Vous êtes sur la page 1sur 13

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/223831935

Passive films on stainless steelschemistry,


structure and growth
ARTICLE in ELECTROCHIMICA ACTA APRIL 2003
Impact Factor: 4.5 DOI: 10.1016/S0013-4686(02)00841-1

CITATIONS

DOWNLOADS

VIEWS

242

227

239

2 AUTHORS, INCLUDING:
Dieter Landolt
cole Polytechnique Fdrale de Lausanne
227 PUBLICATIONS 5,325 CITATIONS
SEE PROFILE

Available from: Dieter Landolt


Retrieved on: 08 August 2015

Electrochimica Acta 48 (2003) 1093 /1104


www.elsevier.com/locate/electacta

Passive films on stainless steels* chemistry, structure and growth


/

C.-O.A. Olsson *, D. Landolt


cole Polytechnique Federale de Lausanne, CH-1015 Lausanne EPFL, Switzerland
Departement des Materiaux, Laboratoire de Metallurgie Chimique, E

Abstract
The outstanding corrosion resistance of stainless steels results from the presence of a thin oxide */passive */film on the metal
surface, typically 1 /3 nm thick. The characterisation of the composition and structure of such thin films and the study of their
interaction with corrosive environments requires a combination of sophisticated experimental techniques. This paper reviews
progress in the characterisation and understanding of passive films on stainless steels achieved over the past two decades. During
this period, ex situ surface analysis methods have made substantial progress and new in situ methods for the study of passive films
with atomic resolution have been introduced, giving real time information on film chemistry and growth. It has been found that
whereas passive film growth occurs in seconds or minutes, long range film ordering is a considerably slower process that takes
several hours. In situ investigations indicate that at short times, charge transfer at the metal/film or the film/solution interface limits
the rate of film growth on stainless steels. In situ estimates of film composition confirm previous data obtained with ex situ
techniques.
# 2003 Elsevier Science Ltd. All rights reserved.
Keywords: Passive films; Stainless steel; Passivity; X-ray photoelectron spectroscopy; Scanning tunnelling microscope; X-ray absorption near edge
structure; Electrochemical quartz crystal microbalance

1. Introduction
Stainless steels were invented almost a century ago by
Monnartz [1]. Today, their usage shows an average
increase rate of about 5% per annum. They are divided
into four main groups based on their microstructure:
ferritic, austenitic, martensitic and austeno /ferritic
(duplex). Among the ferritic steels, one finds a series
of Fe /Cr alloys, e.g. Fe /17Cr (AISI 430), but also
superferritics containing high fractions of molybdenum.
Austenitic stainless steels are usually alloyed with Ni, for
example 18Cr /9Ni (AISI 304); superaustentic stainless
steels containing high fractions of Mo and N can in
certain environments surpass the corrosion resistance of
Ni-base alloys. Stainless steels are subject to continuing
development; some recent inventions are listed in Table
1. Thermodynamic modelling has made it possible to
develop new grades with high contents of molybdenum
and nitrogen (654 SMO), that are extremely resistant to
pitting attacks induced by halides[2]. Another example
* Corresponding author. Fax: /41-21-693-3946.
E-mail address: claes.olsson@epfl.ch (C.-O.A. Olsson).

of the utility of thermodynamics is the super duplex


stainless steel SAFUREX, that has been developed to
withstand highly oxidative environments [3]. With
thermodynamic models, it is possible to find compositions with high contents of Cr, Ni and N that give the
desired phase balance, normally 50/50 austenite/ferrite.
Extreme strength levels can be obtained by introducing
high fractions of nitrogen and manganese. Further
strength increase can, for austenitic materials, be
obtained by cold deformation. One example is the
high Mn, high N steel PANACEA [4]. Other developments include surfaces with structures tailored to avoid
bio fouling (UGICLEAN). Copper alloying in combination with a heat treatment is used to produce a special
line of anti-microbial stainless steels (NSS AM: 1 /4).
The literature shows that the corrosion resistance of
stainless steels can be drastically improved by suitable
alloying. This is possible because the composition and
properties of the passive films depend on the alloy
composition. Numerous investigations have contributed
to our present understanding of the processes governing
film growth and break down as a function of alloy
composition and environment. Today, it is known that
the passive film adapts to changes in potential or anion

0013-4686/03/$ - see front matter # 2003 Elsevier Science Ltd. All rights reserved.
doi:10.1016/S0013-4686(02)00841-1

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

1094

Table 1
Examples of recent commercial developments in the field of stainless steel with high impact on the surface behaviour of the material
Product

Significance

Producer

654 SMO
SAFUREX

First 7% Mo austenitic stainless steel. Very high resistance to pitting corrosion at elevated temperatures
Avesta Polarit
First 29Cr duplex stainless steel. Very high resistance to oxidising acids and localised corrosion, in particular stress Sandvik Steel
corrosion cracking.
UGICLEAN A special surface structure modified to resist bacterial adhesion
Ugine
NSS AM1-4 Cu alloyed stainless steels with special heat treatment giving high resistance to bacterial adhesion
Nisshin steel
Zeron 100
Duplex stainless steel alloyed with W
Weir Materials
DP3-W
W alloyed stainless steel
Sumitomo

concentration in the electrolyte. The dynamic properties


of the passive film provide the key to the high resistance
of stainless steels to corrosive attacks. The state of
knowledge of passive films on stainless steels as of 20
years ago was thoroughly reviewed by Fischmeister and
Roll [5]. The aim of this paper is to present an overview
of new methods and results reported in the past two
decades. One could say that whereas the keyword of the
70s and early 80s was ex situ surface analysis, recent
research has emphasised the use of different in situ
methods.

2. Experimental methods for the study of passive films


Passive films are formed during an exposure of the
bare metal surface to an oxidising environment. Once a
film is formed, the reaction rate between the metal and
the environment will be several orders of magnitude
lower. The original theory of film formation goes back
to Michael Faraday, who in the 19th century studied
iron surfaces and found them altered. A review of the
early days of passive film research has been written by
Uhlig [6]. A general introduction to the theory of
passivity has been published by Sato [7], whereas the
electronic properties of passive films on different
materials have recently been reviewed by Schultze and
Lohrengel [8].

Based on surface analysis, a three-layer model has


been suggested for passive films formed on austenitic
stainless steels in acidic solutions: the outer part of the
film consists of an hydroxide film on top of an oxide
layer. The oxy-hydroxide film is formed on top of a
nickel enriched layer in the metal, the origin of which is
the selective oxidation of Fe and Cr during anodic
polarisation. This picture is obtained essentially from
XPS measurements on passive films, cf. Fig. 1 [9]. The
film is enriched in chromium and the metal closest to the
metal/film interface shows a strong enrichment in nickel.
For an alkaline solution, the solubility is lower for Fe
and higher for Cr; this affects the level of chromium
enrichment in the film. With angular resolved XPS
(ARXPS), it is also possible to obtain information on
the depth distribution of different species and oxidation
states, cf. Fig. 2, which illustrates the angular dependency of different oxidation states of molybdenum in
the passive film after immersion in a ferric chloride
solution [10]. Mo(VI) is enriched in the surface region,
the metal peaks show a buried layer appearance, and the
Mo(IV) oxide and Mo(IV) oxy-hydroxide show an
angular distribution indicating a more homogeneous
distribution through the passive film. The attribution of

2.1. Ex situ methods


Electron spectroscopy has been used since the early
1970s for the study of the composition and thickness of
the passive films, including the determination of oxidation states and of the depth distribution of the different
constituting elements in the film. X-ray photoelectron
spectroscopy (XPS, also known as ESCA, electron
spectroscopy for chemical analysis) and Auger electron
spectroscopy (AES) have been most widely used, but
other methods such as secondary ion mass spectroscopy
(SIMS) and ion scattering spectroscopy (ISS) have also
provided important information.

