Vous êtes sur la page 1sur 6

Journal of Alloys and Compounds 364 (2004) 193198

XPS study of cerium conversion coating on the


anodized 2024 aluminum alloy
Xingwen Yu a, , Guoqiang Li b
a

Department of Interface Chemistry and Surface Engineering, Max Plank Institute for Iron Research, D-40237 Dusseldorf, Germany
b Department of Materials Science and Engineering, Beijing University of Science and Technology, 100083 Beijing, China
Received 3 February 2003; received in revised form 17 March 2003; accepted 7 April 2003

Abstract
Cerium-rich conversion coating was deposited on anodized aluminum alloy 2024 in a solution containing Ce(NO3 )3 . X-ray photoelectron
spectroscopy (XPS) was used as the analysis method. The composition of the Ce conversion coating deposited on the anodized 2024 alloy
was investigated using this method. It was revealed that the coating predominately consisted of three-valent state cerium compound. Some of
the CeIII was oxidized to CeIV in the outer layer coating.
2003 Elsevier B.V. All rights reserved.
Keywords: Ce-rich conversion coating; Aluminum alloy; X-ray photoelectron spectroscopy (XPS)

1. Introduction
For almost 100 years, chromate compounds (Cr6+ ) have
been used as very effective and inexpensive corrosion inhibitors for many alloys systems, including aluminum, zinc
and steel, in a wide range of aqueous environments. However, the recent recognition that chromate are both highly
toxic and carcinogenic has led to extensive worldwide research to develop effective alternative inhibitors.
The idea of cerium conversion coatings as a protective
coating on aluminum alloys was proposed by Hinton et al.
[1] in the mid-1980s. They discovered that additions of
cerium ions to sodium chloride solutions significantly reduced the corrosion rate of 7075 aluminum alloy [1]. The
result inspired a great interest for them and other experts in
investigating the formation processes, formation mechanism
and characterization of cerium conversion coatings for aluminum alloys. These preliminary studies revealed that corrosion protection was attributed to the formation of a hydrated
cerium oxide film on the alloy surface and film deposition
proceeded in terms of a cathodic mechanism [28]. So far,
cerium conversion coating have been developed to meet the
demand of non-toxic coatings processes [919]. It should
Corresponding author. Tel.: +49-211-679-2547;
fax: +49-211-679-2218.
E-mail address: xingwenyu@yahoo.com (X. Yu).

0925-8388/$ see front matter 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0925-8388(03)00502-4

be noted, however, that among these processes almost all


cerium conversion coatings have to be deposited directly on
the matrix of aluminum alloys and little attention was paid
to its application in anodizing of aluminum alloys. Because
anodizing is one of the most widely used techniques for corrosion protection and decoration of aluminum alloys, it is
interesting to explore the possibility of forming cerium conversion coatings on porous film of anodized aluminum.
In our recent work, the cerium conversion coating with a
yellow appearance was deposited successfully as a porous
film on anodized aluminum alloy 2024. The corrosion resistance of the coating was found to be comparable to that
of chromate-sealed anodic coatings, as discussed elsewhere
[20]. The purpose of this paper is to investigate the composition of cerium conversion coating by X-ray photoelectron
spectroscopy (XPS).

2. Experimental
The material studied was 2024 Al alloy and the major
alloying elements are listed in Table 1.
General procedures of degreasing, alkaline etching, and
acid neutralization were used as pretreatment of aluminum
anodizing. The specimens were anodized galvanostatically
in a stirred aqueous solution of 18% (wt%) H2 SO4 at a
current density of 1.5 A dm2 and room temperature. Pb

194

X. Yu, G. Li / Journal of Alloys and Compounds 364 (2004) 193198

Table 1
Major alloying elements of 2024 alloy (wt%)
Cu

Mg

Mn

Fe

Si

Zn

Ni

Ti

Others

Al

3.84.9

1.21.8

0.30.9

0.5

0.5

0.3

0.1

0.15

0.1

Balance

sheets served as a cathode. After being anodized, the specimens were rinsed fully in distilled water for subsequent
use.
The anodized samples were immersed in the solution containing Ce(NO3 )3 and then rinsed with distilled water, and
dried with air. The parameters of the cerium conversion coating process were described in Table 2.
Cerium conversion coating obtained by this process was
about 35 m (examined by eddy-current thickness indicator) with excellent adhesion (examined with bending test)
and yellow colorization. The corrosion resistance of the coating was examined to be comparable to that of chromate sealing anodized coatings for 2024 alloy by salt solution immersion test and electrochemical methods [20].
Surface analysis was performed by X-ray photoelectron
spectroscopy (XPS), using a vacuum generator PHI 5700
ESCA. The X-ray source was unmonochromatized Al K
radiation. The accelerating voltage was 12.5 kV. The target
power was 50 kW. A vacuum generator argon ion gun was
used for depth profiling. The sputter rate was 3 nm min1
in an argon pressure of 0.6 106 at 3 keV energy. The
software packages MULTIPAK was used for the analysis of
XPS data.

