You are on page 1of 8

Met. Mater. Int., Vol. 22, No. 5 (2016), pp.

789~796
doi: 10.1007/s12540-016-6305-1

High Temperature Oxidation Resistance of Ni-(5~13)Co-(10~16)Cr-(5~9)W-5Al-


(1~1.5)Ti-(3~6)Ta Alloys

Hee-Soo Kim1, Si-Jun Park2, Seong-Moon Seo3, Young-Soo Yoo3, Hi-Won Jeong3, and HeeJin Jang1,*

1
Department of Materials Science and Engineering, Chosun University, 309 Pilmundaero, Dong-gu,
Gwangju 61452, Republic of Korea
2
Multi-Material Research Center, Gwangju-Jeonnam Division, Korea Automotive Technology Institute,
Gwangju 62207, Republic of Korea
3
High Temperature Materials Group, Korea Institute of Materials Science, 797 Changwondaero,
Seongsan-gu, Changwon 51508, Republic of Korea
(received date: 4 May 2016 / accepted date: 17 May 2016)

The oxidation behavior of Ni-based superalloys was examined by cyclic oxidation tests at 850 and 1000 °C. The
present study focused on the investigation of two newly developed and three commercial alloys (GTD-111,
IN738LC, and CM247LC). The oxidation resistance of the superalloys were evaluated by the oxidation rate
constants and the mass gains. The oxidation scales mainly consisted of outer Cr2O3 and inner Al2O3 layers at
both temperatures, as well as oxides of minor elements such as TiO2, NiCr2O4, CrTaO4, HfO2, and TaO. The
oxidation resistance of the alloys containing larger amounts of Al, W, and Ta and lower Mo and Ti concentrations
was higher than those of the other alloys. The ranking of oxidation resistance for the alloys corresponded to the
statistical prediction drawn from a response surface analysis. Furthermore, these alloys contained more Ta
oxides, such as CrTaO4 and TaO, suggesting that Ta oxides had an active role in improving the oxidation
resistance.

Keywords: alloys, oxidation, X-ray dif- ments in Ni-based superalloys are Al and Cr. They form thin
fraction, Ni-based superalloys, cyclic oxi- and continuous Al2O3 and Cr2O3 oxide layers, respectively,
dation test on the surface of the alloy. The layers glow slowly and pro-
1. INTRODUCTION vide a good diffusion barrier for additive O atoms intruding
into the underlying alloy [8-11]. Notably, Cr2O3 provides an
Ni-based superalloys have high oxidation resistance and
effective barrier up to 870 °C. Above this temperature, Cr2O3
advanced mechanical properties at high temperature. Their
becomes unstable, and then Al2O3 becomes the dominant
primary practices in high-temperature industrial applications
protective layer [8,11].
include gas turbine blades, jet engines, aerospace compo-
Alloying elements other than Al and Cr may also play
nents, and chemical plants [1-3]. Among them, gas turbine
important roles in enhancing the oxidation resistance of Ni-based
blades must work at very high temperature (over 1000 °C)
superalloys. Klein et al. [12] reported that Si and B enhance
without degradation of oxidation resistance and mechanical
the oxide layer adhesion of Co-based superalloys, while Nb
properties. To overcome this difficulty, the blades are cast as
accelerates the oxidation. Yan et al. [13] reported that Cr, Fe,
a single crystal to prevent creep and fatigue fracture in the grain
and Si in Co-based superalloys at 800 °C improved the oxidation
boundary regions, and the surface is coated with ceramic
resistance, whereas V was detrimental. They also showed that
materials to retard oxidation [2-7]. Nevertheless, the devel-
Cr changed the mechanism of oxide scale formation, and a
opment of alloys with excellent oxidation resistance and mechan-
critical amount of both Al and Cr was necessary to form a pro-
ical properties at high temperature is still required, as the above
tective oxide scale. Several reactive alloying dopants such as
schemes cannot provide perfect protection.
Hf, Zr, Y, Dy, and La improve the oxidation resistance of
The composition of a superalloy is the main factor to con-
β-NiAl by single or co-doping, as the slow outward diffusion
trol its oxidation behavior. The most important alloying ele-
of the reactive element ions inhibits the diffusion of Al ions
*Corresponding author: heejin@chosun.ac.kr [14-18].
KIM and Springer There have been divergent opinions about the effect of Ta
790 Hee-Soo Kim et al.

