of a continuous Al2O3 layer, especially for relatively weak interfacial adhesion of the Al2O3 layer formed near the
the intragranular region, due to its austenitic microstructure usually

7
G. Huang et al. Corrosion Science 195 (2022) 109984

Fig. 9. Cross-sectional microstructure at higher magnification and corresponding elemental distributions of the oxide film formed after oxidation at 600 ℃ for 100 h.

Fig. 10. Cross-sectional microstructure at higher magnification and corresponding elemental distributions of the oxide film formed after oxidation at 750 ℃ for
100 h: (a) the near-grain-boundary region; (b) the intragranular region.

8
G. Huang et al. Corrosion Science 195 (2022) 109984

Fig. 11. Cross-sectional microstructure at higher magnification and corresponding elemental distributions of the oxide film formed after oxidation at 900 ℃
for 100 h.

Fig. 12. XRD patterns of the Al0.3CuCrFeNi2 HEA after oxidation in air at 600, 750 and 900 ℃ for 100 h.

9
G. Huang et al. Corrosion Science 195 (2022) 109984

Fig. 13. TEM observation and analysis of the aluminum oxide layer formed in the intragranular region after oxidation at 900 ℃ for 100 h: (a) STEM image and (b)
EDS mappings for O, Al, Ni, Cr, Fe and Cu.

Fig. 14. TEM observation and analysis of the aluminum oxide layer formed near the grain boundary region after oxidation at 900 ℃ for 100 h: (a) STEM image, (b)
EDS mappings for O, Al, Ni, Cr, Fe and Cu, (c) selected area diffraction patterns of corresponding points 1–3 in (a).

grain boundary region with the lower substrate. In fact, the intrinsic 4. Discussion
microstructure of the formed Al2O3 layer greatly determines its own
adhesion to the substrate. For the Al2O3 layer formed in the intra­ In this study, the surface oxide scales with different microstructures
granular region, the long pegs embedded in the alloy substrate facilitate and compositions are observed on the Al0.3CuCrFeNi2 HEA after
the mechanical anchoring of the Al2O3 layer to the substrate [43,44]. oxidation at 600, 750 and 900 ℃ for 100 h in air. All above results
Also, the pores inside the Al2O3 layer are favorable to the release of the indicate the obvious temperature effect on the oxidation behavior of the
stress generated in its growth process, thereby protecting the oxide scale Al0.3CuCrFeNi2 HEA. In the following, the formation mechanisms of the
from falling off and cracking [45]. These two microstructural factors corresponding surface oxide scales at temperature range from 600 to
work together to improve the interfacial adhesion of the oxide scale 900 ℃ are proposed and discussed in detail.
formed in the intragranular region. However, the pores also provide According to the isothermal oxidation test results (Fig. 3), the
channels for the outward diffusion of metal cations and the inward oxidation kinetics of the Al0.3CuCrFeNi2 HEA are known to follow the
diffusion of oxygen, thereby accelerating the oxidation process. parabolic law at temperature range from 600 to 900 ℃. This suggests
that the oxidation of the Al0.3CuCrFeNi2 HEA is a typical diffusion
controlled process, which consists of the initial rapid oxidation stage
involving the adsorption of oxygen on the alloy surface and its chemical

