Sei sulla pagina 1di 17

CorrosionScience, Vol. 39, No. l&l I, pp. 1897-1913.

1997
0 1997 Elsevier Science Ltd
Printed in Great Britain. All rights reserved
001&938X/97 $17.00+0.00

PII: soolo-938x(97)ooo84-x

A SURFACE ANALYTICAL AND ELECTROCHEMICAL


STUDY
ON THE ROLE OF CERIUM IN THE CHEMICAL SURFACE
TREATMENT OF STAINLESS STEELS
S. VIRTANEN 3* M. B. IVES,*

G. I. SPROULE,+

P. SCHMUKI+

and

M. J. GRAHAM?
W. Smeltzer Corrosion Laboratory, McMaster University, Hamilton, Ontario, Canada
+Institute for Microstructural Sciences, National Research Council, Ottawa, Ontario, Canada

*Walter

Abstract-The mechanism of oxide layer formation and modification during chemical cerium nitrate treatment of
stainless steel has been investigated. The aim of the work was to study the role of cerium in modifying the oxide layer
properties, especially the kinetics of the cathodic reactions. For this, electrochemical and surface analytical studies
were carried out. During exposure to hot (90C) cerium nitrate solution, oxide film formation by chromium
passivation and an accompanying dissolution of iron oxide takes place, leading to an enrichment of chromium in
the oxide layer. Further, insoluble cerium species are precipitated at the cathodic sites of the surface. The oxygen
reduction reaction is inhibited on these films. The effect of the cerium treatment cannot be solely attributed to the
formation of a chromium-rich oxide layer, since the cathodic reactions are more strongly inhibited on the ceriumtreated stainless steel than on passivated pure chromium. Moreover, the cerium treatment is efficient in retarding
the cathodic kinetics on pure chromium. Studies with a redox couple present in the electrolyte clearly show that the
inhibition of the oxygen reduction reaction is not due to a lower electron conductivity of the oxide layer. The
cathodic inhibition effect can be attributed to a high resistance against reductive dissolution. This is partially due to
the chromium enrichment and in addition to the cerium precipitation at the weak sites of the oxide layer which
otherwise under cathodic polarization would lead to reductive dissolution, thus providing current paths for
electrons participating in the oxygen reduction reaction. Treatment parameters such as time, alloy composition,
solution chemistry and potential during treatment were studied. Clearly, all factors leading to a maximum
chromium enrichment and/or cerium precipitation increase the cathodic inhibition efficiency. 0 1997 Elsevier
Science Ltd
Keywords: A. stainless steel, B. galvanostatic, B. SIMS, B. XPS, C. passive film.

INTRODUCTION
treatment
in solutions containing cerium compounds has been widely studied for
the prevention of localized corrosion of aluminum and its alloys.4 In these studies, the
positive effect of cerium has been attributed to inhibition of cathodic reactions. Similarly,
cerium has been found to help in preventing crevice corrosion of stainless steels, either by
cerium implantations or a simple immersion in boiling cerium nitrate solutions.6 Also in the
case of stainless steels, the effect of cerium has been attributed to retardation of cathodic
reactions.5-7 However, in another study, cerium treatment was found to be far less efficient
in improving the corrosion behavior of stainless steel. The precise inhibition mechanism of
the oxygen reduction reaction has yet to be clarified.
Chemical

On leave from the Swiss Federal Institute of Technology, Institute of Materials Chemistry and Corrosion,
ETH-H%ggerberg, 8093 Ziirich, Switzerland.
Manuscript received 20 November 1996; in amended form 21 January 1997.
1897

1898

S. Vlrtanen cl al.

It has been shown by previous surface analytical studies that cerium treatment leads to a
chromium enrichment
in the oxide films. Therefore the question arises as to what is the
specific role of cerium species in the surface modification process. The aim of this work was
to study the processes such as layer growth and changes in oxide composition
taking place
during the cerium treatment
as well as the role of cerium in modifying the oxide layer
properties;
especially the kinetics of cathodic reactions.
For this, the electrochemical
behavior of cerium-treated
stainless steel was studied during cathodic polarization
in
aerated and deaerated borate buffer solution. To obtain information
on the electron transfer
reactions through the oxide films, studies with a redox couple [Fe(CN$/Fe(CN$]
present
in the solution were carried out. Surface analytical data regarding the composition
of the
oxide layers were obtained to try to understand
the special role of cerium in the oxide layer
formation and modification.
Further, the influence of different treatment parameters on the
treatment efficiency was studied.

