Sie sind auf Seite 1von 15

Electrochemical impedance models for molten

salt corrosion
C.L. Zeng
*
, W. Wang, W.T. Wu
State Key Laboratory for Corrosion and Protection, Institute of Corrosion and Protection of Metals, Chinese
Academy of Sciences, Wencui Road 62, Shenyang 110015, People's Republic of China
Received 21 June 1999; accepted 21 June 2000
Abstract
Taking into account the chemical stability and scaling features of metals, four electro-
chemical impedance models were proposed to represent their electrochemical impedance re-
sponses in molten-salt systems at the open-circuit potential. Electrochemical charge transfer
for the non-active metals is the rate-limiting process. For the active metals, the transfer of ions
in the scale and the diusion of oxidants in melts become increasingly important as compared
with the electrochemical process. When a non-protective scale forms on the metal surface, the
impedance diagram may present the characteristics typical of a diusion-controlled reaction,
i.e., a semi-circle at high frequency and a line at low frequency. When a protective scale forms
on the metal surface, the Nyquist plot is composed of double capacitance loops, and the
transfer of ions in the scale is rate limiting. An equivalent circuit of double layer capacitance in
series with oxide capacitance can be used to represent this kind of impedance response. In the
case of localized corrosion, the Nyquist plot also consists of double capacitance loops, which
can be described by an equivalent circuit of double layer capacitance parallel to oxide ca-
pacitance. Impedance measurements of Pt, Ni
3
Al and FeAl intermetallics in molten-salt sys-
tems were conducted to verify the proposed impedance models. 2001 Elsevier Science Ltd.
All rights reserved.
Keywords: Electrochemical impedance; Molten-salt corrosion; Pt, Ni
3
Al, FeAl
www.elsevier.com/locate/corsci
Corrosion Science 43 (2001) 787801
*
Corresponding author. Tel.: +86-24-2390-4553; fax: +86-24-2389-4149.
E-mail address: chaoliu@mail.sy.ln.cn (C.L. Zeng).
0010-938X/01/$ - see front matter 2001 Elsevier Science Ltd. All rights reserved.
PII: S0010- 938X( 00) 00108- 6
1. Introduction
Just as with aqueous corrosion, molten-salt corrosion may be studied by electro-
chemical techniques. Up to now, great progress has been made in the electro-
chemistry of molten-salt corrosion, and some electrochemical methods, such as free
corrosion potential, scanning polarization, cyclic voltammetry, potentiostatic po-
larization, weak polarization curve tting, and linear polarization resistance, have
been developed to study molten-salt corrosion. These techniques have proved to
be eective in simulating basicacidic uxing processes, understanding corrosion
mechanisms, and evaluating corrosion resistance. However, these techniques can
only provide limited information in understanding molten-salt corrosion. Electro-
chemical impedance spectroscopy (EIS) is a technique which has been widely used in
the study of aqueous corrosion, and has proved eective in understanding reaction
mechanisms and kinetics, and by which more information for corrosion processes
may be obtained. Only a limited number of impedance investigations of molten-salt
corrosion by EIS have been reported [111]. Farrell et al. [1] have employed im-
pedance technique to study the corrosion behavior of Nimonic 75 in bulk Na
2
SO
4
and in Na
2
SO
4
1%NaCl at 750C and 900C. They observed that the shape of the
impedance spectra has the characteristics of a diusion-controlled reaction, which
results because of the separation of the specimen from the gaseous environment by a
deep melt. By comparing the impedance at a xed low frequency (50 mHz), the
authors concluded that the corrosion rate was higher at 900C and with the addition
of sodium chloride. Gao et al. [2] also used impedance technique to monitor the hot
corrosion of Ni1%Co and alloy 800 in molten Na
2
SO
4
and Na
2
SO
4
10%NaCl. The
shape of the impedance diagrams also indicated a diusion-controlled reaction.
Wang et al. [3] studied the electrochemical reduction reactions of SO
3
gas by EIS. As
the amount of SO
2
in O
2
increased, and therefore SO
3
increased accordingly, the rate
of charge transfer also increased. However, the models for tting the impedance data
have not been discussed in the above studies [13]. Using the electrochemical im-
pedance technique, Wu and Rapp [4] studied the hot corrosion of preoxidized Ni by
a thin-fused Na
2
SO
4
lm at 1200 K in a catalyzed 0.1%SO
2
O
2
gas mixture. By
varying the specimen purity and preoxidation conditions for Ni, three distinct fea-
tures of hot corrosion (passive, pseudo-passive, and active) were observed. Wu [5]
further measured the double-layer capacitance at the preoxidized Ni/fused Na
2
SO
4
interface. The study of reduction reactions in molten-salt corrosion by EIS has also
been reported [68]. Zheng et al. [6] used EIS to study the cathodic reaction on the Pt
electrode in fused Na
2
SO
4
melt in SO
2
O
2
environments. The shape of the imped-
ance diagrams indicated a diusion-controlled reaction, mainly contributed by
the diusion of S
2
O
2
7
. Nishina et al. [7] made precise measurements of the War-
burg coecient of the impedance of the gold electrode fully immersed in molten
(Li,K)
2
CO
3
, (Li,Na)
2
CO
3
and (Li,K,Na)
2
CO
3
(at rest potential) at 615800C under
the atmosphere of O
2
with CO
2
. By using the RandlesErshler equivalent circuit and
Warburg coecients, the authors analyzed the graphical reaction order, and at last
concluded that the simultaneous diusion of O

