Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Modeling Intergranular Corrosion Spreading of Intergranular

Corrosion on Stainless Steels and AA5083 alloys with


Experimental Verifications
Swati Jaina, Noah D. Budianskyb, J.L. Hudsona, J.R. Scullyb,*
a
Department of Chemical Engineering
b
Department of Materials Science and Engineering
University of Virginia, Charlottesville, VA-22903, USA
* Corresponding author. E-mail: jrs8d@virginia.edu

ABSTRACT
The spreading of intergranular corrosion was investigated by both micrometer scale
simulations and experimental verifications on sensitized AISI 304L [UNS S30403] stainless
steel. The degree of sensitization, presence of a pit, and applied potential all affected
spreading. The parameters/inputs used in a micrometer scale model of a sensitized surface
containing grains and grain boundaries were obtained from model Fe (Bal-X)-Cr-Mo-Ni alloys
simulating various grain boundary Cr depletion levels. Corroding grain boundaries and pits
triggered the corrosion of nearby sensitized boundaries due to Ohmic potential drop. Large
connected clusters of corroding grain boundaries formed at high fractions of Cr depleted grain
boundaries.
This paper has concentrated on the intergranular corrosion spreading on the surface of
sensitized stainless steels. The talk will also concentrate on the IGC spreading on sensitized
AA5083. These alloys become susceptible to intergranular corrosion upon precipitation of
the β phase (Al3Mg2) in the grain boundary region upon natural and artificial aging. The
intergranular corrosion is initiated in several ways including pitting and/or active dissolution of
the β phase and Spreading occurs across surfaces because of the changes in the potential
and concentration fields.
Keywords: Intergranular corrosion, modeling studies, stainless steel

BACKGROUND
Carbon tends to diffuse toward grain boundaries in austentic stainless steels during
material preparation or improper heating in the temperature range of 425oC to 815oC1. This
leads to precipitation of (Fe,Cr)23C62 adjacent to the grain boundaries. The chromium
concentration in the grain boundary region is reduced while the chromium concentration in the

1
interior of the grain remains unchanged This process is referred to as sensitization. The
chromium depleted grain boundary region is less resistant to corrosion than the interior of the
grain. If the chromium concentration falls below a critical value, preferential corrosion takes
place at the grain boundaries while the interior of the grains remains passive under certain
electrochemical conditions. This process is known as intergranular corrosion.
The depletion theory3 described above is supported by ample evidence. With the aid of
extraction replicas and thin metal foils of 05Cr18Ni9 steel (0.05 wt % C,18.48 wt% Cr, 9.34
wt % Ni, 1.5 wt% Mn, 0.62 wt% Si an balance Fe) for electron microscopy, Cihal et al4 proved
that creation of continuous chromium depleted grain boundaries is the primary cause of the
intercrystalline attacks. Theoretical considerations and experiments showed that the chromium
depletion theory is capable of interpreting most of the intergranular corrosion effects5. AEM
(Auger Electron Microscopy) studies and Strauss tests showed that during the intergranular
corrosion in susceptible stainless steel alloys the shape of the corroded region and the depth
of the attack can all be explained by chromium depletion6. A correlation study between
chromium depletion at grain boundaries and IGSCC also showed that the susceptibility to
cracking increases with the extent of the chromium depletion zone7.
The precipitation of the chromium carbides is process in which diffusion of carbon plays
a critical role; therefore, the ageing conditions, i.e. the temperature and time of exposure to
heat, play a critical role in determining the amount of chromium in the grain boundaries after
sensitization 8-11,12,13, 14. The fraction of chromium in the region adjacent to the grain boundary
decreases with the increase in sensitization time in the temperature range where carbide
formation occurs 9, 12, 14.
A study on AISI 304 stainless steel samples, heat treated at 600oC for various times
after recrystallization at 1100oC, showed that the chromium content in the grain boundaries
remains more than 13.2 wt% for short heating times. However, for long heating times the
chromium content in the grain boundary regions can become lower than 10 wt% 9. Finely
distributed carbide precipitates formed on the grain boundaries in Inconel 600 (10 wt% Fe,
73.6 wt% Ni, 16 wt% Cr, 0.02 wt% C) for heat treatments at lower temperatures or for shorter
times of exposure to heat. As the heat treatment was prolonged or the heat treatment
temperature was increased, semi continuous or large discretely distributed carbides were
observed 15, 16.
A minimum critical Cr content is required in the grain boundaries for the material to
remain resistant to intergranular corrosion 9, 17 13. The intergranular corrosion rate increases
sharply if the grain boundary chromium content decreases below 12 wt% in samples of Ni-
16%Cr-9%Fe 17. Bond percolation theory and laboratory experiments showed that if the
fraction of grain boundaries with less than or equal to 13.2 wt% Cr increases above a critical
level in AISI 304 (UNS S30400) stainless steel, it may become susceptible to intergranular
corrosion and IGSCC 8, 9.
Intergranular corrosion due to chromium depletion can be explained by electrochemical
principles. Anodic polarization behavior of Fe-Cr alloys with different chromium contents, that
simulate various grain boundary depletion levels, in reducing acids show that the primary
passivation potential, the Flade potential, the anodic current density, and the passive current
density are all a function of the Cr content 1, 18-24. Alloys with high Cr content have lower Flade
potentials, low critical current densities, and a wider range of passivity. Therefore, sensitized
grain boundaries with lower concentrations of chromium will be more susceptible to corrosion
and the rate of grain boundary dissolution will be higher.
Several environmental factors can also influence the susceptibility to intergranular
corrosion. The bulk chemical composition, the oxidizing power or reducing acid strength, and
the applied potential are some factors that can determine whether IGC ((intergranular