10
G. Huang et al. Corrosion Science 195 (2022) 109984

reaction with metal cations, and the subsequent stable oxide growth grain boundaries.
stage controlled by the diffusion transportation of oxygen anions and/or Considering that the Cu content of the Al0.3CuCrFeNi2 HEA studied
metal cations [46]. In the initial stage of oxidation, only thermody­ in the present work is significantly higher than that of the Ni-Fe-Cr-Al
namically stable oxides are likely to form on the alloy surface. According type alloy, it can be speculated that it is likely that high content Cu
to the Gibbs free energy-temperature map of metal oxides, the Gibbs free leads to different surface states in the intragranular regions and near the
energies of the oxides related to the involved elements in the grain boundary regions when oxidized at 750 and 900 ℃. Doi et al. [52].
Al0.3CuCrFeNi2 HEA are all negative at temperature range from 600 to revealed the occurrence of Cu segregation at oxide/alloy interface on the
900 ℃, with their magnitudes ordered in Al2O3 < Cr2O3 < FeO < Fe3O4 Ni-Cr alloy containing alloyed Cu just 300 s after oxidation at 923 K in
< NiO < Fe2O3 < CuO [47,48]. This signifies that the alloying elements the dusting environment of 60 vol% CO- 26 vol% H2-11.5 vol%
on the surface of the Al0.3CuCrFeNi2 HEA compete with each other to be CO2-2.5 vol% H2O by hard X-ray photoelectron spectroscopy (HX-PES),
oxidized, with the oxygen affinity (i.e. oxide forming ability) in the which is considered to be able to suppress the dissociative absorption of
order of Al > Cr > Fe > Ni > Cu. In this case, the alloying elements with CO. A quite recent work by Hayashi et al. [53]. found that the addition of
higher oxygen affinity can be oxidized prior to those with lower oxygen Cu into FCC Fe-Ni-Cr-Al and Ni-Cr-Al based alloys is beneficial for the
affinity. However, the formation of the initial oxide film on the alloy formation of external Al2O3 scale. The beneficial effect of Cu is explained
surface is also greatly influenced by the kinetic conditions, which are to be due to the decreased oxygen concentration on the alloy surface. In
highly related to the oxidation temperature and solute concentrations the present study, the alloy surfaces oxidized at 750 and 900 ℃ for 5 h
[47,49]. For the Al0.3CuCrFeNi2 HEA, although Al2O3 is most favored to both show a clear enrichment of Cu, as shown in the Supplementary
form thermodynamically in the initial oxidation stage at 600, 750 and materials (Fig. S1 and Fig. S2). After oxidation at 750 and 900 ℃ for
900 ℃, no external Al2O3 layer is formed in all three cases, indicating 100 h, Cu element is found to be enriched in the subsurface close to the
that the formation of a continuous external Al2O3 layer in the initial grain boundary, and Cu segregation layer forms below the oxide scale
stage is not kinetically supported under all three conditions. This is right above the grain boundary (Fig. 11). It can be anticipated that Cu
mainly due to the low bulk Al content and the slow diffusion of Al from segregation occurs at the air/alloy interface in the early stages of
the substrate bulk to the alloy/oxide interface, which make the surface oxidation at 750 and 900 ℃, which effectively reduces the oxygen
Al content not reach the critical Al concentration required to form a pressure on the alloy surface to the point where the Cr cannot be
continuous external Al2O3 layer. Consequently, high content (~18.8 at oxidized. Thus, the [Fe(Cr, Al)2O4 + NiAl2O4] and NiAl2O4 spinel layer
%) Cr with oxygen affinity second only to Al is expected to preferentially first form near the grain boundary regions in the initial stages of
react with oxygen to form the Cr2O3 layer on the whole alloy surface in oxidation at 750 and 900 ℃, respectively. It has been reported that the
the initial oxidation stage at the three temperatures. This expectation increase of Cu content in the Ni-Cr-Al alloy can lead to the change of
seems to be correct at 600 ℃, but at 750 and 900 ℃, the oxides with initial oxide film formed after 4 h of oxidation in air at 1000 ℃ from the
different morphologies and compositions are initially formed in the Cr2O3 to the NiAl2O4 [53]. It should be mentioned that prior to the
intra-grain regions and near the grain boundary regions, respectively. As formation of spinel-type oxides, such as Fe(Cr, Al)2O4 and NiAl2O4,
shown in the Supplementary materials (Fig. S1 and Fig. S2), just 5 h other oxides, like NiO and Cu2O, may also form in the earlier stage of
after oxidation at 750 and 900 ℃ respectively, the Cr-rich oxides only initial oxidation, which requires further study in the future.