2
and CO
2
is the dominant feature of
the oxygen reduction path in the molten carbonates. Makkus et al. [8] investigated
788 C.L. Zeng et al. / Corrosion Science 43 (2001) 787801
the reduction reaction of oxygen on the porous NiO, LiFeO
2
and LiCoO
2
in molten
carbonate fuel cell. It was observed that the Nyquist plots for all electrode materials
are composed of two arcs, which indicated a nite diusion-controlled reaction, and
was dierent from the impedance diagrams of a diusion-controlled reaction re-
ported by else [6,7]. The authors concluded that the molecular oxygen and carbon
dioxide are the predominant diusing species. Rouquette et al. [11] studied the
corrosion and passivation processes of magnesium in the molten 2HFKF mixture
by EIS. The authors proposed that the reaction pathway involved oxidation, de-
sorption, diusion and precipitation steps, and accordingly established the imped-
ance models.
In the previous studies, diusion was widely proposed as the controlling step of
molten-salt corrosion. However, there are limited reports on the establishment of
models for tting the impedance spectra. Successful application of EIS technique
requires suitable models for tting the impedance spectra besides the impedance
measurements. Taking into account the corrosion features of dierent metals, the
present paper discussed the electrochemical impedance models for molten-salt cor-
rosion.
2. Experimental method
The experimental apparatus used in the present study is shown in Fig. 1. A closed
end and grounded stainless tube was used for reaction chamber so as to eliminate the
possible eect of electric and magnetic eld caused by the furnace on the impedance
measurements. A three-electrode system was used for the impedance measurements.
The working electrodes (WE) used in the present study were pure platinum, inter-
metallic compound FeAl and Ni
3
Al. The chemical composition of the FeAl is Fe
40Al0.5B, and Ni
3
Al is Ni21.2Al0.1B0.3Zr (in atom percent). The working
electrode of Pt was prepared by spot welding a Pt foil with the area of 2 cm
2
to a Pt
wire sealed in an alumina tube. The working electrodes of Ni
3
Al and FeAl were
prepared as follows. Sheet-shape specimens with the size of 5 2 15 mm
3
were cut
from ingots by electric spark cutting machine, followed by grinding up to 600
#
SiC
paper. A FeCr wire was spot welded to one end of the specimens for electric
connection. The welding spot was sealed in an alumina tube. The working electrode
surface was polished again on 600
#
SiC paper, degreased, and dried prior to the test.
The reference electrode (RE) was a silver wire dipped into a 0.1AgCl0.9 working
salts (in mole fraction) contained in an alumina tube. The counter electrode (CE) is a
platinum foil. The impedance measurements of Pt and FeAl were conducted at
650C in (0.62Li, 0.38K)
2
CO
3
(in mole fraction) mixture under air, and Ni
3
Al at
700C in (0.36Li, 0.64(Na,K))
2
SO
4
(in mole fraction) in air. All the impedance
measurements were conducted at the rest potential. Working salts of 100 g were
contained in an alumina crucible. After the mixed salts were dried at 300C for 24 h,
the furnace was heated to the experimental temperatures. The impedance measure-
ments were carried out between 0.01 and 9:9 10
4
Hz by M398 impedance system
which is composed of Princeton Applied Research (PAR) 5210 lock-in amplier,
C.L. Zeng et al. / Corrosion Science 43 (2001) 787801 789
a PAR 263 potentiostat interfaced through an IEEE 488 bus to a compatible
computer. The amplitude of input sine-wave voltage is 10 mV. A Fourier transform
technique was employed for frequencies from 0.011.13 Hz to increase measurement
speed and lower the degree of perturbation to the cell. A software developed by
Zhang [12] was used to t the impedance spectra.
The corroded specimens were analyzed by X-ray diraction (XRD) and SEM
coupled with EDAX.
3. Electrochemical impedance models for molten-salt corrosion
Electrochemically, the molten-salt corrosion is very similar to the aqueous cor-
rosion. Nevertheless, compared with the corrosion in aqueous solution, the corro-
sion in molten salts is much faster, and usually may produce a thick scale. Thus, the
transportation of ions in the scale is also an important process to be considered in
molten-salt corrosion besides the diusion of oxidants and charge transfer through
the double-layer capacitance that are typical of aqueous systems. Moreover, molten-
salt systems operate at much higher temperatures than aqueous systems, thus,
molten-salt corrosion may take dierent forms, e.g., uniform corrosion, localized
corrosion and internal corrosion. Taking into account the chemical stability of
Fig. 1. Schematic diagram for the experimental setup: (1) WE, (2) CE, (3) thermocouple, (4) stainless steel
pot, (5) alumina tube, (6) furnace, (7) alumina crucible, (8) Pt foil, (9) molten salts, (10) specimen and (11)
RE.
790 C.L. Zeng et al. / Corrosion Science 43 (2001) 787801
dierent metals in molten salts, the metals can be divided into two groups, i.e., non-
active metals and active metals. The active metals may suer from molten-salt
corrosion, forming a scale that may be porous or protective. Based on the above
considerations, following models, i.e., non-active metals and active metals respec-
tively undergoing uniform corrosion (forming a porous or a protective scale) and
localized attack, can be used to describe the general situations of metals in molten-
salt systems, as shown in Fig. 2.
3.1. Non-active metals
Some non-active metals may have high enough chemical stability in molten-salt
systems, e.g., Pt in molten carbonates. This fact may mean that these metals do not
have enough activation to react easily with molten salts. In other words, it may also
be considered that the charge transfer through the double-layer capacitance cannot
easily go on owing to the large charge transfer resistance. Thus, the electrochemical
charge transfer may be rate-limiting process for the non-active metal/molten-salt
systems. The equivalent circuit for this impedance response can be described by Fig.
3, where R
s
represents the molten-salt resistance, C
dl
the double-layer capacitance
at the metal/melt interface, and R
t
the electrochemical transfer resistance. Accord-
ingly, the Nyquist plot is composed of a large semi-circle. The centre of semi-circle
lies on the real axis, provided that the double layer is considered as a perfect ca-
pacitor. However, in many cases, the centre of semi-circle does not lie on the real
axis, and can be looked upon as a semi-circle rotated in the clockwise sense around
the origin by a certain angle, a. An inclined semi-circle on the complex plane is as-
sociated with dispersion. The dispersion phenomenon may be attributed to the non-
uniform electric eld at the electrode/solution interface owing to the roughness of
electrode surface [13,14]. In molten-salt corrosion, the non-uniform electric eld may
be related to the roughness of the scale grown on the metal surface and even to the
molten-salt system. It can be proved that the dispersion eect may be represented by
Fig. 2. A schematic diagram for the metals in molten salts: (a) non-active metals, (b) forming a porous
scale, (c) forming a protective scale, (d) localized fast corrosion.
C.L. Zeng et al. / Corrosion Science 43 (2001) 787801 791
Cxcotbp=2, where b 1 a2=p0 < b61, represents the dispersion coecient
[15]. Thus, taking into account the dispersion eect, the electrochemical impedance Z
corresponding to the equivalent circuit of Fig. 3 can be expressed by Eq. (1):
Z R
s