2
corrosion) occurs. The presence of another anodic current source, like a pit 25-27, can also be
associated with the initiation of cracks and intergranular corrosion along the pit walls. This may
occur in part due to formation of a highly aggressive solution concentration around the pit
and/or potential drop due to the current in pit. However, the critical details of this process are
unclear.
A critical question is whether a large connected cluster of corroding grain boundaries
can form and what are the environmental and metallurgical conditions that control such
spreading and penetration. Some simulation studies using the Monte Carlo method 28 29, and
percolation models 30 31, and some experimental studies 8, 32 have concentrated on the growth
of a crack into or through the material as a function of percentage of active grain boundaries,
or the orientation of the grain boundaries. The most common mathematical tool used in these
works is based on bond percolation theory which states that the probability of forming an
infinite cluster of connected bonds will approach to one above a critical percentage of active
bonds 33. A study focusing on the effect of varying the grain shape showed that the percolation
threshold changes from 0.711 in a triangular lattice to 0.653 in a standard hexagonal lattice 34.
Developing a new approach based on percolation theory and consideration of crack paths as a
Markov chain, Gertsman et. al 30 analyzed the length of the cracks as a function of percentage
of active grain boundaries. They found that above a critical fraction of resistant grain
boundaries, the crack length may reach a large but maximum value. The actual critical
percentage may change depending on the direction of growth and the nucleation conditions.
Model and experimental studies also showed that when greater than 23% active grain
boundaries were present, a significant increase in the susceptibility to intergranular fracture
was observed 8. Many theories and model systems presented in literature mostly discuss the
minimum fraction of resistant grain boundaries required for immunity to intergranular stress
corrosion cracking and propagation of cracks.
However, there is a lack of model and experimental studies that focus on initiation and
surface spreading of intergranular corrosion in stainless steel and the effect of several
metallurgical, electrochemical, and environmental parameters such as the percentage
distribution and arrangement of the chromium depleted grain boundaries, the shape and size
of the grain boundary, presence of nearby pits, the conductivity of the solution, and the applied
potential on intergranular corrosion surface spreading.
In this paper we report a study using a micrometer scale model complemented by
experimental verifications on AISI 304L stainless steel (UNS S30403). The aim was to study
the spread of intergranular corrosion and the role of the parameters mentioned above.
Verifying experiments were conducted on AISI 304L stainless steel alloys that were sensitized
to various degrees of sensitization (DOS) using a custom designed dual electrode. A stable pit
(AISI 304 SS (UNS S30400)) was placed in the center of a planar sensitized electrode. The
model for surface spreading presented is based on electrochemical principles. As an input to
the model, E-I data in Fe-XCr-0.43Mo-9Ni (0.43 wt % Mo, 9.0 wt% Ni, and the balance is Cr
and Fe) alloys of different Cr contents were determined experimentally. Data from the literature
on distribution of Cr content in AISI 304 SS after sensitization were also used 9.

EXPERIMENTAL SET-UP AND PROCEDURE

Materials
Intergranular corrosion spreading experiment. Intergranular surface spreading experiments
were conducted on AISI 304 (UNS S30400) and AISI 304L (UNS S30403) stainless steel
alloys (Compositions in Table 1) in the form of an embedded wire electrode consisting of a

3
planar electrode with a electrically isolated center wire. The center wire consisted of a 250 µm
AISI 304 stainless steel with carbon concentrations of 0.06 wt%. The planar electrode
consisted of a 0.635 cm diameter AISI 304L stainless steel with a carbon concentration of 0.02
wt%. All materials were received with no heat treatment. Materials were encapsulated in
quartz glass tubing with a triple argon purge and back filled with 15 mm Hg of argon. Both the
304 stainless steel wire and 304L stainless steel planar electrodes were annealed at 1030oC
for 1 h and then the AISI 304 L SS was sensitized at 621oC for various times while the AISI
304 SS wire was annealed for 4 days at 621oC. Materials were heat treated in a digitally
controlled muffler furnace, removed from the furnace after the specified time and allowed to air
cool for approximately 10 min then water cooled to room temperature. The ASTM grain size of
the planar electrode is 6.9 based on ASTM E-112.
Fe-Cr-Ni-Mo alloys were used to investigate the intergranular corrosion kinetics used for
inputs into the IGC model. Different chromium concentration (Cr concentration = 10 wt%, 12
wt%, 14 wt%, and 16 wt%) Fe-XCr-9Ni-0.43Mo alloys were produced by Ames Material
Synthesis Laboratory, Iowa. The electrodes were in the form of wires of diameter of 0.8 mm
exposed to the solution and were received as annealed in a vacuum.

TABLE 1
CHEMICAL COMPOSITION OF AISI 304 SS (UNS S30400), AISI 304L SS (UNS S30403),
AND Fe-XCr-Mo-Ni. ALL VALUES ARE REPORTED IN WEIGHT PERCENT..
C Si Mn Cr Mo Ni S Fe

304 SS 0.060 0.25 1.0 18.23 0.3 8.08 0.018 Bal


304 L 0.02 0.60 1.67 18.12 0.25 8.56 0.025 Bal
SS
Fe- 0 0 0 10-16 0.43 9 0 Bal
XCr-Mo-
Ni

Electrode configuration for experimental verifications


A flush mounted embedded wire electrode with a AISI 304 (UNS S30400) placed in the
center of a flush mounted rod AISI 304L (UNS S30403) was used (Figure 1) to study the
material parameters associated with the initiation and growth of IGC. The center wire could be
forced to undergo pitting corrosion by controlling its potential above the pitting potential while
the outer electrode can be kept at a different potential, using a multi-potentiostat system
capable of imposing different potentials to different electrodes immersed in the same solution.
The wires were coated with PPG cationic paint to ensure electrical isolation from the
outer planar electrode to allow the separate inventory of currents. The wires were inserted into
a 305 µm (0.012 inch) diameter hole drilled in the center of the outer planar electrode. The
planar electrodes were coated with PPG cationic paint to prevent crevice corrosion around the
outer edge. The completed embedded wire electrodes were flush mounted in epoxy and
polished from #320 grit to #1200 grit grinding paper followed by 3 µm and 1 µm diamond slurry.
The total exposed area of the planar electrode was 0.316 cm2 while the inner wire had an
exposed cross sectional area of 4.91x10-4 cm2.

4
The Fe-XCr-0.43Mo-9Ni alloys in solid solution form were mounted in an epoxy with a
cross section area of 0.0050 cm2 exposed to the solution. The electrodes were polished from
#320 grit to #1200 grit grinding paper.

Figure 1 - Optical micrograph of an embedded wire electrode. The center 250 µm wire is
inserted in a 305 µm drilled center hole in outer electrode. The center wire and the outer planar
electrode are electrically isolated from one another.

Experimental Procedure
Electrochemical Characterization of the Degree of Sensitization. Sensitized AISI 304L
stainless steel planar electrodes with different sensitization times were examined in standard
ASTM G-108 solution (0.5 M H2SO4 + 0.01 M KSCN at 30oC) to determine the degree of
sensitization by the SLEPR method.
Intergranular spreading experiments: Experiments were conducted in 0.0158 M H2SO4 +
0.05 M NaCl + 0.01 M KSCN at 30oC (pH 1.5)(1). The electrochemical sequences consisted of
1) the electrodes were held potentiostatically at –0.6 VSCE for 30 s to remove any air formed
oxide films, 2) both electrodes were held at open circuit potential for 10 min, 3) the sensitized
outer planar electrode was held at +0.1 VSCE for 5 min to passivate the surface while the center
wire remained at OCP, and 4) the center wire was held potentiostatically at +1 VSCE to initiate
and grow large pits across the entire surface while the planar electrode was potentiostatically
held at a potential in the passive range. Intergranular corrosion could be triggered and spread
during this stage. The final step (step 4) was conducted for 2 min, 5 min, and 10 min. The
potential time sequence is displayed in Figure 2.

(1)
The low pH solution was aggressive enough to support intergranular corrosion so that no increase in solution

enhancement was required for intergranular corrosion interactions to occur. Thus only Ohmic potential drop was required for

intergranular corrosion to occur.

5
Figure 2 - Potential-time sequences for the center wire and the planar electrode for an
embedded wire electrode interaction experiment used for investigating the effect of Ohmic
potential drop.