Fig. 1. XPS results from a passive film formed in the high passive
region in a 0.1 M HCl/0.4 M NaCl solution. The oxide film is found
to be strongly enriched in chromium, whereas the metal layers closest
to the metal/film interface, i.e. the apparent metal concentration, are
strongly enriched in nickel. The passive film itself shows close to no
incorporation of nickel. After Olefjord and Elfstrom [9].

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

Fig. 2. Concentration gradients in a passive film for molybdenum,


recorded for a 6Mo superaustenitic stainless steel after immersion in
ferric chloride using XPS. For angles close to grazing, there is a strong
contribution of Mo(VI), whereas metallic Mo dominates for angles
close to perpendicular. The IV-valued oxide and oxy-hydroxide states
show less angular dependence. After Olsson and Hornstrom [10].

oxidation states to binding energy has been detailed by


Brox and Olefjord [11].
Superior depth resolution can be obtained using ISS,
which also gives quantitative results. By choosing the
proper primary ions, a resolution sufficient to distinguish between the Fe and Cr contributions to the signal
can be obtained. With ISS, Calinski and Strehblow
showed strong concentration gradients within the passive film on an Fe /Cr alloy [12], cf. Fig. 3. The outer
region of the passive film consists of a layer enriched in
iron, a feature that has also been observed with XPS
sputter depth profiles [13]. In the centre of the passive

Fig. 3. Enrichment and concentration variations for Cr in a passive


film obtained by ISS. Even for low alloying amounts of Cr (5%), there
is an enrichment of Cr in the film, with the exception of the outer part.
After Calinski and Strehblow [12].

1095

film, a strong enrichment of chromium is found,


whereas for the region closest to the metal/film interface,
the composition corresponds to that of the bulk metal.
This implies that the oxidation at the metal/film interface closely follows the composition of the bulk metal,
and that the enrichment in chromium is caused by
dissolution of iron. The investigation of passive layers
on different metals with XPS and ISS has been reviewed
by Strehblow [14].
In 1984, Fischmeister pointed out that XPS is
generally better suited for analysis of passive films on
stainless steels than AES [5]. AES has a higher lateral
resolution and a higher surface sensitivity for some
elements, e.g. molybdenum. The difficulty in analysing
Cr oxides with AES due to the peak overlap between the
OKLL and CrLMM peaks can be resolved through target
factor analysis (TFA) [15]. It is also possible to facilitate
the analysis of Mo by using krypton as a sputtering gas
instead of argon [15,16]. In spite of these improvements,
Fischmeisters preference is still valid, since the state of
the art resolution of XPS for spectral information has
been improved to about 5 mm. The reproducibility of
AES and XPS measurements between different laboratories has been investigated in a round-robin where two
types of ferritic stainless steels were exposed to acidic
sulphate solutions. In terms of raw data (binding energy
shifts, relative intensities of oxide peaks etc.) the overall
agreement between different laboratories was judged
satisfactory [17].
A very critical issue with ex situ surface analysis is the
sample transfer from electrolyte to UHV. No matter
how much precaution is taken, the surface will alter
somewhat during the transfer. Different ingenious
methods for sample transfer without contact with
atmosphere have been developed; one example of such
a design is given by Haupt et al. [18]. The level of utility
of an in vacuo transfer system is determined by the
amount of sample oxidation that can be avoided. The
amount of extra oxidation during transfer in air can be
estimated by forming the passive film in a normal
electrolyte, and making the transfer in an atmosphere
enriched in 18O. The amount of 18O in the film will then
be a measure of the amount of oxidation that has
occured in atmosphere. The inverse experiment is also
possible, i.e. a passive film formed in an electrolyte
enriched in 18O and subsequent transfer in normal
laboratory atmosphere. The different passive films are
then compared using SIMS that can distinguish between
the 16O and 18O isotopes. Graham et al. polarised FeCr
alloys in water enriched in H18
2 O [19 /21], and found that
for pure iron, there is a small effect of sample transfer on
the passive film for polarisations in the passive region,
whereas a more pronounced thickness change in the film
was found for Fe /26Cr samples; it was suggested that
this difference could be caused by the higher ability of
Fe /Cr films to adapt to different environments. Similar

1096

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

experiments have been performed by Courty et al. [22].


A direct estimate of the film change during transfer was
obtained by comparing films formed in H18
2 O and
transfer in 16O2 with films formed in H16
2 O and transfer
in 18O2. For low passive and cathodic potentials, both
laboratories found a significant effect of sample transfer.
Graham et al. reported a transfer influence also for the
high passive region. However, when Courty et al.
transferred samples in 18O2 atmosphere, no uptake of
18
O in the film was seen, even though enrichment levels
close to those of Grahams were observed for anodisa16
tion in H18
O2. Thus, the direct
2 O and transfer in
measurement showed an absence of oxygen uptake
during transfer for samples polarised to the high passive
region [22].
The results indicate that caution is needed during the
sample transfer. However, by ensuring that the transfer
is made consistently for all samples, it is possible to draw
valid conclusions from comparisons between series of
measurements. Evidently, the only way to eliminate
artefacts associated with sample transfer is to use in situ
techniques.

2.2. In situ techniques


A technique new to the study of passive films is X-ray
absorption near edge structure (XANES), which acquires in situ chemical information. It has been applied
to the investigation of passive films on pure iron [23 /
27], iron /chromium alloys [28] and to the study of
sputter deposited iron /chromium oxides [29]. The
sputter deposited oxide films were used as model
systems for passive films. XANES puts strong demands
on the analysed material. The sample is a thin film of
about 10 nm thickness deposited on top of a gold
contact layer. As substrate, a polymer carrier film that
ensures X-ray transparency is used. Sample characterisation has been described by Stenberg et al. [30]. The
study of reaction kinetics is somewhat limited, since

acquisition times of the order of one hour are used [28].


