You are on page 1of 12

Corrosion Science 48 (2006) 38123823

www.elsevier.com/locate/corsci

Passivity breakdown on AISI Type 403


stainless steel in chloride-containing
borate buer solution
Yancheng Zhang a, Digby D. Macdonald a,*,
Mirna Urquidi-Macdonald b, George R. Engelhardt c,
R. Barry Dooley d
a

Center for Electrochemical Science and Technology, Department of Materials Science and Engineering,
The Pennsylvania State University, University Park, PA 16802, USA
b
Department of Engineering Science and Mechanics, The Pennsylvania State University,
University Park, PA 16802, USA
c
OLI Systems Inc., 108 American Road, Morris Plains, NJ 07950, USA
d
Electric Power Research Institute, 3412 Hillview Avenue, Palo Alto, CA 94303, USA
Received 29 November 2005; accepted 26 January 2006
Available online 30 March 2006

Abstract
Passivity breakdown on AISI Type 403 stainless steel (SS), a commonly employed blade alloy in
low pressure steam turbines, has been studied and the data are interpreted in terms of the point defect
model (PDM). The near normal distribution in breakdown potential measured in deaerated borate
buer solution (pH = 8.1 0.1) with dierent chloride concentrations is in satisfactory agreement
with the quantitative characterization of the breakdown potential distribution using the PDM. The
linear dependence of breakdown potential on the square root of potential scan rate (t1/2), as predicted
by the PDM, yields an estimate of the critical areal concentration of condensed vacancies at the metal/
lm interface (n < 7.0 1014 cm2) that leads to passivity breakdown. This is in excellent agreement
with that calculated from the unit cell dimensions of the substrate FeCr alloy (n  1015 cm2) and the
barrier layer oxide (Cr2O3) (n  1014 cm2) for vacancy condensation on the alloy lattice or on the
cation sublattice, respectively, of the lm. These provide convincing evidences for the validity of
the PDM for modeling passivity breakdown on Type 403 SS.
 2006 Published by Elsevier Ltd.
*

Corresponding author. Tel.: +1 814 863 7772; fax: +1 814 863 4718.
E-mail address: ddm2@psu.edu (D.D. Macdonald).

0010-938X/$ - see front matter  2006 Published by Elsevier Ltd.


doi:10.1016/j.corsci.2006.01.009

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

3813

Keywords: A. AISI Type 403 stainless steel; B. The point defect model; C. Passivity breakdown; Pitting corrosion; Low pressure steam turbine

1. Introduction
Localized corrosion damage (LCD), such as pitting, corrosion fatigue, and stress
corrosion cracking, on the blade and disk alloy surfaces often leads to sudden and
catastrophic failures in low pressure (LP) steam turbines. The initial events in the accumulation of the LCD are passivity breakdown and pit nucleation, which readily occur in the
severely corrosive environments within the concentrated, chloride-containing, thin electrolyte lms that form on steel surfaces in LP steam turbines within the phase transition
region [13]. Subsequently, the pits that grow to critical dimensions act as sites for the
nucleation of cracks. Therefore, passivity breakdown and pit nucleation on the turbine
blades and disks are of particular interest for the lifetime prediction of LP steam turbines.
The point defect model (PDM) provides a deterministic description of passivity breakdown on metals and alloys [411]. The present study on AISI Type 403 stainless steel
(SS), a commonly employed blade alloy in LP steam turbines, is to assess the PDM for
quantitatively characterizing passivity breakdown on the turbine blade alloy, Type 403
SS (UNS S40300).
0
The PDM [411] proposes that cation vacancies V vM generated at the lm/solution
interface move to the metal/lm interface, and are annihilated by cation injection into
the lm from the metal.
0