Experiments on Fe-XCr-9Ni-0.43Mo alloys. The determination of the polarization behavior


and the passivation characteristics of Fe-XCr-9Ni-0.43Mo alloys were done in a solution of
0.0158 M H2SO4 + 0.01 M KSCN + 0.01 M NaCl solution. The sequence to obtain the
polarization curve was: each electrode was 1) held potentiostatically held at –0.6 VSCE for 30 s
to remove air formed oxide films, 2) held at open circuit potential for 10 min, 3) passivated at
0.2 VSCE for 5 min, and 4) scanned at 0.1 mV/s from 0.2 VSCE to -0.6 VSCE. E-I plots were
constructed with minimal Ohmic drop enabled by electrode area of 5 x 10-5 cm2.
The time constant for the reactivation or depassivation of the oxide was determined
using the following time sequence. Each electrode was 1) held potentiostatically at –0.6 VSCE
for 30 s to remove the air formed oxide film, 2) then held at open circuit potential for 10 min, 3)
passivated at 0.2 VSCE for 5 min, 4) then polarized at or below the Flade potential (the Flade
potential was defined as the potential for which the change in current was at least two orders
of magnitude per 1 Volt drop in the potential). and the current density was allowed to reach the
maximum current density, and 5) then at a potential of 0.2 VSCE in order to passivate.

IGC SPREADING MODEL


Sensitized stainless steel electrode surfaces were simulated with a grid consisting of 32
µm x 32 µm square grains with a Cr depleted region width of 2 µm. Each grain boundary was
assigned a Cr concentration depending on the sensitization time for the electrode. The
distribution of chromium depleted grain boundaries that was applied to each simulated
electrode was based on the data obtained through a study on AISI 304 SS that was sensitized
at 1100oC for 24 h + 600oC for various times of sensitization 9. The data report cumulative
distribution of different Cr content grain boundaries as a function of sensitization conditions
and the degree of sensitization.
The rate of dissolution of a local grain boundary is a function of the interface potential
that is defined as the difference between the remote applied potential and the Ohmic voltage,
near the grain boundary. If the potential drops to values lower than the Flade potential for a
given chromium concentration for a local grain boundary, it can
starts to corrode. The interface potential may change if there is Ohmic potential drop due to a
current emitting source like a nearby pit or any other form of corroding surface.
Other factors that control the rate of corrosion or reactivation rate of a sensitized
polycrystalline material could be the concentration of aggressive species (like KSCN35, 36, NaCl
1, 37
), change in pH20, 38, grain shape, grain size, and orientation21, 32. The model was derived

6
for data obtained at very low pH where solution enhancement was not required; therefore, the
effect of variability of concentration of aggressive species on the surface of the electrode was
ignored in this first investigation.

Physical Setup of Model

Ohmic Potential drop. The potential in the solution will drop if a current carrying source
like a nearby pit is present. It should be noted that if highly susceptible grain boundaries start
to corrode, the current they emit will also contribute to the total Ohmic potential drop in the
solution.

a) Presence of a Pit
Localized pitting events have been shown to cause initiation of cracks and intergranular
corrosion along the pit walls25-27. This could be due to the increase in concentration of the
aggressive ions or Ohmic potential drop around the pit. The potential drop associated with the
pit can lower the remote potential near the grain boundaries. If the potential drop is large
enough such that EOCP ≤ EInterface ≤ EFlade for a particular chromium concentration at a grain
boundary, it will depassivate and corrode.
Using the Newman’s solution for potential drop generated from a flush mounted planar
disk in an insulated surface 39, 40, the potential drop due to a disk with radius a and current I
can be computed as:

I  2 −1 
φ= 1 − tan (ξ )  (1)
4κa  π 

where, κ is the conductivity and ξ is the elliptical rotational distance from the center of the
disk.

b) Potential drop due to the corroding grain boundaries


The corroding grain boundaries will change the solution concentration and the potential
in the solution near the electrode surface. Each grain boundary is modeled as a chain of units
of 2 µm diameter disks. The potential drop due to these disks can be computed using equation
1.
For each unit (x,y) on the surface of the electrode surface the interface potential may be
lower than the applied potential because of the pit and the corroding grain boundaries.

E I int erface ( x, y, t ) = V app − ϕ pit ( x, y ) − φ gb ( x, y, t ) (2)

where,

7
I j (t )  2 
φ gb ( x, y, t ) = ∑ 1 − tan (ξ j ( x , y ) )
−1
(3)
j 4κa  π 

where, Ij is the current in the j-th unit, a = 1 µm, ξj(x,y) is the elliptical rotational distance from
the j-th corroding unit from the point at location (x,y).

Reactivation of the grain boundaries. The electrode was initially kept at a potential at
which it exhibits passivity. Once the interface potential is stepped down to a potential below the
Flade potential of the grain boundaries with the lowest Cr concentration, the oxide starts to
dissolve till a new dynamic equilibrium was reached under active conditions. The time constant
for the reactivation of the grain boundaries was determined from the potential step down
experiments described in the previous section. The current time series after the potential was
stepped down until it reaches a maximum value is fitted with the function

I (t ) = I max (1 − exp(−(t − t d ) / τ )) t ≥ td
if (4)
I =0 t < td

where, td is the time at which the potential is stepped down. The constant “τ” will be referred
to as the reactivation or depassivation time constant. The constant τ is estimated by fitting
above equations, using the least square error minimization method, to the current signals
obtained from step down experiments described in the earlier section.

Inputs and parameters


The Fe-XCr-0.43Mo-9Ni alloys are assumed to simulate the behavior of the respective
chromium depleted grain boundaries distributed across an electrode. The inputs/parameters
required for the simulations are the chromium distribution levels on a grain by grain basis as a
function of sensitization time, E-log I behavior of the Fe-XCr-0.43Mo-9Ni alloys (X= 12, 14, 16,
and 18.2) obtained under the similar conditions as the intergranular corrosion spreading
experiments, reactivation time constants for the alloys, the applied potential, the pit current, the
pit radius, and the conductivity of the solution.
The cumulative percentage of the different concentration chromium levels in the
intergranular region were estimated using the data obtained from 9. For any given DOS there is
a distribution of Cr depleted grain boundaries with particular wt% or lower chromium
concentration. The cumulative probability of finding grain boundaries depleted to or below a
particular wt % Cr, for materials with different degrees of sensitization, that has been used as
an input to the model are shown in Figure 3(a). After heat treatments which cause sensitization
the chromium level falls below 18.2 wt% (the bulk concentration of chromium in the
unsensitized AISI 304 SS). For the most severely sensitized electrode (DOS = 30C/cm2 – 45
C/cm2) a fraction of the grain boundaries is predicted to have Cr concentrations below 10 wt%.

