You are on page 1of 5

Available online at www.sciencedirect.

com

Thin Solid Films 517 (2008) 1301 1305


www.elsevier.com/locate/tsf

Enhancing pitting corrosion resistance of AlxCrFe1.5MnNi0.5 high-entropy


alloys by anodic treatment in sulfuric acid
C.P. Lee a , Y.Y. Chen a , C.Y. Hsu a , J.W. Yeh a , H.C. Shih a,b,
a

Department of Materials Science and Engineering, National Tsing Hua University, Hsinchu 300, Taiwan
b
Institute of Materials Science and Nanotechnology, Chinese Culture University, Taipei 111, Taiwan
Available online 11 June 2008

Abstract
High-entropy alloys are a newly developed family of multi-component alloys that comprise various major alloying elements. Each element in
the alloy system is present in between 5 and 35 at.%. The crystal structures and physical properties of high-entropy alloys differ completely from
those of conventional alloys. The electrochemical impedance spectra (EIS) of the AlxCrFe1.5MnNi0.5 (x = 0, 0.3, 0.5) alloys, obtained in 0.1 M HCl
solution, clearly revealed that the corrosion resistance values were determined to increase from 21 to 34 cm2 as the aluminum content increased
from 0 to 0.5 mol, and were markedly lower than that of 304 stainless steel (243 cm2). At passive potential, the corresponding current declined
with the anodizing time accounting, causing passivity by the growth of the multi-component anodized film in H2SO4 solution. X-ray
photoelectron spectroscopy (XPS) analyses revealed that the surface of anodized Al0.3CrFe1.5MnNi0.5 alloy formed aluminum and chromium
oxide film which was the main passivating compound on the alloy. This anodic treatment increased the corrosion resistance in the EIS
measurements of the CrFe1.5MnNi0.5 and Al0.3CrFe1.5MnNi0.5 alloys by two orders of magnitude. Accordingly, the anodic treatment of the
AlxCrFe1.5MnNi0.5 alloys optimized their surface structures and minimized their susceptibility to pitting corrosion.
2008 Elsevier B.V. All rights reserved.
Keywords: High-entropy alloy; Anodic film; EIS; XPS; Passivity

1. Introduction
This investigation presents a novel multi-component alloy
that breaks away from the traditional alloy design conventions.
Multi-component alloys comprise five or more principal
elements, each at 5 to 35 at.%. This multiple element system
has been called a high-entropy alloy (HEA) because its
configurational entropy (Sconf = 1.95R 2.2R) exceeds that
of ordinary alloy (1.1R) [1]. Studies have demonstrated that
HEAs had simple solidsolution structures, were easy to
amorphize and nanoprecipitate [2,3]. Furthermore, the HEAs
exhibited favorable mechanical properties, such as high

Corresponding author. Department of Materials Science and Engineering,


National Tsing Hua University, Hsinchu 300, Taiwan. Tel.: +886 3 5715131;
fax: +886 3 5710290.
E-mail address: hcshih@mx.nthu.edu.tw (H.C. Shih).
0040-6090/$ - see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.tsf.2008.06.014

hardness [Vickers hardness (HV) = 700] [4] and excellent


resistance to anneal softening [5], wear [1.21.8 m/mm3] [6]
and oxidation [7]. Additionally, coating technology was further
expending their range of application, to include the hard-facing
of golf heads and rollers [8], diffusion barriers [9] and soft
magnetic films [10].
The HEAs (Al0.5CoCrCuFeNi), evaluated previously, were
more resistant to general corrosion than was type 304 stainless
steel, in H2SO4 solution at room temperature [11]. However, the
resistance of HEAs to pitting corrosion in a Cl environment
was inferior to that of the 304 stainless steel, as revealed by the
lower pitting potential and the narrower passive region [12]. In
the authors' recent study, the AlxCrFe1.5MnNi0.5 alloys were
anodized in H2SO4 solution to optimize its surface structure and
minimize its susceptibility to pitting corrosion. The corrosion
performance of the anodized AlxCrFe1.5MnNi0.5 alloys in
chloride environments were elucidated using electrochemical
impedance spectroscopy (EIS) technique and X-ray photoelectron spectroscopy (XPS).