on the superalloy oxidation behavior [5,12,15,19,20]. Yang the form of bars. The specimens used in the oxidation tests were
[20] reported that an addition of Ta as small as 1 at% (~3.0 wt%) 10 mm in diameter and 3 mm in thickness. The specimen
improved the oxidation resistance by promoting early estab- surfaces were finished with sand papers up to #600 grit to keep
lishment of a protective Al2O3 layer. When 3 at% (~8.8 wt%) the oxidation conditions uniform for all the samples. Each
of Ta was added, complex oxides such as NiMoO4, NiTa2O6, specimen was placed in an alumina crucible of 20 × 20 × 15 mm3
TaO2, MAl2O4, and M2CrO4 were formed, causing degrada- (length × width × height) for the oxidation tests. The crucibles
tion of the oxidation resistance. Klein et al. [12] showed that were baked for over 100 h at 1000 °C to remove humidity and
2 at% (~5.2 wt%) of Ta weakened the oxidation resistance of confirm that the weight remained stable in the test environ-
Co-based superalloys. Lu et al. [2] and Guo et al. [15] suggested ment.
that Ta slows down the outward diffusion of Al, suppressing The oxidation tests were performed at 850 and 1000 °C. A
the growth of oxide scales on a single-crystal Ni-based alloy single cycle of oxidation was composed of the following steps.
or β-NiAl. The present author [19] showed that small addi- (1) The initial specimen was put into a furnace previously
tions of Ta up to 5 wt% degraded the oxidation resistance at preheated at 400 °C; (2) the furnace was heated up to the target
850 °C, slowing down the formation of Al2O3 in Ni-based temperature (the heating rates were 3.75 and 5 °C/min at 850
superalloys. When the Ta content was greater than 5 wt%, or the and 1000 °C, respectively); (3) static oxidation was maintained
Al concentration was low, Ta enhanced the oxidation resistance for 15 h at the target temperature; (4) the furnace was cooled
of the alloys by lowering the O activity or restraining the down to ~700 °C at the rate of 1.3–2 °C/min by turning it off;
outward diffusion of Al in the scale. Remarkably, the oxida-
tion rate was reduced by adding ~3–6 wt% of Ta at 1000 °C.
We developed two new Ni-based superalloys and compared
their high-temperature oxidation resistance with those of
commercial alloys such as GTD-111, IN738LC, and CM247LC.
Oxidation tests were conducted in cycle at 850 and 1000 °C.
The oxide scales were examined by X-ray diffraction (XRD)
and scanning electron microscopy (SEM) along with energy
dispersive spectroscopy (EDS) to examine the effect of alloy
composition on scale formation.

2. EXPERIMENTAL PROCEDURES

The chemical compositions of the Ni-based superalloys


investigated here are listed in Table 1. The GTD-11, IN738LC,
and CM247LC alloys are commercially available. The alloys
named as Aa1 and B1 were newly designed for this study. They Fig. 1. Schematic thermal history of a single oxidation cycle used in
were prepared by using vacuum arc melting, and then cast in this study.

Table 1. Chemical compositions of Ni-based superalloys used in this study (wt%)


Alloying elements GTD-111 IN738LC CM247LC Aa1 B1
Co 9.5 8.5 9.3 12.95 4.74
Cr 14 16 8 9.64 16.09
Mo 1.5 1.8 0.5 < 0.01 0.63
W 4 2.5 9.5 8.79 5.3
Al 3 3.5 5.5 5.14 4.87
Ti 5 3.5 0.7 0.98 1.53
Ta 3 1.5 3.2 2.93 5.98
C 0.1 0.1 0.07 0.097 0.07
B 0.01 0.01 0.15 0.016 0.017
Nb 0 1 0 0 0
Hf 0 0 1.4 < 0.01 < 0.01
Zr 0 0.06 0.01 0.007 0.007
Fe 0 0 0 0.022 0.06
Si 0 0 0 0.03 0.03
Ni Bal. Bal. Bal. Bal. Bal.
High Temperature Oxidation Resistance of Ni-(5~13)Co-(10~16)Cr-(5~9)W-5Al-(1~1.5)Ti-(3~6)Ta Alloys 791