8
For the least sensitized electrodes (DOS = 0 C/cm2 – 15 C/cm2) the chromium concentration in
the grain boundaries were predicted to be greater than13.2 wt%. Note that the 13.2 wt% Cr
concentration is close to the critical concentration of 12 wt% required to maintain passivity in
stainless steels 41 42, 43. The proportion of the grain boundaries with less than or equal to 13.2
wt% Cr, which represents the susceptible grains boundaries to intergranular attack, as a
function of DOS is shown in Figure 3(b).
Experimental E-log I behavior of different Fe-XCr-0.43Mo-9Ni alloys (X= 12, 14, 16, and
18.2) were obtained at pH 1.5 (0.0158 M H2SO4 + 0.01 M NaCl + 0.01 M KSCN) at 30oC as
described in the previous section. Using these curves (Figure 3(b) the Cr-concentration
dependent current density has been estimated as a square function of potential as (Schematic
in Figure 3(c)).

I = Ipeak if EOCP ≤ EInterface ≤ EFlade


else I=0 all other conditions

An exponential curve (Equation 4) was fit to the results of the reactivation experiments to
determine “τ”.

9
(b)
(a)
0
10
10 % Cr
12 % Cr
14 % Cr
Current density(A/cm2)

16 % Cr
-2
10
304 SS (18.2 wt % Cr)

-4
10

-6
10

-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2


V (V )
app SCE

(c)
Figure 3 - Inputs to the model: (a) Distribution of grain boundary chromium content as a
function of sensitization time for sensitized AISI 304 stainless steel 9 (1100oC for 24 h + 600oC
for various times). (b) Proportion of active grain boundaries as defined by depletion to less than
13.2 wt % Cr as a function of DOS. (c) Estimated current densities as a function of interface
potential (VSCE) for different concentrations of chromium in Fe-XCr-0.43Mo-9Ni alloys.

RESUTLS

Experiments to determine degree of sensitization of UNS S30400 (AISI 304 SS) and UNS
S30403 (AISI 304L SS)
The calculated degree of sensitization is non-linear with respect to the sensitization time,
as shown in Figure 4(a). Optical micrographs after ASTM G-108 experiments for electrodes
sensitized for 7 and 28 days, as shown in Figure 4(b) and Figure 4(c), respectively. At low
sensitization times (12 h and 24 h) no grain boundary etching was observed and only localized
damage was observed; note this was probably from isolated carbides. A small fraction of
attacked grain boundaries is visible after the 7 days sensitization heat treatment, as shown in
Figure 4(b). A large fraction of attacked grain boundaries were observed at sensitization times
of 10 days. A very large fraction (almost all of grain boundaries) was attacked after 28 days

10
sensitization time. The connectivity of the grain boundaries increases with increasing
sensitization time directly correlating with the measured degree of sensitization.

Figure 4 - Degree of sensitization vs. sensitization time for sensitized (1030oC for 1 h + 621oC
for various times) AISI 304 L (UNS S30403) stainless steel with different sensitization times in
0.5 M H2SO4 + 0.01 M KSCN (ASTM G-108 solution at 30oC, pH 0). The optical micrograph of a
low sensitized (Figure 4(b): DOS = 27.7 C/cm2) and a high sensitized (Figure 4(c): DOS = 36.5
C/cm2) steel.

Characterization of the E-I behavior of Fe-XCr-0.43Mo-9Ni alloys


Figure 5 shows the polarization scan for the Fe-XCr-Mo-Ni alloys obtained in a solution
of 0.0158 M H2SO4 + 0.01 M NaCl + 0.01 M KSCN solution at T = 30oC. The alloy with 10 wt%
Cr did not passivate in this solution and probably does contain IR drop. As the Cr content was
increased, both the Flade potential and the peak current density decreased. The current in the
passive state showed some spikes due to some metastable breakdown events caused by the
modified ASTM G-108 solution with the addition of NaCl. The solution has large amounts of
chloride and thiocyanate ions that dissolve the oxide film and may have induced breakdown
events.
To determine the reactivation time constants potentiostatic experiments were done as
described in the experimental section. The potential time sequence and the current time
sequence, for the alloy Fe-16Cr-9Ni-0.43Mo, thus obtained are plotted in Figure 6(a) and
Figure 6(b), respectively. The approximated values of reactivation time constants, τ for
different Cr concentration alloys are shown in Table 2. The reactivation time constant

11
decreases with decreasing Cr concentration, implying that lower chromium contents affects the
oxide stability. The reactivation time constant was determined for the grain boundaries with 12
wt% Cr, 14 wt% Cr, 16 wt% Cr, and 304 SS and obtained by linear interpolation for any other
compositions of interest.

0
10
10 % Cr
12 % Cr
Current Density (A/cm )
2

-2
14 % Cr
10 16 % Cr

-4
10

-6 304 SS (18.2% Cr)


10

-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2


Vapp (VSCE)

Figure 5 - Polarization scan for different Cr concentration Fe-XCr-0.43Mo-9Ni alloys in a


solution of 0.0158 M H2SO4 + 0.01 M NaCl + 0.01M KSCN at 30oC. Scan rate = 0.01 mV/s.

(a) (b)
Figure 6 - Reactivation of Fe-XCr-0.43Mo-9Ni alloys in a solution of 0.0158 M H2SO4 + 0.01 M
NaCl + 0.01 M KSCN at 30oC. (a) Potential time series to determine the reactivation rate. (b) The
anodic current time series for potential time series in (a) for alloy with 16 wt% Cr.

12
TABLE 2
TIME CONSTANT FOR REACTIVATION OF DIFFERENT CHROMIUM
CONCENTRATION Fe-XCr-0.43Mo-9Ni ALLOYS. THE CONSTANTS ARE DETERMINED
BY FITTING EQUATION 4 TO THE RESULTS OF THE EXPERIEMENTS ON THE ALLOYS

Active wt % of Cr Concentration in the Fe- τ (s)