1302

C.P. Lee et al. / Thin Solid Films 517 (2008) 13011305

2. Experimental
2.1. Test materials
The HEAs samples prepared were non-equal-mole, multicomponent alloys of AlxCrFe1.5MnNi0.5 with various aluminum
contents, with x from 0 to 0.5. Samples were made by a standard
arc-melting technique. Table 1 presents the chemical compositions
of the AlxCrFe1.5MnNi0.5 alloys in weight percentage. The method
followed the same procedure as described in Ref. [12]. The
AlxCrFe1.5MnNi0.5 alloys were used as the substrate specimens,
and cleaned ultrasonically in acetone and distilled water before use.
The specimens were first held at the free corrosion potential for 60 s
before the potential was one-step switched to the desired value in
the passive region and maintained there for 1800 s. Following the
anodic treatments, each specimen was cleaned in distilled water,
and then dried in nitrogen. Immediately thereafter, the morphology and chemical structure of the anodized surface of the
AlxCrFe1.5MnNi0.5 alloy was investigated under a scanning
electron microscope (SEM, JEOL-5410) and X-ray photoelectron
spectroscopy (XPS, Perkin-Elmer model PHI 1600). A Mg K
source was used for XPS and operated at 250 W. Energy
calibration was done using the Au 4f7/2 peak at 83.8 eV, and the
energy resolution was 0.2 eV for the core-level spectra. The
following XPS regions were examined for Al 2p and Cr 2p.

Fig. 1. Effect of aluminum on potentiodynamic polarization curves of the


AlxCrFe1.5MnNi0.5 (x = 0, 0.3, and 0.5) alloys in 15% H2SO4 solution.

which was measured in 15% H2SO4 solution. The potentiodynamic polarization curve exhibited an active-passive corrosion
behavior, and the AlxCrFe1.5MnNi0.5 alloys yielded a wide
passive region (E N 1100 mV). A stable passive state stretched
roughly from 0.2 to 1.1 V. Moreover, as the aluminum content
in the AlxCrFe1.5MnNi0.5 alloys increased from 0 to 0.5 mol, the
passive current density (ipass) increased from 1.28 10 5 to
1.63 10 4 A/cm2.

2.2. Electrochemical measurements


3.2. Chronoamperometric measurements
Both potentiodynamic polarization and EIS were performed in a
typical three-electrode cell that consisted of a specimen as the
working electrode, an Ag/AgCl in 3 M KCl (E = 208 mVSHE) as the
reference electrode, and a Pt sheet (4 cm2) as the counter electrode.
All of the potentials in this paper were presented on the standard
hydrogen scale (SHE). Polarization curves were measured
potentiodynamically by sweeping the potential in the positive
direction at a sweep rate of 1 mVs 1 from an initial potential of
0.5 V to a final potential of 1.5 V relative to the OCP. The potential
was controlled and the current was measured using a potentiostat
(AUTOLAB PGSTAT30). EIS was carried out at the OCP with a
sinusoidal potential amplitude of 10 mV running from 10 kHz to
10 mHz, and using an AUTOLAB PGSTAT30/FRA system from
ECO CHEMIE.

The passivation procedure was performed in 15% H2SO4


solution for a period of 1800 s. Fig. 2 plots the chronoamperometric
curves for the AlxCrFe1.5MnNi0.5 alloys polarized at potentials of
0.7 V. This anodizing was conducted in the passive region, where
neither transpassive dissolution of AlxCrFe1.5MnNi0.5 alloys nor
oxygen evolution were thermodynamically possible, as determined
from the potentiodynamic polarization curve (Fig. 1). The
continuously decaying current density described passivation of
the CrFe1.5MnNi0.5 and Al0.3CrFe1.5MnNi0.5 alloys by the growth

3. Results and discussion


3.1. Potentiodynamic polarization curves
Fig. 1 displays the potentiodynamic polarization behavior of
the AlxCrFe1.5MnNi0.5 alloys with various aluminum contents,
Table 1
Chemical composition (wt.%) of the AlxCrFe1.5MnNi0.5 alloys
Element

Al(26.96)
(%)

Cr(51.99)
(%)

Fe(55.84)
(%)

Mn(54.94)
(%)

Ni(58.69)
(%)

CrFe1.5MnNi0.5
Al0.3CrFe1.5MnNi0.5
Al0.5CrFe1.5MnNi0.5

0.00
3.55
5.78

23.63
22.79
22.26

38.03
36.68
35.83

24.99
24.10
23.54

13.33
12.86
12.56

Fig. 2. Chronoamperometric curves for the AlxCrFe1.5MnNi0.5 (x = 0, 0.3, and 0.5)


alloys at a constant potential of 0.7 VSHE in 15% H2SO4 solution.