m V vM ! M M mm ve0

where
m and mm are a metal atom and a metal vacancy in the metal phase, respectively; and
0
V vM and MM are a cation vacancy and a cation in normal cation sites on the cation
sub0
lattice of the barrier layer. The annihilation rate of the cation vacancies V vM at the
metal/lm interface is depicted as Jm. Aggressive anion absorption (e.g. chloride anion)
 catalyzes the generation of cation vacancies possibly
into surface oxygen vacancies V O
via a Schottky-pair reaction or anion-assisted cation extraction, and leads to an enhanced
ux of the cation vacancies (Jca) towards the metal/lm interface [510]. The cation vacancies condense at the metal/lm interface if Jca exceeds Jm; furthermore, when the areal concentration in the condensate exceeds a critical value (n), local separation of the barrier
layer from the substrate metal occurs. This local separation process continues at the
periphery of the condensate causing the condensate to expand and preventing the lm
from penetrating into the substrate via the generation of oxygen vacancies via
v 
m ! MM V O
ve0
2
2
However, dissolution continues at the outer surface with the result that the barrier layer
over the condensate thins at a rate that is determined by the dissolution rate and, at some
point, the cap (remnants of the barrier layer) over the condensate will rupture to mark a
passivity breakdown event [510]. The critical concentration of cation vacancies that must
condense per unit area (n) for separation to occur between the barrier layer and the metal
substrate is determined by the structures of the metal and the lm, depending on whether
condensation is envisioned to occur on the metal lattice or on the cation sublattice of the

3814

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

lm. According to this mechanism, taking into account only chloride anion adsorption/
Schottky-pair vacancy generation, the following relationship for the critical breakdown
potential (Vc) was derived [69].
 
4:606RT
b
2:303RT
log
logaCl
3

Vc
vaF
D
aF
!
DG0s v2 DG0A  v2 bF pH  v2 F /0f=s
RTJ m X
 exp
b
4
F veN v
RT
/f=s /0f=s aV bpH

where aCl is the chloride ion activity, D is the diusivity of the cation vacancy in the barrier layer, and b is dened in Eq. (4). The parameters v, e, and X are the oxide stoichiometry (MOv/2), the electric eld strength within the barrier layer, and the volume per
mole of cation in the lm. DG0s and DG0A are, respectively,
  the change in standard Gibbs
0
energy for the Schottky-pair reaction Null $ V vM v2 V O
 and for the absorption of
chloride anions into oxygen vacancies at the lm/solution interface. The parameters a
and b are the dependencies of the potential drop across the lm/solution interface (/f/s)
on the applied potential V and pH, respectively. /0f=s is the value of /f/s in the standard
state of V = 0 and pH = 0. R is the gas constant, T is the temperature, F is Faradays constant, and Nv is Avogadros number.
Passivity breakdown is envisioned to occur at sites on the surface that are characterized
by high cation vacancy diusivity [5]. These sites are considered to include the regions of
high disorder in the structure of the barrier layer that might exist, for example, at points of
intersection of the barrier layer with inclusions (e.g., MnS in stainless steels [12]) and other
precipitates (e.g., Cr23C6, also in stainless steels [13]). The potential breakdown sites are
assumed to be distributed normally with respect to the cation vacancy diusivity with
the distribution being characterized by a mean value, D, and a standard deviation, rD.
Based on this postulate, an analytical distribution function for Vc is readily derived as
Eq. (6) [8,9]. Importantly, Vc is predicted to follow a near normal distribution, as has been
observed experimentally [14].
" 
2 #
dN
c0 D
DD
p
exp  p
dV c
2p  rD
2rD
"
#
0
0
2
bc0
ec V c  ec V c b2
exp
p
expc0 V c
v=2
2r2D avCl
2p  rD  aCl

vaF
where c0 2RT
, and V c is the mean critical breakdown potential, which can be calculated
as,


1
b
v=2
 aCl
V c 0 Ln
7
c
D

The cumulative probability in breakdown potential is dened as Eq. (8), where P(Vc)
represents the percentage of breakdown sites at the Vc in all possible breakdown sites
[8,9].

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

3815


dN
dV c
dV c


P V c 100  Z 1
1
dN
dV c
dV
c
1
Z

Vc

Under potentiodynamic polarization conditions, where the potential is swept in the


positive direction linearly with time at a potential sweep rate of t (V/s), the measured
breakdown potential is found to be signicantly higher at high sweep rates than that measured at lower sweep rate [14,15]. The PDM provides an analytical relationship between
the breakdown potential [Vc(t)] and the potential scan rate (t) as Eq. (9) [15], which
can be used to determine the ratio of (n/Jm) [15,16] from the slope. If Jm is estimated from
its relationship with Jca, n can be calculated. Moreover, the derived value for n can be
appraised by structural arguments by considering the crystal structures of the metal substrate and the barrier layer.