(5) the specimen was cooled down to ~400 °C at the rate of test can be a valid criterion to examine the oxidation of the
2.5–3 °C/min by keeping the furnace door open; (6) the specimen specimen. Figure 2 shows the mass gains of the specimens during
was removed from the furnace, and then (7) air cooled to room the oxidation tests at 850 and 1000 °C. For all specimens, the
temperature. The oxidation cycle was repeated twenty times. weights increased with time by oxidation. Notably, the square
A schematic thermal history of the single oxidation cycle is of the mass gain (Δm) per unit area is proportional to time
2 2 2
shown in Fig. 1. (t): Δm = kpt where kp (mg /cm ·s) is the oxidation rate con-
Between the cycles, the weight of the specimen, including stant, which is related to the diffusion of cations or anions across
its container and the spalled scale, was measured using a pre- the oxide layer [21,22]. In this study, every cycle had an
cision balance to estimate the mass gain due to oxidation. After identical thermal history for a designated target temperature,
the oxidation test, XRD was conducted to identify the phases and the thermal histories at 850 and 1000 °C had identical
composing the oxide scale. Cross-sections of the oxidized time durations at the oxidation temperatures. Therefore, a sin-
specimens were observed by using SEM with EDS. gle cycle of the oxidation test may be considered as a time
unit: Δm2 = kcnc where kc (mg2/cm2·cycle) is the oxidation rate
2
3. RESULTS AND DISCUSSION constant and nc is the number of cycles. The plots of Δm as a
function of the cycle at each temperature are shown in Fig. 3. The
When a metal is oxidized, O atoms intrude into the metal, symbols represent the experimental data, whereas the solid
form oxide films, and increase the mass of the matrix. The lines behind the symbols indicate the best-fitted lines. The fig-
evaluation of the mass gain occurred during the oxidation ure shows that the experimental results followed very well the

Fig. 2. Mass gains during oxidation tests of the Ni-based superalloys, as functions of the cycle at (a) 850 °C and (b) 1000 °C.

Fig. 3. Plots of squares of the mass gains as functions of the cycle at (a) 850 °C and (b) 1000 °C. The symbols indicate the experimental results,
while the solid straight lines represent the best-fitted lines.
792 Hee-Soo Kim et al.

differences between 850 and 1000 °C, although they slightly


increased with the temperature. The oxidations of GTD11 and
IN738LC were severe at 1000 °C, compared with those at
850 °C. At 850 °C, B1 showed the best oxidation resistance.
Its oxidation rate was lower than that of CM247LC, which
had the lowest oxidation rate among all the commercial alloys
investigated in this study. The oxidation rate of Aa1 was between
those of CM247LC and IN738LC. At 1000 °C, Aa1 exhibited
the best oxidation resistance, while B1 showed a good oxidation
resistance similar to that of Aa1.
The alloys with a low oxidation rate were characterized by
compositions comprising low quantities of Mo and Ti and high
amounts of Al, W, and Ta, compared with the other alloys, at
both temperatures. The alloys with high Cr concentration showed
excellent or good resistance to oxidation at 850 °C, but their
oxidation rates increased severely at 1000 °C, as Cr oxide
Fig. 4. Variation of oxidation rate constants, kc, with oxidation tem-
perature for the Ni-based superalloys tested in this study.
becomes volatile at temperatures higher than ~900 °C. The
microstructure also can affect the oxidation rate by promote
or inhibit diffusion, generally. However, the microstructure does
aforementioned kinetic equation. The slopes of the best-fitted not seem to have meaningful influence on the oxidation rate
lines indicate the oxidation rate constants kc. These constants, in this study because the alloys does not show remarkable
plotted at the oxidation temperatures for all the alloys inves- difference in the microstructure, as shown in Fig. 5, showing
tigated in this study, are shown in Fig. 4. very high density of γ’ precipitations with the size about 0.3
After 20 cycles of oxidation, the weights of the specimens ~ 0.5 μm. Therefore, the oxidation rate of the alloys are con-
2
increased approximately from 0.14171 to 0.63265 mg/cm at sidered to be determined predominantly by the composition
2
850 °C, and from 0.34988 to 4.11673 mg/cm at 1000 °C, as in this work.
shown in Fig. 2. These mass gains corresponded to oxidation The results of this work are in good agreement with the
−4 −2
rate constants ranging between 2.058 × 10 and 1.979 × 10 statistical study reported by Park et al. [19,23]. Their report
2 −4
mg/cm ·cycle at 850 °C, and between 4.955 × 10 and 8.495 × suggested that Mo and Ti promote oxidation, while Al and Ta
−1 2
10 mg/cm ·cycle at 1000 °C, as shown in Fig. 4. The oxidation significantly decrease the oxidation rate of Ni-based superal-
rates of CM247LC, Aa1, and B1 did not show substantial loys. Mo is harmful to oxidation resistance [24,29] owing to