EXPERIMENTAL

METHOD

The sample material was a high-purity


stainless steel AISI 304 (S < 0.001%) in sheet
form. Prior to treatment
the surfaces were mechanically
ground to a 600 grit finish for
electrochemical
measurements
or to a 0.25 pm diamond finish for surface analytical studies.
After polishing, the samples were rinsed in acetone and ethanol and dried in nitrogen gas.
The treatment solution was prepared from Ce(N0&.6Hz0
or NaN03. When desired, the
pH of the solution was adjusted by additions of diluted HN03 or NaOH. The natural pH of
the nitrate solutions was slightly acidic (pH 4.8 for 0.05 M cerium nitrate). The cerium
chemical treatments
were carried out at 90C in a water bath. After the treatment
the
samples were rinsed in distilled water and dried with nitrogen gas.
The electrochemical
experiments
were carried out with an EG&G Part model 273
potentiostat.
The scan rate for the potentiodynamic
polarization
curves was 0.2 mV s-.
Galvanostatic
reduction
was carried out with a current density of - 5 PA cmp2. The
electrolyte
solution
was borate buffer, pH 8.4 (8.17 g 1-l Na2B407.10H20+
7.07 g l-
H,B03) to which 0.05 M K3Fe(CN)6+ 0.05 M K4Fe(CN)6 was added for the electron
transfer measurements.
The solutions were prepared from reagent grade chemicals and
distilled water.
X-ray photoelectron
spectroscopy (XPS) measurements
were carried out in a PerkinElmer PHI 5500 system with a monochromated
Al K, source. Cr 2p, Fe 2p, Ni 2p, Ce 2p and
0 1s core levels were collected using a pass energy of 29.4 eV and, if not otherwise stated,
with a take-off angle of 75. The background
was subtracted in the integrated mode. The
spectra were deconvoluted
in order to separate the contribution
of metallic and oxidized
species in a similar manner
as described earlier by other authors.,
The relative
concentration
c(A) for an element A in the oxidized layer composed of i constituents
is
defined by c(A) = [~(OX)~/S,]/[CZ(OX)i/~~],
where 1(0x) and S represent the photoelectron
intensities and the relative sensitivities for oxidized species of an element (for Z(Ox) the sum
of the intensities of the different oxidized species was taken). The relative sensitivities were
taken from literature
data.12 Thickness
values were estimated
by using a treatment
described elsewhere. I3
For secondary-ion
mass spectrometry
(SIMS) a Perkin-Elmer
PHI 590 scanning Auger
microprobe
with a SIMS II attachment
was used. The species measured were mass 68
(52Cr0i),
72 (56Fe0+), 74 (NiO+), and 156 (CeO).
Mass 168 (Fe;) was recorded to

Role of cerium in treatment of stainless steels

1899

of the oxide/substrate
interface.
indicate
the position
Further,
masses 69
(52CrOH+ + 53Cr0+) and 73 (56FeOH+ + 57Fe0+) were measured in order to check for
the presence of hydroxide in the oxide films. In all oxide films some hydroxide was found in
the outermost part of the film. Relative sensitivity factors for the different signals were
determined by calibration with the XPS data. Sputtering yas performed with 1 keV Xe
corresponding
to a sputter rate of approximately 4 A min-.4 The sputter time
corresponding to the interface location was determined at 50% of the mass 168 (Fez)
signal. The oxide layer thicknesses determined from SIMS depth profiles were in very good
agreement with the values of the layer thickness calculated from XPS data. A more detailed
description of the experimental procedure is provided in 15.
EXPERIMENTAL