1=2
2nRT
1
V c t
t2 V c t 0
9
J m vaF
where Vc(t = 0) is dened as the breakdown potential at zero potential scan rate.
2. Experimental
Type 403 SS (UNS S40300) specimens were cut from a commercial round bar (diameter
U = 1.27 cm, from Fry Steel Co., CA) that was in the annealed condition, as received. The
chemical composition of the particular heat was reported to be (weight percent): 0.12 C,
0.18 Si, 0.020 P, 0.004 S, 0.42 Mn, 12.34 Cr, 0.22 Ni, 0.15 Mo, 0.05 Cu, 0.05 Co, 0.004
Sn, 0.01 Al, 0.03 N. After welding stainless steel wires as electric contacts to the back
of the specimens and then mounting the specimens in epoxy resin, the specimens were
mechanically ground with emery paper down to 800 grit and polished with alumina powder nally down to 0.3 lm. The working surface area was 1.27 cm2.
The experimental electrolyte was borate buer solution with dierent concentration of
sodium chloride and dierent pH (Table 1). The electrolyte was purged for about 2 h
before use and during each experiment with ultrahigh purity N2 gas (UHP, 99.999%)

Table 1
Composition, measured pH, and conductivity of borate buer solutions containing dierent concentrations of
sodium chloride
Solution composition

Measured
pH

Measured
conductivity
(mS/cm)

0.75 M
0.75 M
0.75 M
0.50 M
0.50 M
0.75 M
0.75 M

8.20
8.13
8.01
8.16
9.25
5.67
4.55

9.4
15
36
47
20
10
9.1

H3BO3 + 0.20 M NaOH + 0.028 M NaCl


H3BO3 + 0.20 M NaOH + 0.10 M NaCl
H3BO3 + 0.20 M NaOH + 0.28 M NaCl
H3BO3 + 0.15 M NaOH + 0.50 M NaCl
H3BO3 + 0.25 M NaOH + 0.10 M NaCl
H3BO3 + 0.020 M NaOH + 0.10 M NaCl
H3BO3 + 0.002 M NaOH + 0.10 M NaCl

3816

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

having a stated oxygen concentration of less than 1 ppm. The conductivities of the solutions were suciently high (>10 mS cm1, Table 1) that the IR drop between the working electrode and the tip of the Luggin probe could be ignored.
After each specimen was placed in the deaerated electrolyte, cathodic polarization was
applied for 5 min as an initial conditioning followed by exposure at open-circuit for half
an hour. Potentiostatic polarization at 1.0 VSCE and galvanostatic polarization at
0.2 mA might be considered to have an indistinguishable pretreatment eect on the
initial surface when pH = 8.1 0.1; but galvanostatic polarization at 0.2 mA was chosen
for the other pH in order to obtain equivalent pretreatment eects under dierent pH
values. The initial cathodic treatment is not able to completely reduce the air-formed lm
on high chromium alloys [13,17], but it did yield a reproducible initial surface. The breakdown potential of Type 403 SS was then measured potentiodynamically by sweeping the
potential in the positive direction at dierent potential scan rates (from 0.1 to 10 mV/s), as
embodied in Eq. (9). The potentiodynamic experiments at a sweep rate of 0.1 mV/s were
typically repeated 20 times to obtain the statistical distribution of the breakdown
potential.
All electrochemical experiments were performed using a Solartron Electrochemical
Measurement System (Model SI 1280B) and the experiments were carried out at ambient
temperature (22 2 C). A coiled nickel wire was used as the counter electrode and a commercial, saturated calomel electrode (SCE), connected to the cell via a Luggin probe, was
used as the reference electrode. The tip of the Luggin probe was placed about 1 cm from
the specimen surface, in order to avoid shielding of the surface and hence avoid locally
inhomogeneous distribution in the potential and current [18]. All potentials are reported
on the SCE scale.
3. Results and discussion
Typical potentiodynamic polarization curves for Type 403 SS in deaerated borate buer
solutions (pH = 8.1 0.1) as a function of chloride concentration at a potential sweep rate
t = 0.1 mV/s are shown in Fig. 1. The polarization curves are characterized by wide

Fig. 1. Typical potentiodynamic polarization curves (t = 0.1 mV/s) for Type 403 SS in deaerated borate buer
solutions (pH = 8.1 0.1) as a function of chloride concentration.

