ArticlePDF Available

In Situ High‐Resolution Microscopy on Duplex Stainless Steels

IOP Publishing
Journal of The Electrochemical Society
Authors:
  • Wood Group Kenny

Abstract and Figures

The aim of the present work is to focus on the characterization of materials in situ by using a modified near‐field scanning optical microscope (NSOM) with a home‐built tuning fork head that allowed us to obtain optical imaging concurrently with topography of some systems. The first example is related to finding precursor sites for pitting corrosion on duplex stainless steels (DSS). Precursor sites for pitting on DSS were found to be inclusions that are complex in structure and where metastable pits develop at temperatures below the critical pitting temperature (CPT), but where stable pits, with large corrosion current, occur at the CPT. These inclusions were analyzed and found to be inhomogeneous in nature and consisting of a mixture of various elements (Si, Al, Mg, Ca, Ti, Mn, and S ). After analysis of the particles, in situ observation of the particles in 3.5% NaCl and 1 M HCl solutions showed that they developed metastable pits. The pits and corrosion products developed in both particles present in the austenite grains and in particles contained within the ferrite matrix. © 2000 The Electrochemical Society. All rights reserved.
Content may be subject to copyright.
Journal of The Electrochemical Society, 147 (7) 2525-2531 (2000) 2525
S0013-4651(99)12-033-0 CCC: $7.00 © The Electrochemical Society, Inc.
Characterization of materials in situ offers a new window of
opportunity for several areas of science, including electrochemistry,
biology, chemistry, physics, and medicine. At the present time, few
research groups are working toward building more versatile micro-
scopes that extend modern microscopy to studies in situ in liquids.
Smyrl’s group, for example, demonstrated that by using a modified
near-field scanning optical microscope (NSOM), with a home-built
tuning fork, it is possible to obtain in situ topography in liquids
together with functional imaging (FI) on some model systems.1
William’s group2modified an atomic force microscope (AFM) to
study the onset of pits on austenitic stainless steels, by using it as a
scanning electrochemical microscope (SECM). They also modified
a confocal microscope to obtain a photoelectrochemical microscopy
(PEM) and reported images with resolution of the order of 0.5 mm.3
Smyrl’s group has also used PEM to study precursor sites for pitting
in polycrystalline Ti with a resolution in the order of 100 nm.4In the
present work, we show that by using a modified near-field scanning
optical microscope (NSOM), with a home-built tuning fork, one can
obtain high resolution optical images concurrently with the topogra-
phy of different systems while the substrate (sample) and the probe
are immersed in the liquid. This technique has been used to show, for
the first time, metastable pitting experiments in a 25% Cr duplex
stainless steel (DSS) UNS S32550.
It is well established that MnS inclusions are the main cause of
pitting in commercial stainless steels.2,3,5,6 However, in highly al-
loyed steels (as in the case of duplex stainless steels), the cause for
pitting is a more complex phenomenon, mainly because of the pres-
ence of different precursors (not only MnS inclusions) and because
of the development of metastable pits preceding stable pitting.7In
the present work we briefly describe the method for obtaining con-
current topography and optical imaging by using an optical fiber
with an apex smaller than 100 nm. The optical image is obtained by
using a three-way connector that allows the focusing of a laser light
through the fiber to illuminate the sample, and at the same time, that
allows the collection of the reflected light. After finding, analyzing,
and isolating the particles in some areas of a DSS sample, prior to
the corrosion tests, they were imaged in situ in both 3.5% NaCl and
1 M HCl solutions. Then, the metastable pitting process was initiat-
ed with further imaging of the selected areas.
Experimental
Tip preparation.—All NSOM probes were prepared from type F-
AS optical fiber (3.7 mm in core diameter with 125 62 mm of
cladding and 245 615 mm of polymer coating), which were obtained
from Newport Corporation. The optical fibers were pulled with a
commercial pulling machine (from Sutter). After pulling the fibers
they were observed under a 63 times magnification microscope to
check the shape of the fiber before it could be mounted in the fork.
The apex of the fibers were 100 nm or less.8Those fibers with either
a large diameter or that had crashed during the mounting were dis-
carded. More details about the preparation of the nanoprobes used to
perform these tests will be reported in future publications.8
Materials.—All samples used in the present research were cut
from commercially available 25Cr DSS of the type UNS S32550,
which were obtained in wrought form (bars of 10 mm diam). Table I
shows the melt composition of the alloy. Two main annealing proce-
dures were selected, involving solutions treated at 1020 or 11408C.
Annealing at 10208C gives a critical pitting temperature (CPT) 208C
higher than annealing at 11408C.9In both cases, the samples were
annealed for 2 h and then water quenched. The resulting grain size
was about 15 mm in both cases. However, in order to obtain a sam-
ple with coarser grain size, one piece of the bar, already annealed
once at 10208C, was reannealed at 13008C for 2 h (in the ferrite re-
gion), followed by 1 h at 10208C (to grow a controlled amount of
austenite phase in the ferrite matrix), and finally water quenched.
The structure is similar to the other two heat-treatments, but the
grain size was about 100 mm. This alloy is referred to as “rean-
nealed” throughout this work. The chemical compositions of both
ferrite and austenite phases for all three heat-treatments were deter-
mined by electron probe microanalysis (EPM), as described before7
and are presented in Table II.
Sample preparation.—The samples were prepared by successive
polishing with 6, 1, and 0.25 mm diamond paste. The polishing pro-
cedure on these samples was similar to the polishing procedure
described before.4,10 That is, all the samples were polished and test-
ed a few times without being in contact with any particulate other
than diamond polishing pastes (to avoid any contamination from Al
or Si contained in a normal polishing procedure). At the end of the
polishing, the samples were lightly electrolytically etched for 3 s in
40% KOH at an applied voltage of 2.5 V, to allow the ferrite and
In Situ High-Resolution Microscopy on Duplex Stainless Steels
L. F. Garfias* and D. J. Siconolfi
Lucent Technologies, Bell Laboratories, Materials Reliability and Component Processing Research Department, Murray Hill,
New Jersey 07974-0636, USA
The aim of the present work is to focus on the characterization of materials in situ by using a modified near-field scanning optical
microscope (NSOM) with a home-built tuning fork head that allowed us to obtain optical imaging concurrently with topography
of some systems. The first example is related to finding precursor sites for pitting corrosion on duplex stainless steels (DSS). Pre-
cursor sites for pitting on DSS were found to be inclusions that are complex in structure and where metastable pits develop at tem-
peratures below the critical pitting temperature (CPT), but where stable pits, with large corrosion current, occur at the CPT. These
inclusions were analyzed and found to be inhomogeneous in nature and consisting of a mixture of various elements (Si, Al, Mg,
Ca, Ti, Mn, and S). After analysis of the particles, in situ observation of the particles in 3.5% NaCl and 1 M HCl solutions showed
that they developed metastable pits. The pits and corrosion products developed in both particles present in the austenite grains and
in particles contained within the ferrite matrix.