3. XPS analysis
The ASF (atomic sensitivity factors) method was used for
element quantitative analysis. Relative atom concentration
CX was calculated using the equation [21]:
nX
IX /SX
= 
CX = 
n
i i
i Ii /Si

(1)

where subscript X and i represent calculated and component


atom, respectively, n is the number of atoms, I is the integral
intensity of photoelectron signal, and S is the relative atomic
sensitivity factor.
XPS spectra of cerium compounds exhibited complex features owing to hybridization with ligand orbitals and partial
occupancy of the valence 4f orbital. Fig. 1 showed the Ce3d
spectra of CeO2 [22]. The  peak, which arose from a
transition from the 4f initial state to the 4f final state, was
a satellite line of the Ce3d3/2 line. It had been reported in
the literature that the  peak arose exclusively from Ce4+

Fig. 1. Ce3d spectrum of CeO2 .

and was absent from the Ce3d spectra of pure Ce3+ species
[22,23]. The major reason for this was thought to be the lack
of the 4f configuration in the case of Ce3+ . Hence the 
peak could be used as valence state analysis of cerium compounds and quantitative measure of the amount of Ce4+ .
Shyu et al. [22] have demonstrated that the integral area of
the  peak with respect to the total Ce3d area could be
translated into percentage of Ce4+ with the relative error of
being in the range of 10%. In the case of pure Ce4+ , the 
peak should constitute around 14% of total integral intensity. According to the linear dependence of percentage 
on percentage Ce4+ reported in the literature, percentage of
Ce4+ was calculated by:
Ce4+ % =

 %
100%
14

(2)

where  % is percentage of  peak area with respect to


the total Ce3d area.

4. Results and discussion


Fig. 2 shows the XPS survey spectra of the cerium conversion coating at different sputter times. Besides Al, O, Ce
and S, significant amounts of C were present in the outer
layer of the conversion coating. It was reasonable that the
presence of C and partial O may be due to the accumula-

Table 2
The cerium conversion coating process
Ce(NO3 )3 (g l1 )

H2 O2 (g l1 )

H3 BO3 (g l1 )

pH

Temperature ( C)

Time (h)

3.0

0.3

0.5

5.0

30

X. Yu, G. Li / Journal of Alloys and Compounds 364 (2004) 193198

195

Fig. 2. Survey spectra of the Ce conversion coating at different sputter times. (a) Without sputter; (b) 12; (c) 36; (d) 51; (e) 60 min.

tion of contaminants during exposure to air [22]. The element depth profile corresponding to the survey spectra in
Fig. 2 was shown in Fig. 3. It shows that the depth of contamination of C is less than 30 nm. If changes in elements

Fig. 3. XPS depth profile corresponding to the survey spectra in Fig. 2.

proportion due to the adsorption of adventitious C and O


were considered to be neglected, it was clear that the amount
of Ce atoms decreased with increasing depth, which suggested that deposition rate of the cerium conversion coating
increased with immersion time.
The XPS high-resolution spectra of Ce3d obtained from
different sputter times, which correspond to various depths
in Fig. 2, are shown in Fig. 4. When compared with Fig. 1,
it was found that only surface Ce3d spectra of cerium conversion coating was substantially similar to that of CeO2 in
structure. This indicates that surface cerium was predominantly in the four-valent state and enjoyed a similar chemical
environment to that in CeO2 . On the other hand, all inside
Ce3d spectra were quite different to that of surface Ce and
had almost identical peak structure except for the small differences in spectral intensity of the  peak. According to
Eq. (2), Table 3 shows the percentage of Ce4+ to total Ce at
different depth (sputter time). (It should be noted here that
there might be some errors of the data in Table 3 due to the

196

X. Yu, G. Li / Journal of Alloys and Compounds 364 (2004) 193198

Fig. 4. High resolution Ce3d spectra of Ce conversion coating at different sputter times. (a) Without sputter; (b) 12; (c) 36; (d) 51; (e) 60 min.

possibility of the reduction of Ce4+ to Ce3+ during argon


sputtering. But the tendency of decreasing Ce4+ with depth
of the coating is correct. This conclusion can be drawn from
the fact that the increase in Ce3+ is continuous. If the Ce3+
was totally from the reduction of Ce4+ by sputtering, the
Ce3+ content should remain the same after sputtering). From
Table 3, it can be seen that cerium of bulk conversion coating
was predominantly in the three-valent state. The presence of
Ce4+ compounds in the inner Ce conversion coating layer
suggested that some Ce3+ was oxidized to Ce4+ during the
formation of the film. The majority of Ce4+ compounds in
Table 3
Changes in Ce4+ concentration at different sputter times
Sputter time (min)

Ce4+ concentration (%)