Cr-Ni-Mo alloy

10 % 1.00
12 % 15.87
14 % 166.67
16 % 400.00
18 % 1000.00

Intergranular spreading experiments from an embedded wire pit


Induced interaction experiments (described in experimental set up section) were
conducted on embedded wire electrodes surrounded with a rod with degrees of sensitization of
16.7, 27.6, 27.7, 29.6, and 36.5 C/cm2 in 0.0158 M H2SO4 + 0.05 M NaCl + 0.01 M KSCN (pH
1.5). A pit was initiated on the center wire as indicated in Figure 1. The applied potential time
sequence was shown in Figure 2. The currents for the planar electrodes sensitized to different
degrees of sensitization, after the center wire was initiated and the planar electrode was kept
at Vapp = -0.075 VSCE, are shown in Figure 7. The pit current density is indicated by the dotted
line.
For the 16.7 C/cm2 degree of sensitization heat, the anodic current from the
intergranular current dissolution (IIGC) increased slowly with time and after 10 min was
approximately the same as the current for the pitting wire, IPit, as shown in Figure 7. Similar
results were observed for the alloy with a degree of sensitization of 27.7 C/cm2. The current
time series indicates that IIGC increased slowly with time and after 10 min it was approximately
the same as IPit. The optical photomicrographs of the IGC growth for the electrode with DOS =
27.7 C/cm2 after 2 min and 10 min of exposure are shown in Figure 8(a) and Figure 8(b),
respectively. Very fine grain boundary etching, as a small ring around the initiating wire, is
visible after 2 min of exposure and the IGC damage zone did not grow larger with time.
Small bursting events in the current were observed for the 29.6 C/cm2 DOS electrode.
IIGC increased gradually with time and exceeded IPit. Large bursting events and very rapid
increase in anodic current were observed for the 36.5 C/cm2 DOS material. The intergranular
current-time curve for 2 min experiments showed a very rapid increase in anodic current to
approximately 5 x 10-3 A/cm2 before increasing at a slower rate. Optical photomicrographs for
the electrode with DOS = 36.5 C/cm2 at t = 5 min and t = 10 min are shown in Figure 8(c) and
Figure 8(d), respectively. The zone of IGC damage grew very quickly across the outer
electrode to approximately half the electrode radius at t = 5 min. As the time was increased,
the intergranular damage zone reached the edge of the electrode portion that was exposed to
the solution.
Current-time series (Figure 7) curves for 10 min experiments show three distinct
behaviors. Low intergranular corrosion currents were observed for alloys with 16.7 C/cm2 and
27.7 C/cm2 DOS. IGC growth distance vs. time curves in Figure9 (a) indicates that the zone of

13
IGC damage didn’t increase with time. Higher IIGC current densities and slight growth of
damage regions were observed for 27.6 and 29.6 C/cm2 DOS materials (Figure9 (b)). Bursting
events and rapid increases in anodic current were observed for the 36.5 C/cm2 DOS. The
maximum current was reached within a short period of time after interaction experiments were
started (200 s) which is directly related to the rapid growth of IGC damage zone across the
electrode.

-2
10

-3
10
I (A)

-4
10
2
16.7 C/cm (4 days)
2
27.7 C/cm (7 days)
-5
10 2
27.6 C/cm (10 days)
2
29.6 C/cm (14 days)
2
36.5 C/cm (28 days)
Passive
-6 Pitting
10
1000 1100 1200 1300 1400 1500
Time (Sec)

Figure 7. Intergranular spreading experiments (potential time sequence in Figure 2): Vapp = -
0.075 VSCE. Current vs. time for the sensitized steel electrodes with difference degree of
sensitization; the sensitization time is shown in the bracket.

14
Figure 8 - Intergranular spreading experiments (potential time sequence in Figure 2): Vapp = -
0.075 VSCE. (a) and (b) Optical micrographs for the intergranular damage on an electrode with
DOS = 27.7 C/cm2 after 2 min and after 10 min exposure, respectively. (c) and (d) Optical
micrograph for the intergranular damage on an electrode with DOS = 36.5 C/cm2 after 5 min and
after 10 min exposure, respectively. Other parameters are same as in Figure 7.

15
3000
2
16.7 C/cm (4 days)
2
27.7 C/cm (7 days)
2500 2
27.6 C/cm (10 days)
IGC Damage Distance (µ m)

2
29.6 C/cm (14 days)
2
36.5 C/cm (28 days)
2000

1500

1000

500

0
0 100 200 300 400 500 600
Time (sec)

Figure 9 - Intergranular spreading experiments (potential time sequence in Figure 2) Radius of


intergranular damage as a function of time for different levels of sensitized steel at Vapp = -0.075
VSCE. Other parameters are same as in Figure 7.

Simulation Results
Conditions for the simulations. The degree of sensitization of experimental samples was
simulated by applying a distribution of chromium concentrations on grain boundaries as
determined in previous studies 9. Simulated electrodes with DOS levels of 10 C/cm2, 15 C/cm2,
20C/cm2, 25 C/cm2, 30 C/cm2, 37 C/cm2, and 45 C/cm2 were generated. The grain boundaries
locations associated with each particular DOS were assigned randomly. The distribution of Cr
depletion levels for these DOS was previously established (Figure 3(a)). Figure 10(a) and
Figure 10(b) show the positions and number of highly depleted grain boundaries indicated in
black (i.e., grain boundaries with less than or equal to 13.2 wt% Cr) for a low sensitized
electrode (DOS = 10 C/cm2) and a highly sensitized electrode (DOS = 37 C/cm2), respectively.
The simulated electrode size was 800 µm x 800 µm. Chromium depleted grain boundaries
tend to be connected in the case of highly sensitized electrode, whereas they are farther apart
for low sensitized material.

IGC in the presence of a pit. Simulations for an active pit in the center of the electrode and
were conducted for Vapp = -0.087 VSCE. The intergranular corrosion current time series for
different degrees of sensitization are plotted in the Figure 11. The broken horizontal line marks
the pit current. The grain boundary current was zero at t = 0 while the pit had attained the
maximum and constant current density. For the slightly sensitized electrodes with DOS = 10
C/cm2 and 15 C/cm2 and the medium sensitized (DOS > 15 C/cm2 and DOS ≤ 30 C/cm2 the
current from corroding grain boundaries was low as compared to that for the pit The current
rose initially and then reached a maximum value in less than 200 s. A sharp rise in the
intergranular corrosion current was observed for highly sensitized electrodes (DOS is greater
than or equal to 37 C/cm2). The current rose sharply and increased to values greater than that
for the pit in the simulation.
The intergranular damage at t = 0 min and after 10 min of simulation time for a low
degree of sensitized material (DOS = 10 C/cm2) are shown in Figure 12(a) and Figure 12(b),
respectively. In these figures all the grain boundaries are marked using gray lines and the

16
black lines represent the grain boundaries that have started to corrode or have corroded at that
time (The grain boundaries with EFlade ≥ EInterface in equation 2). The dark gray circle in the
center represents the pit. Initially, a circular region of corroded grain boundaries was seen near
the pit due to the Ohmic potential drop. Slight growth in this circular region was observed at t =
10 min; no growth was observed after this time. Figure 12(c) and Figure 12(d) show the
intergranular damage for a highly sensitized electrode (DOS = 37 C/cm2) at t = 0 min and after
10 min of exposure. Initially, similar to the case of the low sensitized electrode, a dense
circular region of corroded grain boundaries formed around the pit; note that all grain
boundaries with 10 wt% Cr should be corroding through the surface of the electrode. Therefore,
we mark them as black lines to understand the IGC growth. The dense region of corroding
grain boundaries grows with time till it reaches the edge of the electrode (Figure 12(d)).
A radius of intergranular corrosion damage (Rt) was defined as the distance from the center
of pit within which 95% of all the corroding grain boundaries are present at a time t. The plot of
radius of damage as a function of time for different levels of sensitization of steel is shown in
Figure 13(b). A sharp rise in the radius is observed at DOS = 37 C/cm2.
In the case of mildly sensitized electrodes (DOS = 10 - 15 C/cm2) the radius grew to about
10% of its initial value. The maximum radius was reached in less than 200 s. Some growth
was seen in the case of medium sensitized electrodes (DOS = 20 - 30 C/cm2) but the damage
did not grow to the edges of the electrode and stopped after some time. However, in the case
of highly sensitized electrodes, the damage grew until it reached the edges of the electrode.
Also, the rate of spreading was much higher in this case.