form in the intra-grain regions, whereas the oxides rich in Al, Cu and Ni Once the oxide layer is initially formed over the entire alloy surface,
tend to form near the grain boundary regions. At 750 ℃, Fe is present the oxidation process will be kinetically controlled by the diffusion
throughout the whole surface oxide scale, with slight enrichment in the through the oxide. After the initial formation of Cr2O3 layer at 600 ℃,
Cr-rich oxides formed in the intragranular regions, while at 900 ℃ it is the Fe cations diffuse outward through the Cr2O3 layer more rapidly due
mainly enriched in the intragranular regions where the Cr-rich oxides to its much higher diffusion coefficient in the Cr2O3 layer than the other
are located, especially along their edges. This suggests that the surface elements of the Al0.3CuCrFeNi2 HEA, resulting in the formation of an Fe-
states of the Al0.3CuCrFeNi2 HEA are distinctly different in the intra­ rich oxide layer on the top of Cr2O3 layer [54]. The early establishment
granular regions and near the grain boundary regions in the early stages of the protective Cr2O3 layer is well known to greatly reduce the oxygen
of oxidation at 750 and 900 ℃, thereby leading to the occurrence of pressure at the Cr2O3/alloy interface, thereby promoting the selective
different oxidation reactions. Here, for the Al0.3CuCrFeNi2 HEA oxidized oxidation of Al beneath it [51]. If the Al content near the Cr2O3/alloy
at 750 and 900 ℃, a question that how the different surface states in the interface is above a certain critical level, a continuous, protective Al2O3
intragranular regions and near the grain boundary regions are generated layer can be formed, otherwise an internal oxide zone of Al2O3 will be
naturally arises. produced [55]. Here, no internal oxidation of Al was found after oxi­
For the M-Cr type alloys (M is Ni, Fe or Ni + Fe) with the Cr content dization at 600 ℃ for 100 h, which is mainly due to that the growth rate
close to 18 wt%, it is well known that an exclusive Cr2O3 scale can be of the duplex-layer oxide scale is higher than that of oxygen penetration.
easily formed on the alloy surface in the initial stage of high-temperature Finally, the duplex-layer oxide scale consisting of an outer Fe2O3&Fe3O4
oxidation [50]. Specially, the Cr2O3 is preferentially formed at the grain layer and an inner Cr2O3 layer is formed on the Al0.3CuCrFeNi2 HEA
boundaries, followed by sideways diffusion of Cr and spreading of surface after oxidization at 600 ℃ for 100 h.
Cr2O3, which is caused by the significantly high Cr flux due to the rapid In the case of 750 and 900 ℃, with the initial formation of the Cr2O3
diffusion of Cr along grain boundaries [35,49]. Adding a certain amount and the spinel-type oxides (Fe(Cr, Al)2O4 and NiAl2O4) in the intra-grain
of Al into the M-Cr type alloys may lead to the formation of a continuous regions and near the grain boundary regions respectively, the oxygen
external Al2O3 scale above a certain temperature. However, in the pre­ pressure at the oxide/alloy interface is effectively reduced. Thus, the
sent study, only internal Al2O3 layer was formed near the grain temperature-dependent selective oxidation of Al begins to occur in the
boundary regions at both 750 and 900 ℃. This indicates that the Cr2O3 intragranular regions and near the grain boundary regions at both
is still expected to form first, which helps to reduce the oxygen pressure temperatures. At 750 ℃, a thin continuous Al2O3 inner layer is formed
and thereby promotes the formation of Al2O3 below it [51]. Meanwhile, beneath the (Fe(Cr, Al)2O4 + NiAl2O4) layer near the grain boundary
it is reasonable to assume that higher Cr content near the grain regions, while Al is only internally oxidized below the Cr2O3 layer in the
boundaries should be more favorable for the formation of Cr2O3, but this intra-grain regions. This is mainly attributed to the faster Al trans­
is not the case. The unexpected absence of Cr2O3 near the grain portation flow near the grain boundary than in the grain interior,
boundaries (see Figs. 5, 6, Figs. S1 and S2) suggests that the surface resulting in the Al content in the near-surface region adjacent to the
conditions of the Al0.3CuCrFeNi2 HEA near the grain boundaries do not grain boundary exceeding the critical Al concentration required to form
support the formation of Cr2O3. Therefore, there must be other factors a continuous Al2O3 layer. Moreover, the lower oxygen pressure beneath
affecting the surface conditions of the Al0.3CuCrFeNi2 HEA near the the [Fe(Cr, Al)2O4 + NiAl2O4] layer, which requires a lower critical Al