Fig. 5. Scanning electron microscopy images of (a) GTD-111, (b) IN738LC, (c) CM247LC, (d) Aa1, and (e) B1 alloys.
High Temperature Oxidation Resistance of Ni-(5~13)Co-(10~16)Cr-(5~9)W-5Al-(1~1.5)Ti-(3~6)Ta Alloys 793

the volatility of its oxide; besides, Mo inhibits the formation ments, i.e., Cr in GTD-111, IN738LC, and B1, was close to, or
of Cr2O3 and promotes oxide spallation [26,27]. Previous works even above, the concentration boundaries of the response
showed that TiO2 is typically found in alloys with high oxida- surface analysis. Park et al. [23,34] used the Box-Behnken
tion rates [19,23]. Sato et al. [32] thermodynamically calcu- scheme for the response surface analysis to minimize the
lated the defect chemistry of oxide scales, and suggested that number of experimental runs; however, this might cause inac-
the addition of Al and Cr facilitates the formation of a con- curate results near the marginal values of the variables. Fur-
tinuous Al2O3 layer, whereas W, Ta, Ti, and Mo inhibit it. In thermore, IN738LC and CM247LC included Nb or Hf, which
addition, other researchers have proposed that Ta retards the were not considered in Park’s model, although researchers have
growth of Al by reducing the O activity [32,33] as well as reported that minor elements such as Nb and Hf have a con-
simultaneously suppressing the Al outward diffusion [4,15] siderable importance in oxidation resistance [13,18]. The
and O inward diffusion [32]. These reports imply that Ta may specimen microstructures may also have an impact; con-
be detrimental to oxidation resistance because it impairs the versely to the model alloys used in Park’s previous work, the
formation of a highly protective Al2O3 layer. However, Ta can alloys used in this study were directionally solidified.
improve the oxidation resistance by retarding the diffusion Note that Aa1 alloy exhibited a decreased oxidation rate
processes required for the oxidation to occur. In practice, the constant value measured at 1000 °C in comparison with that
effect of Ta on the oxidation rate seems to change depending of 850 °C. Aa1 has much a higher Co content than the other
on the oxidation temperature and concentrations of Al and Ta, alloys studied in this work. Park et al. [34] showed that Co
which determine the possibility to form a continuous Al2O3 has little effect on the oxidation resistance of the superalloy
layer [19]. W and Ti might have similar roles to Ta in oxidation. at 850 °C but seemed to improve the oxidation resistance at
Park et al. [23,34] reported the regression coefficients for 1000 °C, especially when it was added more than 7.5 wt%.
the mass gain of Ni-based superalloys after 20 cycles of oxidation Therefore Aa1 can be protected by the presence of Co more
in terms of the main alloying elements by using the response effectively at 1000 °C that it is at 850 °C.
surface analysis for the effects of Co (~0–15 wt%), Cr (~8– The XRD spectra of the oxidized layers of the alloys are
15 wt%), Mo (~0–5 wt%), W (~0–10 wt%), Al (~3–8 wt%), Ti shown in Fig. 7. Various types of oxides were formed during
(~0–5 wt%), and Ta (~0–10 wt%). According to their work, Al, the oxidation tests. At both temperatures of 850 and 1000 °C,
W, Cr, and Ta reduced the mass gain of the alloy at 850 °C. the spectra included peaks of NiO, Ni3Al, Al2O3, NiCr2O4,
At 1000 °C, Al, W, Co, and Ta improved the oxidation resis- Cr2O3, CrTaO4, and TiO2. At 1000 °C, GTD-111 exhibited more
tance, while Cr and Ti had a detrimental effect. Figure 6 shows peaks of Cr2O3 and TiO2 than any other alloy.
the mass gains from the present work and the calculation for Figure 8 shows the cross-sectional SEM images of the
the compositions of the alloys in this study. Although there alloys after 20 cycles of oxidation at 850 °C. Notably, the oxide
were some differences between the present results and the scales consisted of Cr oxide layers, such as Cr2O3 and NiCr2O4,
estimated values, the overall oxidation tendencies were similar. near the surface, and Al2O3 layers underneath the Cr oxide
The differences between the two sets of data may be caused layers. For most alloys, Al2O3 was in the form of discontinu-
by several reasons. The concentration of some alloying ele- ous agglomerates elongated in a direction perpendicular to