C.P. Lee et al. / Thin Solid Films 517 (2008) 13011305

1303

3.3. SEM photomicrographs of anodized surface


Fig. 3 presents the SEM microstructures of the anodized
AlxCrFe1.5MnNi0.5 alloys following anodic treatment at 0.7 V
in 15% H2SO4. Localized dissolution was observed on the
surface of the anodically passivated CrFe1.5MnNi0.5 alloy, as
presented in Fig. 3a. The surface of anodically treated
Al0.3CrFe1.5MnNi0.5 alloy has been formed an oxide film, as
displayed in Fig. 3b. Fig. 3c shows the selective dissolution of
the anodized Al0.5CrFe1.5MnNi0.5 alloy, which was larger and
deeper in size than the anodized CrFe1.5MnNi0.5 alloy. These
dissolution processes on the surface depleted some of the oxide
constituents, such as iron oxide, reducing the relative amount
of chromium, which was the main element involved in
passivation [15].
3.4. XPS analysis of anodized surface
The chemical composition of the AlxCrFe1.5MnNi0.5 alloys
(x = 0 and 0.3 mole) was investigated by XPS. Fig. 4 shows the
Al (2p) core-level spectra of the Al0.3CrFe1.5MnNi0.5 alloy
following anodic treatment in 15% H2SO4 solution. The
spectrum of the Al 2p was fitted with two peaks for the
anodically-formed oxides under the Al0.3CrFe1.5MnNi0.5 alloy,
located at 72.8 eV for aluminum metal and 74.7 eV for
aluminum oxide structure (Al2O3) [16,17]. The intensity and
peak area of the right shoulder peak corresponding to Al metal
was much lower than that of the Al2O3. Therefore, the surface of
anodically treated Al0.3CrFe1.5MnNi0.5 alloy has been formed
an aluminum oxide film. Frangini et al., studied on FeAl
intermetallic alloy containing 28 wt.% Al, which has shown that
the presence of Al is beneficial both in improving the ability to
passivate iron in sulfuric acid [18] and in increasing the
resistance to chloride breakdown [19].
Fig. 5 shows the Cr (2p) core-level spectra of the
CrFe1.5MnNi0.5 and Al0.3CrFe1.5MnNi0.5 alloy following
anodic treatment in 15% H2SO4 solution. Both spectra of the
CrFe1.5MnNi0.5 and Al0.3CrFe1.5MnNi0.5 alloy could be fitted
with four peaks at 574.4 (2p3/2) and 583.6 eV (2p1/2) for Cr
Fig. 3. SEM images of (a) CrFe1.5MnNi0.5, (b) Al0.3CrFe1.5MnNi0.5 and (c)
Al0.5CrFe1.5MnNi0.5 alloys following anodic treatment at a potential of 0.7 VSHE in
15% H2SO4 solution.

of a stable oxide film. The passive current density of the


Al0.5CrFe1.5MnNi0.5 alloy was significantly higher than that of
the CrFe1.5MnNi0.5 and Al0.3CrFe1.5MnNi0.5 alloys, which was
related to the surface dissolution of the Al0.5CrFe1.5MnNi0.5 alloy. It
can be noted that a stable passive state of the Al0.3CrFe1.5MnNi0.5
alloy was established after 500 s in 15% H2SO4 solution. The
smoothness of the curve for the Al0.3CrFe1.5MnNi0.5 alloy implied
that the passive films on their surfaces were very stable, with no
breakdown throughout the period of the passivation treatment [13].
However, the CrFe1.5MnNi0.5 alloy showed current transients that
may be expected if any breakdown in passivation occurs in the
electrolyte [14].

Fig. 4. XPS spectra for Al 2p of the Al0.3CrFe1.5MnNi0.5 alloy following anodic


treatment at a potential of 0.7 VSHE in 15% H2SO4 solution.