RESULTS

Inhibition of the cathodic reactions


The first set of experiments was aimed to clarify what different reactions take place on
cerium-treated
stainless steel samples during cathodic polarization; specifically to
distinguish between oxide reduction and oxygen reduction reactions. This is important in
understanding the cathodic inhibition mechanism, since oxygen reduction takes place in the
potential range where reduction of the passive film is also possible. For this, cathodic
polarization experiments were carried out on as-treated samples in deaerated and aerated
solutions. Figure l(a) shows a comparison for samples after treatment for 1 h in 0.05 M
cerium nitrate solution at 90C. It is evident that in the potential region between the opencircuit potential and z - 700 mV (SCE), a cathodic peak is present in both solutions. This
potential region corresponds to reduction of passive films on iron and therefore the cathodic
peak can be ascribed to reduction of ferric species in the film. An indication of reduction of
the oxide film on stainless steel during cathodic polarization was also found in an earlier
rotating disc study by Lu and Ives6 showing the existence of a mass-transport-independent
contribution to the total cathodic current.
The polarization curves thus suggest that the cathodic current for as-treated samples
stems from two reactions: oxygen reduction and reduction of Fe3+ species present in the
oxide film. Hence, cathodic polarization experiments for as-treated samples do not give
direct information on the kinetics of the oxygen reduction reaction, since it is difficult to
separate the contributions of the oxide and oxygen reduction. This is a particular
disadvantage as the amount of ferric species in the oxide layer can vary depending on the
process of oxide formation. On the other hand, if the cathodic polarization measurements
are carried out on cerium-treated samples after total reduction of the ferric species present in
the film, then the cathodic current stems only from reduction of redox species in the
electrolyte, e.g. from the oxygen reduction reaction. Figure l(b) shows a comparison of the
cathodic polarization curve of a cerium-treated sample [ 1 h, 0.05 M Ce(N03)s, 90C] prior
to and after galvanostatic reduction treatment (2 h, - 5 uA cm-*) compared with the
corresponding polarization curves of an untreated sample. It is clear that after reduction,
the current peak corresponding to reduction of ferric ions is absent. Nevertheless, before
reduction of the samples it is clearly evident that the rate of oxygen reduction is retarded on
the cerium-treated sample in comparison with the untreated sample. The same effect can be
seen if the potential decay curves during galvanostatic reduction experiments in aerated
borate buffer of the untreated and treated samples are compared (Fig. 2). Clearly, the lowerend potential of the cerium-treated sample indicates an inhibition of cathodic reactions.

S. Virtanen

1900

10-s
-1400-1200-1000

-LL-800

YI al

-600

-400

-200

200

200

Potential (mV SCE)

-1400-1200-1000

-800

-600

-400

-200

Potential (mV SCE)


Fig. I. Potentiodynamic
cathodic polarization
curves in borate buffer, pH 8.4. (a) AISI 304
samples treated for 1 h in 0.05 M Ce(NO& in deaerated and aerated solutions. (b) Cerium-treated
[I h in 0.05 M Ce(NO&]
and untreated AISI 304 samples measured directly from open-circuit
potential or after galvanostatic
reduction treatment (2 h. -5 PA cme2).

The influence qf treatment parameters


In order to understand
the role of cerium in the layer formation process, the influence of
various parameters
such as time, solution chemistry, alloy composition,
and potential
during treatment on the properties of the oxide layer was investigated.
Treatment
time. Figure 3 shows the potential
decay curves during galvanostatic
reduction in aerated borate buffer for samples treated for different times. A remarkably
lower end potential is found for samples treated for longer times. In agreement with this, the
cathodic current density in potentiodynamic
experiments in aerated borate buffer solution
decreased with increasing treatment time. In Fig. 4, SIMS profiles for samples treated for
various times in 0.05 M cerium nitrate/90C are compared with the untreated sample and a
sample treated in pure water at 90C. The amount of oxidized nickel was found to be

Role of cerium in treatment

-800

-2000

2000

of stainless steels

4000

6000

1901

8000

Time (s)
Fig. 2.

Potential decay curves during galvanostatic


reduction (- 5 uA cm-*) for untreated
cerium-treated
[l h in 0.05 M Ce(NO,)s] AISI 304 samples in borate buffer.

-500

>
5

-600

and

xj -700
\
,\

s
j$ -800

_._.-._.
_.X

,.*
.\.__-.

-900

-2000

2000

4000

6000

8000

Time (s)
Fig.

3.

Potential
decay curves during
galvanostatic
reduction
in aerated
borate
(-5 uA cm-*) for AISI 304 samples treated in 0.05 M Ce(NO&) for various times.

buffer

negligible in the oxide layers and is therefore not shown in the graphs. Treatment in hot
water [Fig. 4(b)] leads to only minor changes in the oxide composition compared with the
air-formed film on the untreated sample [Fig. 4(a)]. During exposure to cerium nitrate,
however, very significant modification of the oxide film takes place. It is clearly evident in
Figs 4(c), 4(d), 4(e) and 4(f) that cerium treatment leads to a gradual chromium enrichment
and iron depletion in the oxide layer. Both SIMS and XPS data clearly show that a longer
exposure does not lead to a thickening of the film, the thickness remaining being x40 A.
After all treatments in cerium nitrate, cerium was found to be present in the oxide film, but
the amount of cerium does not change as a function of treatment time. The SIMS profiles
indicate that cerium is not present only as a surface contamination but is incorporated in the
(Fe,Cr) oxide film.

