Você está na página 1de 8

Journal of The Electrochemical

Society

Corrosion Pits in Thin Films of Stainless Steel


To cite this article: M. P. Ryan et al 1999 J. Electrochem. Soc. 146 91

View the article online for updates and enhancements.

This content was downloaded from IP address 187.61.230.127 on 28/10/2021 at 15:09


Journal of The Electrochemical Society, 146 (1) 91-97 (1999) 91
S0013-4651(98)04-038-5 CCC: $7.00 © The Electrochemical Society, Inc.

Corrosion Pits in Thin Films of Stainless Steel


M. P. Ryan,a,d,* N. J. Laycock,b,* H. S. Isaacs,a,* and R. C. Newmanc,*
aDepartment of Applied Science, Brookhaven National Laboratory, Upton, New York 11973, USA
bMaterials Performance Technologies, Lower Hutt, New Zealand
cCorrosion and Protection Centre, University of Manchester, Institute of Science and Technology, Manchester M60 1QD,
United Kingdom

Thin stainless steel films were prepared by sputter deposition onto silicon substrates using a 304 stainless steel target. The film
composition was essentially that of 304 stainless steel, but they had a body-centered cubic structure and were free of sulfide inclu-
sions. Potentiodynamic polarization curves were obtained for thin films in various chloride-containing solutions and compared to
results from conventional stainless steel samples. In addition, video images of two-dimensional pits in thin films were used to deter-
mine the anodic pit current density as functions of potential and chloride concentration. Thin stainless steel films were found to be
significantly more resistant to pit initiation than their bulk counterparts, but pit propagation was possible at relatively low poten-
tials. A diffusion-controlled growth regime was identified at high potentials, with a transition to mixed activation/ohmic control at
lower potentials (just above that required for repassivation).
© 1999 The Electrochemical Society. S0013-4651(98)04-038-5. All rights reserved.

Manuscript submitted April 14, 1998; revised manuscript received July 28, 1998.