11
G. Huang et al. Corrosion Science 195 (2022) 109984

level to form a continuous Al2O3 layer, is also beneficial for the estab­ formation of the spinel-type oxide is restricted to closer to the grain
lishment of the continuous Al2O3 layer near the grain boundary. At boundaries, and the subsequent lateral growth covers the adjacent Cr2O3
900 ℃, a thicker continuous and compact Al2O3 inner layer is formed layer. With the formation of the initial spinel-type oxide layer and Cr2O3
near the grain boundary, whereas an inner Al2O3 layer that looks dense layer, the resulting reduced oxygen pressure at oxide/alloy interface
but has micropores inside, is developed in the intergranular regions, triggers the selective oxidation of Al. Due to the relatively high Al
with significant internal oxidation of Al occurred beneath it. It is concentration caused by the faster Al transportation near the grain
apparent that the higher temperature, capable of accelerating the boundary, a continuous and dense Al2O3 inner layer is formed below the
transportation of Al to the alloy surface, contributes to the development initial spinel-type oxide layer at both temperatures, with its thickness
of a continuous Al2O3 inner layer below the initially formed oxide layer, thicker at 900 ℃ than at 750 ℃. Unlike this, because of the slowing of Al
especially near the grain boundary regions. Nevertheless, it seems that at supply in the intragranular regions towards the near-surface region of
900 ℃, the Al supply to alloy surface in the grain interiors is still the alloy, the Al concentration near the Cr2O3/alloy interface cannot
insufficient to establish a continuous and dense Al2O3 inner layer. Thus, reach the critical Al level to form continuous and dense Al2O3 layer,
at either 750 or 900 ℃, the continuous and dense Al2O3 inner layer leading to the formation of an internal oxidation zone of Al below the
formed near the grain boundary can better impede the oxygen pene­ Cr2O3 layer at 750 ℃; at 900 ℃, the elevated temperature effectively
tration and the outward diffusion of metal cations, while the formation accelerates the transport of Al and an visually continuous but actually
of the Al2O3 internal oxidation zone and the non-dense Al2O3 inner layer non-dense Al2O3 inner layer with an internal Al2O3 zone immediately
in the intragranular regions allows Fe and Ni cations to diffuse up and below is generated beneath the Cr2O3 layer. Due to the non-dense nature
through the upper Cr2O3 layer and form the outward-growing NiFe2O4 of the oxidation products below the Cr2O3 layer in the intragranular
layer at 750 and 900 ℃, respectively. As a consequence, the thickness of regions, the Fe and Ni cations can easily diffuse up and through the
the surface oxide scale in the intragranular regions is significantly higher Cr2O3 layer, with the NiFe2O4 layer formed above it at 750 and 900 ℃
than that near the grain boundary, especially at 900 ℃. It should be respectively. In the meantime, the initial Cr2O3 layer covered by the
mentioned that more obvious enrichment of Fe in the surface oxide scale NiAl2O4 layer at 900 ℃ is developed into Fe(Cr, Al)2O4. Thus, after
formed near the grain boundary at 750 ℃ than at 900 ℃ is likely due to oxidation at 750 and 900 ℃ for 100 h, a thin duplex-layer oxide scale
the longer required time to form a continuous and dense Al2O3 inner consisting of an outer spinel-type oxide layer and an inner Al2O3 layer is
layer caused by the relatively low temperature, which allows more Fe formed near the grain boundary regions, with the more obvious overlap
cations to diffuse into the upper oxide layer. between the NiAl2O4 layer and the Cr2O3 layer at 900 ℃, while in the
From the above discussion, the oxidation mechanisms of the intragranular regions, a thick oxide scale composed of a thin NiFe2O4
Al0.3CuCrFeNi2 HEA after isothermal oxidation at 600, 750 and 900 ℃ layer, a thick Cr2O3 layer, and an internal Al2O3 zone (Fig. 15b) and of a
for 100 h are schematically summarized in Fig. 15. At 600 ℃, a rela­ thin NiFe2O4 layer, a thick Cr2O3 layer, and a non-dense Al2O3 inner
tively uniform Cr2O3 layer is first formed on the alloy surface, followed layer with an internal Al2O3 zone below it (Fig. 15c) is formed at 750
by the formation of Fe2O3&Fe3O4 layer above it due to the faster and 900 ℃ respectively.
diffusion of Fe in the Cr2O3 layer. As a result, a duplex-layer oxide scale On the basis of the in-depth understanding on the oxidation mech­
consisting of an outer Fe2O3&Fe3O4 layer and an inner Cr2O3 layer is anism of the Al0.3CuCrFeNi2 HEA at different temperatures, it can be
formed after oxidation for 100 h, as shown in Fig. 15a. At 750 and asserted that fine grains are favorable to its oxidation resistance since
900 ℃, the surface segregation of Cu near the grain boundaries leads to high grain boundary density is beneficial to supply more Cu and Al to the
different oxidation behavior near the grain boundary regions and in the alloy surface, which enables the initial oxide layer to change from the
intragranular regions. The spinel-type oxide layer (Fe(Cr, Al)2O4 Cr2O3 layer to the NiAl2O4 layer, and facilitates the subsequent estab­
+ NiAl2O4 at 750 ℃ and NiAl2O4 at 900 ℃) is first formed near the lishment of a continuous and dense Al2O3 inner layer. Therefore, the
grain boundary regions, while the Cr2O3 layer is formed in the intra­ thermomechanical processing for the Al0.3CuCrFeNi2 HEA is recom­
granular regions. In fact, the initial formation and subsequent growth of mended to optimize grain size in the future work. In addition, the
both the spinel-type oxides near the grain boundary regions and the adhesion between the Al2O3 inner layer and the alloy substrate requires
Cr2O3 in the intragranular regions compete against each other, which is further amelioration. Addition of the rare earth elements (REs), such as
obviously affected by temperature. At higher temperatures, the initial Y, has been proved to be an effective strategy to improve the adhesion of
the oxide scale with alloy substrate and slow down the growth rate of the
scale [56–58]. Thus, alloying the Al0.3CuCrFeNi2 HEA with REs can be
applied for improving both the adhesion and the oxidation resistance. In
summary, the oxidation resistance of the Al0.3CuCrFeNi2 HEA can be
further improved in the future by refining grains or adding trace REs
elements.