1304

C.P. Lee et al. / Thin Solid Films 517 (2008) 13011305

metal, 576.9 (2p3/2) and 586.7 eV (2p1/2) for chromium oxide


(Cr2O3) [15,16], respectively. However, the intensity and peak
area for Cr2O3 in the anodically-passive oxides of CrFe1.5MnNi0.5
alloy was much lower than that of Al0.3CrFe1.5MnNi0.5 alloy. In
reverse, the content of Cr metal of anodized CrFe1.5MnNi0.5 alloy
was more than twice of anodized Al0.3CrFe1.5MnNi0.5 alloy. In
1911, Monnartz [20] first noted the outstanding passivity of
stainless steel that contained at least 12% Cr. The relevant studies
[21] also indicated that Cr2O3 is responsible for maintaining
passivity of the stainless steel. Hence, the result of XPS analysis
may suggest that the anodized Al0.3CrFe1.5MnNi0.5 alloy was
more resistant to corrosion than was the anodized CrFe1.5MnNi0.5
alloy.
3.5. Electrochemical impedance of the AlxCrFe1.5MnNi0.5
alloys
Fig. 6 presents the effect of aluminum on the Nyquist plot of
the un-anodized AlxCrFe1.5MnNi0.5 (x = 0, 0.3, and 0.5) alloys
and 304 stainless steel in 0.1 M HCl solution. The charge
transfer resistance (Rt) value was 21 cm2 for the un-anodized
CrFe1.5MnNi0.5 alloy, and increased to 31 and 34 cm2 for the
alloys containing 0.3 and 0.5 mol of aluminum respectively.

Fig. 5. XPS spectra for Cr 2p of (a) CrFe1.5MnNi0.5 and (b) Al0.3CrFe1.5MnNi0.5


alloy following anodic treatment at a potential of 0.7 VSHE in 15% H2SO4 solution.

Fig. 6. Nyquist plots of the un-anodized AlxCrFe1.5MnNi0.5 (x=0, 0.3, and 0.5) alloys
and 304 stainless steel in 0.1 M HCl.

The Rt of the 304 stainless steel (243 cm2) exceeded those of


the AlxCrFe1.5MnNi0.5 alloy. The Nyquist plot of the Al-free
CrFe1.5MnNi0.5 alloy was characterized by a depressed
capacitive semicircle from high to medium frequencies and an
inductive loop at low frequencies. The inductive loop at low
frequencies may represent the commencement of the incubation
period for pitting corrosion [22]. The Al0.3CrFe1.5MnNi0.5 and
Al0.5CrFe1.5MnNi0.5 alloys with high aluminum content
exhibited a different behavior, in that the EIS diagrams
comprised two capacitive loops, which were related to the
presence of a corrosion product and charge transfer across the
metal and electrolyte interface [23].
The electrochemical and corrosion properties of the anodic
passive films were evaluated from measurements of electrochemical impedance. Fig. 7 displays the Nyquist plot of the anodized
AlxCrFe1.5MnNi0.5 (x =0, 0.3, and 0.5) alloys in 0.1 M HCl solution
after anodic treatment at a potential of 0.7 V in 15% H2SO4
solution. The Nyquist plot of the anodized CrFe1.5MnNi0.5 alloy

Fig. 7. Nyquist plots of anodized AlxCrFe1.5MnNi0.5 (x=0, 0.3, and 0.5) alloys in
0.1 M HCl solution following anodic treatment at a potential of 0.7 VSHE in 15%
H2SO4 solution.

C.P. Lee et al. / Thin Solid Films 517 (2008) 13011305

consisted of a capacitive arc from high to medium frequencies and a


slight inductive loop at low frequencies. The capacitive arc of the
anodically treated CrFe1.5MnNi0.5 alloy was significantly larger
than that of the untreated alloy, because of the formation of a
passive film on the alloy. The capacitive loop was related to the
capacitance of the oxide film and the presence of the inductive loop
revealed that the surface area was partially or totally active [24]. The
EIS measurements of the anodized Al0.3CrFe1.5MnNi0.5 alloy
demonstrated that the Nyquist plot comprised two capacitance arcs,
the second of which had a much larger diameter than that of the unanodized Al0.3CrFe1.5MnNi0.5 alloy, because of the gradual
formation of a protective scale on the Al0.3CrFe1.5MnNi0.5 alloy
[25]. Therefore, the anodic treatment increased the corrosion
resistance in the EIS measurements of the CrFe1.5MnNi0.5 and
Al0.3CrFe1.5MnNi0.5 alloys by two orders of magnitude in the Cl
containing environments.
4. Conclusions
The current transient of the chronoamperometric curve
indicates that the CrFe1.5MnNi0.5 alloy is susceptible to
localized corrosion throughout the period of the passivation
treatment. In contrast, the smoothness of the curve for the
Al0.3CrFe1.5MnNi0.5 alloy implies that a passive film on the
surface is very stable. SEM and XPS analyses reveal that the
surface of anodically treated Al0.3CrFe1.5MnNi0.5 alloy is
covered with a passive oxide film, which effectively protects
against pitting corrosion.
Electrochemical impedance spectroscopy demonstrates that
the impedance of the un-anodized AlxCrFe1.5MnNi0.5 alloys
increases with the alloying aluminum content, and is significantly lower than that of the 304 stainless steel. The anodized
CrFe1.5MnNi0.5 and Al0.3CrFe1.5MnNi0.5 alloys have higher
corrosion resistance than the untreated ones, indicating that the
anodic treatment in 15% H2SO4 solution effectively minimizes
susceptibility to the pitting corrosion.
Acknowledgments
The authors would like to thank the National Science
Council of the Republic of China for the financial support of
this research under the contract of NSC 95-2221-E-007-098.

