You are on page 1of 12

Oxid Met (2009) 72:179–190

DOI 10.1007/s11085-009-9154-1

ORIGINAL PAPER

Electrochemical-Impedance-Spectroscopy (EIS) Study


of Corrosion of Steels 12CrMoV and SS304 Beneath
a Molten ZnCl2–KCl Film at 400 °C in Air

T. J. Pan Æ W. M. Lu Æ Y. J. Ren Æ W. T. Wu Æ
C. L. Zeng

Received: 21 August 2007 / Revised: 8 January 2009 / Published online: 8 April 2009
Ó Springer Science+Business Media, LLC 2009

Abstract The corrosion resistance of 12CrMoV and SS304 steels in contact with a
molten mixture of (55–45) mol.% ZnCl2–KCl, similar to that found in waste-
incineration plants, has been examined in air at 400 °C by the electrochemical-
impedance-spectroscopy (EIS) technique. The initial Nyquist plots of 12CrMoV
showed a semicircle at high frequency and a line at low frequency indicating a
diffusion-controlled reaction. At a later stage, the Nyquist plots are composed of a
small semi-circle at high frequency, a line at medium frequency and a large semi-
circle at low frequency, similar to that shown by SS304 during the whole experi-
mental test. The larger impedance of SS304 as compared to 12CrMoV may be
attributed to the presence of Ni and to the higher Cr content of SS304. Equivalent
circuits representing the features of the corrosion of 12CrMoV and SS304 are
proposed to fit the corresponding impedance spectra, and the electrochemical
parameters in the equivalent circuits are calculated.

Keywords SS304  12CrMoV  ZnCl2–KCl  Hot corrosion  Impedance models

Introduction

Incineration has been adopted as an effective method for disposal of the ever-
increasing voluminous industrial and municipal solid waste in the recent years.
However, the corrosion problems during incineration are usually very severe

T. J. Pan  W. M. Lu  Y. J. Ren  W. T. Wu  C. L. Zeng (&)


State Key Laboratory for Corrosion and Protection, Institute of Metal Research,
Chinese Academy of Sciences, 62 Wencui Rd, Shenyang 110016, China
e-mail: clzeng@imr.ac.cn

T. J. Pan
School of Materials Science and Engineering, Jiangsu Polytechnic University,
Changzhou 213164, China

123
180 Oxid Met (2009) 72:179–190

because heavy-metal (Sn, Pd, Zn) and alkali-metal (Na, K) chlorides and sulfates
frequently exist as molten deposits on the surface of metallic components in the
operation temperature range of the furnace wall tubes and superheater tubes [1].
Molten-salt attack has been identified as the main degradation mechanism in waste
incineration. In the presence of such contaminants, the service life of the furnace
wall tubes and superheater tubes is generally determined by their resistance to hot
corrosion rather than by their stress-rupture resistance. This type of corrosion is
actually an accelerated oxidation in a high-temperature gaseous environment of a
material whose surface is coated by a thin fused-salt film and differs largely from
the corrosion of the materials in deep salts. This attack is mainly electrochemical in
nature and thus may be investigated by electrochemical techniques.
Generally, the use of electrochemical methods such as free-corrosion potential,
scanning polarization, cyclic voltammetry, and linear-polarization resistance may
not be sufficient to elucidate the complicated reaction processes occurring in the
presence of molten salts [2–4]. Conversely, electrochemical-impedance spectros-
copy (EIS) is increasingly used in corrosion research in this field, because it may
provide useful information for clarifying complicated electrode processes, so that
this technique may also be used for corrosion monitoring [5–7]. An additional
advantage of the EIS techniques is that the use of only very weak signals during
measurement does not disturb the corrosion system to be investigated.
Although EIS is a powerful tool for elucidating reaction processes involving
multiple steps, the origin and physical meaning of EIS parameters have been
perplexing problems for many years. Zeng et al. [8] proposed four models to
describe the molten-salt corrosion behavior by considering the scaling features of
metallic materials. A successful application of the EIS technique requires suitable
models for fitting the impedance spectra. EIS has been widely used in the study of
aqueous corrosion. Up to now some papers concerning application of EIS technique
in the study of hot corrosion have also been published [9–13]. A two-electrode
system seemed suitable for studying the hot-corrosion behavior in the presence of a
thin salt film by using EIS [12, 13].
In the present work, the corrosion behavior of two commercial steels, 12CrMoV
and SS304, covered with a thin film of a molten ZnCl2–KCl mixture at 400 °C in air
has been evaluated by electrochemical-impedance spectroscopy (EIS). This paper
focuses mainly on the impedance measurement and the establishment of impedance
models for corrosion induced by thin films of molten chlorides.