A short introduction to XANES applied to passivity
related problems has been written by Schmuki and
Virtanen [31].
A XANES spectrum for Cr and Fe taken from an (Fe,
Cr)2O3 sample is shown in Fig. 4. The height of the
plateau at the high-energy side of the adsorption edge is
proportional to the respective amount of the element in
the sample, and the peak position reflects the oxidation
state. Quantification of XANES spectra is less straightforward than for AES and XPS. Davenport et al.
introduced a formalism to separate the oxide and metal
signals from an oxidised sample [26]. This method was
further developed by Oblonsky et al. for the quantification of film composition on Fe /Cr alloys [28], and later
Ni /Cr [32]. The measurements confirm XPS and AES
results, i.e. for a certain polarisation and bulk composition, the amount of Cr in the passive film surpasses 50%
of cations [33]. The XANES data also support the
enrichment of Cr in the passive film by selective
dissolution of iron. It has also been possible to
demonstrate the presence of Cr(VI) in the film [34].
In situ analysis of binary alloys can also be performed
by more conventional techniques. By solution analysis
with ICP-AES/MS (inductively coupled plasma atomic
emission spectrometry/mass spectrometry), it is possible
to deduce composition changes in the passive film
resulting from potential changes. This approach was
successfully employed by Castle and Qiu [35] and Hamm
et al. [36]. For an Fe /Cr alloy in the passive region, both
papers concluded that the dissolved species consisted
almost solely of iron. Similarly, rotating ring-disc
electrode experiments with Fe /Cr alloys showed that
close to no Cr was emitted from the sample during a
potential change in the passive region [37], in good
agreement with the ICP results. Annergren et al. used
the rotating ring-disc electrode in combination with
electrochemical impedance to study the dissolution and
passivation behaviour of Fe /Cr alloys in sulphuric acid

Fig. 4. XANES spectra (Fe and Cr K edges) of a mixed (Fe, Cr)2O3 sample containing 50% Cr2O3 during anodic potential steps in 0.1 M H2SO4. The
potential was increased in 100 mV steps/15 min, which is illustrated with selected raw spectra. From Schmuki et al. [29].

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

solution with additions of chloride [38 /40]. They


calculated that ferrous ions were leaving the electrode
with a current efficiency even higher than unity for
intermediate frequencies. The behaviour was explained
by postulating a concurrent chemical dissolution path of
Fe(II) [39].

2.3. Environmental influence on the film composition


A passive film is in constant exchange of species with
the electrolyte and consequently alters in thickness and
composition with the environment. Among the factors
that influence the passive film, one finds the potential,
the presence of halides in the electrolyte, the pH, and the
temperature.

2.3.1. Potential
The thickness of passive films on stainless steels grows
linearly with the applied potential. This is illustrated in
Fig. 5 for a Fe15Cr alloy in an acidic sulphate [41] as
well as Fe10Cr and Fe20Cr alloys in a sodium hydroxide
solution [42]. The basic solution gives considerably
thicker films, since dissolution is less pronounced. The
observed increase in film thickness with potential in an
acidic solution is due mainly to the oxide part; the
hydroxide part of the film is approximately independent
of potential [43,44]. Development of the passive film
with time has been investigated by, among others,
Maurice et al. [45]. Between 20 min and 20 h of
polarisation, they found only a small change in the
total film thickness; the oxide part of the film was found
to grow at the expense of the hydroxide part.

Fig. 5. Oxide layer thickness on a stainless steel as a function of


potential for a Fe15Cr alloy in 0.5 M H2SO4, and for Fe10Cr and
Fe20Cr alloys in 1 M NaOH, estimated using XPS. The film growth
region is considerably wider in the basic medium, which also gives
thicker films. After Haupt and Strehblow [41] and Hoppe et al. [42].

1097

The composition and chemistry of the film also vary


with potential. For Fe /Cr alloys, enrichment of chromium occurs in the film in the low passive region, but in
the high passive region the chromium content decreases,
since the stability of iron surpasses that of chromium.
Chromium, in turn, may oxidise to its soluble hexavalent
state. At high passive potentials, Fe has a stabilising
action on the trivalent Cr, making Fe /Cr alloys stable
over a wider potential range than pure Cr. Higher
oxidation numbers are also more dominant towards the
higher end of the passive region, as confirmed by
XANES [34].
2.3.2. Chlorides and sulphates
The presence of certain anions adsorbed on the
surface or incorporated in the passive film can be
detrimental to film stability and lead to pit initiation.
For chloride, there is a vast literature attempting to
explain its effects on pitting. Three different models are
frequently quoted: adsorption leading to local film
dissolution, penetration of anions in the film leading
to weakening of the oxide bonds, and film break down
at defects such as cracks and dislocations [46]. Conclusive analysis of the chloride content of passive films
has proven difficult, since it is rather low; frequently
quoted values range from 1 to 5%. Due to the very thin
passive film and the possible presence of structural
defects, it is not easy to distinguish adsorbed from
incorporated anionic species. Using AES and XPS, it
has been shown that sulphates are present at or close to
the film surface [47 /50]. Chlorides are enriched at the
passive film surface, but the depth distribution is more
homogeneous than for sulphates [51], especially at
elevated temperatures [43]. Mischler et al. have shown
that alloying with molybdenum reduces the surface
enrichment of chlorides and sulphates [52]. MitrovicScepanovic et al. [53] found a higher tendency for
chlorides than sulphates to penetrate the passive film
on Fe /Cr alloys. Chloride penetration into an intact
passive film can be investigated by forming the film in a
non-chloride solution, followed by exposure to chlorides. Such a series of experiments has been performed by
Hubschmid et al. for Fe /25Cr /X alloys. No incorporation was found when chlorides were injected after an
initial hour of film formation in a sulphuric acid
solution [54]. For films formed in chloride electrolytes,
on the other hand, chlorides were found distributed
through the film.
2.3.3. pH
The main effect of an increased pH is a lower
dissolution rate. This leads to a thicker passive film,
and a larger fraction of iron in the film, as iron oxides
are more stable in basic solutions. The pH effect is
illustrated by the data of Schmutz and Landolt, who
compared the response for two FeCr and FeCrMo

1098

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

alloys to a potential change in basic and acidic solutions


using the electrochemical quartz crystal microbalance
(EQCM) [55]. Increasing the potential in the passive
region lead to a mass loss in the acidic solution (0.1 M
H2SO4/0.4 M Na2SO4) and to a mass increase in the
basic solution (0.1 M NaOH). A mass increase corresponds to a net film growth and a low dissolution rate.
The effect of pH on the passive layer of Fe /Cr alloys
have also been studied by Strehblow et al. [41,42]. The
films were found to be thicker in basic solutions. In
addition, there was a marked increase of the amount of
Fe(III) oxide.
2.3.4. Temperature
The effect of temperature on the passive film was
studied by Jin and Atrens using XPS [56]. They found
no differences in composition and thickness between
passive films formed at room temperature and at 90 8C
in a 0.5 M NaCl solution. Mischler et al. quoted slightly
thicker values for films formed on Fe /Cr /Mo alloys at
65 8C when compared to room temperature [52]. The
temperature effect on film thickness for a 6Mo stainless
steel has been quantified by Wegrelius and Olefjord
using XPS [43]. They compared film formation at 22 and
65 8C in an acidic chloride solution. In spite of clear
electrochemical indications of pitting attacks during the
exposure period at the higher temperature, the film was
thicker for the higher temperature.
found to be only 2 A
2.4. Influence of microstructure and alloying elements on
passive film composition
2.4.1. Microstructure
With classic metallurgical techniques, the composition
range of the bulk material is limited by the precipitation
of intermetallic phases. In addition, inclusions and
segregation cause defects in the passive film. For
detailed studies of passive films, the amount of intermetallic phases and other imperfections can be minimised by utilising thin film materials produced with
PVD (physical vapour deposition). Since PVD is a cold
technique */a typical deposition temperature is
400 8C */it is possible to obtain nanocrystalline or
amorphous material that is virtually free from intermetallic phases over a wide composition range.
Inturi and Smialowska compared a nanocrystalline
thin film alloy of the same composition with conventional 304 material [57]. The breakdown potential of the
nanocrystalline material in a near neutral NaCl solution
was found to be shifted anodically by about 850 mV, a
behaviour that was attributed to the smaller grain size
and the high purity of the deposited thin film material.
Kraack et al. found a shift in the pitting potential to
more noble values when molybdenum was added to thin
film Fe/Cr alloys in agreement with the well-known
behaviour of bulk material [58]. A wide range of systems