© 2000 The Electrochemical Society. S0013-4651(99)12-033-0. All rights reserved.
Manuscript submitted December 7, 1999; revised manuscript received March 22, 2000. This was Paper 572 presented at the Hon-
olulu, Hawaii, Meeting of the Society, October 17-22, 1999.
* Electrochemical Society Active Member.
Table I. UNS S32550 alloy used in the present work.
Composition (wt %)
Cr Ni Mo Cu N Mn Si C S P
25.92 5.9 3.19 1.62 0.2 0.96 0.47 0.03 0.007 0.024
2526 Journal of The Electrochemical Society, 147 (7) 2525-2531 (2000)
S0013-4651(99)12-033-0 CCC: $7.00 © The Electrochemical Society, Inc.
austenite to be distinguished in all the microscopes. The etching was
done in order to remove the air-formed oxide (native oxide), to pre-
pare a “fresh” surface, and to find the grains and inclusions for each
test. After etching, the whole sample was covered with an insulating
lacquer except on the area of interest (normally square regions with
an area smaller than 200 3200 mm). No crevices were ever ob-
served at the edges of the lacquer, and this procedure allowed us to
isolate a small part of the surface and consequently to keep the back-
ground passive current very low.4,10
Optical microscopy and atomic force microscopy (AFM).—Opti-
cal microscopy with a Zeiss microscope with different magnifica-
tions up to 1000 times was done in order to select the area or areas
in the sample where the inclusions are located. After location of the
particles of interest, they were transferred to an AFM (Topometrix,
Explorer) to identify the topography of the selected areas and to con-
firm the size of the inclusions and adjacent grains (see Fig. 1).
Particle analysis.—Following optical and AFM imaging, the
grains and matrix in the samples together with the inclusions were
mapped by using a scanning electron microscope (SEM) Cambridge
250, equipped with an energy dispersive spectrometer (EDS), which
provides X-ray mapping and particle analysis (see Fig. 2 and 3).
In situ near-field scanning optical microscopy (NSOM) with con-
current topography.—The topography and optical images in situ in
the liquid were performed on a Topometrix Aurora NSOM with a
custom modified head incorporating a tuning fork transducer, which
has been described in earlier publications.1,4 The experimental setup
is shown in Fig. 4. An optical fiber nanoprobe (described above) was
used to monitor the surface topography and to focus the 480 nm laser
beam onto the DSS surface. The nanoprobe was glued onto one of
the arms of the tuning fork following the feedback procedure devel-
oped by Karrai and Grober.11,12 The fork is rigidly attached to a
piezoelectric vibrator (the driving piezo or dither), which vibrates at
a frequency of 33 kHz with an amplitude of nearly 0.01 nm.11,12 The
shear-force feedback method (a detailed description of our setup can
be found in Ref. 4) was then used to obtain the topography of the
sample concurrently with the optical image in the near-field. We
used the shear-force feedback to maintain a constant distance be-
tween the nanoprobe and the surface.13 The distance is normally reg-
ulated by monitoring the damping of the oscillations of the nano-
probe, while it is in resonance, as it encounters viscous damping near
the sample surface. The method is basically noncontact with a typi-
cal interaction distance between 10 to 30 nm in air.14 After the cell
containing the sample was fixed to the scanner, the other two elec-
trodes (reference and counter) were placed near the surface of the
cell avoiding any contact with either the sample or the cell (see
Fig. 4). All electrodes were connected in order to avoid stirring the
liquid while the scanning was performed. In some cases a “fast”
scan in air was necessary to find the area of interest. Then, the solu-
tion (either 3.5% NaCl or 1 M HCl) was added into the cell, and the
images were obtained.
Electrochemical experiments.—After the first in situ images
were obtained, an EG&G potentiostat/galvanostat model 273 was
used in all our electrochemical experiments (see Fig. 5). We used a
three-electrode electrochemical cell using a saturated calomel elec-
trode (SCE) as the reference and Pt wire as the counter electrode. All
potentials mentioned in this work are vs. the SCE. Both the sample
potential and current were collected through a general purpose inter-
face bus (GPIB) interface into a personal computer. After the meta-
stable pitting experiments, a second set of images was obtained with
the NSOM setup described above.
Results and Discussion
Brief description of metastable pitting in DSS.—The first observa-
tion since 1995 of metastable pitting in highly alloyed steels was made
by one of the authors.15-17 In several studies with DSS containing 22,
25, and 28% Cr content, a number of metastable pits were observed in
large electrodes. (An extensive description on the metastable pitting of
DSS can be found in Ref. 7.) Until now all metastable pitting studies
have been made only on microelectrodes.18-23 While testing the DSS
Table II. Average composition of ferrite and austenite for each
heat treated DSS sample.
Annealing Composition (wt %)
temp. (8C) Phase % Cr Ni Mo NaCu P
1020 Ferrite 47 26.36 4.63 3.69 0.05 1.27 0.043
Austenite 53 23.14 7.36 2.35 0.33 1.95 0.025
1140 Ferrite 64 26.37 5.05 3.36 0.05 1.36 0.038
Austenite 36 24.04 7.15 2.33 0.47 1.83 0.026
Reannealed Ferrite 50 27.27 4.58 3.89 0.05 1.33 0.052
Austenite 50 23.44 7.58 2.36 0.35 1.98 0.023
aNitrogen in ferrite is taken as fixed at the saturation value of ,0.05%,
the remaining portions are attributed to austenite.
Figure 1. Optical micro-
graph (center) from the DSS
sample showing the austenite
grains (dark) and the ferrite
matrix (light color). Notice
that this area contained three
different particles, two in the
ferrite and one in the austen-
ite. The respectvie AFM
images from the areas are
shown in the boxes next to
them (the austenite grains are
also darker than the lighter
ferrite matrix).