Experimental Procedures

The materials employed in this study were two commercial steels with actual
chemical composition (mass %) (12CrMoV—11.2 Cr, 0.2C, 0.3Si, 0.5Mn, 0.5Ni,
1.0Mo, 0.3V balance Fe; SS304—19.28 Cr, 8.83Ni, 1.67Mn, 0.45Si balance Fe). A
two-electrode (two working electrodes, WE) system was used for the impedance
measurement. Top views of the two-electrode arrangement and the experimental
assembly are shown in Fig. 1. The working electrodes of the two steels (12CrMoV
and SS304) were prepared as follows. Samples of the two steels were machined into

123
Oxid Met (2009) 72:179–190 181

Fig. 1 Schematic diagram of test apparatus for corrosion induced by a thin salt film and top view of the
two-electrode arrangement

a size of 10 9 10 9 5 mm3, and then each sample was ground down to 600# SiC
paper. A Ni–Cr wire was spot welded to one end of the specimens for electric
connection. Two specimens very closely spaced were sealed in an alumina tube by
high-temperature cement, with one side of 10 9 5 mm2 uncovered. The exposed
surfaces of the specimens were in the same plane. After the cement was dried at
room temperature, it was further solidified at 200 °C for 24 h. The working
electrode surface was polished again on 600# SiC paper, degreased and dried.
A mixture of 0.55ZnCl2–0.45KCl (molar ratio) was applied in the present study.
A total of 150 g of the salt mixture were dried in an alumina crucible at 200 °C for
24 h in air. Then, the furnace was heated to 400 °C. In order to produce a thin fused-
salt film on the electrode surface, the electrode arrangement was immersed briefly
into the melt, then withdrawn and kept above the melt, as shown in Fig. 1. This
process was repeated after 12 h to ensure the presence of a thin salt film on the
electrode. The actual amount of the thin salt film was about 2 mg/cm2 in this present
study.
Electrochemical-impedance measurements up to 24 h were performed at open-
circuit potential in air between 0.01 and 1 9 105 Hz with a M398 impedance
system, composed of a Princeton Applied Research (PAR) 5210 lock-in amplifier
and a PAR 263 potentiostat interfaced through an IEEE 488 bus to a compatible
computer. A fast Fourier-transform (FFT) technique was employed for frequencies
from 0.01 to 1.13 Hz to increase measurement speed and lower the degree of

123
182 Oxid Met (2009) 72:179–190

perturbation to the cell. The amplitude of input sine signal was 10 mV. A
commercial software (ZSimpWin) developed by Princeton Applied Research was
used to fit the impedance spectra.
The corroded samples were examined by X-ray diffraction (XRD) and Scanning
Electron Microscopy (SEM) coupled with an Energy-Dispersive X-ray microanal-
ysis system (EDX).

Experimental Results

EIS Measurements

The corrosion processes of 12CrMoV beneath a molten ZnCl2–KCl film at 400 °C


consist of two stages with different impedance-spectra features, as shown in Fig. 2.
The Nyquist plots during the initial stage are composed of a depressed semi-circle at
high frequency and a line at low frequency indicating that the corrosion is controlled
by the diffusion of oxidants in the melt, as shown in Fig. 2a. The corresponding
Bode plots show only one time constant. However, the impedance spectra of
12CrMoV changed with time. The Nyquist plots at the second stage are composed