Thin films of metals and alloys can be deposited onto substrates Bulk stainless steel pits tend to grow, at least initially, by under-
by sputtering or evaporation techniques such that the deposited films mining the surface such that a cover is created over the pit cavi-
are essentially free of the nonmetallic particles which are always ty,15,16 often described as having a “lace-like” appearance; these pit
incorporated in bulk materials. Thin films have been used to study covers play an important role in pit stability.17,18 Isaacs and Kissel19
pitting corrosion by Frankel and co-workers in studies of Al, 1 Al were able to increase metastable pit lifetimes by thermally treating
alloys,2 and Fe-Ni, 3 while others have worked with stainless steel samples to create thicker oxide films or by electropolishing rather
thin films.4-6 In all cases the thin film materials have exhibited a than mechanically polishing (so as to reduce surface stress). For Al
much greater resistance to pitting corrosion than the corresponding thin film pits, Frankel1 has observed overhanging strips of passive
bulk material. Inturi and Smialowska4 suggested, for 304 stainless film at the pit edges which increase the ohmic potential drop within
steel, that this was due to the nanocrystalline nature of thin films lead- the pit and help stabilize the dissolution.
ing to a lower average concentration of adsorbed chloride at each
defect and therefore increasing the driving force required for local- Experimental
ized acidification. Betts et al.6 used sputter deposition to investigate The thin films used in this work were prepared by sputter depo-
the effect of nitrogen incorporation in thin films of 316 stainless steel, sition onto chemically polished silicon wafer substrates using a tar-
referring to the resultant layers as s-phase coatings. At low nitrogen get of 304 stainless steel. The composition of the sputter deposited
levels the coatings had a body-centered cubic (bcc) structure, but with layer is essentially that of the bulk target, but the thin film is free
increasing nitrogen levels they became face-centered cubic (fcc). In from nonmetallic inclusions (such as MnS). In addition, X-ray dif-
localized corrosion tests they noted the increased resistance of the s- fraction showed that the thin films have a ferritic (bcc) structure,
phase coatings compared to bulk 316 steel, a beneficial effect of rather than the austenitic (fcc) structure of the bulk steel. The
nitrogen, and the apparent absence of current/potential fluctuations deposited film thickness was approximately 100 nm as monitored by
prior to the onset of stable pitting. The pitting of binary Fe-Cr thin a quartz crystal microbalance during deposition.
film alloys has also been studied and shows a sharp dependence on Specimens for electrochemical tests, approximately 1 3 1 cm,
alloy composition at around 16% Cr, above which pitting was not were cleaved from a wafer with a diamond scribe and glued on to an
observed, that was interpreted using percolation theory.7 This is dis- acrylic holder. Electrical contact was established by connecting cop-
cussed in more detail later in this paper. per wire to the alloy surface with silver conducting paint. To isolate
The kinetics of pit growth in stainless steels have been studied this connection from the electrolyte, the specimen edges were cov-
using artificial pit electrodes,8,9 and by using conventional elec- ered with epoxy resin, leaving an approximately circular electrode
trodes in solutions intended to simulate local pit chemistries.10,11 surface area of 0.25-0.5 cm2 (which was measured accurately for
These methods have found that the critical factor in pit stability is the each electrode). An electrochemical cell was constructed which
concentration of dissolved metal ions in the pit solution; for 300 allowed the samples to be held facing upward in the test solution,
series steels a solution greater than about 70% saturated in metal and to be viewed through an optically flat quartz window using a
chloride is required to sustain rapid dissolution and prevent repassi- Sony CCD video microscope at up to 100 times magnification. Con-
vation. Since such a high concentration is necessary as a minimum, tinuous video images of the sample surface were recorded through-
many pits are likely to become supersaturated and precipitate salt out each experiment.
films on their corroding surfaces. Microscopic examinations of Experiments were initially carried out in a variety of naturally
stainless steel pits have found a mixture of crystallographic etch pits aerated chloride solutions from 1-5 M NaCl or LiCl. The first set of
and polished pits, suggesting salt-free and salt-filmed dissolution experiments was carried out in solutions with constant concentra-
respectively,12,13 although it has been suggested that only salt-filmed tions, but in a second series of tests the solution was diluted at vari-
growth is possible.14 Frankel1 has studied the growth kinetics of pits ous times during the experiment. For example, an experiment was
in thin films of Al and showed diffusion-controlled (salt-filmed) started in 5 M LiCl and corrosion pits were initiated; a syringe was
growth at high potentials, with a transition to ohmically controlled then used to inject distilled water into the cell such that the concen-
(salt-free) growth at lower potentials, sometimes accompanied by a tration was reduced to either 2.5, 1.7, 1.3, or 0.8 M LiCl.
change in the pit perimeter from a smooth circle to a convoluted, A saturated silver/silver chloride reference electrode and plat-
fractal-like surface.1 inum counter electrode were used. An EG&G PAR 173 potentiostat
* Electrochemical Society Active Member. and ramp generator were used for electrochemical control and the
d Present address: Department of materials, Imperial College of Science, Technolo- current and potential data were recorded digitally at a sampling rate
gy and Medicine, London SW7 2BP, UK. of 0.5 Hz. After immersion of the samples in solution, the potential
92 Journal of The Electrochemical Society, 146 (1) 91-97 (1999)
S0013-4651(98)04-038-5 CCC: $7.00 © The Electrochemical Society, Inc.