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

3817

Fig. 2. Apparent mean breakdown potentials, V c t 0:1 mV/s, for Type 403 SS in deaerated borate buer
solutions (pH = 8.1 0.1) as a function of the logarithm of the chloride activity, logaCl .

passive ranges terminating in a sharp rise in the current density at a potential that depends
on the chloride concentration, due to passivity breakdown. The apparent breakdown
potential of Type 403 SS is seen to decrease with increasing chloride concentration, as
is normally observed for passivity breakdown on iron- and nickel-based alloys [5].
The apparent mean breakdown potential, V c t 0:1 mV/s, for Type 403 SS in deaerated borate buer solutions (pH = 8.1 0.1) is found to decrease linearly with the logarithm of chloride activity, logaCl (Fig. 2). The mean molar activity coecient (c) for
Cl in binary NaCl solutions of dierent concentrations [19,20] was used to estimate
the activity, aCl , of chloride ion in the buer solutions. We assume v = 3 for the cation
oxidation state within the barrier layer, since Cr2O3 is considered to comprise the inner
layer (barrier layer) of the passive lm on stainless steel surfaces [21,22]. The possible reasons for the signicant enrichment of Cr2O3 within the barrier layer are the preferential
oxidation of Cr prior to the onset of Fe oxidation [23] and that chromium cations have
a lower mobility than iron cations in the barrier layer lattice, resulting in the preferential
transmission of iron cations, probably as metal interstitials [24,25]. Consequently, the
value of a (the polarizability of the barrier layer/solution interface) can be determined
from the linear dependence of V c t 0:1 mV/s on logaCl , as indicated by Eq. (3) [5
10]. The obtained result is a = 0.23, which is signicantly lower than that found for Type
316 SS [26] (a = 0.690.71) and for Type 304 SS [27] (a = 0.8) in chloride-containing solutions. The lower a value might result from the adsorption of borate anion into surface oxygen vacancies, in competition with Cl. Thus, adsorption of the negatively charged oxygen
on the oxyanion may reduce the positive charge at the barrier layer/solution interface and
hence diminish the electric ux passing through this interface according to Gausss law
[28]. Because of this, /f/s decreases and so does the a value in Eq. (5).
The breakdown potential on Type 403 SS was measured at dierent pH (9.3, 8.1, 5.7
and 4.6) in deaerated borate buer solutions with 0.10 M NaCl by potentiodynamic

3818

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

Fig. 3. Apparent breakdown potentials, Vc(t = 0.1 mV/s), for Type 403 SS in deaerated borate buer solutions
with 0.10 M NaCl as a function of pH.

polarization (at a potential sweep rate t = 0.1 mV/s). As a result, the measured breakdown
potential, Vc(t = 0.1 mV/s), appears to increase with increasing pH of the buer solution
(Fig. 3). From the linear slope in Fig. 3 and the a value determined previously (a = 0.23),
the b value is computed to be 0.01 V.
Potentiodynamic polarization at dierent potential scan rates (from 0.1 to 10 mV/s)
was applied to Type 403 SS in borate buer solution containing dierent chloride concentrations (0.28, 0.10 and 0.028 M NaCl with pH = 8.1 0.1) or dierent pH (8.1, 5.7 and
4.6 with 0.10 M NaCl). The measured breakdown potential [Vc(t)] changes linearly with
the square root of potential scan rate (t1/2), as shown in Fig. 4. The linear slopes appear
to be essentially independent of chloride concentration and pH, which demonstrates the
validity of Eq. (9). The average slope is 1.62 (V s)1/2, from which the ratio of n/Jm is calculated to be 35 s. In this regard, the value of n is estimated as follows after the approximation of Jm from its relationship with Jca.
According to the PDM [411], when inducing passivity breakdown, the enhanced ux
of the cation vacancies (Jca) through the barrier layer towards the metal/lm interface
must be greater than the ux of the cation vacancies submerging into the metal phase
(Jm). The barrier layer on Type 403 SS is known to be an n-type semiconductor [29], which
is the same semiconductor type of passive lm on stainless steel or FeCr alloys, as
reported by Cho, Simoes, and Tsuchiya et al. [3032], implying that the dominant defects
are cation interstitials (most likely) and/or oxygen vacancies, not the cation vacancies that
are postulated to be responsible for passivity breakdown. Accordingly, the fraction of the
current density carried by the cation vacancies must be less than that indicated by the measured current density (Iss), so that, at breakdown.
J m J ca <

