Вы находитесь на странице: 1из 7

ISSN 1023-1935, Russian Journal of Electrochemistry, 2007, Vol. 43, No. 7, pp. 837843. Pleiades Publishing, Ltd. 2007.

Published in Russian in Elektrokhimiya, 2007, Vol. 43, No. 7, pp. 878885.

Corrosion Behavior of Magnesium and Its Alloy


in NaCl Solution*
F.-H. Caoa,z, V.-H. Lenb, Z. Zhanga, and J.-Q. Zhanga,b
a

Zhejiang University, Hangzhou 310027, P.R. China


Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, P.R. China
Received January 11, 2006; in final form, November 13, 2006

AbstractThe electrochemical behavior of cast Mg, AZ91, and cast AZ91 in 0.1 M NaCl solution is investigated by measuring open-circuit potential (OCP), steady-state currentpotential, and electrochemical impedance spectra (EIS). The similar electrochemical impedance behavior is found of three corrosion electrodes.
There are two capacitances in high- and medium-frequency domains and one inductive loop or component in
low-frequency domain. From equivalent circuit simulation, cast AZ91 has the worst corrosion resistance. The
EIS results are in good agreement with those obtained by OCP and polarization curves. Based on the Cao theory, a simple corrosion mechanism is put forward, supplying a possible explanation for low-frequency inductive
behavior for Mg and its alloy in NaCl solution at OCP.
Key word: magnesium, corrosion, electrochemical technology, inductive behavior
DOI: 10.1134/S1023193507070142

INTRODUCTION
The employment of magnesium as an anodic material in different types of batteries has been the subject
of much attention in recent years. Magnesium offers
several advantages as negative electrode material [1, 2],
e.g. a high electrode potential (2.37 V vs. NHE), a low
electrochemical equivalent (0.45 g per A h), good stability towards corrosion, relative abundance from sea
bitterns, low toxicity. This has prompted the authors to
investigate magnesium in conjunction with certain
novel N-halogen organic compounds for use in battery
devices [3, 4].
Magnesium-based alloys exhibit an attractive combination of low density and high strength/weight ratio
associated with good castability and workability. Thus
they offer substitutes to ferrous alloys and aluminum in
automotive and aeronautical applications [5, 6]. Among
the magnesium alloys, AZ91D is one of those having
the best mechanical properties/corrosion resistance balance and the present study is a continuation of a previous one devoted to the influence of the elaboration process of the AZ91D alloy on its corrosion behavior in
corrosive environments [7, 8]. It was demonstrated that
the corrosion resistance of AZ91D elaborated through
the semi-solid process (SSP) was somewhat higher than
that of the same alloy manufactured by high pressure
die casting (DC). As the SSP process induces strong
modifications of the alloy microstructure, in terms of
distribution, composition and volume fraction of the
* The text was submitted by the authors in English.
z Corresponding author, e-mail: nelson_cao@zju.edu.cn

constituting phases, the better corrosion resistance of


the alloys was attributed, in part, to the high aluminum
content of the a-phase (2.7 at %), (the a-phase from the
DC process only contains 1.6 at %).
Electrochemical technology (Electrochemical
impedance spectroscopy and polarization) has been
widely utilized in the last ten years as a tool to investigate the process of corrosion of magnesium alloy [7, 9
11]. It is evident that to obtain this result it is necessary
to distinguish the impedance values as a function of the
frequency in the single contribution due to the components of the system, for example distinguishing the part
of the impedance due to initial surface and from that
part due to corrosion message.
The present work constitutes part of a more extensive project on the corrosion of magnesium alloys. The
aim of this work is to clarify the approach of electrochemical technology, especially electrochemical
impedance spectroscopy, on magnesium alloy. We try
to gain a better understanding of the corrosion process
of cast pure magnesium and AZ91 magnesium alloy
(commercial and cast) in diluted natural NaCl solution
environments.
EXPERIMENTAL
Materials
Cast Mg, AZ91 (MgAlZn), and cast AZ91 alloys
were used as test materials. Their chemical compositions are shown in Table 1. Specimens were cylinders
with a diameter of 3 mm. Before use, the electrodes
were ground with 600 grit SiC paper, rinsed with ace-

837

838

CAO et al.