was stepped to around 2400 mV (Ag/AgCl) and then swept in the ing that pit propagation is much easier than pit initiation in thin films.
noble direction at 1 mV s21. When pitting was seen to initiate on the However, the low surface roughness and lack of inclusions in thin
sample, the potential was held and pit growth monitored at constant films means that pits must initiate without the degree of occlusion
potential. At intervals during pit growth (usually about 30-200 s) the normally provided by those factors in bulk steels,18 and therefore
potential was stepped to a different value and the pit growth rate higher current densities are required to maintain a concentrated local
monitored at the new potential. In this way the pit growth rate was chemistry. A critical variable in pit propagation is xi, where x is the
measured as a function of applied potential in a variety of different pit depth an i is the current density, and an estimate of the current
chloride concentrations. A number of pitting potentials were also density before the pit penetrates through to the substrate can be made
measured in 5 M LiCl using bulk 304 samples with a surface area of using the critical value of xi, determined for hemispherical pits18 as
,1 cm2, polished to a 600 grit finish. 3 mA cm21. For a pit just as it reaches the substrate, x is 100 nm, and
Pits in the thin film samples tended to undermine the original sur- the corresponding current density is therefore 300 A cm22. Once a pit
face such that a very thin layer of this surface was left behind after has reached the substrate it propagates as a two-dimensional disk
pitting had stopped. A sample of this layer was carefully removed with an overhanging lip of the original surface at the edge of the
and placed on a copper grid for imaging and analysis in the trans- disk.1 In this stage of growth, current densities of 80 A cm22 were
mission electron microscope (TEM). measured in thin film pits on Fe-Cr alloys,7 suggesting that pit initia-
Results and Discussion tion at >100 A cm22 may indeed be possible in thin films.
The pitting potentials of thin films were measured in 1, 3, and
Pit initiation.—Typical anodic polarization data for 304 SS thin
5 M NaCl and in 5 M LiCl for comparison with the pitting resistance
films in 5 M NaCl and LiCl solution are shown in Fig. 1a and b. Both
of bulk 304 stainless steel. In 1 M NaCl, only 40% of the 304 thin
figures show a noisy current signal at the start of the experiment, but
film samples pitted at any potential and even in 5 M LiCl, where
the signal becomes less noisy as the test progresses. The onset of pit-
100% of samples pitted, some samples required transpassive poten-
ting is indicated by an almost vertical increase in current, making it
tials for pit initiation. Table I shows the measured thin film pitting
easy to define a pit initiation potential to an accuracy of 610 mV. In
potentials in all solutions and for bulk samples in 5 M LiCl. It is dif-
Fig. 1a the direction of the potential sweep was reversed once pitting
ficult to draw quantitative conclusions from the comparison between
had initiated, and then held at 350 mV before further changes were
thin film and bulk samples, due partly to the impossibility of com-
made (not shown in the figure). In Fig. 1b, pitting did not begin until
paring equivalent surface finishes and partly to the relatively small
after the onset of transpassive Cr dissolution, indicated by the small
surface area in these tests, which leads to higher measured pitting
increase in anodic current above ,850 mV.
potentials and greater variation.21 However, it is clear that thin films
In neither of the tests shown in Fig. 1a or b did repassivation of
resist pit initiation to a much greater degree than do bulk samples of
the pits occur, despite the potentials being lowered by 300 and
the same composition, as has been found by others.4-6
600 mV, respectively, from the apparent pit initiation potential; show-
The most important difference between pit initiation in bulk
stainless steels and in thin films is due to the absence of sulfide
inclusions from thin films. It is possible that some nonmetallic par-
ticles are incorporated in thin films during deposition, but the size
and number of inclusions is greatly reduced compared to even the
cleanest bulk steels. The results show that in the absence of these
inclusions, pit initiation is inhibited significantly, i.e., relatively high
chloride activities and applied potentials are required for initiation.
The obvious implication is that pit initiation in conventional stain-
less steels is almost always caused by inclusions, consistent with the
results of many experimental studies.22,23 When considering the role
of inclusions in pitting, effects on both pit initiation and on the first
stage of propagation must be considered. That is, the inclusion may
cause localized breakdown of passivity, but may also enhance the
localized dissolution kinetics via the action of sulfur species, such as
thiosulfate ions or adsorbed sulfur.24 This effect on kinetics can only
be transient, as any stable pit must eventually outgrow the supply of
the activating species.
However, it does not necessarily follow that pit initiation occurs
through completely different mechanisms in thin films and bulk sam-
ples. For example, Macdonald25 proposed that inclusions act by
locally increasing the defect density in the passive film, such that a
lower potential across the film is required for breakdown while the

Table I. Pitting potentials for thin film and bulk 304 samples in
various solutions. Values are in millivolts vs. Ag/AgCl.

Experiment 5 M LiCl
number 1 M NaCl 5 M NaCl 5 M LiCl (bulk 304)

1 No pits 141 12100 1240


2 1400 430 12866 2104
3 1370 No pits 12213 2144
Figure 1. (a, top) Cyclic anodic polarization curve for thin film 304 in 5 M 4 No pits 11217 2136
LiCl solution, showing pit initiation in the passive region and hysteresis on 5 No pits 12981 1236
the reverse sweep. (b, bottom) Anodic polarization curve for thin film 304 in 6 710 12630
5 M LiCl solution, showing pit initiation in the transpassive region. The 7 600 21100
reverse potential data was obtained from a sequence of potential steps. 8 700 21200
Journal of The Electrochemical Society, 146 (1) 91-97 (1999) 93
S0013-4651(98)04-038-5 CCC: $7.00 © The Electrochemical Society, Inc.