IGC with no pit. The above results demonstrate that the radius of intergranular corrosion
“zone” grows from a smaller radius to a large radius for a highly sensitized electrode when a pit
is present in the vicinity of the electrode. The next simulation covers the case where there is no
pit and few grain boundaries are initiated initially. The issue is whether the intergranular
damage will still grow without a pre-existing pit. The applied potential on the electrode was set
to lower values such that initially (t = 0) all the grain boundaries with EFlade ≥ Vapp start to
corrode at t = 0.
The result of such a simulation on an electrode of size 736 µm x 736 µm is shown in the
Figure 14. The growth of intergranular damage for a highly sensitized (DOS = 45 C/cm2)
electrode at Vapp = -0.133 VSCE (i.e. all the grain boundaries with 11.5 wt% Cr or less are
turned on initially). Figure 14(a) and Figure 14(b) show the intergranular damage for this
electrode at t = 0 min and t = 20 min, respectively. Initially some grain boundaries started to
corrode (Figure 14(a)), but, at t = 20 min a large number of grain boundaries were corroding
(Figure 14(b)). Several clusters of corroding grain boundary form originating from the initially
corroding grain boundaries formed on the surface. Figure 14(c) shows the normalized length of
the largest connected cluster, L at t = 20 min, for different DOS electrodes. The x-axis mark the
fraction of grain boundaries with less than or equal to 13.2 wt% Cr in these electrodes; note
that as the DOS increases the grain boundaries with less than or equal to 13.2 wt% Cr
increases (Figure 3(b)). The y axis show the normalized length of largest connected cluster, Lt
= 20 min - Lt = 0 min, and the error bars show the standard deviation of the mean length for 10
simulations. A gradual rise in the cluster length was observed if the fraction of grain boundaries
with less than or equal to 13.2 wt% Cr was increased.
The normalized mean largest cluster length, L t = 20 min – L t = 0 min, for three different applied
potentials and various DOS levels as shown in Figure 14 (d). The circles mark the length for
Vapp = -0.132 VSCE (Vapp ≥ EFlade ( 12 wt% Cr) ), the triangles line mark the length for Vapp = -
0.133 VSCE (Vapp ≥ EFlade ( 12 wt% Cr) ) and the squares mark the normalized length for Vapp =
-0.137 VSCE (Vapp ≥ EFlade ( 12.6 wt% Cr) ). At two applied potentials there is a large rise in

17
cluster length if the fraction of depleted boundaries with less than or equal to 13.2 wt% Cr grain
boundaries is above 22 %. The normalized cluster length is lower for lower potential because
more susceptible grain boundaries will be turned on initially so the depleted grain boundaries
to which intergranular corrosion can propagate will be lower in number.

(a) ( b)

Figure 10 - Distribution of the Cr depletion levels for different DOS of sensitized steel. In the
above figures the grain boundaries are marked with light gray color lines and the grain
boundaries with less than or equal to 13.2 wt% Cr are marked with a black lines for (a) DOS = 15
C/cm2 with 1.4 % grain boundaries containing less than or equal to 13.2 wt% Cr. (b) DOS = 37
C/cm2 with 22 % grain boundaries containing less than or equal to 13.2 wt% Cr.

-2
10

-3
10
10 C/cm2
Current (A)

15 C/cm2
20 C/cm2
-4 25 C/cm2
10 30 C/cm2
37 C/cm2
45 C/cm2

-5
10

0 200 400 600 750


Time (s)
Figure 11 - Simulation results with a pit in the center: Total current vs. time for different levels
of sensitized resulting in different distributions of Cr depleted grain boundaries. A square
electrode of size 0.2624 cm x 0.2624 cm was examined with a pit of diameter 250 µm in the
center. Pit current = 0.374 mA and Vapp = -0.087 VSCE.

18
Figure 12 - Simulation results with a pit in the center: (a) and (b) intergranular damage for a
low sensitized material with DOS = 10 C/cm2 at t = 0 min and after 10 min, respectively. (c) and (d)
intergranular damage for a highly sensitized material with DOS = 37 C/cm2 at t = 0 and after 10
min, respectively. The respective currents are shown in the Figure 11.

19
1000 1000
10 C/cm2
15 C/cm
2 900
800 20 C/cm2
25 C/cm2 800
30 C/cm2
600 700

R (µ m)
R (µ m)

37 C/cm2
45 C/cm2
600
400
500

200 400

300
0 10 20 30 40 50
0 200 400 600 750
Time (s) DOS (C/cm2)

( a) (b)

Figure 13 - Simulation results with a pit in the center: (a) Radius of intergranular damage zone
as a function of time for different level of sensitization in type 304 stainless steel. A comparison
can be made with Figure9. (b) Radius of intergranular damage as a function of DOS.

20
Figure 14 - Simulation results with no pit. (a) Initial (t = 0) intergranular damage for a highly
sensitized polycrystalline material (DOS = 45 C/cm2) at Vapp = -133 mVSCE. (b) Intergranular
damage after 20 min for electrode in (a) and same applied potential. (c) Mean of the normalized
length of the largest connected cluster, Mean (L t = 20 min – L t = 0 min)as a function of fraction of
grain boundaries with less than or equal to 13.2 wt% Cr at Vapp = -133 mVSCE. Error bars show the
standard deviation of the mean of the cluster length; 10 repetitive simulations were made. (d)
Mean of normalized length of largest connected cluster, Mean (L t = 20 min – L t = 0 min), as a function
of fraction of grain boundaries with less than or equal to 13.2 wt% Cr at three different
potentials.

DISCUSSION
The degree of sensitization (based on SLEPR experiments on AISI 304L SS) increases in a
non-linear fashion as the sensitization time is increased (Figure 4(a)). Optical micrographs
indicate that the fraction of susceptible grain boundaries range from almost no grain
boundaries attacked (Figure 4(b)) to almost all grain boundaries attacked in ASTM standard
testing (Figure 4(c)). For alloys with degree of sensitization greater than 36.5 C/cm2 bursting
events and rapid growth in grain boundary anodic current were observed (Figure 7). The
simulations (Figure 13(a)) strongly mimic the experimental results (Figure9) suggesting that
the model captures the important governing phenomena.