S. Virtanen

I902

et al

Sputter time (min)

Sputter time (min)

-1
Sputler time (min)

Sputter time (min)

Sputter time (min)


Fig. 4.

SIMS

profiles

of AISI

(b) 30 min in Hz0,90C:

Sputter time (min)

304 stainless steel after various

(c f, 0.05 M Ce(N0&/90C:

treatments.

(a) Polishing

(untreated);

(c) I5 min; (d) 30 min; (e) 1 h; (f) 3 h.

The results of the SIMS analysis were confirmed by XPS studies showing a continuous
increase of chromium content as a function of time (Fig. 5). The cerium content, on the
other hand, is more or less constant with treatment time. In all cases the cerium content is
relatively small (max. 5-6 at%). In agreement with the SIMS data, the amount of oxidized
nickel in the oxide layer determined by XPS is very small and was thus not included.
Alloy composition.

the cerium

treatment,

Because of the finding that chromium enrichment takes place during


it must be considered whether the main effect of the treatment is to

Role of cerium in treatment of stainless steels

1903

at-% Fe
at-% Ce
0.05 M Ce(No,), 190C

80
z
$

60

5
40
20
0

Fig. 5.

0 min

15 min

30 min

Ih

3h

Composition of the oxide layers determined by XPS.

increase the chromium content of the oxide film. Since chromium oxides are very stable and
will not be reductively dissolved under the conditions used in this work, the presence of a
chromium oxide layer could provide a more efficient barrier to the oxygen reduction
reaction than a passive film on stainless steel. Further, it has been shown earlier that prepassivation of stainless steel in acidic solutions can lead to a strong improvement in the
resistance to localized breakdown and this was interpreted to be due to a high chromium
content of the passive film formed in acidic solutions.6
A comparison of the cathodic reduction characteristics of a cerium-treated stainless steel
with those of an untreated pure chromium sample is shown in Fig. 6(a). The slow potential
decay in the case of the stainless steel is due to reduction of ferric species in the passive film.
After this reduction wave, however, a higher overpotential for the cathodic current is needed
for the cerium-treated stainless steel than for passivated chromium. This indicates that
cathodic reactions (oxygen reduction and/or hydrogen evolution) are less inhibited on the
passive film on chromium than on the oxide layer of cerium-treated stainless steel [Fig. 6(a)].
This finding was confirmed in potentiodynamic experiments. In addition, cerium treatment
of pure chromium leads to a retardation of cathodic reaction kinetics on the chromium
passive film [Fig. 6(b)]. Clearly, cerium plays a specific role in the inhibition mechanism of
chromium-rich oxide films.
Another question concerns the effectiveness of the treatment on Fe-Cr alloys of varying
chromium content. For this, Fe-Cr alloys with a chromium content from 5 to 30 at% were
treated for 1 h in 0.05 M cerium nitrate and in 0.15 M NaN03, and the samples were
subsequently galvanostatically reduced in aerated borate buffer. Figure 7 shows a summary
of the results. Presented is the end potential of galvanostatic reduction (after 5000 s) as a
function of the chromium content for untreated, NaNOs-treated and Ce(NO&-treated
samples. This end potential qualitatively represents the inhibition efficiency of cathodic
reactions. Clearly, for low chromium contents in the alloy, neither nitrate treatment is
particularly efficient in inhibiting cathodic reactions. On the other hand, for higher
chromium contents, the cerium nitrate treatment leads to a more significant retardation of
the cathodic reactions than sodium nitrate. SIMS profiles of Fe-l 5Cr and Fe-30Cr samples
treated for 1 h at 90C in 0.05 M Ce(NOs)s and in 0.15 M NaNOs are shown in Fig. 8. As
expected, in both solutions, the oxide layer on Fe-30Cr [Figs 8(b) and 8(d)] contains
significantly more chromium oxide than the layer on Fe-15Cr [Figs 8(a) and 8(c)]. The

S. Virtdnen et ul.

1904

-AISI 304
\~--j

Ce-treated

----...-- Cr (non-treated)

2000

4000

Time (s)

-700
-750 t

-800
-850
-900 t
-2000

2000

4000

6000

8000

Time (s)
Fig. 6. Galvanostatic
in 0.05 M Ce(NO&]

reduction (- 5 PA cm ) in aerated borate buffer for: (a) cerium-treated


AISI 304 and unWedted passive film on pure chromium; (b) untreated
cerium-treated
[30 min in 0.1 M Ce(NO&] chromium.