Fig. 6. Comparison of mass gains after 20 cycles of oxidation at (a) 850 °C and (b) 1000 °C with the data estimated by response surface analysis.
The line segments between the data points were drawn for ease of comparison.
794 Hee-Soo Kim et al.

Fig. 7. X-ray diffraction spectra of Ni-based superalloys after 20 cycles of oxidation at (a) 850 °C and (b) 1000 °C.

Fig. 8. Cross-sectional scanning electron microscopy images of (a) GTD-111, (b) IN738LC, (c) CM247LC, (d) Aa1, and (e) B1 alloys after 20
cycles of oxidation at 850 °C.

the surface. In the case of Aa1, a partially continuous Al2O3 CM247LC) and TaO (in Aa1) were also found. A discontin-
layer was also found. GTD-111 exhibited a TiO2 layer near the uous Al2O3 layer may have been the cause of the high oxida-
surface, while CM247LC had some NiO. CrTaO4 was found tion rate of GTD-111. In the case of IN738LC, Al2O3 existed
in the depths, similar to NiCr2O4, in CM247LC, Aa1, and B1, as a thin layer or was in a mixed form with Cr2O3. The alloys
which exhibited slower oxidation rates than the other alloys, with a continuous Al2O3 layer, i.e., CM247LC, Aa1, and B1,
as shown in Figs. 2 and 3. Probably, CrTaO4 enhanced the exhibited lower oxidation rates, as shown in Figs. 2–4.
oxidation resistance of the alloys. Giggins and Pettit [35] suggested the oxidation mechanism
Figures 9 and 10 show the cross-sectional SEM images of of Ni-Cr-Al alloys as follows. The first oxide layer on the sur-
the alloys after 20 cycles of oxidation at 1000 °C. The overall face of the alloy is predominantly made of Ni(Cr,Al)2O4 and
cross-sectional structures were analogous to those observed NiO. The formation of Cr2O3 and Al2O3 is controlled by the
for the samples oxidized at 850 °C. The Al2O3 layer was con- diffusion of Cr, Al, and O. As the activity of O required to
tinuous, except in the case of GTD-111. Traces of HfO2 (in oxidize Cr is smaller than that required for Ni(Cr,Al)2O4 and
High Temperature Oxidation Resistance of Ni-(5~13)Co-(10~16)Cr-(5~9)W-5Al-(1~1.5)Ti-(3~6)Ta Alloys 795

Fig. 9. Cross-sectional scanning electron microscopy images of (a) GTD-111, (b) IN738LC, and (c) B1 alloys after 20 cycles of oxidation at 1000 °C.

Fig. 10. Cross-sectional scanning electron microscopy images of (a, b) CM247LC and (c, d) Aa1 alloys after 20 cycles of oxidation at 1000 °C.