Table 1. Chemical analysis of the materials


Chemical analysis, wt %

Material
Cast Mg
AZ91
Cast AZ91

Al

Zn

Mn

Si

Cu

Ni

Fe

Mg

0.004
8.1
8.5

0.003
0.6
0.6

0.004
0.25
0.25

0.005
0.03
0.05

<0.001
0.003
0.25

<0.001
<0.001
0.001

0.002
0.003
0.004

>99.95
balance
balance

tone, washed with distilled water, and dried by flowing


warm air. All 0.1 M NaCl solutions were made with
A.R. chemicals and deionized water. The pH value was
adjusted to the neutral with 1 M HCl and 1 M NaOH.

with a 5210 lock-in analyzer over the frequency range


105102 Hz using a 10 mV peak-to-peak ac excitation.
The experiments were done at least two times and
the solutions were neither deaerated nor stirred because
the presence of O2 does not influence the corrosion of
magnesium [12].

Electrochemical measurements
The experiments were performed in a classical
three-electrode electrochemical cell. The working electrode was cylindrical samples and had a 3-mm crosssectional diameter. The auxiliary electrode was a large
surface area platinum sheet and reference electrode was
a saturated Hg/Hg2Cl2 electrode (SCE). The reference
electrode was placed outside the cell and connected to
the specimen surface by a Luggin capillary and a salt
bridge. All potentials presented in this paper are
referred to SCE (E = 0.224 vs. NHE).
All electrochemical measurements were carried out
at 25 0.5 using a thermostat. The variation of open
circuit potential with immersion time was carried out
using powerlab/4sp recorder with a GP amplifier. The
potentiodynamic polarization curves and the electrochemical impedance measurements were performed
using an EG&G potentiostat (Model 273A) driven by a
computer. Impedance measurements were carried out
OCP, V

1.4
1
2
3

1.5
1.6
1.7
1.8

16

32

48
Time, h

Fig. 1. Evolution of OCP with immersion time in 0.1 M


NaCl.

RESULTS AND DISCUSSION


Open-circuit potential
Although the standard potential of the magnesium
electrode is 2.37 V, the steady-state working potential
is generally about 1.50 V. This deviation in potential is
due to the formation of a magnesium hydroxide film on
the metal surface. The hydroxide film results from the
precipitation of Mg2+ ions in the solution wherein OH
ions are produced by the reduction of water according
to [1315].
Typical open-circuit potential (OCP) curves with
immersion time for cast Mg, AZ91, and cast AZ91 alloy
in 0.1 M NaCl solution at a sampling interval of
0.5 hour are shown in Fig. 1. It can be seen that the OCP
of cast AZ91 is more positive than that of Mg and AZ91
alloy. The OCP of AZ91 and Mg decreases fleetly at
first immersion, and then increases slightly, and keeps
relatively stable at the last 30-h immersion, while OCP
of cast AZ91 increases punily during the whole immersion time. There are some peaks in AZ91 curve, which
indicates localized corrosion, such as pitting corrosion,
is main corrosion type. Pure Mg decreases apace firstly,
then maintains a relatively stable value. The OCP of
cast AZ91 increases slowly during whole immersion
time, which shows that uniform corrosion is main corrosion type accompanied by localized corrosion (pitting corrosion) through metallurgical microscope
investigation.
The EIS study
Since a simple electrochemical system consists of
double layer capacitance, a solution resistance and a
charge transfer resistance, the use of ac signal can provide more information, than that obtained from open
circuit potential and dc polarization method. The
impedance locus diagram is obtained by applying a
10-mV (rms) potential between the working and
counter electrode. We can acquire some interface information through frequency response characteristic.

RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 43

No. 7

2007

CORROSION BEHAVIOR OF MAGNESIUM


ImZ, cm2

839

ImZ, cm2

1
2
3
4
5

8000

1
2
3
4
5

4000

2000

4000

4000

2000
0

10000

5000

10000
ReZ, cm2

20000
ReZ, cm2
Fig. 3. Nyquist plot for AZ91 in 0.1 M NaCl.

Fig. 2. Nyquist plot for cast Mg in 0.1 M NaCl.