Journal of The Electrochemical Society, 147 (7) 2525-2531 (2000) 2527
S0013-4651(99)12-033-0 CCC: $7.00 © The Electrochemical Society, Inc.
below the CPT in neutral NaCl solutions, a region of very large peaks
associated with metastable pits was observed at low potentials
(between 200 and 350 mVSCE), followed by a decrease of activity at
higher potentials.7Above the CPT, stable pits were initiated at the
same low potentials where metastability occurs but preceded by only
a few metastable events. We found observable differences between the
growth of metastable pits and its transition to stable pitting in high-
chromium DSS from that seen in single-phase austenitic steels.
Inclusions as precursors sites for pitting in DSS.—Figure 3
shows typical EDX spectrum for the particles found in the ferrite
matrix and in the austenite grains in 25% Cr DSS. Table III shows a
Figure 2. EDX spectra from
the DSS sample showing the
characteristic peaks for the
austenite grains and those for
the ferrite matrix. Notice that
the ferrite matrix contains a
larger Mo peak and smaller
Ni peak than the austenite
spectrum.
Figure 3. EDX spectra from
two inclusions in the DSS.
The inclusion in the ferrite
matrix contains large Al and
Mn peaks and smaller Si, Ca,
and Ti peaks; the Fe and Mo
peaks are smaller than in the
matrix. In the austenite inclu-
sion the large peaks are from
Al, Si, Ca, and Mn with small
peaks corresponding to Mg
and Ti. The peaks corre-
sponding to Fe, Ni, and Cu
are smaller than in the nor-
mal austenite grains.
2528 Journal of The Electrochemical Society, 147 (7) 2525-2531 (2000)
S0013-4651(99)12-033-0 CCC: $7.00 © The Electrochemical Society, Inc.
more detailed distribution of the peaks in some of the DSS samples.
The inclusions were inhomogeneous in nature and consisted of a
mixture of various elements (Si, Al, Mg, Ca, Ti, Mn, and S); each
inclusion was normally different from the others. However, in most
of the inclusions a large Al peak was always present and very often
the Mo and Fe peaks were recessed (compared to the ferrite or
austenite spectrum).
Most of the inclusions found in the ferrite matrix, showed rela-
tively large Al and Mn peaks, denoting the presence of both elements
in the particle. In comparison the Mo and Fe peaks were recessed. In
a previous investigation,16 with a different microprobe that could
resolve and differentiate the Mo and S peaks, we observed that in
this kind of steel a high Mn peak was not always associated with a
high S content in the inclusion. Those inclusions with Mn and S may
appear similar to the ones observed in conventional austenitic stain-
less steels.2,3,5,6 However, the presence of other, different elements
shows the complexity of these inclusions. Indeed, in the present
research, only in one inclusion did we observe the Mo peak to be
larger than in the normal austenite or ferrite spectrum (see Table III).
This can only happen if Mo is enriched in the inclusion or if S is
enriched in the inclusion, we believe that (in this case) sulfur enrich-
ment in the particle is the correct explanation.
The number of inclusions found in the austenite grains was nor-
mally less than in the case of the ferrite matrix. We estimated that
about 30% of the inclusions found were located in the austenite (see
Table III). The austenite particles showed a more complex and dis-
tinct composition than the inclusions found in the ferrite (different
relevant peaks). However, they also can trigger metastable activity
(see below).
We propose that the initiation sites on DSS can be attributed to
four distinct causes: inclusions on the ferrite matrix, inclusions in the
austenite grains, inclusions in the boundary between austenite grains
and matrix and (of less relevance) defective oxide growth over active
grain boundaries or other surface features. In the case of the inclu-
sions, they appear to be complex particles that in some cases can be
found near the grain boundaries, most probably because of the
migration of these contaminants during heat-treatment (see, for ex-
ample, Fig. 6). Although most of them contain Al, Si, Ca, Ti, and
Mn, they have other impurities in lower concentrations, which may
also contribute to the developing of the metastable pits.
After location and analysis of the inclusions, in situ imaging in
the area with the inclusion or inclusions desired was done first in air
and then after the solution was introduced in the cell. We did not find
any difference between the dry images and the wet images (topogra-
phy and optical imaging) before the electrochemical experiments.
Metastable pitting in DSS.—Most of the electrochemical tests
done have already been reported in the same type of DSS.4,15-17 In
most of the experiments we try to isolate a small working area sam-
ple, with the least number of inclusions, typically on the order of
0.1 mm2. In some cases, depending of the location of the inclusions,
we used rectangles with approximate lengths between 300 to
400 mm. By using a small area we ensure that the passive current will
be relatively low and that we will be measuring the current directly
related to the metastable events. Figure 5 shows a typical anodic
polarization in 3.5% NaCl of a DSS sample (scan rate 1 mV/s and test
temperature around 218C). The current transient show metastable
Figure 4. Optical imaging with concurrent topography setup.
Figure 5. Anodic polarization in 3.5% NaCl (room temperature, scan rate 1
mV/s) of DSS (sample 3) showing metastable pitting. The largest event
strarted at 150 mVSCE followed by two events, one starting at 300 mVSCE and
the other at 450 mVSCE.
Table III. Detailed distribution of peaks obtained by EDX in the inclusions on DSS samples used in the present work.
Sample Inclusion
no. no. Phase Large peaks Small peaks Changed
1 2 Austenite Al, Si, Mn, Ca Mg, Ti None
2 Ferrite Al, Mn Si, Ca, Ti FefMof
4 2 Ferrite Al, Mn, Ti(Med) Si FefMof
(A2)
2 Ferrite Al, Mn Ti Mof
3 Ferrite Al, Mn Ti Mof
5 1 Ferrite Al, Si, Mg, Ca Ti CrfFefMof
2 Ferrite Al, Si, Mg, Ca Ti CrfFefMof
3A Austenite Si, Ca Ti, Al Crf
3B Between Al, Mn, SaTi MoaF
4 Between Al, Si, Ca Ti CrfFefMof
aThe Mo peak increased because of the significant contribution from sulfur.