(a) 5 20
2h 2h
15
4 3h 3h
10
, deg
Zim, ohm.cm2

5
3
0.01Hz
2
|Z|, ohm.cm

0
9
2 0.01Hz 6

1
3

0 -2 -1 0 1 2 3 4
0 2 4 6 8 10 10 10 10 10 10 10 10

Zre, ohm.cm2 Frequency, Hz

(b) 5 15
10h 10h
10 14h
4 14h
0.01Hz
, deg

24h
Zim, ohm.cm2

24h 0.01Hz 5
3
2
|Z|, ohm.cm

20 -2 -1 0 1 2 3
2 10 10 10 10 10 10
15
0.01Hz
10
1

0 5 -2
0 5 10 15 20 10 10
-1
10
0 1
10
2
10
3
10
2
Zre, ohm.cm Frequency, Hz

Fig. 2 Nyquist and Bode plots for the corrosion of 12CrMoV beneath a molten ZnCl2–KCl film at
400 °C in air after different exposure times. Symbol: experimental data; line: simulated data

123
Oxid Met (2009) 72:179–190 183

(a) 90 15
1h 1h
0.01Hz
5h 0.03Hz 10 5h

, deg
60
2

5
Zim, ohm.cm

4000

2
|Z|, ohm.cm
300
30 200

100

0
0 100 200 300 400 10
-2
10
-1
10
0
10
1
10
2
10
3
10
4

2
Zre, ohm.cm Frequency, Hz

(b) 40 15 10h
10h 0.05Hz
10 13h
13h 0.03Hz
30 , deg
24h
24h 0.03Hz 5
2
Zim, ohm.cm

20 0
|Z|, ohm.cm

150

100
10

50
0
-1 0 1 2 3 4
0 50 100 150 200 10 10 10 10 10 10
2 Frequency, Hz
Zre, ohm.cm

Fig. 3 Nyquist and Bode plots for the corrosion of SS304 beneath a molten ZnCl2–KCl film at 400 °C in
air after different exposure times. Symbol: experimental data; line: simulated data

of a depressed semicircle at high frequencies and a large semicircle at low


frequencies, connected by a line at intermediate frequencies, as shown in Fig. 2b.
The corresponding Bode plots show clearly two time constants.
During the whole experimental test the impedance spectra of SS304 are similar to
those of the second stage for 12CrMoV, but with much larger impedance. Typical
Nyquist and Bode plots for SS304 are shown in Fig. 3. The high-frequency portion
is a small capacitive loop, while the low-frequency portion is a depressed semicircle
with a much larger radius, again connected by a line at intermediate frequencies.
The corresponding Bode plots show clearly the presence of two time constants. The
impedance values decrease with time, indicating an enhanced corrosion of SS304.

Scale Morphology

Figure 4 shows the typical morphology of the scale formed on the salt-coated
samples after 24 h. The external layer of the scale formed on 12CrMoV consisted
mainly of iron oxide (Fe3O4) with the presence of some chromium (1–2 at.%), zinc
(3–4 at.%), potassium (1–2 at.%), and chlorine (2–3 at.%). The inner layer,
relatively thinner, had a much larger Cr content (about 18 at.%) and had again
similar contents of Zn, K, and Cl. Furthermore, the scales were significantly porous

123
184 Oxid Met (2009) 72:179–190

Fig. 4 Micrographs of cross sections of 12CrMoV (a) and SS304 (b) corroded beneath a ZnCl2–KCl salt
film at 400 °C for 24 h

and poorly adherent, as shown in Fig. 4a. Conversely, the scale grown on SS304
was composed of a complex mixture of Fe, Ni (about 6 at.%), and Cr (about
24 at.%) oxides, plus some zinc (5 at.%), potassium and chlorine. A small amount
Zn-Cr double oxide (ZnCr2O4) was also detected by XRD in the scales. Chlorine
was also detected in the innermost region of the scale close to the alloy/scale
interface, as shown in Fig. 4b.