21
The spreading of IGC is substantial at DOS ≥ 37 C/cm2, when the fraction of grain
boundaries with less than or equal to 13.2 wt% Cr is greater than or equal to 22 %. This is also
nearly the 1-D percolation limit predicted by 8 for susceptibility to IGSCC. The spreading
phenomena can be explained by analyzing the interface potential (equation 2) near the surface
of the electrode in the case where a pit is present. The interfacial potential computed (Figure
15) using equation 2 at the center axis of electrode (y = 1328 µm) as a function of distance
from the edge of the electrode at different times of simulation for a highly sensitized electrode
(DOS = 37 C/cm2; all other parameters are same as in Figure 11, Figure 12, and Figure 13).
The broken horizontal lines in the figure report the Flade potential for the grain boundaries with
different Cr concentration. The applied potential is also marked with a horizontal line.
Initially, at t = 0, the applied potential is greater than the Flade potential of all the grain
boundaries. Therefore, all grain boundaries remain passive and grain boundary corrosion will
not occur without the presence of a pit. In the presence of a corroding pit, the local interface
potential on the electrode drops below the applied potential. Grain boundary corrosion occurs
when EOCP ≤ EInterface ≤ EFlade. The interface potential curve at t = 0 in Figure 15 that is equal
to Vapp - φpit. The horizontal line that reports the Flade potential for a given n % Cr
concentration grain boundary intersects the potential curve at t = 0 at a distance R0n (marked
in the Figure for n = 12% wt Cr, 16 wt% Cr and 304 SS) from the center of the electrode. Since
the Flade potential increases with decreasing chromium concentration, R016 < R014 < R012 <
R010. Within the distance R0n from the center all the grain boundaries with Cr concentration less
than or equal to n % depassivate and corrode. Inside a very small region around the pit, the
interface potential drops below the Flade potential of the 18.2 wt% Cr grain boundaries and the
matrix; therefore, some of the matrix may also start to corrode. Corrosion of the matrix is not
considered in the present model since the overall effect will be small due to very low current
densities on the 18.2 wt% Cr alloys (Figure 5).
The role of IR drop being greater than a critical potential difference, ∆φ∗ (∆φ∗ is the
difference between the applied potential and the electrode potential of the active passive
transition or the Flade potential), in the initiation of the IGC in stainless steels has been
discussed earlier by 44. However, the mechanism described so far does not account for IIGC as
well as IPit which acts together to increase IR drop.
After the initiation of some grain boundaries, the Ohmic voltage in the solution drops
further due to the corroding grain boundaries which emit a current greater than the current
from the center pit. At t = 0, when the potential drops due to the pit only, the Ohmic potential
drop near the edge of the initially corroding region is nearly 35 mV. This corroding region
grows because of the additional Ohmic drop due to the corroding grain boundaries. The
current of the corroding grain boundaries becomes larger than the current density of the pit
after 10 min. The Ohmic potential drop due only to the grain boundaries is 39 mV near the
edge of the electrode which is more than that provided by the pit.
The radii Rn (Figure 15) increases only if the Ohmic potential drop is larger than ∆φ∗ and
increases with time such that interfacial potential drops below the Flade potential of the grain
boundaries adjacent to the already corroding grain boundaries. In other words, if the effective
distance of IR drop is greater than the average distance, ρsusceptible, between the
susceptible/active grain boundaries, i.e. dIR > ρsusceptible, IGC propagation will occur. Therefore,
the two factors 1) the current density (equation 3) of the already corroding region and 2) the
probability that the Flade potential of the non corroding outer zone of grain boundaries will be
higher than the true interfacial potential, dictate whether the radius associated with the IGC will
grow. In the case of the highly sensitized electrodes, there is a higher fraction of lower
concentration chromium grain boundaries. Thus, the current densities are higher and the
probability of finding a high Flade potential grain boundary that can depassivate is higher. A

22
similar argument can be made regarding an initial corroding grain boundary with no pit except
the geometry and currents differ.
Moreover, the effective distance for the IR drop, dIR, will depend on the applied potential.
If the applied potential is low, the threshold in terms of required IR drop will be lower and the
effective distance will be larger. The threshold in terms of degree of sensitization, required for
formation of large clusters, can change slightly with Vapp (Figure 14(d)). Thus there is no
threshold based on material parameters alone. However, the situation may change with a
change in solution chemistry which was reasonably ignored by performing studies in acid.
The average distance between the susceptible/active grain boundaries, ρsusceptible, is
also dependent on the shape and size of the grain boundaries. In these calculations a square
grain structure was assumed. In the case the grains boundaries were given the same size but
the grains were made triangular, ρsusceptible, will be smaller and in the case of hexagonal grains,
ρsusceptible, will be larger.

Figure 15 - Simulation Results with a pit in the center: Interface potential as a function of
distance from the edge of the electrode at y = 1328 µm for the intergranular damage shown in
Figure 12 (b) for a highly sensitized electrode with a pit in the center. The Flade potential for the
different chromium depleted levels and the applied potential are marked. The broken vertical
lines show the radius of corrosion, Rtn of the different chromium concentration alloys at t = 0, 1,
5 and 10 min.

CONCLUSION
A micrometer scale model and an experimental study were presented to study the
surface spreading of intergranular corrosion. The model presented here was built on
electrochemical concepts and selected material parameters. The model developed gives a
qualitative description of the mechanism of intergranular corrosion spreading seen in the
experiments and also explains the role of various factors that control the spreading. Trends
seen in the experiments were duplicated in the model.
The degree of sensitization or fraction of Cr depleted grain boundaries dictates whether
intergranular corrosion can grow across an electrode. In the conditions described here, the
IGC spreading extends across an entire electrode when the DOS was greater than or equal to
37 C/cm2 or if the fraction of grain boundaries with less than or equal to 13.2 wt% Cr is more
than 22 %.
The propagation of IGC occurs when the Ohmic potential drop in the solution due to the
corroding grain boundaries is large enough to drop the interface potential to that below the
Flade potential of the nearby non-corroding but Cr-depleted grain boundaries. The Ohmic
potential drop is directly proportional to the current density of the corroding region and will be

23
larger for electrodes with high DOS. Also, the electrodes with high DOS have larger fraction of
grain boundaries with high Flade potentials. Therefore, the surface spreading of IGC occurs is
significant for high DOS alloys. Finally, both pitting and corrosion of sensitized grain
boundaries or just corrosion of sensitized grain boundaries alone can cause IGC spreading.

ACKNOWLDGEMENTS
The United States Department of Energy under the contract DEFG02-00ER45825 with Dr.
Jane G. Zhu as contact monitor. The Office of Naval research under the contract ONR:
N00014-08-10315 with Dr. Airan Perez as contact monitor. Ames Laboratory, Iowa for
supplying the materials required for the project.