12

36

51

60

81.36

20.17

17.68

15.15

9.54

the outer coating layer indicates that some of the remnant


Ce3+ were also oxidized to Ce4+ during the course of exposing the specimens in the open air.
The high-resolution O1s spectra at different sputter times
measured on the sample with Ce conversion coating are
shown in Fig. 5. From Fig. 3, it could be seen that there was
no Al in the survey spectra of the un-sputtered surface. For
the samples after sputtering at different times, the coating
consisted of aluminum oxide and cerium oxides/hydroxides.
The assumption being that the O1s spectrum was attributed
to the AlO group (531.6 eV), CeO group (529.5 eV) and
the CeOH group (532 eV) (no AlO on the surface without sputter). Based on the three groups above, O1s spectra were fixed as in Fig. 5. According to area of the peaks,
[OCeOH ]/[OCeO ] and [Al]/[Ce] were obtained as shown
in Table 4. Apparently, cerium oxide was the predominant
composition of the coating. The cerium hydroxide content
decreased with increasing depth of the film while the aluminum content increased with increasing depth of the film.

X. Yu, G. Li / Journal of Alloys and Compounds 364 (2004) 193198

197

Fig. 5. High resolution O1s spectra of Ce conversion coating at different sputter times. (a) Without sputter; (b) 12; (c) 36; (d) 51; (e) 60 min.

Table 4
Changes in [OCeOH ]/[ OCeO ] and [Al]/[Ce] at different sputter times

ions. During exposure of the specimens to open air, some


Ce3+ on the surface of the coating were oxidized to Ce4+ .

Sputter time (min)

[OCeOH ]/[ OCeO ]


[Al]/[Ce]

12

36

51

60

0.910

0.284
0.439

0.225
0.453

0.196
0.579

0.129
1.003

Acknowledgements
This work has been carried out with the support of The
Chinese Postdoctoral Science Fund and The Special Funds
for the Major State Basic Research Projects G19990650.

5. Conclusion
Cerium conversion coating on anodized 2024 aluminum
alloy consisted of Al oxide, Ce oxide, and Ce hydroxide. The
Ce state exhibited a mixture of Ce3+ and Ce4+ . The outer
layer of the coating consisted of mostly Ce4+ compounds.
This indicates that some of the Ce3+ was oxidized to Ce4+
during coating formation in the solution containing Ce3+

References
[1]
[2]
[3]
[4]

B.R.W. Hinton, D.R. Arnott, N.E. Ryan, Metals Forum 7 (1984) 211.
D.R. Arnott, B.R.W. Hinton, N.E. Ryan, Corrosion 45 (1989) 12.
B.R.W. Hinton, D.R. Arnott, N.E. Ryan, Mater. Forum 9 (1986) 162.
F. Mansfeld, S. Lin, S. Kim, H. Shih, Corrosion 45 (1989) 615.

198

X. Yu, G. Li / Journal of Alloys and Compounds 364 (2004) 193198

[5] D.R. Arnott, N.E. Ryan, B.R.W. Hinton, Appl. Surf. Sci. 2223
(1985) 236.
[6] B.R.W. Hinton, J. Alloys Comp. 180 (1992) 15.
[7] A.J. Davenport, H.S. Isaacs, M.W. Kendig, Corros. Sci. 32 (1991)
653.
[8] A.J. Aldykiewicz, A.J. Davenport, H.S. Isaacs, J. Electrochem. Soc.
143 (1996) 147.
[9] F. Mansfeld, S. Lin, K. Kim, H. Shih, Corros. Sci. 27 (1987) 997.
[10] F. Mansfeld, Y. Wang, H. Shih, J. Electrochem. Soc. 138 (1991) L74.
[11] F. Mansfeld, Y. Wang, H. Shih, Electrochim. Acta 37 (1992) 2277.
[12] B.R.W. Hinton, D.R. Arnott, Mater. Aust. 19 (1987) 18.
[13] F. Mansfeld, Br. Corros. J. 29 (1994) 192.
[14] F. Mansfeld, Y. Wang, Mater. Sci. Eng. A 198 (1995) 51.
[15] C. Chen, F. Mansfeld, Corros. Sci. 39 (1997) 1075.

[16] J. Li, L. Gao, C. Lu, B. Luo, Corros. Sci. Protect. Technol 8 (1996)
139, in Chinese.
[17] J. Li, L. Gao, C. Lu, B. Luo, Corros. Sci. Protect. Technol 8 (1996)
271, in Chinese.
[18] X. Yu, D. Zhou, Z. Yin, Y. Zhou, Chin. J. Nonferr. Metals 9 (1999)
73, in Chinese.
[19] D.R. Arnott, B.R.W. Hinton, N.E. Ryan, Mater. Perform. 26 (1987)
42.
[20] X. Yu, C. Cao, Z. Yao, Mater. Sci. Lett. 19 (2000) 1907.
[21] S. Liu, The XPS Analysis, Vol. 94, Science Press, Beijing, 1988.
[22] J.Z. Shyu, K. Otto, W.L.H. Watkins, G.W. Graham, J. Catal. 114
(1988) 22.
[23] A.E. Hughes, J.D. Gorman, P.J.K. Paterson, Corros. Sci. 38 (1996)
1957.

Vous aimerez peut-être aussi