Impedance Models

The Nyquist plots for the corrosion of 12CrMoV in the initial stage showed clearly
the features typical of a diffusion-controlled reaction. Therefore, the impedance
spectra for the corrosion of 12CrMoV may be described by the equivalent circuit of
Fig. 5a, where Rs represents the electrolyte resistance, Cdl the double-layer
capacitance, Rt the charge-transfer resistance, and Zw the diffusion-induced
Warburg resistance. Taking into account the dispersion effect, a constant phase
angle element (CPE) Q was used to describe the parameter Cdl in the fitting
procedure. Thus, the total electrochemical impedance for the circuit of Fig. 5a can
be expressed by Eq. 1

123
Oxid Met (2009) 72:179–190 185

Fig. 5 Equivalent circuit repre-


senting the corrosion of 12CrMoV
and SS304 beneath a molten
ZnCl2–KCl film at 400 °C in air

1
Z ¼ Rs þ ndl 1
ð1Þ
Ydl ðjxÞ þ Rt þZ w

where Ydl and ndl are constants representing the element Qdl. The Warburg resis-
tance Zw can be expressed by Eq. 2 [14]:
Zw ¼ Aw ðjxÞnw ð2Þ
where Aw is the modulus of Warburg resistance and nw the Warburg coefficient,
ranging between -0.5 and 0, related to the direction of the oxidants diffusion. When
nw is equal to -0.5, the diffusion direction of the oxidants is parallel to their
concentration gradient in molten-salts and accordingly the slope of the line at low
frequency in Nyquist plot is equal to 1. When nw [ -0.5, the diffusion direction of
the oxidants deviates from their concentration gradient, a situation denoted as
‘‘tangential diffusion’’, and the slope of the line at low frequency in Nyquist plot is
smaller than unity. The tangential diffusion’’ in molten-salt corrosion may be related
to the porosity of the scale, which has been observed in some studies [15].
With extended time, the impedance spectra of 12CrMoV changed and became
similar to those of SS304 during the whole duration of the experimental test. These
spectra are composed of a depressed semicircle at high frequency and a large
semicircle at low frequency, connected by a line at intermediate frequencies. The
change of impedance of 12CrMoV may be related to the growth of porous scales.
The molten salt can penetrate inward along the micropores in the scale to the metal
to induce fast corrosion. This typical corrosion can be represented by the equivalent
circuit of Fig. 5b, where Rs represents again the electrolyte resistance, Cdl, and Rt
represent the double-layer capacitance and the charge-transfer resistance at the
electrolyte/metal interface formed by the inward penetration of molten salts along

123
186 Oxid Met (2009) 72:179–190

pores, respectively, while Zw is the diffusion-induced Warburg resistance, related to


the diffusion of species along the micropores. Cps and Rp represent the porous scale
capacitance and the pore resistance, respectively. Considering the dispersion effect,
(CPE) Q is used to describe the parameters Cdl and the Cps in the fitting procedure.
The electrochemical impedance Z can be expressed by Eq. 3:
1
Z ¼ Rs þ ð3Þ
Yps ðjxÞnps þ R þZ 1
1
p wþ
Ydl ðjxÞndl þ 1
Rt

where Ydl and ndl, and Yps and nps are constants representing the parameters Qdl and
Qps, respectively.
The electrochemical parameters in Eqs. 1–3 can be obtained by fitting the
impedance spectra based on the equivalent circuit of Fig. 5, and are listed in
Tables 1 and 2, respectively. The Nyquist plots in Figs. 2 and 3 show clearly that
the fitting is fairly good, indicating that the proposed equivalent circuit is
reasonable. From Table 1 it can be seen that the medium resistance Rs underwent
slight changes with time, possibly related to changes of the salt film on the corroded
samples. The small values of Rt indicate that the electrochemical-reaction process is
rapid. The small values of Rp indicate poor corrosion resistance of the scale.
The gradual decrease of the charge-transfer resistance Rt with time shown in
Table 2 indicates that the electrochemical-reaction process becomes faster. Tables 1

Table 1 Fitting results of the impedance spectra of 12CrMoV beneath a ZnCl2–KCl salt film in air at
400 °C
Time Rs Rt Ydl ndl Aw,dl - Rp Yps nps
(hour) (X cm2) (X cm2) (X-1 cm-2 S-n) (X cm2 Sn) nw (X cm2) (X-1 cm-2 S-n)

2 2.96 4.08 5.79e-3 0.57 5.36 0.5


3 2.39 3.74 6.51e-3 0.62 2.92 0.5
6 5.84 1.60 1.343 0.73 0.70 0.37 2.47 2.53e-2 0.53
10 5.50 1.90 1.196 0.76 1.11 0.31 2.18 2.09e-2 0.56
14 5.30 5.77 0.506 0.73 0.92 0.46 3.36 3.01e-2 0.47
20 5.45 3.21 0.730 0.70 0.56 0.41 2.71 2.74e-2 0.50
24 6.64 5.70 0.519 0.72 1.02 0.42 3.20 2.78e-2 0.48