[3 h
and

chromium content in the oxide layers is very similar for samples treated either in sodium or
cerium nitrate. The cerium content does not seem to depend on the alloy composition.
These
findings indicate that a high chromium
content in the oxide film is necessary for the
achievement of the cerium effect on the kinetics of the cathodic reactions.
Solution chemistry. Since the above findings show that chromium
enrichment
is an
essential factor to achieve a good surface treatment, it is possible to try to maximize the
chromium
enrichment,
for instance by decreasing
the pH of the treatment
solution.
Therefore, a comparison
of the galvanostatic
reduction behavior was carried out on AISI
304 and chromium samples treated for 1 h in 0.1 M Ce(NO& or in 0.3 M NaNOs at various
pH values. The results are summarized
in Fig. 9 (end potential
of the galvanostatic
reduction as a function of solution pH). In the case of stainless steel [Fig. 9(a)], in nearneutral solutions
treatment
in cerium nitrate leads to a retardation
of the kinetics of

Role of cerium in treatment

-400

::
2

-500
-600

1
1

A
W

1905

of stainless steels

w
3
z
2
g

-700

-800

-900

-1000

-~~~~~.~~~~~(~((~~~
5
IO
15
20
25

30

35

% Cr
Fig. 7. The end potential (after 5000 s) of the galvanostatic
reduction (-5 uA cm-*) in aerated
borate buffer for Fe-Cr alloys with a varying chromium content for untreated samples as well as for
samples treated for 1 h in 0.05 M Ce(NOs)s/90C
or in 0.15 M NaNOs/90C.

0.8

J5
0
5
.Y
E
3
m

0.6

0.8

0.4
0.2

10

15

20

10

15

20

15

20

Sputter time (min)

Sputter time (min)

0.8

IO
Sputter time (min)

15

20

10
Sputter lime (min)

Fig. 8. SIMS profiles of: (a) Fe-1SCr treated in O.iS M NaNOs; (b) Fe-30Cr treated in 0.15 M
NaNO,; (c) Fe-1SCr treated in 0.05 M Ce(NO&; (d) Fe-30Cr treated in 0.05 M Ce(NO&.

S. Virtanen et al.

1906
-400

-5oo-
c

_600_1

::
>

-700-r

non-treated
------------_
- -

,.D Na

(a)

j.

:
.I.
.1
.:::/-

.E
,,, -800-r
_9oo_i

ii::.
,

-1000.1

,&

,I.

,,.

.. ..._.._......._. Ce
/

PH

-840

:: -860
>
J_ -880 i 1

w -900.~
b Ce
2

PH
Fig.

9.

Effect

of treatment

solution

pH

on

the end

potential

of galvanostatlc

reduction
or in

( - 5 PA cm -*) in aerated borate buffer for samples treated for 1 h at 90C in 0.1 M Ce(NO&
0.3 M NaNO?: (a) AISl 304 stainless steel; (b) chromium.

reactions,
whereas
the sample treated in NaN03 shows a behavior identical to that
of the untreated
sample. In both nitrate solutions, the inhibition
effect is stronger after
treatments in lower pH solutions. At pH 2, treatment in both nitrate solutions leads to an
identical galvanostatic
reduction behavior. For pure chromium [Fig. 9(b)], a different pH
dependence is found. In this case the cathodic inhibition effect increases with increasing pH,
but again the largest difference between the cerium and sodium nitrate can be found at pH 6.
AISI 304 samples treated at pH 2 and pH 6 were further analysed by SIMS and the
profiles are shown in Fig. 10. Clearly, chromium enrichment is significantly stronger after
treatment in the solutions of low pH [Figs 10(a) and 10(b)]. On the other hand, very little
cerium is found on the surface of the sample treated at pH 2 in cerium nitrate [Fig. 10(b)],
and the composition
of the oxide layer on this sample is thus almost identical to that of the
sample treated in NaN03 at the same pH [Fig. IO(a)]. Therefore it is not surprising that the
samples showed an almost identical
electrochemical
behavior
during galvanostatic
reduction.
Samples treated in solutions
of pH 6 [Figs 10(c) and 10(d)] show a lower
chromium content, as can be expected. In this case, treatment in cerium nitrate leads to
significant amounts of incorporated
cerium [Fig. 10(d)].
cathodic

Treatment under polarization.