Journal of The Electrochemical Society, 147 (7) 2525-2531 (2000) 2529
S0013-4651(99)12-033-0 CCC: $7.00 © The Electrochemical Society, Inc.
activity at three different potentials. The first metastable event started
at 150 mVSCE and ended at 250 mVSCE, the maximum current
achieved was nearly 70 nA. After the test, the transient for that event
was isolated and its passive current subtracted (in that range of poten-
tial, between 150 and 250 mVSCE, was around 20 nA). By using the
same kind of analysis described by Pistorius and Burstein,22 and
assuming hemispherical pit geometry, it was demonstrated that
metastable activity in DSS shows distinct transients from that of con-
ventional austenitic steels.7In this case, the transient was analyzed
and a final pit radius of about 10 mm was calculated. The relevance
of this result is discussed below. The other two transients in Fig. 5
showed relatively lower currents, one of them (most probably the one
started at 450 mVSCE) could be related to a different metastable pit,
whereas the one started at 300 mVSCE could correspond to a smaller
pit (not found), or as a part of any of the other two events.
In situ NSOM and topography after metastable pitting.—In the
present investigation, we have confirmed that in most cases the
inclusions found showed metastable activity and growth of the pits
under a cover, mainly formed of corrosion products (see Fig. 7 and
9). Figure 6 shows a side-by-side concurrent topography and NSOM
images of the same area in a DSS sample while immersed in 1 M
HCl solution (before the anodic polarization in that solution). Both
of the images on the top show a square area of 60 360 mm. From
the topography images, one can measure the size of the inclusion,
which is approximately 8 mm across and 135 nm in height (see scale
bar in the left of the topography image). The NSOM image shows
the light reflection signal at the same area, with the inclusion having
very little reflection of the light. In this case, higher resolution
images (30 330 mm) are shown on the bottom of each image. The
resolution of the images in both cases was probably better 100 nm.
One important feature in those images corresponds to dark areas of
nearly 1 mm, near the inclusion in the grain boundary. The size of
these dark areas cannot be fully estimated from the topography, but
only from the optical images. Furthermore, they appear to be more
similar to the inclusions than to any of the two adjacent phases. Fig-
ure 7 shows images of the same inclusions before (top) and after
(bottom) the metastable test in 1 M 1 HCl. The bottom images (after
the test) show two important features. First, the inclusions have de-
veloped into a pit of larger size that contains corrosion products,
most probably redeposited outside the pit where the inclusion was
originally located. Some small holes can be seen together with a rel-
atively large aperture of the cavity. It is also important to notice the
change in the Z scale, which turned from 135 nm before the test to
1154 nm after the test. The main contribution in the Zdirection is
from the corrosion products found outside of the pit. The second rel-
evant piece of information is related to the partial “etching” of the
austenite grains (at 2200 mVSCE). The austenite grains, which be-
fore the test were higher in the Zdirection, now appear to be re-
cessed after the (selective) anodic dissolution at lower potentials.
Nevertheless, we have observed in the past that metastable pits de-
veloped with more open cavities in HCl solutions than in the NaCl
solution.7,9,15-17 After the tests, the samples were slightly polished,
and etched. Figure 8 shows the topographical images of two inclu-
sions before (top) and after corrosion tests and repolishing (bottom).
In the case of particle 3 (left side images), the pit was not sufficient-
ly deep and can be seen only as a small hole in the ferrite matrix.
However, in some cases as in the case of particle 2 (right side
images), the pit is still deep after the mild polishing. This information
was used to associate a transient with its correspondent inclusion
when testing an area with multiple inclusions. Evidently, a larger
transient was always associated with those large pits. Figure 9 shows
the topography (3D) and NSOM images of the inclusion associated
with the large transient described above (in the NaCl solution). In this
case, after the metastable pitting experiment, the inclusion developed
into a pit with the typical corrosion products protruding outside the
pit (after the test). Similarly to the cases described above, the Zdirec-
tion in the image below (after the corrosion test) shows an increase
related to the corrosion products. The inside of the pit before and after
metastable pitting shows a different contrast. Before metastable pit-
ting the cavity basically does not reflect the light (therefore the cavi-
ty appears dark), but after metastable pitting the cavity shows a bet-
ter contrast in and outside the pit (despite the fact that the bottom of
the pit is about 1 mm deep in reference to the top of the corrosion
products). This is most probably due to the reflection of the light by
the corrosion products inside the pit. The size of this inclusion was
Figure 6. Concurrent topography and
optical imaging in 1 M HCl of inclusion
P3 (in the ferrite matrix). Notice that in
some cases some features are easily seen
in the optical image. However, both
images contain useful information that
complement each other.
2530 Journal of The Electrochemical Society, 147 (7) 2525-2531 (2000)
S0013-4651(99)12-033-0 CCC: $7.00 © The Electrochemical Society, Inc.
found to be around 15 mm. As mentioned in the previous section, the
calculated transient corresponded to a hemispherical pit of 10 mm
radii. The main difference in the calculated size lies in the depth of
the pit. While assuming hemispherical pits is a reasonable approxi-
mation, in the case of DSS this may be an overestimation, mainly
because an inclusion may develop into a pit as deep as the inclusion.
We believe that metastable events that develop within an inclusion
will eventually find a more resistant material underneath and near the
walls of the pit that can hinder the further growth. However, some
other circumstances may impede metastable growth.24 We believe
Figure 8. Topography of both particles (2
and 3) before (top) and after (bottom) cor-
rosion tests. In both cases a pit developed
in the inclusion. The bottom images were
obtained after the sample was polished,
cleaned, and etched.
Figure 7. Concurrent topography and opti-
cal imaging (in 1 M HCl) of inclusion P3
(in the ferrite matrix), before (top) and after
(bottom) anodic polarization between
2300 mVSCE and 1500 mV. The topogra-
phy shows that the particle developed into
a pit, whereas the austenite grains are par-
tially etched. The optical image shows
some of the deposits created in the surface
near the particle.
Journal of The Electrochemical Society, 147 (7) 2525-2531 (2000) 2531
S0013-4651(99)12-033-0 CCC: $7.00 © The Electrochemical Society, Inc.
that those conditions were not achieved in our experiments, particu-
larly since we have not found any evidence of salt films in the bottom
of the pits, but rather crystallographic etch pits.25,26 Those pits grow
with a cover that, as the pit developed, enhanced the aggressive envi-
ronment within the pit and most likely attacked the cover from inside,
as described in Ref. 26.