Table 2 Fitting results of the impedance spectra of SS304 beneath a ZnCl2–KCl salt film in air at 400 °C
Time Rs Rt Ydl ndl Aw,dl - Rp Yps nps
(hour) (X cm2) (X cm2) (X-1 cm-2 S-n) (X cm2 Sn) nw (X cm2) (X-1 cm-2 S-n)

1 58.58 36.85 1.13e-2 0.96 73.69 0.38 67.65 2.89e-4 0.61


5 55.91 47.68 4.38e-2 0.95 100.10 0.24 21.43 1.32e-4 0.64
7 59.12 26.69 6.23e-2 0.96 81.63 0.31 29.71 2.11e-4 0.66
10 44.45 22.46 1.48e-2 0.94 55.86 0.29 13.04 3.93e-4 0.60
13 46.98 9.38 7.35e-2 0.95 57.04 0.27 15.11 3.04e-4 0.64
24 42.75 15.16 1.58e-2 0.95 58.51 0.21 7.23 2.75e-4 0.65

123
Oxid Met (2009) 72:179–190 187

and 2 indicate that Rp of SS304 is larger than that of 12CrMoV, suggesting a better
corrosion resistance of SS304 as compared to 12CrMoV. The charge-transfer
resistance Rt in Table 1 is much smaller than that in Table 2, which also means that
the electrochemical process proceeds much faster for 12CrMoV than for SS304.
With extended exposure, the values of Rp and Rt of SS304 tend to decrease,
indicating an easier diffusion of species along micropores and a faster electro-
chemical reaction process. In addition, the fact that the values of ndl and nps in
Tables 1 and 2 deviate significantly from unity demonstrates clearly the existence of
dispersion effects, which are a notable feature for molten salt corrosion.
Furthermore, nw deviates from -0.5, indicating the presence of ‘‘tangential
diffusion’’ during corrosion, as observed in other studies [15]. Though 12CrMoV
formed a much thicker scale than SS304, the values of the pore resistance Rp for
12CrMoV were much smaller than those for SS304, relating to the formation of a
more porous scale, through which the melt could penetrate inward more easily.
Finally, Tables 1 and 2 show that the values of Rs of SS304 were much larger than
those of 12CrMoV. This significant difference may be ascribed to the following
aspects. The salts deposited on the alumina cement separating the two working
electrodes could evaporate instantly due to the high vapor pressure of ZnCl2 besides
partial consumption in the corrosion of the alloys. The fast evaporation of ZnCl2
would increase the melting point of the deposits. On the other hand, however, the
growth of corrosion products on the alloy surface could also affect obviously the
medium resistance Rs. The fast lateral growth of corrosion products on 12CrMoV
with respect to SS304 could give rise to a smaller value of Rs.

Discussion

The salt mixture used in the present study is molten at the reaction temperature
because its melting point is only about 250 °C. Thus, the accelerated corrosion is
mainly induced by the presence of molten salts. The molten salt acts as an
electrolyte to take part in the corrosion process at high temperature. In this case, iron
is oxidized according to the following anodic reaction:
Fe ¼ Fe2þ þ 2e ð4Þ
2?
Fe formed during corrosion procedure may diffuse outward through the molten
salt film and then be further oxidized to Fe3? at the sites where the oxygen pressure
is high enough according to the following anodic reaction:
Fe2þ ¼ Fe3þ þ e ð5Þ
At the same time, the corresponding cathodic reaction will occur as follows:
1
O2 þ 2e ¼ O2 ð6Þ
2
At the gas/melt interface, O2- and Fe3? will probably precipitate as a porous
oxide by the following reaction:

123
188 Oxid Met (2009) 72:179–190

2Fe3þ þ 3O2 ¼ Fe2 O3 ð7Þ


During corrosion, the iron oxide formed on the surface may dissolve into the melt
at the steel/salt interface, forming iron chlorides, likely to be FeCl3, by the following
reaction [16, 17]:
Fe2 O3 þ 6Cl ¼ 2FeCl3 þ 3O2 ð8Þ
The iron and oxygen ions dissolved in the melt may diffuse outwards towards the
salt/gas interface and produce iron oxide precipitates due to higher oxygen pressure
according to the above reaction (7) or the following reaction (9):
3
2FeCl3 þ O2 ¼ Fe2 O3 þ 3Cl2 ð9Þ
2
The oxide scale formed this way is rather porous, and thus the alloy suffered
rapid corrosion. At the same time, reaction (9) releases free chlorine, which will
penetrate down to the surface of the metal and react with iron and chromium from
the alloy substrate at the alloy/scale interface to form new metal chlorides according
to the following reactions:
Fe þ Cl2 ðgÞ ¼ FeCl2 ð10Þ
Cr þ Cl2 ðgÞ ¼ CrCl2 ð11Þ
FeCl2 and CrCl2 will dissolve in the melt and then diffuse rapidly towards the
scale/gas interface due to their high vapor pressures at 400 °C. At sites of
sufficiently large oxygen pressure Fe2O3 and Cr2O3 are precipitated again according
to the following reactions:
3
2FeCl2 þ O2 ¼ Fe2 O3 þ 2Cl2 ð12Þ
2
3
2CrCl2 þ O2 ¼ Cr2 O3 þ 2Cl2 ð13Þ
2
The chlorine released this way diffuses again inward to form new chlorides. This
process is repeated cyclically and may be sustained even in the presence of small
amounts of chlorides. The mechanism of formation of porous scales by reaction
with oxygen and chlorine, denoted as ‘‘active oxidation’’ [1, 18], involves the
participation of chlorine and of volatile metal chlorides in complex processes of
matter transport through the scales. Moreover, according to a fluxing mechanism
proposed by Spiegel [17], the chlorine, which takes part in ‘‘active oxidation’’
procedure, may also be produced according to the following reactions (14–17),
because iron and chromium oxides may dissolve in the molten ZnCl2–KCl salt:
4KCl þ 2Fe2 O3 þ O2 ¼ 2K2 Fe2 O4 þ 2Cl2 ð14Þ
2ZnCl2 þ 2Fe2 O3 þ O2 ¼ 2ZnFe2 O4 þ 2Cl2 ð15Þ
2ZnCl2 þ 2Cr2 O3 þ O2 ¼ 2ZnCr2 O4 þ 2Cl2 ð16Þ
4KCl þ 2Cr2 O3 þ O2 ¼ 2K2 Cr2 O4 þ 2Cl2 ð17Þ

123
Oxid Met (2009) 72:179–190 189

The oxide scales formed by the above reactions are rather porous and can hardly
provide any effective protection. The existing micropores in the scale allow a rapid
access for oxygen and chlorine. For the initial stage of corrosion of the steel
12CrMoV, because the diffusion of oxidants to the metal/salt interface is not fast
enough to meet its consumption in the corrosion reaction, the corresponding
impedance spectra exhibit the typical features of a diffusion-controlled reaction.
The dissolved chlorides and the oxide ions diffuse outwards through the molten salt
to the salt/gas interface to react with oxygen to precipitate in the scale as porous
oxides. After formation of porous scales on the samples surface, the oxidants need
to diffuse along the micropores in the scale to react with the metal. The corrosion
process during this stage is controlled by the diffusion of oxidants through the
micropores in the scale. Therefore, the impedance spectra for the corrosion of
12CrMoV after the initial stage present a depressed semicircle at high-frequency
and a large semicircle at low frequency, connected by a line at intermediate
frequencies. This change of impedance is closely related to the growth of porous
scales.
In this study, the higher Cr level in SS304 as compared to 12CrMoV resulted in
larger volume fractions of chromium oxides in the scales which acted as an effective
diffusion barrier against the transport of chlorine and volatile chlorides, thus
retarding corrosion. Therefore, the porous scale resistance of Rp and impedance of
SS304 are larger than those of 12CrMoV. Moreover, the inner layer rich in
chromium is also possibly associated with the fact that gaseous chromium chlorides
can be oxidized closer to the scale/matrix interface than iron and nickel chlorides.
However, this conversion from chlorides to the corresponding oxides can produce
more porosity and larger growth stresses in the scale. This is also an important
reason for the poor adherence of the oxide scales formed on Cr-rich steels.
On the other hand, the better corrosion resistance of SS304 as compared to
12CrMoV is also possibly related to the presence of nickel in the former material.
Actually, NiCl2 has a higher thermodynamic stability and a lower vapor pressure
compared with iron and chromium chlorides at 400–500 °C [19]. The solubility
measurements of oxide scales in molten NaCl–KCl mixtures by Ishitsuka et al. [20]
showed that NiO is less soluble than the Fe and Cr oxides. Thus, it is easier to reach
saturation in molten chlorides for NiO than for Fe and Cr oxides. Therefore, nickel-
base alloys are better candidate materials than iron-base materials for good
resistance against chlorides deposits.