At first sight, there is one capacitive arc and an shown in Fig. 5 is applied to analysis EIS data by
inductive component in Nyquist plot for cast Mg in ZView program. Rs, Rp, and Rt are the solution, polar0.1 M NaCl solution (Fig. 2). Based on simulation pro- ization, and charge transfer resistances, while constant
cess and evolvement of EIS, we consider that there are phase elements, Chf and Clf, are high- and low-fretwo capacitive arcs above real axes. The radius of mid- quency capacitive loops corresponding to initial and
dle-frequency capacitive arc decreases with immersion corrodent interface, respectively. All the capacitances
time at the first 24 hours, then maintains a relatively sta- shown in the equivalent circuits are mathematically
ble value, which indicates that the oxide film formed in modeled using a constant phase element in order to
air improves the corrosion resistance, and its localized consider also the electrochemical behavior of systems
rupture is main process at the first 24 hours and repas- which do not correspond exactly to a pure capacitance;
sivation translate an important process at the following in this way we have two parameters describing the ele24 hours.
ment: the value of admittance (Q) and the exponent of
Typical Nyquist plot of AZ91 alloy in 0.1 N NaCl is angular frequency (n). The exponent n is usually related
shown Fig. 3. It can be seen that there are two capaci- to diffusion phenomena, surface morphology and dissitive loops above real axis, while some scattered points pative process. The simulated results of different elecunder real axis. The radius of second capacitive arc has trode are shown in Table 2. There is a need to point out
a litter change during whole immersion time what indicates corrosion process is relative stable. The scattered
points in low frequency domain show the slow reac- ImZ, cm2
tions on metal oxide film/solution interface are not stable enough.
1500
Electrochemical impedance behavior of cast AZ91
1
in 0.1 M NaCl (Fig. 4) is similar to cast Mg. There are 1000
2
two capacitive loops and an inductive loop during the
3
whole frequency domain. The inductive component is
4
more obvious than that of the above two situation.
500
5
Since the Randle circuit model does not apply to
alloys with this additional constant, a new approach
0
was necessary for determining the corrosion rate based
on the complex impedance behavior described above.
When corrosion is not controlled by one-step, the
500
polarization resistance may no longer serve as a measure of the corrosion rate. In many such cases, e.g. mul2000
0
4000
tistep electron transfer, diffusion control, charge transReZ, cm2
fer resistance could determine the corrosion rate, which
may be used in a SternGeary expression [16, 17].
Fig. 4. Nyquist plot for cast AZ91 in 0.1 M NaCl.
Based on the above discussion, equivalent circuit
RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 43

No. 7

2007

840

CAO et al.
Chf

Rs

Clf

Rp

Rt
Fig. 5. Equivalent circuit of Mg and its alloy in 0.1 M NaCl.

that the equivalent circuit shown in Fig. 5 does not analyze impedance component under real axis. We omit
the inductive analysis in equivalent-circuit method.
More information will be given by mathematics
method of EIS developed by C.N. Cao [18, 19].
Double layer capacitance Chf increases form 0.28 to
240.01 F cm2 in turns of cast Mg, AZ91, and cast
AZ91 alloys, while corresponding dispersion coefficientnhf decreases from 1 to 0.333. The low-frequency
capacitive behavior takes on complex variation rule.
The Qlf value of cast Mg and AZ91 increases during the
first 24 h, then decreases, while Qlf of cast AZ91
increases during the whole immersion time. At the
same time, nlf is equal or close to unity, indicating that
the second capacitive loop is well-defined. The polarization resistance Rp is defined as real part of the impedance as frequency tends to zero, while charge transfer
resistance Rt is the real part of the impedance as frequency tends to infinite. Resistance Rt is generally considered as a corrosion rate parameter, rather than Rp, in
systems that contain more than one time constant. It is
easy to find that the corrosion resistance of cast Mg is

much larger than that of two alloys, especially than cast


AZ91 alloy.
The EIS spectra of Mg and its alloys are similar
except for the difference in the diameters of the loops.
This means that the corrosion mechanisms of these
alloys are the same, but their corrosion rates could be
different. Form the equivalent circuit (Fig. 5), it can be
seen that the highmiddle capacitive arc originates
from the original corroded surface, while the low-frequency capacitive arc from the newly formed interface
by aggressive ion attacking and Rt is charge transfer
resistance. Figure 6 shows the variation of parameters
Qlf and 1/Rt with immersion time. From Fig. 6, it can be
seen that the 1/Rt increases gradually with immersion
time for first immersion time of cast pure Mg and
AZ91, then slightly decreases, while for cast AZ91 it
increases fleetly during the whole immersion time. At
the same time, the corrosion resistance of cast AZ91 is
much smaller than that of other two electrodes. At the
end of a 24-h immersion, the AZ91 surface has very
severe corrosion and Qlf has the same change trend.
From variation of Qlf, it can be believed that localized
corrosion behavior increases, while corrosion production covers on the electrode surface at the end of
immersion time.
Corrosion reactions
Magnesium is oxidized to oxide, hydroxide or the
divalent ion in the presence of Cl ions. The total anodic
reaction is
Mg2+ + 2e.