5. Conclusions

In the present study, the oxidation behavior of the Al0.3CuCrFeNi2


HEA at 600, 750 and 900 ◦ C in air was systematically investigated. The
isothermal oxidation kinetics of the Al0.3CuCrFeNi2 HEA follow the
parabolic law at all three temperatures, with the oxidation rate
increasing with the temperature, especially at 900 ℃. After oxidation at
the three temperatures for 100 h, the corresponding mass gains per unit
area of the Al0.3CuCrFeNi2 HEA are 0.11, 0.23 and 1.06 mg/cm2,
respectively. Raising the temperature induces progressively different
oxidation behavior in the intragranular regions and near the grain
boundary regions. This can be attributed to the surface segregation of Cu
near the grain boundary in the initial oxidation stage, which leads to
different surface states in the intragranular regions and near the grain
boundary regions, and thereby affects the initial oxide formation as well
Fig. 15. The schematic illustration of the proposed oxide growth. as the subsequent oxide development. At 600 ℃, the entire surface of

12
G. Huang et al. Corrosion Science 195 (2022) 109984

the alloy is uniformly oxidized, resulting in the formation of a uniform [9] Y.L. Zhao, T. Yang, Y. Tong, J. Wang, J.H. Luan, Z.B. Jiao, D. Chen, Y. Yang, A. Hu,
C.T. Liu, J.J. Kai, Heterogeneous precipitation behavior and stacking-fault-
and continuous oxide scale consisting of an outer Fe2O3& Fe3O4 layer
mediated deformation in a CoCrNi-based medium-entropy alloy, Acta Mater. 138
and an inner Cr2O3 layer; at 750 and 900 ℃, the oxide scale near the (2017) 72–82.
grain boundary typically comprises an external spinel-type oxide layer [10] Z.G. Wang, W. Zhou, L.M. Fu, J.F. Wang, R.C. Luo, X.C. Han, B. Chen, X.D. Wang,
and an internal Al2O3 layer at both temperatures, while that in the Effect of coherent L12 nanoprecipitates on the tensile behavior of a fcc-based high-
entropy alloy, Mater. Sci. Eng. A 696 (2017) 503–510.
intragranular regions is composed of an outer NiFe2O4 thin layer, an [11] W. Lu, X. Luo, Y. Yang, K. Yan, B. Huang, P. Li, Nano-precipitates strengthened
intermediate Cr2O3 thick layer and an internal Al2O3 zone at 750 ℃/a non-equiatomic medium-entropy alloy with outstanding tensile properties, Mater.
non-dense Al2O3 inner layer with an internal Al2O3 zone immediately Sci. Eng. A 780 (2020), 139218.
[12] X. Huang, L. Huang, H. Peng, Y. Liu, B. Liu, S. Li, Enhancing strength-ductility
below at 900 ℃. synergy in a casting non-equiatomic NiCoCr-based high-entropy alloy by Al and Ti
combination addition, Scr. Mater. 200 (2021), 113898.
CRediT authorship contribution statement [13] J.Y. He, H. Wang, H.L. Huang, X.D. Xu, M.W. Chen, Y. Wu, X.J. Liu, T.G. Nieh,
K. An, Z.P. Lu, A precipitation-hardened high-entropy alloy with outstanding
tensile properties, Acta Mater. 102 (2016) 187–196.
Guoqiang Huang: Conceptualization, Methodology, Validation, [14] Y. Tong, D. Chen, B. Han, J. Wang, R. Feng, T. Yang, C. Zhao, Y.L. Zhao, W. Guo,
Investigation, Formal analysis, Writing – original draft. Jie Wu: Inves­ Y. Shimizu, C.T. Liu, P.K. Liaw, K. Inoue, Y. Nagai, A. Hu, J.J. Kai, Outstanding
tensile properties of a precipitation-strengthened FeCoNiCrTi0.2 high-entropy
tigation, Formal analysis, Writing – review & editing. Rui Yuan: alloy at room and cryogenic temperatures, Acta Mater. 165 (2019) 228–240.
Investigation, Formal analysis. Yingxi Li: Investigation, Formal anal­ [15] W.H. Liu, Z.P. Lu, J.Y. He, J.H. Luan, Z.J. Wang, B. Liu, M.W. Chen, C.T. Liu,