1
jxC
dl
xC
dl
cotbp=2
1
R
t
; 1
where b represents the dispersion coecient of the capacitance loop, and
xC
dl
cotbp=2 the resistance caused by the dispersion eect.
The electrochemical parameters in Eq. (1) can be evaluated by tting the im-
pedance spectra based on the equivalent circuit of Fig. 3. It can be expected that the
charge transfer resistance R
t
for non-active metals should be much larger than that
for active metals in molten salts.
3.2. Active metals forming a porous scale
Diering from aqueous corrosion, the charge transfer reaction of active metals
during molten-salt corrosion can easily occur, and may not be rate limiting. Instead,
the transportation of ions in the scale and the diusion of oxidants in the melts may
become rate-limiting process, which will further depend on the protection of the scale
formed on the metal surface. When a porous scale forms on the metal surface, the
corrosion may occur at relatively high rate. The diusion of oxidants to the scale/
melt interface is not fast enough to meet their consumption in the reaction. In this
case, the anodic and cathodic charge transfers are fast process, and are not rate-
limiting. The cathodic process mainly including the cathodic charge transfer and
Fig. 3. Equivalent circuit representing non-active metal electrode in molten salts and its schematic im-
pedance spectrum.
792 C.L. Zeng et al. / Corrosion Science 43 (2001) 787801
diusion of oxidants is controlled by the diusion of oxidants in the melts. Thus, the
overall corrosion will be controlled by the diusion of oxidants in the melts. This
kind of diusion-controlled reaction has the typical features of impedance spectrum,
i.e., a semi-circle at high frequency and a line at low frequency, and can be described
by the equivalent circuit of Fig. 4, where R
a
represents the anodic charge transfer
resistance; R
c
, the cathodic charge transfer resistance; and Z
w
, the Warburg resis-
tance. Considering dispersion eect, the electrochemical impedance Z can be ex-
pressed by Eq. (2):
Z R
s