NiO, O diffuses into the matrix and form Cr2O3 below the In this study, TiO2 was found in GTD-111, and a mixture of
surface layer. Moreover, the O activity to form Al2O3 is even Cr2O3 and TiO2 existed in B1 at 1000 °C, although all the alloys
smaller than that for Cr2O3, and Al2O3 precipitates beneath used in this work had Ti as an alloying element. Seal et al. [38]
Cr2O3. For all alloys in this study, Cr2O3 was found near the reported that TiO2 did not form on the surface after a short-time
surface, and Al2O3 existed below the Cr2O3 layer at both 850 oxidation (~10 h) at 900 °C, possibly owing to its low activity
and 1000 °C, as shown in Figs. 8–10. However, NiCr2O4 in the IN-738LC alloy, compared with that of Cr and Ni. A long
appeared on the surface, although only in CM247LC and Aa1 time, over 100 h, was necessary to form TiO2. Similarly, in this
at 1000 °C. NiO was found in CM247LC and Aa1 at 850 °C study, TiO2 was expected to form after more cycles of oxidation
only. In all other cases, a little amount of NiO or NiCr2O4 was also in alloys other than GTD-111 and B1.
found as a mixed form between the Cr2O3 and Al2O3 layers. In CM247LC, Aa1, and B1, Ta oxides such as CrTaO4 and
Most of NiO seemed to be spalled off during the oxidation TaO were observed. Although GTD-111 and IN738LC had Ta
test, owing to the stress arisen from the large mismatch between as an alloying element, no Ta oxide was found. GTD-111 and
the thermal expansion coefficients of NiO and Cr2O3, which IN738LC had a higher Ti content than the other alloys. Ti likely
was formed just underneath the NiO layer [36]. impeded the formation of Ta oxides. This was consistent with
TiO2 is known to typically form a layer on the surface [4,37,38]. the fact that Ti weakens the effect of Ta at 850 °C, as previously
796 Hee-Soo Kim et al.