24
Appendix A: Simulation Algorithm
An electrode of size L µm x L µm was simulated. The grain size is g µm (in this case g = 32)
and the shape is square, with thickness of the depleted region = 2 µm. The cumulative
probability of finding a grain boundary with n wt% Cr or less is assumed to be function of the
degree of sensitization (DOS) or the sensitization time. The algorithm steps are:
1) Divide an electrode of size L µm x L µm into (L/32+1) x (L/32 + 1) uniformly distributed
squares of size 2 µm x 2 µm. Assign a very low value of Flade potential and current
density = 0 to all these points. In the case of simulations with the pit. The grid points with
(x-L/2)2 +(y-L/2)2 < (125)2 were assumed to represent a hole in the electrode and the
position occupied by the pit.
2) Set y = 1
3) Choose first 16 points that form one grain boundary.
4) Assign a Cr concentration to it for the given degree of sensitization alloy. This is done by
generating a random number using the ran2 45 function in FORTRAN77. The value of the
random number is compared to the probability of finding n wt % Cr concentration grain
boundary calculated using the data presented in Figure 3. The grain boundary is assigned
the properties (the Flade potential, current density, and reactivation time constant) of the
n wt% Cr concentration grain boundary if the random number is less than the computed
probability.
5) At the same y, move to next 16 points and repeat the same step 4 if the coordinates are
not covered by the pit.
6) Step 5 is repeated till the end of the electrode is reached.
7) Now change y as, y = y + 16 µm and repeat the steps 3, 4, and 5. Keep changing y till the
end of the electrode is reached.
8) Similar to moving in the y direction we move in the x-direction and repeat steps 2, 3, 4, 5,
6, and 7.
9) Set t = 0.
10) Compute the potential at each grid point on the electrode. This is done by calculating the
distance of each point from the center of the pit and then calculating it through equation 2.
11) Find the points with the EOCP ≤ EInterface≤ EFlade and assume that they start to corrode at
this time. The current density for these points can be calculated using Equation 4.
12) Time is changed to t = t + dt and the step 10 is repeated. dt = 5 s and dt = 15 s for the
case with no pit and a pit at the center, respectively.

REFERENCES
1. D. A. Jones, "Principles and Prevention of Corrosion" (London, England: Prentice-Hall, Inc.,1996)
2. A. J. Sedriks, "Corrosion of Stainless Steels" (New York: John Wiley & Sons, Inc.,1996)
3. E. C. Bain, R. H. Aborn, J. J. B. Rutherford, Transactions of the American Steel Treating Society
(1933), p. 481
4. V. Cihal, I. Kasova, Corrosion Science (1970), p. 875
5. A. Bäumel, H.-E. Bühler, H.-J. Schüller, P. Schwaab, W. Schwenk, H. Ternes, H. Zitter, Corrosion
Science (1964), p. 89
6. C. L. Briant, E. L. Hall, Corrosion (September 1986), p. 522
7. S. M. Bruemmer, B. W. Arey, L. A. Charlot, Corrosion (January 1992), p. 42
8. M. A. Gaudett, J. R. Scully, Metallurgical and Materials Transactions A-Physical Metallurgy and
Materials Science (April 1994), p. 775

25
9. M. A. Gaudett, J. R. Scully, Journal of the Electrochemical Society (December 1993), p. 3425
10. M. Casales, M. A. Espinoza-Medina, A. Martinez-Villafane, V. M. Salinas-Bravo, J. G. Gonzalez-
Rodriguez, Corrosion (November 2000), p. 1133
11. C. L. Briant, C. S. O'Toole, E. L. Hall, Corrosion (January 1986), p. 15
12. H. Sahlaoui, K. Makhlouf, H. Sidhom, J. Philibert, Materials Science and Engineering A-
Structural Materials Properties Microstructure and Processing (May 2004), p. 98
13. W. E. Mayo, Materials Science and Engineering A-Structural Materials Properties Microstructure
and Processing (July 1997), p. 129
14. Y. F. Yin, R. G. Faulkner, Corrosion Science (December 2007), p. 2177
15. G. P. Airey, Metallography (1980), p. 21
16. M. Thuvander, M. K. Miller, K. Stiller, Materials Science and Engineering A-Structural Materials
Properties Microstructure and Processing (September 1999), p. 38
17. G. S. Was, V. B. Rajan, Metallurgical Transactions A-Physical Metallurgy and Materials Science
(July 1987), p. 1313
18. K. Osozawa, H.-J. Engell, Corrosion Science (1966), p. 389
19. K. Osozawa, K. Bohnenkamp, H.-J. Engell, Corrosion Science (1966), p. 421
20. M. S. Basiouny, S. Haruyama, Corrosion Science (1976), p. 529
21. V. Cihal, "Intergranular corrosion of steels and alloys" (Amsterdam-Oxford-New York-Tokyo:
Elsevier,1984)
22. H. Kaesche, "Metallic Corrosion" National Associate of Corrosion Engineers,1985)
23. M. Keddam, O. R. Mattos, H. Takenouti, Electrochimica Acta (September 1986), p. 1147
24. J. A. L. Dobbelaar, E. C. M. Herman, J. H. W. De Wit, Corrosion Science (May 1992), p. 765
25. Z. Szklarska-Smialowska, "Pitting and Crevice Corrosion" (Houston, Texas: NACE Press,2005)
26. V. Guillaumin, G. Mankowski, Corrosion (January 2000), p. 12
27. K. Kowal, J. DeLuccia, J. Y. Josefowicz, C. Laird, G. C. Farrington, Journal of the
Electrochemical Society (August 1996), p. 2471
28. D. B. Wells, J. Stewart, A. W. Herbert, P. M. Scott, D. E. Williams, Corrosion 1989), p. 649
29. C. A. Schuh, R. W. Minich, M. Kumar, Philosophical Magazine (February 2003), p. 711
30. V. Y. Gertsman, K. Tangri, Acta Materialia (October 1997), p. 4107
31. A. P. Jivkov, N. P. C. Stevens, T. J. Marrow, Computational Materials Science (December 2006),
p. 442
32. V. Y. Gertsman, S. M. Bruemmer, Acta Materialia (May 2001), p. 1589
33. D. Stauffer, A. Aharony, "Introduction to percolation theory" (London: Taylor & Francis
Inc.,1992)
34. D. T. Fullwood, J. A. Basinger, B. L. Adams, Acta Materialia (March 2006), p. 1381
35. V. S. Kuzub, A. L. Anokhin, V. S. Novitskii, Protection of metals (1981), p. 246
36. T.-F. Wu, W.-T. Tsai, Corrosion Science (February 2003), p. 267
37. C.-O. A. Olsson, D. Landolt, Electrochimica Acta (April 2003), p. 1093
38. G. B. Reartes, P. J. Morando, M. A. Blesa, P. B. Hewlett, E. Matijevic, Langmuir (June 1995), p.
2277
39. J. Newman, Journal of the Electrochemical Society (1966), p. 501
40. J. Newman, "Electrochemical Systems" (New Jersey: Prentice Hall,1974)
41. R. C. Newman, F. T. Meng, K. Sieradzki, Corrosion Science (1988), p. 523
42. K. Sieradzki, R. C. Newman, Journal of the Electrochemical Society (September 1986), p. 1979
43. S. Fujimoto, R. C. Newman, G. S. Smith, S. P. Kaye, H. Kheyrandish, J. S. Colligon, Corrosion
Science (1993), p. 51
44. H. W. Pickering, Journal of the Electrochemical Society (May 2003), p. K1
45. W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, "Numerical recipes in Fortran"
(Cambridge, England: Cambridge University Press,1992)

26

You might also like