To study cerium incorporation
on the stainless steel
surface, the treatment was carried out under cathodic or anodic polarization
and the current

Role of

ceriumin treatmentof stainlesssteels

1907

0.8
0.8
0.4
0.2
0
5

10

15

20

10

10

15

20

15

20

Sputter time (min)

Sputter time (min)

15

Sputter time (min)

20

10
Sputter time (min)

Fig. IO. SIMS profiles of AIS1304 stainless steel after various treatments for 1h at 90C: (a) 0.3 M
NaN03, pH 2; (b) 0.1 M Ce(NO&, pH 2; (c) 0.3 M NaNOJ, pH 6; (d) 0. I M Ce(NO&, pH 6.

was monitored. Figure 11 shows the current density as a function of time for both
cathodically (a) and anodically (b) polarized samples. If cerium nitrate is added to the
solution during exposure to NaNOs under cathodic polarization, the current slowly
decreases indicating a gradual blocking of the cathodically active surface. Smaller currents
are also observed, if the sample is initially exposed to cerium nitrate. If the corresponding
experiment is carried out under anodic polarization, addition of cerium even increases the
passive current density. Since no decrease of the current densities can be observed even in
higher concentrated cerium nitrate solutions, it can be concluded that under anodic
polarization in cerium-containing solutions no blocking of the anodically active surface
takes place.
Electronic properties of the oxide luyers

Since cerium nitrate treatment clearly leads to a retardation of the oxygen reduction
reaction, it is interesting to investigate the electronic properties of the oxide layers. To study
electron transfer reactions, polarization curves were measured with a Fe(CN)i-/Fe(CN)iredox system present in the borate buffer solution. This redox system is well known and
widely used to study electron transfer reactions.* Fig. 12(a) shows a comparison of the
polarization curves of an untreated sample with a sample treated in hot water and in hot
cerium nitrate. Clearly, electron transfer reactions are accelerated after both surface

S. Virtanen et al

1908

(a):

E =

i ,.

-500 mV SCE

90%

1J

w/
+if-

.A=,-

addhon of Ce(NO,),

.-0.1

M NaNO,

0.05 M Ce(NO,),

-500

500

1000

1500

Time

25

2000

2500

3000

90C

(s)

,,,,,,,,,,,.,,/,,,.,~,,/
E = +400

mV SCE

--0.1M

NaNO,

0.05 M Ce(NO,),

addition of Ce(NO,),

ot,
0

I.1

I*lL,..*

1000

500

Time
Fig.

I 1.

Current

1500

2000

2500

(s)

density as a function of time during treatment under polarization


stainless steel: (a) E=

~ 500

mV (SCE);
(b)

E=

for

AISI304

400mV (WE).

treatments
and even more strongly by the cerium nitrate treatment.
The finding of an
increased electron conductivity
after film formation in high-temperature
water is in good
agreement with earlier work by other authors.9720 Further, if the redox system studies are
carried out on samples which have been treated for various times in cerium nitrate, an
identical behavior is found for all treatment
times [Fig. 12(b)]. This indicates that the
retardation
of theoxygen reduction reaction after longer treatment times cannot be due to a
hindered electron transfer through the oxide film. A comparison
of the electron transfer
kinetics on the untreated and treated passive film of chromium shows only a very slight
change in the kinetics after the cerium treatment.
Further, electron transfer is faster on
passive chromium than on untreated stainless steel surfaces. This can be attributed to the
thinner oxide film in the case of chromium,
which increases the probability
of electron
transfer via a tunneling mechanism through the oxide.18

DISCUSSION
Oxide layerjbrmation
The role of cerium on the oxide layer formation can be best understood
if the surface
analytical data are considered in more detail. Clearly, during exposure to hot cerium nitrate

Role of cerium in treatment of stainlesssteels


~
- -

-400

10-2 4
-400

non-treated
H,O - treated

...-

1909

, a.