breakdown process remains the same. Similarly, Williams et al.26


suggested that the passive current in stainless steels is made up from
transient bursts of current, leading to localized fluctuations in pH,
chloride activity, and potential, which will occasionally increase in
magnitude through a feedback mechanism and become pits. Williams
et al.27 went on to propose a mechanism for local fluctuations in pas-
sive current density based on the existence of randomly distributed
iron-rich clusters within the metal lattice. Rapid, picoamp-sized cur-
rent transients have since been observed and identified as pit initia-
tion events.28,29 More recently, Zhu and Williams30 have used a scan-
ning electrochemical microscope to image the sites of these events
and show the progression from initiation event to propagating pit; ini-
tiation events were seen all over the surface, but a propagating pit
only formed at an inclusion site.
We have recently provided further evidence in support of
Williams’ model27 for pit initiation by testing the pitting resistance
of Fe-Cr alloys in dilute hydrochloric acid solutions.7 At Cr contents
greater than 16 atomic %, pitting did not occur in either potentiody-
namic or potentiostatic tests. The results were interpreted in terms of
achievement of a high density percolation condition for iron, i.e., at Figure 2. Pit radius vs. time for two pits growing simultaneously on one 304
a critical iron concentration there exists an infinitely connected clus- thin film sample at 200 mV in 5 M NaCl.
ter of Fe atoms that have only iron nearest neighbors; these are the
initiation sites for pitting. At higher Cr contents only relatively small
iron-rich clusters are present and, in the absence of nonmetallic the presence of favorable defects, are able to propagate at lower
inclusions, pit initiation is not possible. Given the lack of pitting in potentials without the passivating influence of chromate ions.
thin film Fe-Cr alloys with greater than 16% Cr, it is perhaps sur- Pit propagation.—The growth rate of individual pits was deter-
prising that pits were initiated in 304 thin films. A possible explana- mined under potentiostatic conditions at each step. An example of
tion is the presence of Ni in the 304 which may somehow produce a these measurements is shown in Fig. 1b. Since the pits initially prop-
more defective passive oxide, perhaps related to enrichment of Ni at agate as 2D disks, and the thin film has a known thickness, video
the metallic surface beneath the oxide.31 Alternatively, the passive images can be used to determine the actual anodic current density in
oxide in the 304 could be “stronger” than that on the binary Fe-Cr the pit, regardless of cathodic reactions within the pit, such as hydro-
alloys and so provide a more effective diffusion barrier during the gen evolution. The corrosion rate at any particular potential was cal-
very early stages of growth, in a manner similar to that described by culated from the rate of propagation, dr/dt, of a corroding pit edge,
Isaacs and Kissel.19 (This is consistent with diffusion length calcu- usually measured over a time interval of 20-60 s. The measurement
lations discussed later in this paper.) A simpler explanation is that was made using at least three points from around the pit perimeter,
the 304 films are rougher than the FeCr films they are compared with from which the average value was taken, and the standard deviation
here, or that more defects (such as dust particles) were incorporated was given as the error. This method of calculation assumes a con-
during deposition of the 304. Other pit initiation mechanisms have stant current density over the time interval of each measurement; an
been proposed by Macdonald et al.32 and by Mattin and Burstein.33 assumption which is supported by the results of successive measure-
In the point-defect model of Macdonald, the apparent effect of Ni in ments over a long time period at a fixed potential. Figure 2 shows
our data could possibly be explained by an injection of lattice defects plots of r vs. t for two pits on the same 304 thin film sample at
into the passive film due to incorporation of Ni. However, many 200 mV in 5 M NaCl. The linear variation of r with t indicates a con-
authors have reported the absence of Ni from the passive film on stant current density, and the different gradient of the two plots indi-
stainless steels34 based on surface analytical techniques. In the initi- cates a real variation in current density between pits initiated under
ation scheme proposed by Mattin and Burstein, a metal salt forms identical conditions.
beneath the passive film and causes the oxide above it to rupture. To calculate a current density for these pits, Eq. 1-2 are applied,
The metal salt is assumed to be essentially ferrous chloride, but the where a is the experimentally measured rate of propagation and
solubility of this salt could be lowered by the incorporation of Ni, Eq. 2 is an application of Faraday’s law
thereby encouraging pit initiation.
dr/dt 5 a [1]
The results obtained for FeCr7 and 304 thin films are consistent
with the following hypothesis: pit initiation is caused by the selec- i 5 anFr/Z [2]
tive dissolution of iron-rich clusters which can make the transition to
become propagating pits only when the cluster is above a critical i is the anodic current density in the pit, n is the average number of
size (defined by a high density 3D percolation threshold for iron), or equivalents in the dissolution reaction, F is Faraday’s constant, Z is the
if extra stability is conferred by the geometry of the initiation site. In average atomic number, and r is the density of the thin film. Stoichio-
bulk steels, MnS inclusions provide both a favorable geometry and
reduced sulfur species to catalyze anodic dissolution; in 304 thin
films, the undermined passive film itself can support the stabilization
of initiation events below the critical size, although relatively high Table II. Mean ionic activity coefficients and calculated
activities for solutions used in this work.41
applied potentials are required.
Within the set of results for 304 in 5 M LiCl, there appear to be Solution Ionic activity coefficient Activity
two separate groups of data; one group of pits initiated at relatively
low potentials (220 to 1210 mV) and repassivated below
2200 mV, while a second group initiated at >600 mV and repassi- 1.0 M NaCl 0.66 10.7
5.0 M NaCl 0.87 14.4
vated near 300 mV. As a tentative explanation of this bimodal distri- 0.8 M LiCl 0.76 10.6
bution, we suggest that the higher potential subset are possibly influ- 1.3 M LiCl 0.80 11.0
enced by production of Cr61 ions at the initiation stage, which are 1.7 M LiCl 0.90 11.5
subsequently reduced at lower potentials, encouraging repassivation. 2.5 M LiCl 1.03 12.6
Pits which initiate without transpassive chromium oxidation, due to 5.0 M LiCl 2.02 10.1
94 Journal of The Electrochemical Society, 146 (1) 91-97 (1999)
S0013-4651(98)04-038-5 CCC: $7.00 © The Electrochemical Society, Inc.