produced using PVD has been investigated by Hashimoto et al. [59 /64] who underlined the extremely high
corrosion resistance of amorphous alloys.
A showcase for the importance of microsctructure is
the austeno /ferritic (duplex) stainless steels. Their twophase structure makes them particularly sensitive to the
formation of intermetallic phases. AES has been used to
investigate the homogeneity of the passive film on the
austeno /ferritic stainless steel 2205 (EN 1.4462) [65]. A
typical phase region for a commercial duplex stainless
steel covers 10/20 mm. It has been demonstrated that
the passive film composition shows a comparatively
homogeneous composition, virtually independent of the
underlying phase structure. This homogeneity continues
ngstroms down in the metal phase where a nickel
some A
and nitrogen enrichment is found for the austenitic and
the ferritic phases [65]. This enrichment would be
sufficient to produce a fully austenitic structure in a
bulk material. Significant differences between phase
regions were found by Hubschmid et al. who studied a
Fe /25Cr /11Nb model alloy with a considerably coarser
dendritic microstructure than for commercial 2205 [66].
It appears the lateral homogeneity of the passive film is
linked to the lateral extension of different phases in the
microstructure.

2.4.2. Fe
In acid solution, anodic polarisation of a Fe /Cr alloy
in the passive potential region leads to selective dissolution of iron, leaving chromium enriched in the passive
film. This behaviour is governed by the difference in
diffusion rates of iron and chromium in the passive film
[13,67]. When performing sputter depth profiles with
AES or XPS, there is normally an enrichment of iron in
the outer region of the passive film. This is possibly a
transfer artefact related to the higher diffusion rate of
iron through the film. At about 0.58 VSHE, iron changes
valence from 2 to 3 [41] for an acidic solution; for a basic
solution, the corresponding potential is about/0.3
VSHE [18].

2.4.3. Cr
For moderate anodic polarisations in acidic solutions,
the passive film consists essentially of chromium in its
trivalent state. As the potential increases above the
stability limit of chromium III, (about 0.6 /0.8 VSHE),
the passive film will start to change composition and the
fraction of trivalent iron in the film will increase. For
acidic solutions, the cation fraction of Cr in the passive
film normally amounts to 50/70%. For basic solutions,
the solubility of Cr increases, resulting in a higher
fraction of iron. The surface chemistry of pure chromium has been studied by Bjornkvist and Olefjord
[68,69]

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

2.4.4. Ni
Nickel is less readily oxidised than iron and chromium. Consequently, there is an enrichment of Ni in its
metallic state in the metal closest to the oxide/metal
interface. This enrichment could assist in the formation
of a nickel nitride [70]. Nickel could also bring down the
overall dissolution rates of Fe and Cr. Recent scanning
tunnelling microscope (STM) results indicate that the
passive film formed on austenitic stainless steels shows a
higher degree of long-range order within a shorter time
span than the corresponding ferritic alloy [45].
2.4.5. Mn
Recently, addition of Mn to stainless steel has been
used to increase the solubility of nitrogen and molybdenum, both of which have a strong beneficial influence on
the pitting resistance. Examples include the low nickel
high strength austenitic material PANACEA [4] and the
commercial superaustenitic stainless steel 654 SMO [2].
2.4.6. Mo
Molybdenum is an alloy element with a strong
beneficial influence on the pitting resistance of a
stainless steel. Pure molybdenum does not form a
three-dimensional passive film [71]. It appears that the
passive region of pure molybdenum is associated with a
diffusion controlled oxidation of hydrogen released
from the metal rather than with the formation of a
passive oxide film [72]. On the other hand, when
included as an alloying element in a stainless steel,
molybdenum is incorporated into the passive film,
showing a complex oxide chemistry with different states
of oxidation. An example of the concentration gradients
of molybdenum in a passive film on a stainless steel is
given in Fig. 2. Hexavalent Mo is found to be enriched
at the surface, whereas tetravalent states show a more
homogeneous distribution through the film [10]. This
has been shown also for the two phases of a duplex
stainless steel [73]. There are two possible hexavalent
states: MoO3, which is soluble in acidic eletrolytes and
MoO2
which shows a higher stability. Distinguishing
4
between these two states is difficult even with a high
resolution XPS spectrometer. Attempts have been made
by Lu and Clayton, [70,74 /78], who discussed the
interaction between nitrogen and molybdenum. The
passivity of stainless steels was interpreted with a
bipolar model. Another subject of discussion is the
presence of a pentavalent state, cf. Kim et al. [79].
Different molybdenum oxides have been thoroughly
characterised by Brox et al. [11,71], who found the XPS
peak of the MoO2 compound to be considerably less
shifted than expected for an ionic bond. This effect was
not seen for a tetravalent molybdenum oxy-hydroxide
and was explained by core-level screening. The literature
on the effects of Mo on the corrosion/pitting resistance

1099

has been reviewed by Jargelius-Pettersson and Pound


[80].
2.4.7. W
Tungsten is fairly recent as a major alloying element
in commercial stainless steels. It has been attributed
properties similar to those of molybdenum [81]. The
surface properties of ferritic alloys containing Mo and/
or W have been thoroughly investigated by Landolt et
al. [52,54,82 /84]. An important difference between
tungsten and molybdenum oxides lies in the different
stability of their oxides in acid solution. While hexavalent molybdenum oxide dissolves at potentials well
below oxygen evolution, the stability of hexavalent
tungsten oxide extends to anodic potentials of several
tens of volts [85].
2.4.8. N
Nitrogen is the element attributed the strongest
beneficial influence on localised corrosion in the pitting
resistance equivalent formula (PREN). Different types
of synergism with molybdenum have been proposed, as
listed in [77]. As for molybdenum, nitrogen also shows
strong concentration gradients in the passive film. It has
been suggested that ammonia or ammonium ions react
with free chlorine to form combined chlorine species
that are less effective oxidants, thereby inhibiting
chlorination enhanced localised corrosion [86]. Another
possibility is the formation of a nitride at the metal/film
interface which brings down the dissolution rates for the
individual elements in the alloy [70,87]. Nitride has also
been put forward as a possible mechanism for synergy
between nitrogen and molybdenum [70].
2.4.9. Others
Copper is added to highly corrosion resistant austenites (904L, 254SMO ) to further boost the corrosion
resistance. The influence of copper in Fe /Cr alloys has
been studied by Postrach et al. [88,89]. Copper additions
are also used to give the surface anti-bacterial properties, cf. Table 1. Another class of alloying elements with
intriguing electronic properties are the rare-earth metals
(REMs). They are used to enhanced the high temperature corrosion resistance by improving the adhesion of
oxide scales.