I ss N v
vF

10

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

3819

Fig. 4. Measured breakdown potentials Vc(t) at dierent potential scan rates (t) for Type 403 SS in deaerated
chloride-containing borate buer solutions. Above: dierent chloride concentrations with pH = 8.1 0.1; below:
dierent pH values with 0.10 M NaCl.

where Nv is Avogadro constant, v is the valence state of the cation vacancy, and F is Faradays constant. This calculation is necessarily approximate in the absence of transport
number data for the defects in the barrier layer, considering the contribution from the
transport of oxygen vacancies and/or Cr3+ interstitials due to the n-type semiconductor
behavior, as noted above. The error incurred by selecting the passive current density to
represent the current density due to cation vacancies is countered by the fact that the local
current density at a breakdown site is considerably higher than the average current density
measured in the external circuit. On the other hand, it is possible that the vacancies that
condense on the metal lattice at the metal/barrier layer interface, and which lead to
separation of the barrier layer from the substrate, are those that form by the generation
of cation interstitials
0
m ! M v
i vm ve

11

The enhanced generation of metal vacancies (vm) at the metal/barrier layer interface in the
 M v , where the
presence of chloride ion adsorbed into surface oxygen vacancies ClO
i

3820

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

species is written in KrogerVink notation, could be accounted for by chloride-catalyzed


extraction of metal interstitials at the barrier layer/solution interface, as indicated by the
reaction
Cl M v ! M v V  Cl
12
O

v+

where M and Cl are species in the solution. Theoretical analysis of this case [33] yields
an expression of the form of Eq. (3) for the breakdown potential. In this case, at the point
of breakdown, Eq. (10) is modied to read
J m J ca

I ss N v
vF

13

and, hence, it is possible to specify, more or less exactly, the rate of annihilation of the condensing vacancies at the metal/barrier layer interface, provided that metal interstitials are
overwhelmingly the dominant defect in the barrier layer. Thus, considering Eqs. (10) and
(13), we write the relationship between Jm and Iss as
Jm 6

I ss N v
vF

14

The largest measured current density that was recorded just before passivity breakdown
on Type 403 SS in borate buer solutions of various chloride concentrations and pH was
found to be about 10 lA/cm2. Assuming that cation vacancies V Cr30 are the condensing
species and, hence using Eq. (10), Jca is estimated to be less than 2.0 1013 cm2 s1. Thus,
Table 2
Parameter values used in calculating cumulative probabilities in the breakdown potential for Type 403 SS in
deaerated borate buer solution (pH = 8.1 0.1) with dierent chloride concentrations at 22 C
Parameter

Value

Dimensions

Source

F, Faradays constant
R, the gas constant
T, the absolute temperature
Nv, Avogadros number
v, the barrier layer
stoichiometry (Cr2O3)
X, molar volume of
Cr2O3 per cation
e, the electric eld strength
Jm, the rate of annihilation
of cation vacancies at the
metal/lm interface
n, the critical areal cation
vacancy concentration
D; the mean cation vacancy
diusion coecient
rD, the standard deviation for
the cation vacancy diusivity
w v2 DG0A DG0s  v2 F /0f=s
a, the dependence of /f/s on
the applied voltage
vaF
c0 2RT
b, the dependence of /f/s on pH

96,487
8.3144
295.15
6.023 1023
3

C/equiv
J/(mol K)
K
no./mol

Fundamental constant
Fundamental constant
Dened
Fundamental constant
Assigned

14.59

cm3/mol

From density

5 106
1013

V/cm
no./(cm2 s)

Assigned [29]
From Fig. 4

3.5 1014

no./cm2

From Fig. 4

5 1020

cm2/s

Assigned [8,9]