Mg

(1)

Table 2. EIS simulation results of Mg and its alloy in 0.1 M NaCl


Alloy
Cast Mg

AZ91

Cast AZ91

Time, h Qhf, 1 cm2 sn

nhf

Rp, cm2

Qlf, 1 cm2 sn

nhf

Rct, cm2

0.33

0.28

1.0

338.3

0.81

0.836

21510

0.50

0.988

799.2

0.52

0.992

10023

24

1.00

0.978

178.9

2.08

0.959

3233

30

1.02

0.966

223.6

1.90

0.939

2854

48

1.18

0.999

708.2

1.69

0.930

3164

0.5

1.05

0.8889

14 270

1.09

1.0

12592

1.46

0.848

13276

1.34

1.0

9803

24

2.71

0.742

16521

1.91

1.0

5468

30

3.72

0.682

1671

0.13

1.0

17122

48

4.39

0.691

681.6

0.22

1.0

12600

7.76

0.611

117.4

0.82

1.0

12.19

0.601

163.7

0.82

1.0

27.37

0.504

238.8

1.23

1.0

618.8

12

117.58

0.276

126.9

4.56

0.979

200.5

24

240.01

0.333

3.19

0.965

72.91

RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 43

618.8
4207

40.97
No. 7

2007

CORROSION BEHAVIOR OF MAGNESIUM


Qlf, F cm2

In 1954 and 1955, J. Kleinberg et al. calculated


valence of dissolved magnesium, and considered that
the valence was in the range form 1.33 to 1.66 [2022],
which indicated monovalent magnesium ion existing
when magnesium/alloy dissolved. So in the solution
containing Cl ions, the corrosion reaction will be distinguished into different steps. First, the fine particles
of metal may become separated from the electrode surface [23, 24]:
K1

Mg

K1

Mg+

+ e.

(a)
4

K2

(2)

Mg2+ + OH + 1/2H2.

(3)

1/Rt, 1/

From our experiment, the hydrogen evolution process on the magnesium/alloy was very visible during
the whole immersion time. Otherwise, Eq. (3) supports
the negative difference effect (NDE) mechanism that
hydrogen evolution takes place on working electrode
by monovalent magnesium ion. The results are the formation of the stable valence of magnesium and the evolution of hydrogen. In neutral solution, the cathodic
reaction is Eq. (4) because the OH ions will be reacted
with Mg2+ ions:
2H2O+ 2e

K3

2OH +

1
2
3

The Mg+ ions have short lifetime with high reactivity. In the neutral solution, the monovalent magnesium
ion is oxidized by water rather than H+ in acidic solution [25]:
Mg+ + H2O

841

H2.

0.02

(7)

Let I1 and I3 be the current densities corresponding


to the electrochemical reactions (2) and (4), respectively. Then,
(8)
I1 = FK1 FcmK1,
I3 = 2FK3,

(9)

IF = (I1 I3).

(10)

The faradaic admittance of the electrode is as follows [18, 19]:


Y F = 1/R t
+ ( I F / X 1 ) ( d X 1 / dE )/ [ 1 j/ ( X '1 / X 1 ) ]

(11)

+ ( I F / X 2 ) ( d X 2 / dE )/ [ 1 j/ ( X '2 / X 2 ) ].
RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 43

0.0002

0.01

0.0001
0

12

24

36

48

0
0

The rate constants of the other two elementary steps,


Ki, obey the Tafel law:
Ki = ki xp(E/i),
(5)

X2 = cm.

1/Rt, 1/
0.0003

(4)

where ki is the potential-independent preexponential


rate constant, i is the ith Tafel constant. The reaction
sequence (2)(5) involves two independent surface
state variables: , which is the film free area of the surface, and cm, which is the Mg+ concentration. Based on
[24, 26], we consider that there are two state variables
of the above dissolution process.
(6)
X1 = ,

(b)

1
2
3

12

24

36

48
Time, h

Fig. 6. Variation of Qlf and 1/Rt with immersion time.