3. Atomic arrangement
Besides chemistry and elemental composition of the
passive films, their structure is also a vital parameter
that influences the corrosion resistance. The first investigation on the crystal structure of passive films on
Fe /Cr alloys was performed by McBee and Kruger in
1972 [90] using electron diffraction in a transmission
electron microscope. They found that the films formed

1100

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

on alloys with 24% Cr were amorphous, whereas 0/10%


Cr in the bulk resulted in a spinel-like structure. With
the invention of the STM in the early 1980s, it became
possible to study the crystal structure of the passive film
in situ with atomic resolution.
3.1. Scanning tunnelling microscopy
One of the first studies on the atomic arrangement of
active and passive stainless steel surfaces was performed
by Ryan et al. [91,92]. By studying iron in a borate
buffer solution and Fe /Cr alloys in sulphuric acid
solutions, they confirmed the conclusion of McBee
and Krueger [90] that the passive film is amorphous
for Cr concentrations above 12 /19%. Marcus et al. have
made detailed investigations on a series of single crystal
surfaces: Ni [93], Cr [94], Fe /Cr [48] and Fe /Cr /Ni
[45]. The passive films were found to grow close to
epitaxially (Cr (110)) in small (3 nm diam) crystalline
regions on iron /chromium alloys. Aging under polarisation led to growth of the crystallised regions, i.e. an
enhanced long range order. The ordering process was
observed to be slower for austenitics, which also showed
a lower production rate of Cr2O3 on the surface. It was
suggested that these effects were caused by the enrichment of nickel at the metal/film interface which limited
the access of chromium. An example of an STM image
is given in Fig. 6, which shows the atomic arrangement
of a passive film on an austenitic Fe /Cr /Ni alloy [45].
The STM can also be used to obtain depth information. This was demonstrated by Zuili et al. on Cr
surfaces [94]. By varying the voltage of the scanning
tip, it is possible to control the depth from which the

Fig. 6. Topographic STM image of a passive film on a (100) Fe /


18Cr /13Ni substrate after passivation in 0.5M H2SO4 at 500 mVSHE
for 2 h. An almost hexagonal lattice is superimposed to show the
directions of the close-packed rows. From Maurice et al. [45]

major part of the response is acquired. The structure


found was not a flat hydroxide layer, but resembled a
buried island structure, much like the one suggested by
Wegrelius and Olefjord based on XPS results [95]. STM
studies of passive films on different materials have
recently been reviewed by Marcus [96].
3.2. Geometrical considerations
Conjectures based on the geometrical arrangement of
atoms in or near the passive film have been developed as
a possible explanation for some aspects of passivity. The
simplest geometrical approach is to consider a bcclattice where every atom has eight neighbours. For
iron /chromium, this means that if 1/8 (12.5%) of the
atoms are Cr, at least 50% of the Cr atoms will have at
least one Cr as their nearest neighbour. This corresponds to the onset of a continuous network of
chromium atoms in the bulk material. The fraction 1/8
also equals the lower limit of chromimum needed to
make a steel stainless. Percolation theory is related to
the 1/8 concept. It has previously been used in materials
science to treat different phenomena in amorphous
materials, e.g. phase transformations. A general treatment of percolation has been presented by Zallen [97].
Percolation is used to describe network connectivity.
For a one-dimensional system (e.g. a string with node
points), the percolation limit is trivial, since the first
node cut will cause a break in connectivity between the
end points. For a 2D structure, it is possible to find an
exact percolation limit which will vary with the number
of connections of each node (atom or molecule) in the
network. For a 3D structure, the issue is more complex
and percolation limits become less well defined, since the
number of possible connections are much higher.
The connectivity of Fe /Cr matrices was studied for a
series of different compositions by Newman et al. For
their Monte Carlo simulations, they initially considered
a square two/dimensional lattice in which they randomly placed iron and chromium atoms [98,99]. By
assuming a high dissolution rate for iron and zero
dissolution for chromium, they derived percolation
limits of 12% for onset and 17% for completed passivity.
These percentages are in good agreement with practical
experience.
Later, the percolation model was modified to incorporate pit initiation [100]. By assuming a low dissolution
rate for Cr, it was shown that an intermittent dissolution
of iron clusters at discrete times could be obtained. If
large enough, these clusters could serve as initiation sites
for pitting. Recently, the percolation results have been
compared to experimental data for Fe /Cr PVD films
exposed to acidic chloride media [101,102]. The percolation approach has also been modified by assuming
higher dissolution rates at defect sites on the surface, e.g.
kinks, or if one alloy component is restrained to the

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

surface by thermodynamic limitations. These aspects


have been reviewed by Heusler [103]. The percolation
model can be used to investigate how different dissolution rates of iron and chromium affect the surface
roughness in the nano range [99].
Newman et al. considered the surface to be passive
when a hydroxide or oxide chain can bridge the gap and
form a link between chromium atoms at the surface
[104]. Another approach is to consider the connectivity
between cations within the oxide film. This concept has
been applied to Fe /Cr [105] and other binary Cr alloys
[106] by McCafferty, who considered a series of binary
systems following a formalism developed by Megirditchian [107] and Randic [108]. For a certain ratio of Fe/Cr
atoms in the oxide matrix, a connectivity between the
chromium atoms is obtained. Thus, for a certain
amount of iron, the continuous chromium network
through the oxide will be lost. The connectivity parameter was successfully compared with parameters such
as passivation currents, passivation potentials or experimental data on passive film compositions [105].

@d
@t

rg k1hf eBU=d

1101

(1)

where d is the film thickness, rg is the growth fraction,


khf
1 a high field model constant containing the specific
volume and valence, B is an oxide-specific constant, and
U is the applied potential. For the HFM, the thickness
of the passive film is found in the denominator of the
exponential.
The alternative to high field rate control is to consider
a rate limiting reaction at either the metal/film or the
film/electrolyte interface. A rate limiting reaction at the
metal/film interface is the base for the point defect
model (PDM) that was introduced in the early 80s. The
PDM concept has since then been developed and
discussed extensively in a series of publications by
Macdonald et al. [110,112/117]. A growth rate limiting
reaction at the film/electrolyte interface has been proposed by Vetter and Gorn [118]. It can be shown that
the form of the rate limiting reaction is independent of
which interface is considered [119]. Thus, the film
growth equation for the interface type models (IFM:s)
takes the form
@d

k1if ek2 [DUE0

Dd]

(2)

4. Film growth kinetics

@t

Passive films are constantly changing and adapting to


the environment. Thus, understanding film growth and
dissolution is essential to the understanding of passive
film stability in different and changing environments.
High reaction rates and small dimensions of the film (1 /
3 nm thickness) make it difficult to obtain unique results
for passive films on stainless steels. Some valve metals
show thicker films and are thus more easily studied.
Numerous attempts to model film growth based on
different concepts have been made; a recent review has
been written by Macdonald [109]. More details on
earlier models for film growth dynamics have been
given by Chao et al. [110].
For a passive film covering the entire surface, two
types of rate-limiting processes can be distinguished:
high field assisted ionic transport through the film or
charge transfer at either the metal/film or film/electrolyte interface. These models will be referred to as high
field (HFM) and interface models (IFM), respectively.
Most models presented in the literature give growth
equations that reduce to either the IFM or the HFM
type.
If the electric field is assumed to increase upon a
change in potential, the film growth will be limited by
high field ion conduction through the oxide (HFM). An
early formalisation of this concept was made by Verwey
[111]. The HFM is in widespread use for the modelling
of passive currents, and can be respresented by an
equation of the form

where kif1 and k2 are constants with different meanings


depending on which model is represented [119]. For the
IFM, the electric field E0 is usually assumed constant
during the applied external potential change. The film
will show a thickness change Dd until the potential
applied across the film is balanced. For the IFM, the
thickness change is found in the nominator and the
growth rate is not explicitly dependent on the absolute
film thickness.
An attempt to unify the IFM and HFM conjectures
was made by Kirchheim in 1987 [120], focussing on
galvanostatic experiments on iron in acid environments.
Kirchheim divided the current into a growth and a
corrosion part: itot /icorr/igrowth. The growth current
was assumed to be controlled by an interface equilibrium, as previously suggested by Vetter and Gorn
[118,121], and the total current was assumed to be
limited by the HFM. The corrosion part can be
estimated by solution analysis. Kirchheim showed that
neither the interface equilibrium nor the high field
limitation for the total current on its own could
successfully explain all experimental observations [120].
To understand steady-state dissolution in the passive
region, Heusler introduced different diffusion rates for
the iron and chromium cations in the film. He found an
agreement with experimental data for diffusion rates of
iron eight times higher than for those of chromium [67].
A similar approach was argued by Kirchheim et al., who
adopted different transport rates for cations in combination with the concepts developed for iron [120].