Conclusions

The electrochemical impedance spectra for the corrosion of 12CrMoV beneath a


molten ZnCl2–KCl film at 400 °C in air are composed of two stages with different
features. The initial spectra consist of a small semicircle at high frequency and a line
at low frequency indicating a diffusion-controlled reaction. The impedance spectra
of 12CrMoV at longer times are similar to those of SS304 during the whole duration
of the corrosion tests. Their Nyquist plots are composed of a depressed semicircle at
high frequency and a large semicircle at low frequency, connected by a line at

123
190 Oxid Met (2009) 72:179–190

intermediate frequency. The changes of the impedance spectra may be associated


with the growth of porous scales on the surface of these steels. The corrosion
resistance of 12CrMoV and SS304 can be evaluated by comparing the electro-
chemical parameters Rp.

Acknowledgments A financial support by the National Natural Scientific Foundation of China (NSFC)
under the research grant No. 50671114 is gratefully acknowledged.

References

1. H. J. Grabke, E. Reese, and M. Spiegel, Corrosion Science 37, 1023 (1995).


2. M. Keddam, O. R. Mattos, and H. Takenouti, Journal of the Electrochemical Society 128, 257
(1981).
3. E. Barcia and O. R. Mattos, Electrochimica Acta 35, 1601 (1990).
4. L. Bai and B. E. Conway, Journal of the Electrochemical Society 137, 3737 (1990).
5. A. Nishikata, Y. Ichihara, and T. Tsuru, Corrosion Science 37, 897 (1995).
6. A. Nishikata, Y. Ichihara, and T. Tsuru, Electrochimica Acta 41, 1057 (1996).
7. A. Nishikata, Y. Yamashita, H. Katayama, T. Tsuru, A. Usami, K. Tanabe, and H. Mabuchi, Cor-
rosion Science 37, 2059 (1995).
8. C. L. Zeng, W. Wang, and W. T. Wu, Corrosion Science 43, 787 (2001).
9. G. Gao, F. H. Stott, J. L. Dawson, and D. M. Farrel, Oxidation of Metals 33, 79 (1990).
10. X. J. Zheng and R. A. Rapp, Journal of the Electrochemical Society 140, 2857 (1993).
11. Y. M. Wu and R. A. Rapp, Journal of the Electrochemical Society 138, 2683 (1991).
12. C. L. Zeng and T. Zhang, Electrochimica Acta 49, 1429 (2004).
13. C. L. Zeng and J. Li, Electrochimica Acta 50, 5533 (2005).
14. F. Mansfeld and M. W. Kendig, ASTM STP 866, 122 (1985).
15. C. L. Zeng, P. Y. Guo, and W. T. Wu, Electrochimica Acta 49, 1445 (2004).
16. Y. S. Li, M. Al-Omary, Y. Niu, and K. Zhang, High Temperature Materials Processes 21, 12 (2002).
17. M. Spiegel, Materials and Corrosion 50, 373 (1999).
18. M. J. McNallan, W. W. Liang, J. M. Oh, and C. T. Kang, Oxidation of Metals 17, 371 (1982).
19. A. Zahs, M. Spiegel, and H. J. Grabke, Materials and Corrosion 50, 561 (1999).
20. T. Ishitsuka and K. Nose, Materials and Corrosion 51, 177 (2000).

123

You might also like