Conclusions
Metastable pitting in DSS was studied in situ in both 1 M HCl
and 3.5% NaCl solutions. Topography and NSOM images of the
selected areas containing inclusions showed that they developed into
pits that repassivate (metastable pits). The inclusions are complex in
structure, where metastable pits develop at temperatures below the
CPT, while stable pits, with large corrosion currents, occur at the
CPT. These inclusions were analyzed and found to be inhomoge-
neous in nature and consisting of a mixture of various elements (Si,
Al, Mg, Ca, Ti, Mn, and S). After analysis of the particles, in situ
observation of the particles in 3.5% NaCl and 1 M HCl solutions
showed that they developed metastable pits. Pits and corrosion prod-
ucts developed in both particles present in the austenite grains and in
particles contained within the ferrite matrix. The resolution and
accuracy of the topographic information in liquid is retained (com-
pared to conventional scanning probe methods in air). At the present
time the lateral resolution in this kind of test is around 100 nm.
Acknowledgments
We acknowledge useful discussions with M. Buechler, W. H.
Smyrl, R. B. Comizzoli, J. D. Sinclair and R. P. Frankenthal.
Lucent Technologies, Bell Laboratories, assisted in meeting the publica-
tion costs of this article.
References
1. P. I. James, L. F. Garfias-Mesias, P. J. Moyer and W. H. Smyrl, J. Electrochem.
Soc., 145, L64 (1998).
2. Y. Zhu and D. E. Williams, J. Electrochem. Soc., 144, L43 (1997).
3. D. E. Williams, T. F. Mohiuddin, and Y. Y. Zhu, J. Electrochem. Soc., 145, 2664
(1998).
4. L. F. Garfias-Mesias and W. H. Smyrl, J. Electrochem. Soc., 146, 2495 (1999).
5. G. Eklund, J. Electrochem. Soc., 121, 467 (1974).
6. J. E. Castle and R. Ke, Corros. Sci., 30, 409 (1990).
7. L. F. Garfias-Mesias and J. M. Sykes, Corros. Sci., 41, 959 (1999).
8. L. F. Garfias-Mesias, J. Weaver, M. Buechler, and W. H. Smyrl, In preparation.
9. L. F. Garfias-Mesias, J. M. Sykes, and C. D. S. Tuck, Corros. Sci., 38, 1319
(1996).
10. L. F. Garfias-Mesias, M. Alodan, P. I. James, and W. H. Smyrl, J. Electrochem.
Soc., 145, 2005 (1998).
11. K. Karrai and R. D. Grober, Appl. Phys. Lett., 66, 1842 (1995).
12. K. Karrai and R. D. Grober, Ultramicroscopy, 61, 197 (1995).
13. E. Betzig, P. L. Finn, and J. S. Weiner, Appl. Phys. Lett., 60, 2484 (1992).
14. M. A. Paesler and P. J. Moyer, Near-Field Optics: Theory, Instrumentation, and
Applications, John Wiley & Sons, Inc., New York (1996).
15. L. F. Garfias, J. M. Sykes, and C. D. S. Tuck, Paper presented at “Electrochem ’95,
Bangor, Wales, Sept 10-14, 1995.
16. L. F. Garfias-Mesias, Ph.D. Thesis, Oxford, U.K. (1996).
17. L. F. Garfias and J. M. Sykes, in Proceedings of the 13th International Corrosion
Congress, Melbourne, Australia, Nov 25-29, 1996.
18. Y. Hisamatsu, T. Yoshii, and Y. Matsumura, in Localized Corrosion, NACE-3,
B. F. Brown, J. Kruger, and R.W. Staehle, Editors, p. 427, NACE, Houston, TX
(1974).
19. N. Sato, J. Electrochem. Soc., 129, 255 (1982).
20. N. Sato, J. Electrochem. Soc., 129, 260 (1982).
21. G. S. Frankel, L. Stockert, F. Hunkeler, and H. Boehni, Corrosion, 43, 429 (1987).
22. P. C. Pistorius and G. T. Burstein, Philos. Trans. R. Soc. London, Series A, 341, 531
(1992).
23. G. T. Burstein, P. C. Pistorius, and S. P. Mattin, Corros. Sci., 35, 57 (1993).
24. H. S. Isaacs, J. Electrochem. Soc., 120, 1456 (1973).
25. H. W. Pickering and R. P. Frankenthal, J. Electrochem. Soc., 119, 1297 (1972).
26. R. P. Frankenthal and H. W. Pickering, J. Electrochem. Soc., 119, 1304 (1972).
Figure 9. Concurrent topography and
optical imaging (in 3.5% NaCl at room
temperature) of an inclusion in the ferrite
matrix (sample 3), before (top) and after
(bottom) anodic polarization from 2300
mVSCE to 1850 mVSCE. The topography
image shows how the partially recessed
particle developed into a pit with sur-
rounding corrosion products, whereas the
optical image shows the deposits created
in the surface near the particle.
... For conventional process stainless steel, inclusions are one of the significant contributors to the deterioration of its corrosion resistance. During the conventional smelting process, unavoidable steps like deoxidation introduce inclusions such as Al2O3, MnO, and others [36,37]. These inclusions have higher melting points than the alloy, and their stable nature makes it challenging to eliminate them through subsequent processing steps. ...
... For conventional process stainless steel, inclusions are one of the significant contributors to the deterioration of its corrosion resistance. During the conventional smelting process, unavoidable steps like deoxidation introduce inclusions such as Al 2 O 3 , MnO, and others [36,37]. These inclusions have higher melting points than the alloy, and their stable nature makes it challenging to eliminate them through subsequent processing steps. ...
Article
Full-text available
Duplex stainless steels are widely used in many fields due to their excellent corrosion resistance and mechanical properties. However, it is a challenge to achieve duplex microstructure and excellent properties through additive manufacturing. In this work, a 0.09% N 25Cr-type duplex stainless steel was prepared by additive manufacturing (AM) and heat treatment, and its corrosion resistance was investigated. The results show that, compared with S32750 duplex stainless steel prepared by a conventional process, the combination value of film resistance and charge transfer resistance of AM duplex stainless steel was increased by 3.2–5.5 times and the pitting potential was increased by more than 100 mV. The disappearance of residual thermal stress and the reasonable distribution of Cr and N elements in the two phases are the reasons for the improvement of the corrosion resistance of AM duplex stainless steel after heat treatment. In addition, the extremely high purity of AM duplex stainless steel with no visible inclusions resulted in a higher corrosion resistance exhibited at lower pitting-resistance-equivalent number values.
... Nonmetallic inclusions, such as MnS inclusions, are known to be initiation sites for pitting on austenitic, ferritic, and martensitic stainless steels in chloride solutions [37][38][39][40][41][42][43][44][45][46][47] . It has also been highlighted that duplex stainless steels may play an important role in pit initiation [48][49][50] . Garfias and Siconolfi found precursor sites for pitting corrosion in duplex stainless steels using a modified near-field scanning optical microscope 48 . ...