metric dissolution within the pit was assumed to produce Fe21, Ni21, FeCl2) within the pit. The solubility of FeCl2 is known to be 4.2 M
and Cr31 such that the values of n and Z were 2.2 and 55.5, respec- in water at 238C 36; assuming this value is reasonable in the less con-
tively, while r was taken to equal that of the bulk 304, 7.9 g cm23. centrated LiCl solutions, and approximating the diffusion as one-
Using this method, pit current densities were determined over a dimensional, Fick’s first law can be used to calculate an effective dif-
range of applied potentials in solutions of various chloride activity fusion length of ,10 mm. Figure 4 shows the assumed pit geometry
from 0.5 to 10 (in LiCl solutions the mean ionic activity coefficient (after Frankel1), with an overhanging lip of undermined material at
increases to higher values than in NaCl solutions of similar concen- the pit edge, which provides a diffusion barrier for the dissolution
tration, Table II). Figures 3a-d show the variation of anodic current products and helps to stabilize the pit. The effective diffusion length
density within individual pits, ia, as a function of potential in a range is large compared to the 100 nm thickness of the film, and the real
of solutions. Within experimental error, ia is independent of potential situation is undoubtedly more complex than that shown in Fig. 4. In
over a wide range in all solutions, indicating that the dissolution is all conditions, pits propagate beneath the original surface leaving a
under diffusion control. The diffusion-limited current density, ilim, is transparent film over the corroded area; this film can be seen clearly
noticeably higher in the lower chloride solutions than in the 5 M as a convex dome over each of the two pits in Fig. 5. The overhang-
LiCl solution (see Fig. 3a-d), as would be expected35 since it de- ing lip of passive film, found by Frankel on pits in thin film Al, is
pends directly on the solubility of metal chloride salts (essentially extended in these 304 thin films to cover the entire pit. Hydrogen

Figure 3. Pit anodic current density, ia, as a function of applied potential for single pits in (a) 5 M LiCl, (b) 2.5 M LiCl (c) 1.7 M LiCl, and (d) 1.3 M LiCl. Each
point is the average of at least three measurements from different sites around the pit perimeter. Error bars represent 6 the sample standard deviation.
Journal of The Electrochemical Society, 146 (1) 91-97 (1999) 95
S0013-4651(98)04-038-5 CCC: $7.00 © The Electrochemical Society, Inc.

during growth of a single pit, initially in 5 M LiCl, but with dilution


of the solution to 1.3 M NaCl after ,100 s growth (as indicated in
the figure). The applied potential is also plotted on the same figure.
At first, the current increases as the pit grows due to the increasing
surface area. Dilution of the solution causes a step increase in the
current, due to the increased solubility of metal ions as the chloride
activity decreases from that in the original solution. After allowing
some time for pit growth to stabilize in the new solution, the applied

Figure 4. Schematic diagram of ideal 2D pit geometry.

bubbles produced at the corroding pit edge (visible beneath the cover
in Fig. 5) must also provide a significant resistance to current flow.
In the previous study of FeCr alloys7 the pits were found to grow at
a much higher rate, implying an effective diffusion length of only
,2 mm. It is possible, then, that the cover over 304 pits is inherent-
ly more stable (stronger) than those on the Fe-Cr alloys, thus help-
ing an incipient pit in 304 to establish the local chemistry required
for pit propagation.
At the lower end of the potential range tested, the pit current den-
sity varied approximately linearly with potential (Fig. 3); this poten-
tial range, between diffusion control and repassivation, is relatively
small. Figure 6 shows the measured current as a function of time

Figure 5. Video images of 304 thin film pits in 5 M NaCl as a function of


time (a-d). The potential was stepped from 200 to 400 mV between images
(b) and (c) as indicated.