1
jxC
dl
xC
dl
cotbp=2
1
R
a

1
R
c
Z
w
: 2
The Warburg resistance Z
w
may be expressed by Eq. (3) [16]:
Z
w
A
w
jx
N
w
; 3
where A
w
is the modulus of Warburg resistance associated with the solubility and
diusion coecient of oxidants in the melt; and N
w
, the Warburg coecient
0:5 6N
w
< 0. The parameter N
w
is related to the diusion direction of oxidants.
When N
w
is equal to 0.5, the diusion direction of oxidants is parallel to the
concentration gradient of oxidants. When N
w
> 0:5, the diusion direction of
oxidants deviates from the concentration gradient of oxidants, i.e., ``tangential dif-
fusion''. The typical feature of tangential diusion is that the slope of the line at low
frequency is less than 1. A porous scale formed on the metal surface may be con-
sidered to be permeable to the molten salts, and thus may inuence the diusion
direction of oxidants. At present, a quantitative correlation of N
w
with the porosity
Fig. 4. Equivalent circuit representing the corrosion of metals forming a porous scale in molten salts and
its schematic impedance spectrum.
C.L. Zeng et al. / Corrosion Science 43 (2001) 787801 793
of the scale is not available. The values of the dispersion coecient b and the
Warburg coecient N
w
may reect some common properties of the scale to some
extent.
According to Eqs. (3) and (2), it can be seen that R
c
is much greater than Z
w
at
high frequency. Thus, Eq. (2) may be simplied to
Z R
s