ysis. Fanqiang Meng: Supervision, Writing – review & editing. Penghui Ductile CoCrFeNiMox high entropy alloys strengthened by hard intermetallic
phases, Acta Mater. 116 (2016) 332–342.
Lei: Investigation, Formal analysis. Chenyang Lu: Supervision, Writing [16] D. Li, C. Li, T. Feng, Y. Zhang, G. Sha, J.J. Lewandowski, P.K. Liaw, Y. Zhang, High-
– review & editing. Fujun Cao: Investigation, Formal analysis. Yifu entropy Al0.3CoCrFeNi alloy fibers with high tensile strength and ductility at
Shen: Investigation, Formal analysis, Writing – review & editing. All ambient and cryogenic temperatures, Acta Mater. 123 (2017) 285–294.
[17] Y.J. Liang, L.J. Wang, Y.R. Wen, B.Y. Cheng, Q.L. Wu, T.Q. Cao, Q. Xiao, Y.F. Xue,
authors read and contributed to the manuscript. G. Sha, Y.D. Wang, Y. Ren, X.Y. Li, L. Wang, F.C. Wang, H.N. Cai, High-content
ductile coherent nanoprecipitates achieve ultrastrong high-entropy alloys, Nat.
Commun. 9 (2018) 1–8.
Declaration of Competing Interest [18] C. Ng, S. Guo, J. Luan, Q. Wang, J. Lu, S. Shi, C.T. Liu, Phase stability and tensile
properties of Co-free Al0.5CrCuFeNi2 high-entropy alloys, J. Alloy. Compd. 584
The authors declare that they have no known competing financial (2014) 530–537.
[19] B. Gwalani, V. Soni, D. Choudhuri, M. Lee, J.Y. Hwang, S.J. Nam, H. Ryu, S.
interests or personal relationships that could have appeared to influence H. Hong, R. Banerjee, Stability of ordered L12 and B2 precipitates in face centered
the work reported in this paper. cubic based high entropy alloys-Al0.3CoFeCrNi and Al0.3CuFeCrNi2, Scr. Mater. 123
(2016) 130–134.
[20] G. Laplanche, U.F. Volkert, G. Eggeler, E.P. George, Oxidation behavior of the
Data Availability CrMnFeCoNi high-entropy alloy, Oxid. Met. 85 (2016) 629–645.
[21] Y.K. Kim, Y.A. Joo, H.S. Kim, K.A. Lee, High temperature oxidation behavior of Cr-
The raw/processed data required to reproduce these findings cannot Mn-Fe-Co-Ni high entropy alloy, Intermetallics 98 (2018) 45–53.
[22] C. Stephan-Scherb, W. Schulz, M. Schneider, S. Karafiludis, G. Laplanche, High-
be shared at this time as the data also forms part of an ongoing study. temperature oxidation in dry and humid atmospheres of the equiatomic
CrMnFeCoNi and CrCoNi high- and medium-entropy alloys, Oxid. Met. 95 (2021)
Acknowledgements 105–133.
[23] W. Kai, F.C. Chien, F.P. Cheng, R.T. Huang, J.J. Kai, C.T. Liu, The corrosion of an
equimolar FeCoNiCrMn high-entropy alloy in various CO2/CO mixed gases at 700
This study was financially supported by the National Natural Science and 950◦ C, Corros. Sci. 153 (2019) 150–161.
Foundation of China (Grant No. 12075179). Jie Wu acknowledges the [24] A. Erdogan, K.M. Doleker, S. Zeytin, Effect of Al and Ti on high-temperature
oxidation behavior of CoCrFeNi-based high-entropy alloys, JOM 71 (2019)
financial support of Natural Science Foundation of Jiangsu Province 3499–3510.
(Grant No. BK20200746), General Projects Of Natural Science Research [25] Z.Y. Ding, B.X. Cao, J.H. Luan, Z.B. Jiao, Synergistic effects of Al and Ti on the
in Universities of Jiangsu Province (Grant No. 20KJB460008) and Start- oxidation behaviour and mechanical properties of L12-strengthened FeCoCrNi
high-entropy alloys, Corros. Sci. 184 (2021), 109365.
Up Fund from Nanjing University of Posts and Telecommunications
[26] W. Kai, C.C. Li, F.P. Cheng, K.P. Chu, R.T. Huang, L.W. Tsay, J.J. Kai, Air-oxidation
(Grant No. NY220077). of FeCoNiCr-based quinary high-entropy alloys at 700-900 ℃, Corros. Sci. 121