Figure 6. Measured electrochemical current, I, and potential E, as a function


of time for a 304 thin film sample in 5 M LiCl with dilution to 1.3 M LiCl as Figure 7. Pit perimeter change as a function of time and potential for a 304
indicated. thin film pit in 1.3 M LiCl.
96 Journal of The Electrochemical Society, 146 (1) 91-97 (1999)
S0013-4651(98)04-038-5 CCC: $7.00 © The Electrochemical Society, Inc.

potential was decreased in a series of steps: decreasing the potential the pit edges and collapsing back onto the surface. The pits contin-
from 630 to 580 mV had little effect on the measured current, but ued to propagate without the cover, although some areas were repas-
between 580 and 370 mV, the current decreased with potential. A sivated. This event suggests that hydrogen evolution increases with
further potential step from 370 to 310 mV caused this particular pit increasing applied potential, which has been previously reported in
to repassivate. Figure 7 shows the outline of this pit as a function of
time as the potential was lowered from 630 to 310 mV. It was a fea-
ture of this and other pits, that the corrosion front changed from a
smooth curve at higher potentials to a more convoluted structure at
lower potentials, with some repassivated regions. This effect made it
more difficult to measure accurate propagation rates in these condi-
tions, but is in itself an indicator that pit propagation was no longer
under diffusion control. Frankel has reported a similar change for
pits in thin film Al as the growth regime changes from diffusion con-
trol to ohmic control,1 and in bulk stainless steels many authors have
found a transition from polished pit surfaces at high potentials to
crystallographic pits at lower potentials.12,13
The potential dependence of the current density can be interpret-
ed using the results of a number of artificial pit experiments.8,9 At
high potentials, a metal chloride salt film is precipitated on the cor-
roding surface and growth is diffusion controlled, but at lower poten-
tials, the salt film is not present and growth is under mixed activa-
tion/ohmic control. In this region, the current density varies approx-
imately linearly with potential. Pit propagation requires a balance
between anodic dissolution and diffusion of corrosion products away
from the pit, such that steady states of propagation exist only where
diffusion and dissolution rates are equal. At potentials just below the
diffusion-controlled region, some parts of the pit repassivate (where
dissolution and diffusion are both effectively zero), while others
adopt a steady state where no salt film is present, with a current den-
sity just below the diffusion-limited value. This is possible because
of the strong, nonlinear dependence of the dissolution rate on the
local concentration of corrosion products.8
An additional feature of these experiments was the apparent
absence of metastable pitting; no short lived current transients were
detected (at least within the resolution limit of ,50 nA) and no vis-
ible pits formed but then repassivated (without a decrease in the
applied potential). This is consistent with the earlier observation that
the potential required for pit initiation in thin films is much higher
than that necessary to sustain 2D pit propagation, but is in contrast
with bulk stainless steels, where metastable pits are found at poten-
tials17,18 and temperatures37 where stable propagation is not possi-
ble. For bulk stainless steels, picoamp-sized pit initiation events have
been detected at all potentials in the passive range,33 and higher res-
olution experiments on thin films are required to confirm that pit ini-
tiation is indeed restricted to high potentials. Pitting events with peak
currents up to the noise limit of the present experiments would pen-
etrate to the substrate in less than 0.1 s, which is too fast to have been
resolved in this work.

Pit covers.—In bulk stainless steels, pits are well known to


undermine the original metal surface and propagate beneath a perfo-
rated pit cover, which is thicker than the passive oxide layer.38 As
shown in Fig. 5, the cover over thin film pits was a transparent layer
which usually stayed in place throughout the entire experiment. The
cover formed a convex dome, with a wrinkled appearance where it
was joined to the pit edges. Often, the cover would tear open to leave
a central hole through which hydrogen bubbles would immediately
escape in a stream (Fig. 8). Mankowski and Smialowska10 have
observed convex pit covers on bulk steel samples and found that col-
ored streams of pit solution sometimes escaped from holes in this
cover during growth. They proposed that the curvature of the cover
is caused by osmotic pressure due to a decreased water concentra-
tion in the pit relative to the bulk solution. In the present experi-
ments, it seems more likely that the pressure of gaseous hydrogen
evolved within the pit is responsible for forcing the pit cover to tear.
In one experiment (Fig. 5), two pits were growing beneath complete
covers in 5 M NaCl at a potential of 200 mV, which was then stepped Figure 8. Successive images (a and b) of a 304 thin film pit in 5 M LiCl
to 400 mV. Immediately following the potential step, the pit cover showing formation of a hole in the pit cover and escape of hydrogen bubbles.
expanded like a balloon and became taut before breaking away from Pit continues to grow after disruption of the film.
Journal of The Electrochemical Society, 146 (1) 91-97 (1999) 97
S0013-4651(98)04-038-5 CCC: $7.00 © The Electrochemical Society, Inc.