1
jxC
dl
xC
dl
cotbp=2
1
R
t
; 4
where R
t
R
a
R
c
=R
a
R
c
represents the charge transfer resistance.
The parameters in Eqs. (3) and (4) can be calculated by tting the impedance
diagrams based on the equivalent circuit of Fig. 4. Just as mentioned above, the
modulus of Warburg resistance, A
w
, is related to the solubility and diusion coef-
cient of oxidants in the melt, and under conditions of semi-innite diusion can be
given by Eq. (5) [17]:
A
w

X
RT
n
2
i
F
2

2
p
1
C
i

D
i
p

; 5
where i denotes the diusing species, n
i
, the electron number per molecule of species
i, F, the Faraday constant, and C
i
and D
i
are the solubility and the diusion coef-
cient of species i in the melts. Based on Eq. (5), by varying the composition of gases
the measurements of A
w
may contribute to understanding the diusion and reduc-
tion mechanism of oxidants in the melts.
3.3. Active metals forming a protective scale
When a protective scale forms on the metal surface, the corrosion may be slowed
down. In this case, the transportation of ions in the scale, not the diusion of oxi-
dants in the melt may be rate limiting. If the scale is considered as a capacitor, it has
series relation with the double-layer capacitance at the scale/melts interface, as
presented by the equivalent circuit of Fig. 5, where, C
ox
represents the oxide ca-
pacitance, R
ox
the transfer resistance of ions in the scale. Correspondingly, the
Nyquist plot is composed of double capacitance loops as shown in Fig. 5. The
electrochemical impedance for this equivalent circuit may be expressed by Eq. (6):
Z R
s

1
jxC
dl
C
dl
xcotb
dl
p=2
1
R
t

1
jxC
ox
C
ox
xcotb
ox
p=2
1
R
ox
; 6
where b
dl
and b
ox
represent respectively the dispersion coecient of the rst and
second capacitance loop. Accordingly, C
dl
xcotb
dl
p=2 and C
ox
xcotb
ox
p=2 are
impedance elements caused by the dispersion eect. Because the corrosion is con-
trolled by the transportation of species in the scale, the radius of the second ca-
pacitance loop should be larger than that of the rst loop. Moreover, in this case the
corrosion resistance of the metals in molten salts may be represented by the pa-
rameter R
ox
.
794 C.L. Zeng et al. / Corrosion Science 43 (2001) 787801
3.4. Active metals suering from localized corrosion
Molten-salt corrosion often exhibits localized fast attack besides uniform attack.
Here, the so-called localized fast attack just means the local fast growth of scale, not
internal oxidation and/or suldation. For example, when a protective scale suers
from partial failure, the alloy may go through localized fast corrosion, provided that
the scale can not reheal. The localized corrosion zone may be covered with a non-
protective scale or directly exposed to the molten salts, while the rest (slow corrosion
zone) is covered with a more protective scale. Obviously, the reaction along the slow
corrosion site and that along the fast corrosion site occur coordinately. Both pro-
cesses are physically parallel. At the slow corrosion zone, the charge transfer resis-
tance at the scale/melts interface may be neglected compared with the transportation
resistance of ions in the scale. Thus, this kind of localized corrosion can be repre-
sented by the equivalent circuit of Fig. 6, where C
dl
and R
t
, respectively represents
double-layer capacitance and charge transfer resistance along the localized corrosion
zone, and C
ox
and R
ox
oxide capacitance and transfer resistance of ions in the scale
along the slow corrosion zone. The Nyquist plot corresponding to the equivalent
circuit of Fig. 6 also consists of two capacitance loops. The electrochemical im-
pedance Z can be expressed by Eq. (7):
Z R
s