1305

Reference
[1] R.A. Swalin, Thermodynamics of Solids, 2nd ed., Wiley, New York, 1991,
p. 21.
[2] J.W. Yeh, S.K. Chen, J.Y. Gan, S.J. Lin, T.S. Chin, T.T. Shun, C.H. Tsau, S.
Y. Chang, Metall. Mater. Trans. A 35A (2004) 2533.
[3] J.W. Yeh, S.K. Chen, S.J. Lin, J.Y. Gan, T.S. Chin, T.T. Shun, C.H. Tsau, S.Y.
Chang, Adv. Eng. Mater. 6 (2004) 299.
[4] C.J. Tong, M.R. Chen, S.K. Chen, J.W. Yeh, T.T. Shun, S.J. Lin, S.Y.
Chang, Metall. Mater. Trans. A 36A (2005) 1263.
[5] H.Y. Chen, C.W. Tsai, C.C. Tung, J.W. Yeh, T.T. Shun, C.C. Yang, S.K.
Chen, Ann. Chim.-Sci. Mat. 31 (2006) 685.
[6] C.Y. Hsu, J.W. Yeh, S.K. Chen, T.T. Shun, Metall. Mater. Trans. A 35A
(2004) 1465.
[7] P.K. Huang, J.W. Yeh, T.T. Shun, S.K. Chen, Adv. Eng. Mater. 6 (2004) 74.
[8] Y.S. Huang, L. Chen, H.W. Lui, M.H. Cai, J.W. Yeh, Mater. Sci. Eng. A
457 (2007) 77.
[9] T.K. Chen, T.T. Shun, J.W. Yeh, M.S. Wong, Surf. Coat. Technol. 188189
(2004) 193.
[10] H.F. Kuo, W. Chin, T.W. Cheng, W.K. Hsu, J.W. Yeh, Appl. Phys. Lett. 89
(2006) 182503.
[11] C.P. Lee, Y.Y. Chen, C.Y. Hsu, J.W. Yeh, H.C. Shih, J. Electrochem. Soc.
154 (2007) C424.
[12] Y.Y. Chen, T. Duval, U.D. Hung, J.W. Yeh, H.C. Shih, Corros. Sci. 47
(2005) 2257.
[13] J.E. Castle, J.H. Qiu, J. Electrochem. Soc. 137 (1990) 2031.
[14] G.T. Burstein, R.M. Souto, Electrochim. Acta 40 (1995) 1881.
[15] A.A. Hermas, M. Nakayama, K. Ogura, Electrochim. Acta 50 (2005) 3640.
[16] J.F. Moulder, Handbook of x-ray photoelectron spectroscopy: a reference
book of standard spectra for identification and interpretation of XPS data,
Perkin-Elmer, Eden Prairie, MN, 1995.
[17] V.S. Rao, V. Raja, Corrosion 59 (2003) 575.
[18] S. Frangini, N.D. Cristofaro, J. Lascovich, A. Mignone, Corros. Sci. 35
(1993) 153.
[19] N. De Cristofaro, S. Frangini, A. Mignone, Corros. Sci. 38 (1996) 307.
[20] P. Monnartz, Metallurgie 8 (1911) 161.
[21] I. Olefjord, B. Brox, in: M. Froment (Ed.), Passivity of Metals and
Semiconductors, Elsevier Science Publishers B.V., Amsterdam, 1983, p. 561.
[22] T. Zhang, C.L. Zeng, Electrochim. Acta 50 (2005) 4721.
[23] A. Aballe, M. Bethencourt, F.J. Botana, J. Cano, M. Marcos, Corros. Rev.
18 (2000) 1.
[24] L.F. Li, M. Daerden, P. Caenen, J.P. Celis, J. Electrochem. Soc. 153 (2006)
B145.
[25] J. Masalski, J. Gluszek, J. Zabrzeski, K. Nitsch, P. Gluszek, Thin Solid
Films 349 (1999) 186.

You might also like