(2017) 116–125.
[27] J. Lu, Y. Chen, H. Zhang, L. Li, L.M. Fu, X.F. Zhao, F.W. Guo, P. Xiao, Effect of Al
Appendix A. Supporting information content on the oxidation behavior of Y/Hf-doped AlCoCrFeNi high-entropy alloy,
Corros. Sci. 170 (2020), 108691.
Supplementary data associated with this article can be found in the [28] N.K. Adomako, J.H. Kim, Y.T. Hyun, High-temperature oxidation behaviour of low-
entropy alloy to medium-and high-entropy alloys, J. Therm. Anal. Calorim. 133
online version at doi:10.1016/j.corsci.2021.109984.
(2018) 13–26.
[29] M. Vilémová, K. Illková, Š. Csáki, F. Lukáč, H. Hadraba, J. Matějíček, Z. Chlup,
References J. Klečka, Thermal and oxidation behavior of CoCrFeMnNi alloy with and without
yttrium oxide particle dispersion, J. Mater. Eng. Perform. 28 (2019) 5850–5859.
[30] J. Lu, H. Zhang, L. Li, A.H. Huang, X.Z. Liu, Y. Chen, X.C. Zhang, F.W. Guo, X.
[1] Y.F. Ye, Q. Wang, J. Lu, C.T. Liu, Y. Yang, High-entropy alloy: challenges and
F. Zhao, Y-Hf co-doped Al1.1CoCr0.8FeNi high-entropy alloy with excellent
prospects, Mater. Today 19 (2016) 349–362.
oxidation resistance and nanostructure stability at 1200 ◦ C, Scr. Mater. 203 (2021),
[2] E.P. George, D. Raabe, R.O. Ritchie, High-entropy alloys, Nat. Rev. Mater. 4 (2019)
114105.
515–534.
[31] Y. Wang, M. Zhang, J. Jin, P. Gong, X. Wang, Oxidation behavior of CoCrFeMnNi
[3] Y. Zhang, Y.J. Zhou, J.P. Lin, G.L. Chen, P.K. Liaw, Solid-solution phase formation
high entropy alloy after plastic deformation, Corros. Sci. 163 (2020), 108285.
rules for multi-component alloys, Adv. Eng. Mater. 10 (2008) 534–538.
[32] Z.P. Tong, H.L. Liu, J.F. Jiao, W.F. Zhou, Y. Yang, X.D. Ren, Laser additive
[4] Y. Zhang, T.T. Zuo, Z. Tang, M.C. Gao, K.A. Dahmen, P.K. Liaw, Z.P. Lu,
manufacturing of CrMnFeCoNi high entropy alloy: microstructural evolution, high-
Microstructures and properties of high-entropy alloys, Prog. Mater. Sci. 61 (2014)
temperature oxidation behavior and mechanism, Opt. Laser Technol. 130 (2020),
1–93.
106326.
[5] D.B. Miracle, O.N. Senkov, A critical review of high entropy alloys and related
[33] B. Gwalani, D. Choudhuri, V. Soni, Y. Ren, M. Styles, J.Y. Hwang, S.J. Nam, H. Ryu,
concepts, Acta Mater. 122 (2017) 448–511.
S.H. Hong, R. Banerjee, Cu assisted stabilization and nucleation of L12 precipitates
[6] T. Yang, Y.L. Zhao, B.X. Cao, J.J. Kai, C.T. Liu, Towards superior mechanical
in Al0.3CuFeCrNi2 fcc-based high entropy alloy, Acta Mater. 129 (2017) 170–182.
properties of hetero-structured high-entropy alloys via engineering
[34] S. Pour-Ali, M. Weiser, N.T. Nguyen, A.R. Kiani-Rashid, A. Babakhani, S. Virtanen,
multicomponent intermetallic nanoparticles, Scr. Mater. 183 (2020) 39–44.
High temperature oxidation behaviour of AISI 321 stainless steel with an ultrafine-
[7] D. Choudhuri, B. Gwalani, S. Gorsse, M. Komarasamy, S.A. Mantri, S.G. Srinivasan,
grained surface at 800 ◦ C in Ar-20 vol% O2, Corros. Sci. 163 (2020), 108282.
R.S. Mishra, R. Banerjee, Enhancing strength and strain hardenability via
[35] J.H. Kim, D.I. Kim, S. Suwas, E. Fleury, K.W. Yi, Grain-size effects on the high
deformation twinning in fcc-based high entropy alloys reinforced with
temperature oxidation of modified 304 austenitic stainless steel, Oxid. Met. 79
intermetallic compounds, Acta Mater. 165 (2019) 420–430.
(2013) 239–247.
[8] K. Ming, X. Bi, J. Wang, Precipitation strengthening of ductile
Cr15Fe20Co35Ni20Mo10 alloys, Scr. Mater. 137 (2017) 88–93.