1
jC
dl
x C
dl
xcot b
dl
p=2
1
R
t

1
jC
ox
xC
ox
xcot b
ox
p=2
1
R
ox
: 7
Based on the above discussions, it can be seen that the electrochemical impedance
diagrams for metals respectively undergoing uniform corrosion with the formation
Fig. 5. Equivalent circuit representing the corrosion of metals forming a protective scale in molten salts
and its schematic impedance spectrum.
C.L. Zeng et al. / Corrosion Science 43 (2001) 787801 795
of a protective scale, and localized corrosion all present the characteristics of double
capacitance loops. Thus, the analysis of the corrosion products is needed for the
establishment of a suitable impedance model.
4. Experimental results and discussion
4.1. Electrochemical impedance for platinum
Fig. 7 shows the Nyquist plot of the Pt electrode in molten (Li,K)
2
CO
3
at 650C in
air. The impedance spectrum is composed of a large capacitance loop. The semicircle
at the tested frequency scope can be attributed to the impedance of the charge
transfer reaction at the Pt/melts interface. This kind of impedance behavior can be
well described by the equivalent circuit of Fig. 3. The capacitance and the charge
transfer resistance at the Pt/melts interface were determined by tting the impedance
spectrum based on the equivalent circuit of Fig. 3, and are listed in Table 1. The
electrochemical charge transfer of the Pt electrode is the rate-limiting process in the
melts. As shown in the following sectors, the charge transfer resistance for the Pt
electrode is much larger than that for active metals.
4.2. Electrochemical impedance for Ni
3
Al
After the electrode of Ni
3
Al was immersed in the melt, its free corrosion potential
dropped rapidly to around 900 mV, and changed little with extended time. This
fact indicates that the alloy underwent a steady corrosion. The Nyquist plot for the
Fig. 6. Equivalent circuit representing metals suering from localized corrosion in molten salts and its
schematic impedance spectrum.
796 C.L. Zeng et al. / Corrosion Science 43 (2001) 787801
corrosion of Ni
3
Al in molten (Li,Na,K)
2
SO
4
at 700C is composed of a small
semicircle at high frequency and a line at low frequency (Fig. 8), which is the typical
features of a diusion-controlled reaction. The emergence of Warburg resistance at
the low frequency may indicate that the alloy suers from fast corrosion due to non-
formation of a protective scale. The small semicircle at the high frequency may mean
that the charge transfer reaction at the scale/melt interface can easily occur, and is
not the rate-limiting process. The corrosion of the alloy Ni
3
Al is controlled by the
diusion of oxidants.
The analysis of the corrosion products by XRD and SEM coupled with EDAX
revealed the formation of a non-protective layer with the thickness of more than 1
mm after corrosion for 11 h. The corrosion layer is mainly composed of un-corroded
Ni with some Ni
3
S
2
, NiO and Al
2
O
3
, which are dispersively distributed in the cor-
rosion layer. This non-protective corrosion layer is in agreement with the features of
impedance diagram to some extent.
The diusion-controlled reaction of Ni
3
Al can be described by the equivalent of
Fig. 4. Some electrochemical parameters were evaluated by tting the impedance
spectra based on the equivalent circuit of Fig. 4, and are listed in Table 2. The value
of the charge transfer resistance R
t
has the same magnitude as that of the modulus of
Warburg resistance A
w
. This fact indicates that the Warburg resistance contribution
Fig. 7. Nyquist plot for the electrode Pt in molten (K,Li)
2
CO
3
at 650C.
Table 1
Fitting results of the impedance spectrum of platinum in molten (Li,K)
2
CO
3
at 650C
Time (h) R
s
(Xcm
2
) R
t
(Xcm
2
) C
dl
(Fcm
2
) b
1 2 175.6 4.13E3 0.7025