a transition to mixed activation/ohmic control and rougher, more


convoluted pit perimeters.
4. Thin film pits propagate beneath the undermined passive film
even up to pit diameters of a few millimeters. TEM examination
reveals this film to have a crystalline structure.
Acknowledgments
This research was performed under the auspices of the U.S.
Department of Energy, Division of Materials Sciences, Office of Basic
Energy Sciences under contract no. DE-AC02-76CH00016 and by the
New Zealand Ministry of Research, Science and Technology under
contract CO8621, with assistance from the Royal Society of New
Zealand as part of the ISAT linkages program. The authors thank G. S.
Frankel for useful discussions.
Brookhaven National Laboratory assisted in meeting the publication
costs of this article.
References
1. G. S. Frankel, Corros. Sci., 30, 1203 (1990).
2. G. S. Frankel, J. O. Dukovic, V. Brusic, B. M. Rush, and C. V. Jahnes, J. Elec-
Figure 9. TEM micrograph with diffraction pattern (inset) from a section of trochem. Soc., 139, 2196 (1992).
the undermined pit cover of a 304 thin film pit in 5 M LiCl. 3. G. S. Frankel, R. C. Newman, C. V. Jahnes, and M. A. Russak, J. Electrochem. Soc.,
140, 2192 (1993).
4. R. B. Inturi and Z. Szklarska-Smialowska, Corrosion, 48, 398 (1995).
5. M. Kraack, H. Böhni, and W. Muster, Mater. Sci. Forum., 192-194, 165 (1995).
the pitting of aluminum.39 Initially, the pits were growing without a 6. A. J. Betts, K. Dahm, P. Dearnley, and G. A. Wright, in Advanced Materials Devel-
opment and Performance, W. G. Ferguson and W. Gao, Editors, p. 110, The Uni-
salt film, the subsequent potential step caused precipitation of a thick versity of Auckland, New Zealand (1997).
salt film and a minimum in anodic current density due to the poten- 7. M. P. Ryan, N. J. Laycock, H. S. Isaacs, and R. C. Newman, J. Electrochem. Soc.,
tial drop across the salt layer.40 Immediately following precipitation, 145, 1566 (1998).
before the salt film reached an equilibrium thickness, the potential at 8. H. S. Isaacs, J. Electrochem. Soc., 120, 1456 (1973).
9. G. T. Gaudet, W. T. Mo, J. Tilly, J. W. Tester, T. A. Hatton, H. S. Isaacs, and R. C.
the metal/salt interface reached a minimum and hydrogen evolution Newman, AIChE J., 32, 949 (1986).
was sufficient to force the cover off the two pits in Fig. 5. It was 10. J. Mankowski and Z. Szklarska-Smialowska, Corros. Sci., 15, 493 (1975).
apparent in most experiments that at the lowest potentials, where pit 11. T. Hakkarainen, in Corrosion Chemistry Within Pits, Crevices and Cracks, A. Turn-
growth was in the salt-free state, there appeared to be less hydrogen bull, Editor, p. 17, HMSO (1987).
12. W. Schwenk, Corrosion, 20, 129t (1964).
bubbles at the pit edge than were present at higher potentials. 13. N. Sato, Corros. Sci., 37, 1947 (1995).
The apparent toughness of the passive film on stainless steel 14. K. J. Vetter and H. H. Strehblow, in Localized Corrosion, R. W. Staehle, B. F.
means that the pits grow with large, tenacious covers in place that act Brown, J. Kruger, and A. Agrawal, Editors, p. 240, NACE, Houston, TX (1974).
as diffusion barriers and so little ohmically control is seen at low 15. I. L. Rosenfeld and I. S. Danilov, Corros. Sci., 7, 129 (1967).
16. H. W. Pickering and R. P. Frankenthal, J. Electrochem. Soc., 119, 1297 (1972).
potentials. This is in contrast to Al in which the passive film is easi- 17. H. S. Isaacs and G. Kissel, J. Electrochem. Soc., 119, 1628 (1972).
ly broken by the large amounts of hydrogen evolution in the pit, and 18. G. S. Frankel, L. Stockert, F. Hunkeler, and H. Böhni, Corrosion, 43, 429 (1987).
where larger ohmic regions were observed.2 19. P. C. Pistorius and G. T. Burstein, Philos. Trans. R. Soc. London A, 341, 531 (1992).
The pit cover from a 304 thin film pit was removed carefully after 20. J. R. Galvele, J. Electrochem. Soc., 123, 464 (1976).
21. G. T. Burstein and G. O. Ilevbare, Corros. Sci., 38, 2257 (1996).