1102

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

Measurements of the passive current [122] as well as


estimates of the amount of dissolved material [123] were
both interpreted as a lower mobility for chromium in the
passive film. Further investigations also included ellipsometry [124], that was used to estimate the film growth
current igrowth. Another possibility is to use solution
analysis to measure the dissolution current icorr. This
approach was adopted successfully by Castle and Qiu
[35], who showed that the dissolved material was heavily
dominated by iron.
A combination of the above ideas was introduced in
an attempt to create a general theory for passive film
growth-the mixed conduction model (MCM), which has
been applied to pure Cr [125] as well as Fe /Cr [126 /
128], Ni /Cr [129], Fe /Mo [130], and Fe /Cr /Mo alloys
[131,132]. The MCM includes a rate limiting step at an
interface as well as transport through the film. It starts
out from a diffusion equation, which for potential
sweeps in the anodic direction essentially reduces to

the HFM. The MCM has been successful in fitting


impedance as well as passive film conductivity data.
Recently, the electrochemical quartz crystal microbalance (EQCM) has been used to investigate passive
film growth [55,133,134]. It gives in situ information
with a time resolution sufficient to provide real time
growth curves of the passive film. This has been
illustrated by Olsson et al., who used the EQCM to
study passive films on Cr [119], Fe /Cr [135], and Fe /
Cr /Ni /(Mo/W) alloys [136]. The film growth during a
potential change in the anodic direction for a Fe /Cr /
Ni /Mo stainless steel PVD alloy is illustrated in Fig. 7a
[136]. The fit parameters for the IFM and HFM were
fixed for a 304L type PVD material and it was assumed
that the addition of Mo would essentially affect the
ability of the film to withstand an increased electric field,
a hypothesis supported by the difference in thickness
change between the alloys. As can be seen, the IFM
produces a satisfactory fit to the growth curve, whereas
less correspondence is found for the HFM. This experiment was repeated for alloys with additions of Cr and
W. The electric fields in the film derived from the HFM
and IFM fits were then compared to that calculated
directly from dividing the applied potential change by
the experimental EQCM film thickness change, cf. Fig.
7b. The IFM fit values of the electric field gave a good
correlation with the direct estimates, whereas an opposite trend was found for the HFM. In addition, it was
found that alloying with Mo and W gave considerably
higher mass losses than alloying with Cr. This indicates
that their presence renders the film more permeable to
ion transport, which would allow the film to reach an
equilibrium composition faster.

5. Concluding remarks

Fig. 7. (a) Experimental EQCM growth curve for a passive film on a


Fe /Cr /Ni /Mo alloy undergoing a potential change in the anodic
direction within the passive region. This curve is compared to growth
curves simulated using Eqs. (1) and (2) for the HFM and IFM. (b)
Values of the electric field yielding a best fit to experimental data is
shown for different alloys for the IFM and HFM:s. The IFM fit values
compare better with the electric fields calculated directly from the
EQCM thickness change. From Olsson et al. [136].

There is no such thing as a static passive film. The


passive film on a given stainless steel changes with the
environment. It can grow or dissolve, and may adsorb
or incorporate anions. A parameter critical to the
stability of a passive film is the time necessary to
respond to an environmental change. With the introduction of in situ techniques, different time dependent
parameters have become accessible.
The passive film thickness changes within a couple of
seconds in response to a potential change. Structural
ordering, on the other hand, follows considerably slower
kinetics. Both these parameters are important factors in
the growth and break down of passive films.

Acknowledgements
We are indebted to Mats Liljas at Avesta Polarit
R&D, and Marie-Gilles Verge for valuable comments to

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

the script. Financial support was provided by Fonds


Nationale Suisse.

References
[1] P. Monnartz, Metallurgie 8 (1911) 161.
[2] B. Wallen, M. Liljas, P. Stenvall, in: P.J. Tunturi (Ed.),
Eurocorr, 1992, Corrosion Society of Finland, Esbo, 1992, p. 81.
[3] T. Thorvaldsson, Sci. J. Met. 26 (1997) 71.
[4] P.J. Uggowitzer, R. Magdowski, M.O. Speidel, ISIJ 36 (1996)
901.
[5] H. Fischmeister, U. Roll, Fresenius Z. Anal. Chem. 319 (1984)
639.
[6] H.H. Uhlig, Corrosion Sci. 19 (1979) 777.
[7] N. Sato, Corrosion Sci. 31 (1990) 1.
[8] J.W. Schultze, M.M. Lohrengel, Electrochim. Acta 45 (2000)
2499.
[9] I. Olefjord, B.-O. Elfstrom, Corrosion 38 (1982) 46.
[10] C.-O.A. Olsson, S.E. Hornstrom, Corrosion Sci. 36 (1994) 141.
[11] B. Brox, I. Olefjord, Surf. Interface Anal. 13 (1988) 3.
[12] C. Calinski, H.-H. Strehblow, J. Electrochem. Soc. 136 (1989)
1328.
[13] R. Kirchheim, B. Heine, S. Hofmann, H. Hofsass, Corrosion Sci.
31 (1990) 573.
[14] H.-H. Strehblow, Corrosion and Prevention 97, Brisbane, Paper
no. 5, Australasian Corrosion Association, Melbourne, 1997.
[15] C.-O.A. Olsson, Surf. Interface Anal. 21 (1994) 846.
[16] R. Goetz, D. Landolt, Electrochim. Acta 27 (1982) 1061.
[17] P. Marcus, I. Olefjord, Corrosion Sci. 28 (1988) 589.
[18] S. Haupt, C. Calinski, U. Collisi, H.-W. Hoppe, H.-D. Speckmann, H.-H. Strehblow, Surf. Interface Anal. 9 (1986) 357.
[19] J.A. Bardwell, G.I. Sproule, M.J. Graham, J. Electrochem. Soc.
140 (1993) 50.
[20] M.J. Graham, Corrosion Sci. 37 (1995) 1377.
[21] M.J. Graham, J.A. Bardwell, G.I. Sproule, D.F. Mitchell, B.R.
Macdougall, Corrosion Sci. 35 (1993) 13.
[22] C. Courty, H.J. Mathieu, D. Landolt, Electrochim. Acta 36
(1991) 1623.
[23] A.J. Davenport, J.A. Bardwell, C.M. Vitus, J. Electrochem. Soc.
142 (1995) 721.
[24] A.J. Davenport, J.A. Bardwell, C.M. Vitus, J. Electrochem. Soc.
142 (1995) 721.
[25] L.J. Oblonsky, A.J. Davenport, M.P. Ryan, H.S. Isaacs, R.C.
Newman, J. Electrochem. Soc. 144 (1997) 2398.
[26] A.J. Davenport, M. Sansone, J. Electrochem. Soc. 142 (1995)
725.
[27] S. Virtanen, P. Schmuki, M. Buchler, H.S. Isaacs, J. Electrochem. Soc. 146 (1999) 4087.
[28] L.J. Oblonsky, M.P. Ryan, H.S. Isaacs, J. Electrochem. Soc. 145
(1922) 1998.
[29] P. Schmuki, S. Virtanen, H.S. Isaacs, M.P. Ryan, A.J. Davenport, H. Bohni, T. Stenberg, J. Electrochem. Soc. 145 (1998) 791.
[30] T. Stenberg, J. Keranen, P. Vuoristo, T. Mantyla, S. Virtanen, P.
Schmuki, M. Buchler, H. Bohni, Vacuum 52 (1999) 477.
[31] P. Schmuki, S. Virtanen, Electrochem. Soc. Interface 6 (1997) 38.
[32] L.J. Oblonsky, M.P. Ryan, J. Electrochem. Soc. 148 (2001)
B405.
[33] K. Asami, K. Hashimoto, S. Shimodaira, Corrosion Sci. 18
(1978) 151.
[34] J.A. Bardwell, G.I. Sproule, B.R. Macdougall, M.J. Graham,
A.J. Davenport, H.S. Isaacs, J. Electrochem. Soc. 139 (1992)
371.
[35] J.E. Castle, J.H. Qiu, Corrosion Sci. 29 (1989) 591.
[36] D. Hamm, K. Ogle, C.-O.A. Olsson, S. Weber, D. Landolt,
Corrosion Sci. 44 (2002) 1443.