C.L. Zeng et al. / Corrosion Science 43 (2001) 787801 797
caused by the diusion of oxidants can not be neglected. In fact, the Warburg re-
sistance is in inverse proportion to frequency x, but in direct proportion to R
t
. The
Warburg resistance coecient N
w
obviously deviates from 0.5, which shows the
existence of tangential diusion associated with the formation of a non-protective
scale on Ni
3
Al. Moreover, the value of the dispersion coecients b also signicantly
deviates from 1, which results from obvious dispersion eect owing to the formation
of a rough scale on Ni
3
Al. Additionally, the charge transfer resistance of Ni
3
Al is
much smaller, compared with that of Pt. This fact further indicates that diering
from non-active metals, the charge transfer reaction of the active metals can easily
take place, and is not rate limiting.
4.3. Electrochemical impedance for FeAl
Fig. 9 shows the typical impedance diagrams for the corrosion of FeAl in the
initial stage and later stage, which were reported earlier [18]. The change of im-
pedance diagrams is related to the change of the scale grown on FeAl. The Nyquist
plot in the initial times has the characteristics of a diusion-controlled reaction,
which can be described by the equivalent circuit of Fig. 4. After a 20-h immersion,
Fig. 8. Nyquist plot for the corrosion of Ni
3
Al after a 2-h immersion in molten (Li,Na,K)
2
SO
4
at 700C.
Table 2
Fitting results of the impedance spectrum of Ni
3
Al after corrosion in molten (Li,Na,K)
2
SO
4
at 700C for
2 h
R
s
(Xcm
2
) C
dl
(Fcm
2
) A
w
(Xcm
2
s
0:5
) R
t
(Xcm
2
) b N
w
0.76 8:7 10
2
0.581 0.5 0.74 0.31
798 C.L. Zeng et al. / Corrosion Science 43 (2001) 787801
Fig. 9. Nyquist plots for the corrosion of FeAl in molten (Li,K)
2
CO
3
at 650C for (a) 5 and (b) 48 h.
C.L. Zeng et al. / Corrosion Science 43 (2001) 787801 799
the impedance measurements showed that the Nyquist plot is composed of two
capacitance loops, which may be ascribed to the gradual formation of a protective
scale containing external LiFeO
2
layer and inner Al
2
O
3
layer. In this stage, the
corrosion is controlled by the transportation of ions in the scale, not by the diusion
of oxidants in the melts. This impedance diagram can be represented by the equiva-
lent circuit of Fig. 5. The free corrosion potential of FeAl also went through great
changes from around 900 to 200 mV in the initial stage, and changed little with
extended time. The change of the free corrosion potential of FeAl is in accord with
the impedance diagrams.
The tting results of the impedance diagrams respectively based on the equivalent
circuit of Fig. 4 and Fig. 5 are reported in Table 3. The value of R
ox
is much bigger
than that of R
t
. This result further indicates that the transfer of ions through the
compact scale is the rate-limiting process.
5. Conclusions
Four theoretical impedance models were proposed by taking into account the
chemical stability and surface scaling features of metals in molten-salt systems. The
charge transfer resistance for non-active metals is relatively large, and the charge
transfer reaction is rate limiting. The corresponding impedance diagram is composed
of a large semi-circle. Compared with the non-active metals, the charge transfer
reaction for active metals can easily occur, and may not be rate-limiting process.
The rate-limiting process and the corresponding impedance diagrams are related to
the scaling on the metal surface. When a porous scale forms on the metal surface, the
corrosion is controlled by the diusion of oxidants in the melt, and the Nyquist plot
consists of a small semi-circle at high frequency and a line at low frequency. On the
contrary, when a protective scale forms on the metal surface, the Nyquist plot is
composed of double capacitance loops, and the transfer of ions in the scale may be