2.5 1020 or 1.25 1020

cm2/s

From tting in Fig. 5

23.4
0.23

kJ/mol

From Fig. 2
From Fig. 2

13.56
0.01

V1
V

Calculated
From Fig. 3

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

3821

we nd: Jm < 2.0 1013 cm2 s1 and hence n < 7.0 1014 cm2. On the other hand,
assuming that metal vacancies are the condensing species, as depicted by Reaction (11),
n = 7.0 1014 cm2.
The critical areal vacancy concentration n (cm2) that leads to passivity breakdown can
be calculated from the unit cell dimension of the metal and/or the barrier layer, depending
upon which lattice condensation is envisioned to occur. Note that the unit cell for the
metal (taken to be CrxFe1x, x = 01) is body-centered cubic with a = 0.2860
0.2897 nm [34] and the unit cell for the barrier layer (Cr2O3) is rhombohedral with
a = 0.493 nm and c = 1.358 nm [35]. Thus, the density of Cr/Fe atoms per unit area in
a monolayer of the unit cell is, respectively, 1015 cm2 for the metal and 1014 cm2
for the barrier layer.
In both cases, the value of n estimated from the experimental relationship between Vc(t)
and t1/2 is in close agreement with that calculated from the geometric structure of the substrate alloy (CrxFe1x, x = 01) and the barrier layer (Cr2O3). Recognizing that the barrier

Fig. 5. Calculated cumulative probabilities in the breakdown potential for Type 403 SS in deaerated buer
solution (pH = 8.1 0.1) with dierent chloride concentrations, compared with the experimentally-determined
distributions. The solid lines are the calculated distributions with rD 0:25D (above) or rD 0:50D (below),
while the marked points are the experimental data.

3822

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

layer is n-type in electronic character, and hence that the dominant defect is likely to be
the cation interstitial, it is likely that condensation occurs on the metal lattice on the
alloy side of the metal/lm interface, although it is strictly not possible to dierentiate
the two possibilities at the present time. However, in either case, the level of agreement
is suciently good to provide a convincing evidence for the validity of the PDM for
describing passivity breakdown on Type 403 SS in chloride-containing borate buer
solutions.
By using Eqs. (6)(8), and the parameter values extracted from the above experiments
(Table 2), the statistical distribution in the breakdown potential for Type 403 SS in deaerated borate buer solution with dierent chloride concentrations and at pH = 8.1 0.1
can be calculated. The calculated distributions using rD 0:25D (above) and rD
0:50D (below) are plotted in Fig. 5 as cumulative probabilities in the breakdown potential,
and are compared with the experimental distributions in the breakdown potential obtained
from the potentiodynamic experiments performed in this study. Typically, these experiments were repeated 20 times, in order to obtain statistically valid distributions. The
marked points are experimental data while the solid lines correspond to the calculated
distributions.
When the chloride concentration is 0.50 or 0.28 M, the experimental distributions are
narrow, and hence in better agreement with the calculated distribution functions with
rD 0:25D than with rD 0:50D. However, in the case of the chloride concentration
being 0.028 M, a better t is obtained with rD 0:50D. Therefore, while the eect is
not large, it seems likely that increasing chloride concentration tends to narrow the distribution in the breakdown potential. In other words, increasing chloride concentration
tends to decrease the standard deviation, rD, in the defect diusivity D at the breakdown
sites. In any event, the experimental distributions in the breakdown potential are in
satisfactory agreement with those calculated from the PDM with rD 0:5D or
rD 0:25D.
4. Summary and conclusions
Experimental relationships between the breakdown potential (Vc) and chloride activity,
pH and potential scan rate have demonstrated the validity of the point defect model
(PDM) for describing passivity breakdown on Type 403 SS in chloride-containing borate
buer solution. Analysis of the dependence of the breakdown potential on chloride activity
yields a value for the polarizability (a) of the barrier layer/solution interface. This value is
found to be signicantly smaller than those previously determined for Type 316 SS
(a = 0.690.71) and Type 304 SS (a = 0.8).
The critical cation vacancy concentration n leading to passivity breakdown on Type 403
SS is of the order of 7.0 1014 cm2, as determined from the expression for the dependence
of the apparent breakdown potential on the potential sweep rate given by the PDM. This
is in excellent agreement with that (10141015 cm2) estimated structurally, assuming
vacancy condensation on either the alloy substrate (CrxFe1x, x = 01) or on the cation
sublattice of the barrier layer (Cr2O3).
The experimentally-determined, near normal distributions in the breakdown potential
for dierent chloride concentrations are found to be in satisfactory agreement with theory
in the quantitative characterization of the breakdown potential distributions for Type 403
SS as calculated from the PDM.