... It has also been highlighted that duplex stainless steels may play an important role in pit initiation [48][49][50] . Garfias and Siconolfi found precursor sites for pitting corrosion in duplex stainless steels using a modified near-field scanning optical microscope 48 . They demonstrated that pits are initiated at inclusions in UNS S32550 (Fe-26Cr-6Ni-3Mo-2Cu-0.2 N). ...
Article
Full-text available
An aging treatment of UNS S32750 super duplex stainless steel at 1173 K for 1.0 ks produced σ and secondary austenite (γ2) phases in the α and γ phase boundary regions. Small amount of Mn-Cr oxide and MnS complex inclusions were present in the steel. No pitting was observed during the potentiodynamic polarization of a small area (200 × 200 µm) without inclusions in 1 M MgCl2 at 348 K. Because the electrode area included α, γ, σ, and γ2 phases, these phases and their boundaries alone could not act as initiation sites for pitting. The electrode area size was gradually reduced from 3 × 3 mm to 1 × 1 mm, and pitting corrosion was observed to occur at the Mn-Cr oxide and MnS complex inclusions in a region that appeared to be the γ2 phase near the σ phase.
... In this context, several researches were performing to deleterious phases characterization by different techniques in this stainless steel family, such as Light Optical Microscopy (LOM), [20][21][22][23] Scanning Electron Microscopy (SEM), 20,24,25 Scanning Kelvin Probe Force Microscopy (SKPFM), [25][26][27] X-ray Diffraction (XRD), 28 hardness test, 29 Eddy current 30 and another magnetic analyses. 31,32 Although there are many efforts for analysis and quantification of deleterious phases, most of the techniques presented are limited to use in the laboratory as a destructive test. ...
Article
Full-text available
Linear sweep voltammetry (LSV) test parameters in alkaline medium were optimized by Doehlert matrix design in order to quantify the deleterious phases in a superduplex stainless steel UNS S32750. The microstructural analysis was performed, in several heat treated specimens, by Light Optical (LOM) and Scanning Electron (SEM) Microscopies and correlated with the electrochemical tests. In these tests, optimized parameters were obtained for tests in aqueous solutions of KOH. The concentration of 3.55 mol l⁻¹, scan rate of 3.42 mV s⁻¹ and initial potential of −0.818 V, showed a good correlation between the deleterious phases precipitated and charge density values. Differently from LOM characterization, chi and sigma deleterious phases can be distinguished by LSV optimized test. Finally, this test can be a non-destructive powerful tool of quality control to detect embrittlement and corrosion resistance decay that commonly affected this stainless steel as consequence of inadequate fabrication processes.
... Different local electrochemical and microscopic techniques have been developed for in-situ corrosion studies, allowing characterization of metal-electrolyte interfaces while monitoring ongoing corrosion processes. Local probing techniques include microelectrode (Garfias-Mesias and Sykes, 1999), microcell (Kobayashi et al., 2000;Perren et al., 2001), local EIS (Annergren, 1996;Bayet et al., 1998), scanning vibration electrode technique (SVET) (Uchida et al., 2001), scanning reference technique (SRET) (Sargeant and Ronaldson, 1996;Lin et al., 1998), scanning tunneling microscopy (STM) (Fan and Bard, 1989;Miyasaka and Ogawa, 1990), atomic force microscopy (AFM) (Rynders et al., 1994;Reynaud-Laporte et al., 1997;Garfias-Mesias and Siconolfi, 2000;Williford et al., 2000), scanning electrochemical microscopy (SECM) Williams, 1997, 2000;Tanabe et al., 1998;Williams et al., 1998;Paik et al., 2000), scanning Kelvin probe (SKP) (Han and Mansfeld, 1997;Chen et al., 1998), and combination of these techniques (Böhni et al., 1995;de Wit et al., 1998;Guillaumin et al., 2001). Microelectrodes and microcells enable probing of small areas. ...
Article
Full-text available
Duplex stainless steels (DSSs) consist of ferrite and austenite phases with approximately equal volume fraction. They exhibit combined superior mechanical strength and corrosion resistance, therefore are increasingly used in various applications. However, under certain conditions, passivity breakdown may occur, leading to corrosion initiation, which is often related to weak points in the heterogeneous microstructure. To understand the influence of microstructure on the passivity and breakdown of DSSs requires local probing techniques that can be used in situ, so that corrosion initiation process can be correlated to the microstructure. Recent studies employing advanced scanning probe microscopy and synchrotron-based techniques, in combination with electrochemical measurements, have contributed to a deep understanding of the passive film, passivity breakdown, and corrosion initiation of DSSs, as well as the influence of microstructural and environmental factors. This mini review presents a short summary of recent literature focusing on the studies utilizing local probing techniques and synchrotron-based analyses.
... Wie vermessen, die eine geringe räumliche und zeitliche Auflösung bieten [145, 146]. Mittels optischer Mikroskopie gelang die in-situ-Beobachtung vergleichsweise großer und stabiler Korrosionsgrübchen (Durchmesser > 10 µm) [147] , und einzelne metastabile Grübchen wurden mit pHempfindlichen Gelen visualisiert [148, 149]. ...
Thesis
Full-text available
In this thesis, two different pattern-forming catalytic reactions are examined by means of a variety of optical methods: CO-oxidation on Pt(110) single crystal surfaces under vacuum conditions, and pitting corrosion on stainless steel in aqueous NaCl-solution. CO-oxidation is a paradigmatic example for pattern-forming non-equilibrium systems. The dynamics of the reaction are widely understood. Thus, CO-oxidation is utilized as a well-suited model system for the analysis of pattern formation under spatial and temporal manipulation. This work and additional theoretical studies put pitting corrosion into a conceptual framework of autocatalytic processes and non-equilibrium pattern formation phenomena in reaction-diffusion systems. CO-Oxidation on Pt(110) is observed by means of a newly designed reflection anisotropy microscope (RAM) with increased lateral resolution, and by imaging interferometry. The former method (RAM) makes it possible for the first time to combine imaging of pattern formation between inactive microstructures (boundaries), which are fabricated by micro-lithography, on a crystal surface with the spatio-temporal manipulation of patterns by focussed laser light. This possibility paves the way for studying new aspects of the dynamics of pattern formation, particularly in regard to stability of concentration patterns. The experimental results are compared to numerical simulations (Qiao Liang, Princeton, NJ). The second experimental method (imaging interferometry) is used for mapping the topography of Pt(110) foils with a thickness of a few 100 nm. The findings suggest that, by using their mechanical properties, ultrathin metal foils may be utilized for the detection of small amounts of heat. The interplay between the dynamics of the catalytic reaction and thermal effects due to the release of reaction heat results in the emergence of spatio-temporal patterns in the topography of the platinum foil. Contrast-enhanced optical microscopy and imaging ellipsometry are employed to visualize pitting corrosion. Using these methods, the onset of pitting corrosion is followed for the first time in situ during an electrochemical experiment. The nucleation and activity of single corrosion pits and the damaging of the protective oxide layer on the surface of stainless steel are monitored with high spatial and temporal resolution. A comparison of the experimental results with numerical simulations (Levent Organ, Charlottesville, VA) substantiates the assumption that the onset of pitting corrosion can be regarded as an autocatalytic process.