rate limiting. An equivalent circuit of double layer capacitance in series with oxide
capacitance can be used to describe this kind of impedance diagram. When a metal
suers from localized corrosion, the impedance diagram is also composed of double
capacitance loops. An equivalent circuit of double layer capacitance at the localized
corrosion zone parallel to oxide capacitance can be set up to describe this impedance
response.
Table 3
Fitting results of the impedance spectra of the FeAl during the corrosion in molten (Li,K)
2
CO
3
at 650C
Time
(h)
R
s
(Xcm
2
)
R
t
(Xcm
2
)
C
dl
(Fcm
2
)
b
dl
N
w
A
w
(Xcm
2
s
0:5
)
R
ox
(Xcm
2
)
C
ox
(Fcm
2
)
b
ox
5 1.31 2.965 9.21E4 0.567 0.32 8.565
20 1.21 5.380 1.73E4 0.543 156.66 4.98E2 0.632
48 1.25 5.357 1.70E4 0.585 161.33 4.74E2 0.613
74 1.40 4.916 2.42E4 0.611 141.66 4.86E2 0.639
800 C.L. Zeng et al. / Corrosion Science 43 (2001) 787801
The practical impedance measurements of Pt, Ni
3
Al and FeAl in molten-salt
systems partially proved the proposed impedance models. The impedance diagram
of Pt in molten (Li,K)
2
CO
3
at 650C is composed of a large semi-circle, and the
charge transfer is the rate-limiting process. The corrosion of the Ni
3
Al in molten
(Li,Na,K)
2
SO
4
at 700C presented the characteristics of a diusion-controlled re-
action owing to the formation of a non-protective scale. The corrosion of the FeAl
underwent two stages corresponding to dierent impedance diagrams. In the initial
immersion, the corrosion showed a typical diusion-controlled reaction. In the ex-
tended immersion, the Nyquist plot is composed of double capacitance loops re-
sulting from the formation of a protective scale containing external LiFeO
2
and
inner Al
2
O
3
, and the corrosion is controlled by the transfer of ions in the scale.
Besides, the tangential diusion and dispersion eects are very obvious in molten-salt
corrosion.
Acknowledgements
This work was supported by the National Natural Science Foundation of China
under contract no. 59401012.
References
[1] D.M. Farrel, W.M. Cox, F.H. Stott, D.A. Eden, J.L. Dawson, G.C. Wood, High Temp. Technol. 3
(1985) 15.
[2] G. Gao, F.H. Stott, J.L. Dawson, D.M. Farrel, Oxid. Met. 33 (1990) 79.
[3] W.C. Fang, R.A. Rapp, J. Electrochem. Soc. 130 (1983) 2335.
[4] Y.M. Wu, R.A. Rapp, J. Electrochem. Soc. 138 (1991) 2683.
[5] Y.M. Wu, J. Electrochem. Soc. 138 (1991) 2342.
[6] X.J. Zheng, R.A. Rapp, J. Electrochem. Soc. 140 (1993) 2857.
[7] T. Nishina, I. Uchida, J.R. Selman, J. Electrochem. Soc. 141 (1994) 1191.
[8] R.C. Makkus, K. Hemmes, J.H.W. De Wit, J. Electrochem. Soc. 141 (1994) 3429.
[9] C.T. Liu, O.F. Devereux, J. Electrochem. Soc. 138 (1991) 3349.
[10] G. Picard, H. Lefebve, B. Tremillon, Molten Salts, in: G. Mamantov, M. Nlander, C. Hussey, C.
Mamantov, M.L. Saboungi, J. Wilkes (Eds.), The Electrochemical Society Softbound Proceedings
Series, PV. 87-7, 1987, p. 1028.
[11] S. Rouquette, D. Ferry, G. Picard, J. Electrochem. Soc. 136 (1989) 3299.
[12] J.Q. Zhang, G.Q. Sun, C.N. Cao, Corros. Sci. Protect. Technol. 6 (4) (1994) 318 (in Chinese).
[13] G.M. Schmid, Electrochim. Acta 15 (1970) 65.
[14] S. Iseki, K. Ohashi, S. Nagaura, Electrochim. Acta 17 (1972) 2249.
[15] J. Wang, C.Y. Shi, S.Z. Song, C.N. Cao, J. Chin. Soc. Corros. Protect. 13 (1989) 12.
[16] F. Mansfeld, M.W. Kendig, in: G.S. Havenes, R. Baboian (Eds.), Laboratory Corrosion Tests and
Standards, ASTM STP 866, American Society for Testing and Materials, vol. 122, Philadelphia, 1985.
[17] M. Sluyters-Rehbach, J.H. Sluyters, in: A.J. Bard (Ed.), Electroanalytical Chemistry, vol. 4, Marcel
Dekker, New York, 1970, p. 16.
[18] C.L. Zeng, W. Wang, W.T. Wu, Oxid. Met. 53 (2000) 289.
C.L. Zeng et al. / Corrosion Science 43 (2001) 787801 801

Das könnte Ihnen auch gefallen