1103

[37] S. Haupt, H.-H. Strehblow, J. Electroanal. Chem. 216 (1987)


229.
[38] I. Annergren, M. Keddam, H. Takenouti, D. Thierry, Electrochim. Acta 38 (1993) 763.
[39] I. Annergren, M. Keddam, H. Takenouti, D. Thierry, Electrochim. Acta 41 (1996) 1121.
[40] I. Annergren, M. Keddam, H. Takenouti, D. Thierry, Electrochim. Acta 42 (1997) 1595.
[41] S. Haupt, H.-H. Strehblow, Corrosion Sci. 37 (1995) 43.
[42] H.-W. Hoppe, S. Haupt, H.-H. Strehblow, Surf. Interface Anal.
21 (1994) 514.
[43] L. Wegrelius, I. Olefjord, 12th International Corrosion Congress, Houston, TX, NACE, Houston TX, 1993, p. 3887.
[44] C.-O.A. Olsson, D. Hamm, D. Landolt, J. Electrochem. Soc. 147
(2000) 2563.
[45] V. Maurice, W.P. Yang, P. Marcus, J. Electrochem. Soc. 145
(1998) 909.
[46] H.-H. Strehblow, Werkst. Korr. 35 (1984) 437.
[47] P. Marcus, J.M. Grimal, Corrosion Sci. 33 (1992) 805.
[48] V. Maurice, W.P. Yang, P. Marcus, J. Electrochem. Soc. 143
(1996) 1182.
[49] I. Olefjord, B. Brox, U. Jelvestam, J. Electrochem. Soc. 132
(1985) 2854.
[50] B. Brox, I. Olefjord, Stainless Steel 1984, Gothenburg, Sweden,
The Institute of Metals, London, 1984, p. 134.
[51] I. Olefjord, L. Wegrelius, Corrosion Sci. 31 (1990) 89.
[52] S. Mischler, A. Vogel, H.J. Mathieu, D. Landolt, Corrosion Sci.
32 (1991) 925.
[53] V. Mitrovic-Scepanovic, B.R. Macdougall, M.J. Graham, Corrosion Sci. 27 (1987) 239.
[54] C. Hubschmid, D. Landolt, H.J. Mathieu, Fresenius J. Anal.
Chem. 353 (1995) 234.
[55] P. Schmutz, D. Landolt, Corrosion Sci. 41 (1999) 2143.
[56] S. Jin, A. Atrens, Appl. Phys. A 45 (1988) 83.
[57] R.B. Inturi, Z. Szklarska-Smialowska, Corrosion 48 (1992) 398.
[58] M. Kraack, H. Bohni, W. Muster, J. Patscheider, Surf. Coat.
Technol. 68/69 (1994) 541.
[59] M.-W. Tan, E. Akiyama, A. Kawashima, K. Asami, K.
Hashimoto, Corrosion Sci. 38 (1996) 1495.
[60] P.Y. Park, E. Akiyama, H. Habazaki, A. Kawashima, K. Asami,
K. Hashimoto, Corrosion Sci. 38 (1996) 1649.
[61] X.Y. Li, E. Akiyama, H. Habazaki, A. Kawashima, K. Asami,
K. Hashimoto, Corrosion Sci. 39 (1997) 1365.
[62] J. Bhattrai, E. Akiyama, H. Habazaki, A. Kawashima, K.
Asami, K. Hashimoto, Corrosion Sci. 40 (1998) 155.
[63] J. Bhattrai, E. Akiyama, H. Habazaki, A. Kawashima, K.
Asami, K. Hashimoto, Corrosion Sci. 39 (1997) 355.
[64] M.-W. Tan, E. Akiyama, H. Habazaki, A. Kawashima, K.
Asami, K. Hashimoto, Corrosion Sci. 38 (1996) 2137.
[65] C.-O.A. Olsson, S.E. Hornstrom, Duplex Stainless Steel IV,
Glasgow UK, Paper no. 68, Abington Publishing, Cambridge,
UK, 1994.
[66] C. Hubschmid, H.J. Mathieu, D. Landolt, Surf. Interface Anal.
20 (1993) 755.
[67] K.E. Heusler, Corrosion Sci. 31 (1990) 597.
[68] L. Bjornkvist, I. Olefjord, Corrosion Sci. 32 (1991) 231.
[69] L. Bjornkvist, I. Olefjord, Proc. Eurocorr 1987, Dechema,
Karlsruhe, 1987, p. 325.
[70] R.D. Willenbruch, C.R. Clayton, M. Oversluizen, D. Kim, Y.
Lu, Corrosion Sci. 31 (1990) 179.
[71] B. Brox, W. Yi-Hua, I. Olefjord, J. Electrochem. Soc. 135 (1988)
2184.
[72] F. Falkenberg, V.S. Raja, E. Ahlberg, J. Electrochem. Soc. 148
(2001) B132.
[73] C.-O.A. Olsson, Corrosion Sci. 37 (1995) 467.
[74] C.R. Clayton, Y.C. Lu, J. Electrochem. Soc. 133 (1986) 2465.