suggested by the response surface analysis [23,34]. As the 9. T. S. Jo, S.-H. Kim, D.-G. Kim, J. Y. Park, and Y. D. Kim,
oxidation of GTD-111 and IN738LC became severe at 1000 °C, Met. Mater. Int. 14, 739 (2008).
Ta oxides likely played a great role in improving the oxida- 10. S. Sharma, F. Li, G. Ko, and K. Kang, J. Nucl. Mater. 405,
tion resistance of the alloys. Park et al. [23] reported that the 165 (2010).
effect of Ta addition on oxidation resistance was clearer at 11. C. Tedmon, J. Electrochem. Soc. 113, 766 (1966).
1000 °C than at 850 °C. Guo et al. [15] pointed out that Ta atoms 12. L. Klein, A. Bauer, S. Neumeier, M. Göken, and S. Virtanen,
are large, and thus their diffusion is slow. Ta oxides may form Corros. Sci. 53, 2027 (2011).
more easily at 1000 °C than at 850 °C, likely enhancing the 13. H. Y. Yan, V. A. Vorontsov, and D. Dye, Corros. Sci. 83,
oxidation resistance of the alloys at 1000 °C. 382 (2014).
14. H. Guo, D. Li, L. Zheng, and S. Gong, Corros. Sci. 88, 197
4. CONCLUSIONS (2014).
15. H. Guo, D. Wang, H. Peng, S. Gong, and H. Xu, Corros. Sci.
The oxidation resistance of Ni-based superalloys was eval- 78, 369 (2014).
uated by cyclic oxidation tests at 850 and 1000 °C. The alloys 16. H. Guo, T. Zhang, S. Wang, and S. Gong, Corros. Sci. 53,
used in this study included the commercially available GTD-111, 2228 (2011).
IN738LC, and CM247LC alloys. Two newly designed superal- 17. D. Li, H. Guo, D. Wang, T. Zhang, S. Gong, and H. Xu,
Corros. Sci. 66, 125 (2013).
loys were also investigated. The oxidation scales mainly con-
18. K. Yan, H. Guo, and S. Gong, Corros. Sci. 83, 335 (2014).
sisted of outer Cr2O3 and inner Al2O3 layers at both 850 and
19. S.-J. Park, S.-M. Seo, Y.-S. Yoo, H.-W. Jeong, and H. Jang,
1000 °C. The oxidation of GTD-111 and IN738LC became
Corros. Sci. 90, 305 (2015).
severe at 1000 °C; the two alloys did not contain Ta oxides
20. S. Yang, Oxid. Met. 15, 375 (1981).
after the tests. Presumably, Ta oxides improved the oxidation
21. D. Monceau and B. Pieraggi, Oxid. Met. 50, 477 (1998).
resistance of the other alloys. The ranking of oxidation resis-
22. J.-S. Oh, M.-C. Shim, M.-H. Park, and K.-A. Lee, Met.
tance for the superalloys corresponded to the statistical predic- Mater. Int. 20, 915 (2014).
tion drawn from a response surface analysis, suggesting that 23. S.-J. Park, S.-M. Seo, Y.-S. Yoo, H.-W. Jeong, and H. Jang,
the presence of Al, W, Co, and Ta remarkably improved the J. Nanomater 2015, 929546 (2015).
oxidation resistance. 24. C.-Y. Bai, Corros. Sci. 49, 3889 (2007).
25. N. Birks, G. H. Meier, and F. S. Pettit, Introduction to the
ACKNOWLEDGEMENTS High Temperature Oxidation of Metals, Cambridge Univer-
sity Press (2006).
This work was supported by a grant from the Fundamental 26. N. Hussain, K. A. Shahid, I. H. Khan, and S. Rahman, Oxid.
R&D Program for Core Technology of Materials (10041233) Met. 41, 251 (1994).
funded by the Ministry of Trade, Industry and Energy of the 27. F. A. Khalid, N. Hussain, and K. A. Shahid, Mater. Sci. Eng.
Republic of Korea. A 265, 87 (1999).
28. P. Kofstad, High Temperature Corrosion, Elsevier Applied
REFERENCES Science, London, UK (1988).
29. Q. Yang, W. Xiong, S. Li, H. Dai, and J. Li, J. Alloys Compd.
1. M. H. Li, X. F. Sun, J. G. Li, Z. Y. Zhang, T. Jin, and H. R. 506, 461 (2010).
Guan, Oxid. Met. 59, 591 (2003). 30. M. E. El-Dahshan, D. P. Whittle, and J. Stringer, Corros.
2. C. T. Liu, J. Ma, X. F. Sun, and P. C. Zhao, Surf. Coat. Sci. 16, 83 (1976).
Technol. 204, 3641 (2010). 31. S. Espevik, R. A. Rapp, P. L. Daniel, and J. P. Hirth, Oxid.
3. Y. Yuan, Z. H. Zhong, Z. S. Yu, H. F. Yin, Y. Y. Dang, X. B. Met. 14, 85 (1980).
Zhao, Z. Yang, J. T. Lu, J. B. Yan, and Y. Gu, Met. Mater. 32. A. Sato, Y. L. Chiu, and R. C. Reed, Acta Mater. 59, 225
Int. 21, 659 (2015). (2011).
4. S. Cruchley, H. E. Evans, M. P. Taylor, M. C. Hardy, and S. 33. G. N. Irving, J. Stringer, and D. P. Whittle, Corros. Sci. 15,
Stekovic, Corros. Sci. 75, 58 (2013). 337 (1975).
5. X. Lu, S. Tian, X. Yu, and C. Wang, Rare Metals 30, 439 34. S.-J. Park, Master’s Thesis, Chosun Univerisity, Korea (2005).
(2011). 35. C. S. Giggins and F. S. Pettit, J. Electrochem. Soc. 118, 1782
6. J. G. Yoon, S. I. Kwon, J. W. Lee, X. P. Tan, B. G. Choi, I. (1971).
S. Kim, C. Y. Jo, and H. U. Hong, Korean. J. Met. Mater. 36. D. L. Deadmore and C. E. Lowell, Oxid. Met. 11, 91 (1977).
52, 397 (2014). 37. J. Litz, A. Rahmel, M. Schorr, and J. Weiss, Oxid. Met. 32,
7. S. W. Nam, Korean. J. Met. Mater. 53, 833 (2015). 167 (1989).
8. D. Caplan and M. Cohen, J. Electrochem. Soc. 108, 438 38. S. Seal, S. C. Kuiry, and L. A. Bracho, Oxid. Met. 57, 297
(1961). (2002).

You might also like