Article
The comparative investigation of the corrosion behavior of S32304 cast lean duplex and CF3M cast austenitic stainless steels was conducted in a chloride environment. The resistance to pitting corrosion of the S32304 alloy was “one” time higher than that of the CF3M alloy because the pitting potential, passive region, and critical pitting temperature of the low Ni-low Mo S32304 alloy were higher than those of the high Ni-medium Mo commercial CF3M alloy. There are two main reasons for the enhancement of the pitting corrosion resistance of high Cr-low Mo-medium N S32304 alloy compared to the low Cr-medium Mo CF3M alloy: First, the pitting resistance equivalent number (PRENδ+γ) value of the S32304 alloy is higher than that of the CF3M alloy. Second, the passive region of the S32304 alloy is larger than that of the CF3M alloy. It indicates that the synergistic effects of the three elements by adding high Cr and low Mo-medium N to the S32304 alloy enhance the passivity of the passive film, thereby increasing the resistance to pitting corrosion. It was verified that based on the PRENγ and PRENδ values calculated using an N-factor of 30, the pitting corrosion of the S32304 alloy was selectively initiated at the γ-phases and the inclusions of Mn oxy-sulfieds because the γ-phase’s PRENγ value was smaller than the δ-phase’s PRENδ, and finally propagated from the γ-phase to the δ-phase.
Article
We demonstrate that fluorogenic molecules that "turn-on" upon redox reactions can sense the corrosion of iron at the single-molecule scale. We first observe the cathodic reduction of nonfluorescent resazurin to fluorescent resorufin in the presence of iron in bulk solution. The progression of corrosion is seen as a color change that is quantified as an increase in fluorescence emission intensity. We show that the fluorescence signal is directly related to the amount of electrons that are available due to corrosion progression and can be used to quantify the catalyzed increase in the rate of corrosion by NaCl. By using modern fluorescence microscopy instrumentation we detect real-time, single-molecule "turn-on" of resazurin by corrosion, overcoming the previous limitations of microscopic fluorescence corrosion detection. Analysis of the total number of individual resorufin molecules shows heterogeneities during the progression of corrosion that are not observed in ensemble measurements. Finally, we discuss the potential for single-molecule kinetic and super-resolution localization analysis of corrosion based on our findings. Single-molecule florescence microscopy opens up a new spatiotemporal regime to study corrosion at the molecular level.
Article
Passive films on duplex stainless steels are heterogeneous, with properties depending on the two-phase microstructure. Accordingly, the interaction of chloride ions with the passive films on either of the two phases ought to vary. These issues have been investigated by some indirect methods, not sufficiently nor directly corroborated by experimental evidence. Using aberration-corrected transmission electron microscopy, we simultaneously investigate the films on the ferrite and austenite phase near the phase boundary, as well as the film evolution in chloride-containing media. This study provides some virtual experimental insights into the features of chloride attack on the passive film of duplex alloys.
Article
Full-text available
A new methodology to study in-situ the pitting corrosion of stainless steels at high temperature and high pressure (HT/HP) was developed. This methodology allowed the real-time in-situ anodic polarization while determining the sites for pit initiation in UNS S32305 (22% Cr Duplex Stainless Steels), in chloride containing environments with additions of H 2S and CO 2 at high temperature and pressure. The methodology allowed measurements at two temperatures 40°C (and 582 psi) and 180°C (and 752 psi). This methodology allowed the in-situ imaging of metastable pitting as well as stable pitting under different conditions. Pitting corrosion developed in inclusions (5-10 micrometer or larger); which are mainly composed of intermetallic particles containing significant amounts of B, N, O, Mg, Al, S, Ca and C.
Article
Full-text available
A Novel Mini-Autoclave where the tests are performed in-situ under an aqueous environment (simulated production brine) saturated with CO/H 2S gas mixtures was developed (with similar performance compared to a normal autoclave). The mini-autoclave ultimate limits are 400 °C and 4000 psi and can stand a mixture of a gases and liquids. The surface of the sample inside the chamber can be photographed and a real time video can be recorded while electrochemical measurements on the Area of Interest (AOI) in the surface of the sample can be done. Typically, small areas (5 μm ≤ AOI ≤ 5 mm) are suitable for imaging using this mini autoclave. In the present study, electrochemical measurements simultaneously with optical imaging and real time videos of UNS S32550 (25% Cr Duplex Stainless Steels) in chloride containing environments at two temperatures 40 °C (and 582 psi) and 180 °C (and 900 psi) were used to confirm (visually and electrochemically) the conditions under which the transition from passivity to corrosion could take place. This methodology allowed the real-time in-situ anodic polarization of the samples and observation of the precursor sites for pitting above the Critical pitting Temperature on UNS S32550 (25% Cr Duplex Stainless Steels).
Article
Full-text available
Metastable pitting is a phenomenon that many passive metals exhibit in aggressive solutions at low potentials. Furthermore, it is accepted that stable pits experience a metastable stage prior to transitioning to stability during which they behave as metastable pits. This perspective summarizes the findings of the first paper that provided a thorough analysis of metastable pitting and describes other papers that contributed to the field. © 2019 National Assoc. of Corrosion Engineers International. All Rights Reserved.