From experiment, the first variable increased with


applied electrode potential E. That is, whenever  0,
then
d/dE > 0.
(12)
At the OCP, IF is larger than zero:
I F / = I F > 0,

(13)

( I F / ) ( d/dE ) > 0,

(14)

so
indicating the Nyquist plot should contain an inductive
loop related to the state variable for cast Mg and AZ
Mg alloy.
For the second state variable cm, from [24]:
( I F / c m ) ( dc m /dE ) < 0.

(15)

This equation means that the intermediate species Mg+


brings about another capacitive loop in the EIS plot for
Mg and AZ Mg alloy.
From the above simple reaction model, we can
acquire the origination of different parts of EIS plot.
AH impedance experiments show that there are two
No. 7

2007

842

CAO et al.

trend: cast AZ91 > AZ91 > cast Mg, which indicates
that cast AZ91's free corrosion current is much higher
than that of the other two corrosion electrodes. The
impedance shows the same trend. The shift of corrosion
potential may be considered to content of zinc and different microstructure of magnesium and its alloy. The
corrosion potential of AZ91 and cast pure Mg is so
closely, indicating that the cast Mg is not very pure. At
the same time, Cl ions improve the localized corrosion
susceptivity through making the surface active and
increasing the broken area of the film.

logi [A]
2

1
2
3

6
2.0

1.5

1.0

0.5

Potential, V
Fig. 7. Typical log i vs. log E plots for 0.1 M NaCl; scan rate
0.1 mV s1.

capacitive loops, distributing at high-frequency (HF)


and medium-frequency (MF) domains, respectively,
and an inductive loop or component in low-frequency
(LF) domain. It is obvious that the HF capacitive is
originated from the film double layer. Form the above
mathematic deduction, the MF capacitive are related to
intermediate species Mg+, while the LF inductive component is originated from coverage of film-free activate
site shown in Figs. 24. The analysis of the three parts
of EIS of cast Mg and AZ Mg alloys supported the corrosion mechanism supposed by G.L. Song [24], who
built up the corrosion mechanism by researching pure
Mg corrosion process in NaCl and Na2SO4 solutions.
Because the MF capacitive loop is originated from
monovalent Mg+, the corrosion rate is dependent on
dissolution of Mg to monovalent Mg ions on the Mg
and its alloy surface, which supports the Rt simulated
from equivalent-circuit method is used to evaluate corrosion rate.

CONCLUSIONS
The results obtained in this study support the idea of
formation of intermediate species Mg+ in the neutral
solution containing Cl ions.
The result of OCP experiment indicates that the
main corrosion type is localized corrosion on the three
magnesium electrodes surface. Analysis of the impedance plots identifies the presence of three time constants at the corrosion potential, two capacitive loops
and one inductive component. It is obvious that the HF
capacitive loop is related the films double layer and
film resistance. The equivalent circuit simulation of EIS
show the cast AZ91 magnesium alloy has the lowest
corrosion resistance, while cast pure Mg has similar
value with AZ91 alloy, implied cast is not very perfect
because corrosion resistance of pure Mg should be
larger than that of AZ91. At the same time, according to
theory of Cao, the MF capacitive is originated by intermediate species Mg+, while the LF inductive component is related to coverage of film-free activate site
based on a simple reaction model. Since the MF capacitive loop is originated from intermediate species Mg+,
the corrosion rate is dependent on dissolution of Mg to
monovalent Mg ions on the Mg and its alloy surface,
supporting the idea of corrosion resistance simulating
from the MF capacitive arc. Moreover, the polarization
curves indicate similar result of EIS.