13
G. Huang et al. Corrosion Science 195 (2022) 109984

[36] Y. Xu, W. Li, L. Qu, X. Yang, B. Song, R. Lupoi, S. Yin, Solid-state cold spraying of [47] D. Zou, Y. Zhou, X. Zhang, W. Zhang, Y. Han, High temperature oxidation behavior
FeCoCrNiMn high-entropy alloy: an insight into microstructure evolution and of a high Al-containing ferritic heat-resistant stainless steel, Mater. Charact. 136
oxidation behavior at 700–900 ◦ C, J. Mater. Sci. Technol. 68 (2021) 172–183. (2018) 435–443.
[37] J.H. Kim, D.I. Kim, J.H. Shim, K.W. Yi, Investigation into the high temperature [48] X. Cheng, K. Du, D. Wang, Effect of doping Al on the high-temperature oxidation
oxidation of Cu-bearing austenitic stainless steel using simultaneous electron behavior of Ni-11Fe-10Cu alloy, Oxid. Met. 93 (2020) 417–431.
backscatter diffraction-energy dispersive spectroscopy analysis, Corros. Sci. 77 [49] J.H. Kim, B.K. Kim, D.I. Kim, P.P. Choi, D. Raabe, K.W. Yi, The role of grain
(2013) 397–402. boundaries in the initial oxidation behavior of austenitic stainless steel containing
[38] H. Shi, R. Fetzer, A. Jianu, A. Weisenburger, A. Heinzel, F.B. Lang, G. Müller, alloyed Cu at 700C for advanced thermal power plant applications, Corros. Sci. 96
Influence of alloying elements (Cu, Ti, Nb) on the microstructure and corrosion (2015) 52–66.
behaviour of AlCrFeNi-based high entropy alloys exposed to oxygen-containing [50] V.P. Deodeshmukh, S.J. Matthews, D.L. Klarstrom, High-temperature oxidation
molten Pb, Corros. Sci. 190 (2021), 109659. performance of a new alumina-forming Ni-Fe-Cr-Al alloy in flowing air, Int. J.
[39] W. Wei, S.J. Geng, D.B. Xie, F.H. Wang, High temperature oxidation and corrosion Hydrog. Energy 36 (2011) 4580–4587.
behaviours of Ni-Fe-Cr alloys as inert anode for aluminum electrolysis, Corros. Sci. [51] M.P. Brady, I.G. Wright, B. Gleeson, Alloy design strategies for promoting
157 (2019) 382–391. protective oxide-scale formation, JOM 52 (2000) 16–21.
[40] W. Wei, S.J. Geng, C. Gang, F.H. Wang, Growth mechanism of surface scales on Ni- [52] T. Doi, K. Kitamura, Y. Nishiyama, N. Otsuka, T. Kudo, M. Sato, E. Ikenaga,
Fe-Cr alloys at 960 ◦ C in air, Corros. Sci. 173 (2020), 108737. S. Ueda, K. Kobayashi, Analysis of Cu segregation to oxide-metal interface of Ni-
[41] S. Zhang, S. Hayashi, S. Ukai, N.H. Oono, Effect of Co addition on the high- based alloy in a metal-dusting environment, Surf. Interface Anal. 40 (2008)
temperature oxidation behavior of oxide-dispersion-strengthened FeCrAl alloys, 1374–1381.
Corros. Sci. 184 (2021), 109391. [53] S. Hayashi, D. Kudo, R. Nagashima, H. Utsumi, Effect of Cu on oxidation behaviour
[42] M.P. Brady, Y. Yamamoto, B.A. Pint, M.L. Santella, P.J. Maziasz, L.R. Walker, On of FCC Fe-Ni-Cr-Al and Ni-Cr-Al based alloys, Corros. Sci. 163 (2020), 108273.
the loss of protective scale formation in creep-resistant, alumina-forming austenitic [54] R.E. Lobnig, H.P. Schmidt, K. Hennesen, H.J. Grabke, Diffusion of cations in
stainless steels at 900 ◦ C in air, Mater. Sci. Forum 595 (2008) 725–732. chromia layers grown on iron-base alloys, Oxid. Met. 37 (1992) 81–93.
[43] D. Pilone, F. Felli, A. Brotzu, High temperature oxidation behaviour of TiAl-Cr-Nb- [55] D.J. Young, High Temperature Oxidation and Corrosion of Metals. Alloy oxidation
Mo alloys, Intermetallics 43 (2013) 131–137. II: Internal Oxidation, Elsevier, 2008, pp. 261–333.
[44] Y. Pan, X. Lu, M.D. Hayat, F. Yang, C. Liu, Y. Li, X.Y. Li, W. Xu, X.H. Qu, P. Cao, [56] Y.X. Xu, J.T. Lu, W.Y. Li, Z. Yang, Y.Y. Dang, X.W. Yang, Effect and role of alloyed
Effect of Sn addition on the high-temperature oxidation behavior of high Nb- yttrium on the fireside corrosion behaviour of Ni-Fe based alloys for 750 ◦ C ultra-
containing TiAl alloys, Corros. Sci. 166 (2020), 108449. supercritical boiler applications, Corros. Sci. 143 (2018) 148–156.
[45] J. Ju, C. Yang, S.Q. Ma, M.D. Kang, K.M. Wang, J.J. Li, H.G. Fu, J. Wang, Effect of [57] S. Wang, Z.B. Zheng, K.M. Zheng, J. Long, J. Wang, Y.Y. Ren, Y.M. Li, High
temperature on oxidation resistance and isothermal oxidation mechanism of novel temperature oxidation behavior of heat resistant steel with rare earth element Ce,
wear-resistant Fe-Cr-B-Al-C-Mn-Si alloy, Corros. Sci. 170 (2020), 108620. Mater. Res. Express 7 (2020), 016571.
[46] C. Wagner, Theoretical analysis of the diffusion processes determining the [58] J. Lu, H. Zhang, Y. Chen, L. Li, X.Z. Liu, W.W. Xiao, N. Ni, X.F. Zhao, F.W. Guo,
oxidation rate of alloys, J. Eelectrochem. Soc. 99 (1952) 369–380. P. Xiao, Y-doped AlCoCrFeNi2.1 eutectic high-entropy alloy with excellent
oxidation resistance and structure stability at 1000 ◦ C and 1100 ◦ C, Corros. Sci.
180 (2021), 109191.

14

You might also like