one experiment, and Fig. 9 shows a diffraction pattern obtained from 22. M. Smialowski, Z. Szlarska-Smialowska, M. Rychick, and A. Szummer, Corros.
this sample, indicating a crystalline structure. EDX analysis revealed Sci., 9, 123 (1969).
the cover to be Cr rich, with some Fe also present, but no Ni. This 23. J. Stewart and D. E. Williams, Corros. Sci., 33, 457 (1992).
composition suggests that the thin film pit cover is simply the pas- 24. A. J. Kucernak, R. Peat, and D. E. Williams, J. Electrochem. Soc., 139, 2337
(1992).
sive (hydr)oxide layer and that the metal has been corroded away 25. D. D. Macdonald, J. Electrochem. Soc., 139, 3434 (1992).
while the passive film above it has remained intact.31,34 This is in 26. D. E. Williams, C. Westcott, and M. Fleischmann, J. Electrochem. Soc., 132, 1804
contrast with pit covers in bulk steels, which can be more than a (1985).
micron thick and consist of both metal and oxide.38,41 Unfortunate- 27. D. E. Williams, R. C. Newman, Q. Song, and R. G. Kelly, Nature, 350, 216 (1991).
28. A. M. Riley, D. B. Wells, and D. E. Williams, Corros. Sci., 32, 1307 (1991).
ly, the conditions of formation of the passive film in these experi- 29. G. T. Burstein and S. P. Mattin, Philos. Mag. Lett., 66, 127 (1992).
ments are not well defined; the (hydr)oxide would have formed 30. Y. Zhu and D. E. Williams, J. Electrochem. Soc., 144, L43 (1997).
immediately on exposure to the air after the sputter deposition 31. C. R. Clayton and I. Olefjord, in Corrosion Mechanisms in Theory and Practice, P.
process and could not be removed by cathodic treatment in the near- Marcus and J. Oudar, Editors, pp. 175-200, Marcel Dekker, Inc., New York (1995).
32. L. F. Lin, C. Y. Chao, and D. D. Macdonald, J. Electrochem. Soc., 128, 1194 (1982).
neutral pH conditions of these tests. The film structure could also 33. G. T. Burstein and S. P. Mattin, in Critical Factors in Localized Corrosion II, R. G.
have been affected by the anodic polarization applied to control the Kelly, P. M. Natishan, G. S. Frankel, and R. C. Newman, Editors, PV 95-15, p. 1,
pitting process. Nevertheless, the TEM data shown here are consis- The Electrochemical Society Proceedings Series, Pennington, NJ (1995).
tent with recent surface-sensitive X-ray diffraction experiments 34. I. Olefjord, B. Brox, and U. Jelvestam, J. Electrochem. Soc., 132, 2854 (1985).
35. J. W. Tester and H. S. Isaacs, J. Electrochem. Soc., 122, 1438 (1975).
showing that the air-formed oxide film on 304 stainless steel has a 36. H. C. Kuo and D. Landolt, Electrochim. Acta, 20, 393 (1975).
nanocrystalline spinel structure.42 Further discussions on the oxide 37. N. J. Laycock, M. H. Moayed, and R. C. Newman, J. Electrochem. Soc., 145, 1101
structure will be presented elsewhere.43 (1998).
38. P. Ernst, N. J. Laycock, M. H. Moayed, and R. C. Newman, Corros. Sci., 39, 1133
(1997).
Conclusions 39. T. R. Beck, Electrochim. Acta, 29, 485 (1984).
40. H. S. Isaacs and R. C. Newman, in Corrosion and Corrosion Protection, R. P.
1. Sputter deposited thin films of 304 stainless steel have signif- Frankenthal and F. Mansfeld, Editors, PV 81-8, p. 120, The Electrochemical Soci-
icantly increased resistance to pit initiation compared with bulk 304 ety Proceedings Series, Pennington, NJ (1981).
stainless steel. 41. N. J. Laycock, S. P. White, J. S. Noh, and R. C. Newman, J. Electrochem. Soc., 144,
2. Pits can be initiated in thin films in solutions of high chloride 1101 (1998).
activity and then grow at a high rate (up to 15 A cm22). 42. M. P Ryan, L. J. Oblonsky, and M. F. Toney, Abstract 233, p. 288, The Electro-
chemical Society Meeting Abstracts, Vol. 97-1, Montreal, Quebec, Canada, May 4-
3. At high applied potentials, pits propagate as smooth 2D disks 9, 1997.
with a diffusion-limited current density, but at low potentials there is 43. M. P. Ryan and N. J. Laycock, In preparation.

Você também pode gostar