The Tafel study


Figure 7 presents the steady-state polarization
curves obtained for different Mg alloys in an aerated
0.1 M NaCl solution. The curves show a difference of
behavior, depending on the composition and microstructure of magnesium alloy. It is well known that corrosion potential of Mg is 1480 mV, while cast AZ91
has more positive corrosion potential, and AZ91
closely positive corrosion potential from Fig. 7. The
presence of zinc in the AZ91 alloy causes a shift in the
potential towards a more positive value. In the anodic
range, a current plateau is observed for three corrosion
electrodes. In 0.1 M NaCl solution, its height is also
dependent on the composition and structure of magnesium alloy. The plateau current density of AZ91 is
about 13 mA cm2, for 0.1 M NaCl solution, which is
larger than that of other alloys. From curves simulation,
the exchange current density, i0, has the following

ACKNOWLEDGMENTS
The authors acknowledge the financial support of
the National Natural Science Foundation of China
(no. 50 471 043) and the State Key Laboratory for Corrosion and Protection (China).
REFERENCES
1. Gregory, T.D., Hoffman, R.J., and Winterton, R.C.,
J. Electrochem. Soc., 1990, vol. 137, p. 775.
2. Udhayan, R., Muniyandi, N., and Mathur, P.B., Br. Core,
1992, vol. 27, p. 68.
3. Udhayan, R., Muniyandi, N., and Mathur, P.B., J. Power
Sources, 1991, vol. 34, p. 303.
4. Udhayan, R. and Bhatt, P.B., Electrochim. Acta, 1992,
vol. 37, p. 1971.

RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 43

No. 7

2007

CORROSION BEHAVIOR OF MAGNESIUM


5. Barzl, G. and Pebre, N., Corros. Sci., 2001, vol. 43,
p. 471.
6. Makar, G.L. and Kruger, J., J. Electrochem. Soc., 1990,
vol. 37, p. 414.
7. Mathieu, S., Rapin, C., Steinmetz, J., and Steinmetz, P.,
Corros. Sci., 2003, vol. 45, p. 2741.
8. Song, G.L., Atreas, A., and Dargusch, M., Mater. Sci.
Eng., A, 2004, vol. 346, p. 74.
9. Song, G.L., Atrens, A., Stjohn, D., Nairn, J., and Li, Y.,
Corros. Sci., 1995, vol. 39, p. 855.
10. Pebere, N., Riera, C., and Dabost, R., Electrochim. Acta,
1990, vol. 35, p. 555.
11. Udhayan, R. and Bhatt, D.P., J. Power Sources, 1996,
vol. 63, p. 103.
12. Gencvieve, B. and Nadine, P., Corros. Sci., 2001, vol. 43,
p. 471.
13. Bradford, P.M., Case, B., Dearnaley, G., Rurner, J.F., and
Woolsley, I.S., Corros. Sci., 1976, vol. 16, p. 747.
14. Cowan, K.G. and Harrison, J.A., Electrochim. Acta,
1980, vol. 25, p. 899.
15. Yamaguchi, S., J. Appl. Phys., 1954, vol. 25, p. 1437.

RUSSIAN JOURNAL OF ELECTROCHEMISTRY

Vol. 43

843

16. Stern, M. and Geary, A., J. Electrochem. Soc., 1957,


vol. 104, p. 59.
17. Epelboin, I., Gabrielli, C., Keddava, M., and
Takenouti, H., Electrochemical Corrosion Test, Standard ASTM-727, Mansfeld, F. and Bertocci, U., Eds.,
Philadelphia (PA): ASTM, 1981, p. 150.
18. Cao, C.N., Electrochim. Acta, 1990, vol. 35, p. 831.
19. Cao, C.N., Electrochim. Acta, 1990, vol. 35, p. 837.
20. Petty, R.L., Davidson, A.W., and Kleinberg, J., J. Am.
Chem. Soc., 1954, vol. 76, p. 363.
21. Rausch, M.D., McEwen, W.E., and Kleinberg, J., J. Am.
Chem. Soc., 1954, vol. 76, p. 3622.
22. Rausch, M.D., McEwen, W.E., and Kleinberg, J., J. Am.
Chem. Soc., 1955, vol. 76, p. 2093.
23. Bard, A.J., Encyclopedia of Electrochemistry of the Elements, New York: Marcel Dekker, 1978, vol. VIII.
24. Song, G.L., Atrens, A., John, D.S., Wu, X., and Nairn, J.,
Corros. Sci., 1997, vol. 39, p. 1981.
25. Makar, G.L. and Kruger, J., Int. Mater. Rev., 1993,
vol. 38, p. 138.
26. Cao, F.H., Zhang, Z., Zhang, L.J., Zhang, J.Q., and
Cao, C.N., Werkst. Korros., 2005, vol. 56, p. 318.

No. 7

2007

Вам также может понравиться