1104

C.-O.A. Olsson, D. Landolt / Electrochimica Acta 48 (2003) 1093 /1104

[75] Y.C. Lu, C.R. Clayton, A.R. Brooks, Corrosion Sci. 29 (1989)
863.
[76] Y.C. Lu, C.R. Clayton, Corrosion Sci. 29 (1989) 927.
[77] Y.C. Lu, M.B. Ives, C.R. Clayton, Corrosion Sci. 35 (1993) 89.
[78] C.R. Clayton, Y.C. Lu, Corrosion Sci. 29 (1989) 881.
[79] D. Kim, S.V. Kagwade, C.R. Clayton, Surf. Interface Anal. 26
(1998) 155.
[80] R. Jargelius-Pettersson, B.G. Pound, J. Electrochem. Soc. 145
(1998) 1462.
[81] G.P. Halada, C.R. Clayton, J. Vac. Sci. Technol. A 11 (1993)
2342.
[82] R. Goetz, D. Landolt, Electrochim. Acta 29 (1984) 667.
[83] D. Landolt, S. Mischler, A. Vogel, H.J. Mathieu, Corrosion Sci.
31 (1990) 431.
[84] H.J. Mathieu, D. Landolt, Surf. Interface Anal. 14 (1989) 744.
[85] A. DiPaola, F. DiQuarto, G. Serravalle, J. Less-Common Met.
42 (1975) 315.
[86] M.B. Ives, Y.C. Lu, J.L. Luo, Corrosion Sci. 32 (1991) 31.
[87] G.P. Halada, D. Kim, C.R. Clayton, Corrosion 52 (1996) 36.
[88] B. Postrach, I. Garz, H.-H. Strehblow, Werkst. Korr. 45 (1994)
508.
[89] B. Postrach, H.-H. Strehblow, I. Garz, Werkst. Korr. 45 (1994)
544.
[90] C.L. McBee, J. Kruger, Electrochim. Acta 17 (1972) 1337.
[91] M.P. Ryan, R.C. Newman, G.E. Thompson, J. Electrochem.
Soc. 142 (1995) L177.
[92] M.P. Ryan, R.C. Newman, G.E. Thompson, J. Electrochem.
Soc. 141 (1994) L164.
[93] D. Zuili, V. Maurice, P. Marcus, J. Electrochem. Soc. 147 (2000)
1393.
[94] D. Zuili, V. Maurice, P. Marcus, J. Phys. Chem. B 103 (1999)
7896.
[95] L. Wegrelius, F. Falkenberg, I. Olefjord, J. Electrochem. Soc.
146 (1999) 1397.
[96] P. Marcus, Electrochim. Acta 43 (1998) 109.
[97] R. Zallen, The Physics of Amorphous Solids, Wiley, New York,
1983.
[98] S. Qian, R.C. Newman, R.A. Cottis, J. Electrochem. Soc. 137
(1990) 435.
[99] S. Qian, R.C. Newman, R.A. Cottis, K. Sieradzki, Corrosion
Sci. 31 (1990) 621.
[100] D.E. Williams, R.C. Newman, Q. Song, R.G. Kelly, Nature 350
(1991) 216.
[101] M.P. Ryan, N.J. Laycock, R.C. Newman, H.S. Isaacs, J.
Electrochem. Soc. 145 (1998) 1566.
[102] M.P. Ryan, N.J. Laycock, H.S. Isaacs, R.C. Newman, J.
Electrochem. Soc. 146 (1999) 91.
[103] K.E. Heusler, Corrosion Sci. 39 (1997) 1177.
[104] R.C. Newman, F.T. Meng, K. Sieradzki, Corrosion Sci. 28
(1988) 523.
[105] E. McCafferty, Corrosion Sci. 42 (1993) 2000.
[106] E. McCafferty, Corrosion Sci. 44 (2002) 1393.

[107]
[108]
[109]
[110]
[111]
[112]
[113]

[114]
[115]
[116]
[117]
[118]
[119]
[120]
[121]
[122]
[123]
[124]
[125]
[126]
[127]
[128]
[129]
[130]
[131]
[132]
[133]
[134]
[135]
[136]

J.J. Meghirditchian, J. Am. Chem. Soc. 113 (1991) 395.


M. Randic, J. Am. Chem. Soc. 97 (1975) 6609.
D.D. Macdonald, Pure Appl. Chem. 71 (1999) 951.
C.Y. Chao, L.F. Lin, D.D. Macdonald, J. Electrochem. Soc. 128
(1981) 1187.
E.J.W. Verwey, Physica 2 (1935) 1059.
L.F. Lin, C.Y. Chao, D.D. Macdonald, J. Electrochem. Soc. 128
(1981) 1194.
D.D. Macdonald, L. Zhang, E. Sikora, J. Sikora, in: P.M.
Natishan, H.S. Isaacs, M. Janik-Czachor, V.A. Macagno, P.
Marcus, M. Seo (Eds.), Passivity and its Breakdown, Paris, The
Electrochemical Society, Pennington NJ, 1997, p. 411.
D.D. Macdonald, M. Urquidi-Macdonald, J. Electrochem. Soc.
137 (1990) 2395.
D.D. Macdonald, J. Electrochem. Soc. 139 (1992) 3434.
L. Zhang, D.D. Macdonald, E. Sikora, J. Sikora, J. Electrochem. Soc. 145 (1998) 898.
D.D. Macdonald, M. al-Rifaie, G.R. Engelhardt, J. Electrochem. Soc. 148 (2001) B343.
K.J. Vetter, F. Gorn, Electrochim. Acta 18 (1973) 321.
C.-O.A. Olsson, D. Hamm, D. Landolt, J. Electrochem. Soc. 147
(2000) 4093.
R. Kirchheim, Electrochim. Acta 32 (1987) 1619.
K.J. Vetter, Electrochim. Acta 16 (1923) 1971.
R. Kirchheim, H. Fischmeister, S. Hofmann, H. Knote, U. Stolz,
Corrosion Sci. 29 (1989) 899.
B. Heine, R. Kirchheim, Corrosion Sci. 31 (1990) 533.
J. Hafele, B. Heine, R. Kirchheim, Z. Metallkd. 83 (1992) 395.
M. Bojinov, G. Fabricius, T. Laitinen, T. Saario, G. Sundholm,
Electrochim. Acta 44 (1998) 247.
M. Bojinov, G. Fabricius, T. Laitinen, K. Makela, T. Saario, G.
Sundholm, Electrochim. Acta 45 (2000) 2029.
M. Bojinov, G. Fabricius, T. Laitinen, K. Makela, T. Saario, G.
Sundholm, J. Electrochem. Soc. 146 (1999) 3238.
M. Bojinov, I. Betova, G. Fabricius, T. Laitinen, R. Raicheff, T.
Saario, Corrosion Sci. 41 (1999) 1557.
M. Bojinov, G. Fabricius, P. Kinnunen, T. Laitinen, K. Makela,
T. Saario, G. Sundholm, Electrochim. Acta 45 (2000) 2791.
M. Bojinov, I. Betova, R. Raicheff, Electrochim. Acta 44 (1998)
721.
M. Bojinov, G. Fabricius, T. Laitinen, T. Saario, Electrochim.
Acta 44 (1999) 4331.
M. Bojinov, G. Fabricius, T. Laitinen, K. Makela, T. Saario, G.
Sundholm, Electrochim. Acta 46 (2001) 1339.
P. Schmutz, D. Landolt, Electrochim. Acta 45 (1999) 899.
C. Gabrielli, M. Keddam, F. Minouflet, H. Perrot, Electrochim.
Acta 41 (1996) 1217.
D. Hamm, C.-O.A. Olsson, D. Landolt, Corrosion Sci. 44 (2002)
1009.
C.-O.A. Olsson, D. Landolt, J. Electrochem. Soc. 148 (2001)
B438.

Vous aimerez peut-être aussi