Y. Zhang et al. / Corrosion Science 48 (2006) 38123823

3823

Acknowledgements
The authors gratefully acknowledge the support of this work by the Electric Power
Research Institute, Palo Alto, CA under Contract No. EP-P1150-C434.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]

O. Jonas, Materials Performance 24 (2) (1985) 918.


O. Jonas, J.M. Mancini, Materials Performance 40 (3) (2001) 4853.
S. Zhou, A. Turnbull, Corrosion Engineering, Science and Technology 38 (2) (2003) 97111.
D.D. Macdonald, M. Urquidi-Macdonald, Corrosion 48 (5) (1992) 354367.
D.D. Macdonald, Pure Applied Chemistry 71 (6) (1999) 951978.
L.F. Lin, C.Y. Chao, D.D. Macdonald, Journal of Electrochemical Society 128 (1981) 11941198.
C.Y. Chao, L.F. Lin, D.D. Macdonald, Journal of Electrochemical Society 128 (1981) 11871194.
D.D. Macdonald, M. Urquidi-Macdonald, Electrochimica Acta 31 (8) (1986) 10791086.
M. Urquidi-Macdonald, D.D. Macdonald, Journal of Electrochemical Society 134 (1987) 4146.
M. Urquidi-Macdonald, D.D. Macdonald, Journal of Electrochemical Society 136 (1989) 961967.
D.D. Macdonald, Journal of Electrochemical Society 139 (1992) 34343449.
D.D. Macdonald, D.F. Heaney, Corrosion Science 42 (2000) 17791799.
Z. Szklarska-Smialowska, Pitting Corrosion of Metals, NACE, Houston, Texas, 1986.
T. Shibata, T. Takeyama, Corrosion 33 (7) (1977) 243251.
T. Haruna, D.D. Macdonald, Journal of Electrochemical Society 144 (1997) 15741581.
I.T.E. Fonseca, N. Lima, J.A. Rodrigues, M.I.S. Pereira, Electrochemistry Communications 4 (2002) 353
357.
S. Silverman, G. Cragnolino, D.D. Macdonald, Journal of Electrochemical Society 129 (1982) 24192424.
D.D. Macdonald, Transient Techniques in Electrochemistry, Plenum Press, New York, 1977.
V.M.M. Lobo, Electrolyte solutions: Literature Data on Thermodynamic and Transport Properties,
Coimbra, Portugal, 1984.
D. Dobos, Electrochemical Data: A Handbook for Electrochemists in Industry and Universities, Elsevier
Scientic Publishing Company, AmsterdamOxfordNew York, 1975.
T. Piao, S.-M. Park, Journal of Electrochemical Society 144 (1997) 33713377.
G. Lorang, M.D.C. Belo, A.M.P. Simoes, M.G.S. Ferreira, Journal of Electrochemical Society 141 (1994)
33473356.
A.M.P. Simoes, M.G.S. Ferreira, G. Lorang, M.D.C. Belo, Electrochimica Acta 36 (2) (1991) 315320.
R. Kirchheim, B. Heine, H. Fischmeister, S. Hofmann, H. Knote, U. Stolz, Corrosion Science 29 (1989) 899
917.
G.M. Bulman, A.C. Tseung, Corrosion Science 12 (1972) 415432.
C.B. Breslin, D.D. Macdonald, J. Sikara, E. Sikora, Electrochimica Acta 42 (1) (1997) 127136.
D.D. Macdonald, E. Sikora, M.W. Balmas, R.C. Alkire, Corrosion Science 38 (1996) 97103.
W.H. Hayt Jr., Engineering Electromagnetics, fourth ed., McGraw-Hill Book Company, 1981.
Y. Zhang, Determination of Damage Functions for the Pitting of AISI 403 Blade Alloy and ASTM A470/
471 Disk Alloy, Ph.D. Thesis, The Pennsylvania State University, 2005.
E. Cho, H. Kwon, D.D. Macdonald, Electrochimica Acta 47 (2002) 16611668.
A.M.P. Simoes, M.G.S. Ferreira, B. Rondot, M.D.C. Belo, Journal of Electrochemical Society 137 (1990)
8287.
H. Tsuchiya, S. Fujimoto, Science and Technology of Advanced Materials 5 (12) (2004) 195200.
S. Yang, D.D. Macdonald, Theoretical and Experimental Studies for the Pitting of Type 316L Stainless Steel
in Borate Buer Solution Containing NO
3 , in preparation.
P. Villars, L.D. Calvert, second ed.Pearsons Handbook of Crystallographic Data for Intermetallic Phases,
vol. 2, ASM International, Materials Park, Ohio, 1991.
W.P. Davey, Physical Review 21 (1923) 716.

You might also like