Article
Full-text available
The evolution of corrosion pits on stainless steel immersed in chloride solution occurs in three distinct stages: nucleation, metastable growth and stable growth. This paper describes the growth of metastable corrosion pits on stainless steel immersed in chloride solution, and their transition to stability. The rate of growth of individual corrosion pits is controlled by diffusion of the dissolving metal cations from the pit interior, the surface of which is saturated with the metal chloride. This process is independent of electrode potential. Analysis of the diffusion yields a critical value of the product of the pit radius and its dissolution current density (termed the `pit stability product') below which the pit is metastable and may repassivate, and above which the pit is stable. The critical value of the pit stability product for stainless steel in chloride solution is 0.3 A m-1. All pits, whether metastable, or destined to become stable, grow initially in the metastable condition, with a pit stability product which increases linearly with time, but below the critical value. Metastable growth requires a perforated cover over the pit mouth to provide an additional barrier to diffusion, enabling the aggressive pit anolyte to be maintained. In this state pits grow at a constant mean current density which is maintained by periodic partial rupture of the cover. Stable pit growth is then achieved when the cover is no longer required for continued propagation, and the pit depth is itself a sufficient diffusion barrier; stability is characterized by a constant mean pit stability product above the critical value. If the cover is lost prematurely, before the critical pit stability product is achieved, the pit anolyte is diluted and repassivation is inevitable. In contrast to the growth rate of individual pits, the distribution of pitting current transients is dependent on electrode potential: the pit nucleation site, particularly its geometry, is exclusively responsible for this potential distribution. It is proposed that shallower, more-open sites are activated only at higher potential and higher current density, and are consequently more likely to achieve stability.
Article
Full-text available
An aluminum coated tapered optical fiber is rigidly attached to one of the prongs of a high Q piezoelectric tuning fork. The fork is mechanically dithered at its resonance frequency (33 kHz) so that the tip amplitude does not exceed 0.4 nm. A corresponding piezoelectric signal is measured on electrodes appropriately placed on the prongs. As the tip approaches within 20 nm above the sample surface a 0.1 nN drag force acting on the tip causes the signal to reduce. This signal is used to position the optical fiber tip to about 0 to 25 nm above the sample. Shear forces resulting from the tip-sample interaction can be quantitatively deduced.
Article
The effect of annealing temperature on pitting resistance of 25% Cr duplex UNS S32550 was investigated. CPT and pitting potential were determined for this alloy after annealing at 1020, 1060, 1100 and 1140 degrees C. The higher values of CPT and pitting potential were found after annealing at the lower temperatures. Fitting was always observed preferentially in the ferrite phase. The results can be partially explained by the changes in chemical composition of ferrite and austenite phases. Copyright (C) 1996 Elsevier Science Ltd.
Article
The nucleation and growth of corrosion pits on stainless steels in chloride is described. Pits arise from distinct sites on the surface that are destroyed after reaction. Metastable and stable pits grow at a diffusion-controlled rate. In the metastable state metal dissolution occurs through a perforated cover: the pit only achieves stability when this cover is no longer necessary to maintain the diffusion barrier. The growth of the pit and the transition from metastability to stability is described by the pit stability product. Recent results show that nucleation of the corrosion pit occurs by a microscopically violent event, observed as a sharp tiny current transient which initiates metastable pit growth. Many such nucleation events do not achieve metastability, however, and die immediately after nucleation.
Article
The electrocapillary stability of passive films on metals has been studied in terms of energetics. It is predicted that there is a critical potential above which breakthrough pores are formed in the film. This critical film breakdown potential is less noble, as the surface tension at the metal-electrolyte interface is smaller and hence as the anion adsorption is stronger. It is also predicted from the dissolution kinetics of semiconducting surface films that above a certain critical potential the passive film is electrochemically unstable, undergoing potential-dependent transpassive dissolution. Anion adsorption, which introduces electron acceptor levels in the film, will lower the transpassivation potential at the adsorption sizes. The passive film of n-type oxide appears to be more stable against anodic polarization than that of p-type oxide.
Article
The stability criteria for pitting dissolution of metals are examined on the basis of recent results. Pitting dissolution is stable if the local ion concentration buildup at pit sites exceeds a certain critical value, which for pits at noble potentials appears to correspond to the lowest ion concentration required for the transition from etching dissolution to brightening dissolution. The smallest pit size for stable pitting decreases with anodic potential, and hence the larger defect on the metal surface gives rise to the less noble pit initiation potential. The local pH at pit sites seems to contribute to the initiation of etching pits at less noble potentials, whereas the stability of brightening pits at noble potentials is determined by the local concentration of the aggressive anion.
Article
Equilibrium calculations on some sulfide systems actual in commercial stainless steel have shown that the sulfides cannot thermodynamically exist at the potential of the passive steel. The ions released by the dissolution of the sulfides give rise to an acid solution in microareas. A pitting corrosion mechanism based on the sulfides themselves is suggested. The sulfides are polarized to the potential of the passive steel surface and tend to dissolve. During the dissolution a virgin metal surface is exposed to the environment. When the solution in the microarea thus developed has reached a certain composition, the contacting metal can no longer passivate and the metal starts to dissolve.
Article
The dissolution of austenitic stainless steels in artificial crevices has been studied as a function of potential and chloride ion activity. Impedance measurements have shown the presence of a resistive layer on the steel which limited the rate of the electrode reaction in the localized corrosion of stainless steels. The potential across the layer, equal to the product of the direct current and layer resistance, varied linearly with the electrode potential indicating ohmic ionic conduction across the layer. The growth of the layer was limited by its solubility and the diffusion of metal cations in the aqueous phase in the cavity. The layer had the properties of a metal chloride with a given solubility product. Increased chloride activity decreased the saturation concentration of the metal cation, its concentration gradient, and gave decreased dissolution rates or currents.
Article
A modified model of pit growth is presented which involves growth by active dissolution and includes a high resistance path resulting from a constriction caused by a hydrogen bubble. This model of pit growth is also believed to apply to the propagation of crevices and, in some instances, of intergranular attack. Results show that pit morphology on iron is controlled primarily by solution composition: in sulfate-containing solutions the pits are initially crystallographic, but then become polished; in nonsulfate-containing solutions they always remain crystallographic. On stainless steels the pit morphology is determined by solution composition and by the nature of the initiation site; it depends only secondarily on alloy composition. In stainless steels, both austenitic and ferritic, crevice corrosion is the prevalent corrosion form; pitting is initiated only at significantly higher potentials. The results are in agreement with an electrochemical study that showed that the potential across the interface in all local corrosion cells is in the active part of the anodic polarization curve of the metal and that pits and crevices propagate by the same mechanism.