Ce-treated

-200
0
Potential

-200

200

Potential

400

600

400

600

(mV SCE)

200

(mV SCE)

Fig. 12. Polarization curves in deaerated borate buffer containing the redox couple
Fe(CN)i-/Fe(CN)ifor AISI 304 stainless steel: (a) untreated sample, samples treated for 1 h at
90C in Hz0 or in 0.05 M Ce(NO&; (b) samples treated at 90C in 0.05 M Ce(NO& for various

times.
solution, a gradual enrichment of chromium in the oxide film takes place. The SIMS profiles
show a similar depth distribution of iron and chromium in cerium-treated samples (Fig. 4),
whereas in the air-formed oxide as well as in oxide films formed in hot water, or in NaN03,
chromium is present in the inner part of the layer and a much higher iron content is found in
the outer part. If the amount of hydroxide species is considered in the SIMS profiles, then a
higher Cr(oxide + hydroxide)/Fe(oxide + hydroxide) ratio is found in the outer part of the
cerium-treated samples. This is confirmed by angle-resolved XPS measurements, which
indicate for the cerium-treated samples a higher amount of oxidized Cr(hydroxide + oxide)
in spectra measured at a low angle (157 corresponding to a higher surface sensitivity. These
findings suggest that cerium treatment leads to a gradual dissolution of the iron oxide out of
the air-formed film. Simultaneously, oxide growth by chromium passivation takes place.
The overall layer thickness is only slightly changed. The dissolution of iron is most probably
due to the slightly acidic pH of the cerium nitrate solution (pH 4.8). By decreasing the pH of
the solution, chromium enrichment is stronger due to acceleration of iron dissolution.
Figure 13 shows the average composition of the oxide layer as a function of the
chromium content of the alloy (a) or solution pH for AISI 304 (b) for sodium and cerium

S. Virtanen

1910

120

er al.

(a)

100

80
;.

s 60
.L
m
40

20

Fe- ISCr
NaNO:

Fe- 150
Ce(NO?)?

Fe-30Cr
NaNO 1

Fe-3OCr
Ce(NO-c)j

Fe
120

(b)

100

RO
;:
s

60

2
10

20

pH 2
NaNO;

PH 2
Ce(NO 311

PH 6
NaNOl

PH 6
WNO3)3

Fig. 13. Average composition


of the oxide layer determined from the SIMS data: (a) as a function
of the chromium content of Fe-Cr alloys; (b) as a function of solution pH for AISI 304 stainless steel.

nitrate treatments.
It is clear from the figure that for a fixed pH or fixed alloy composition,
the Cr/(Fe + Ce) ratio is constant for both sodium and cerium nitrate. In the case of samples
treated in cerium nitrate, cerium seems to replace part of the iron oxide in the oxide layer.
The SIMS data further indicate that the oxide layers formed in cerium nitrate are in all cases
thinner than the corresponding oxide layers formed in sodium nitrate. This suggests that
cerium in the oxide film makes the film more protective, thus hindering further film growth.
In this way, the presence of cerium in the solution leads to a higher Cr/Fe ratio in the oxide
film, which is generally beneficial for the stability of passive films on Fe-Cr alloys.

The influence of cerium on the kinetics of cathodic reactions


It is evident from the cathodic polarization measurements that cerium treatment leads to
a retardation of the oxygen reduction kinetics. The electrochemical studies in the presence of
the redox system clearly indicate that this is not due to an increased electron resistivity of the
oxide layer. Since oxygen reduction takes place in the potential range where reduction of the
oxide film can take place as well, it is therefore suggested that the inhibition of the oxygen

Role of cerium in treatment of stainless steels

191I

reduction reaction is associated with the behavior of the oxide layer under cathodic
polarization. It has been shown elsewhere that reduction of thin Fez03 films in borate buffer
leads to a complete dissolution of reduced Fe2+ ,2 whereas in the case of mixed (Fe,Cr)zOs
part or all of the Fe 2+ is trapped in the oxide without dissolution22 depending on the Fe/Cr
ratio. Furthermore, chromium oxide generally cannot be reduced under moderate cathodic
polarization in borate buffer.22 Therefore it can be concluded that the oxide films which are
highly enriched with chromium are more resistant to reductive dissolution than the passive
film on untreated stainless steel.
Accordingly, chromium enrichment itself leads to a higher resistance to reductive
dissolution. Since it was shown that cerium treatment of pure chromium is efficient in
retarding the oxygen reduction kinetics, the effect of cerium is not solely based on chromium
enrichment of the oxide. Even though chromium oxides are very stable it is possible that
weak sites of the oxide are dissolved under cathodic polarization and these sites would then
become low-resistivity paths for the current, leading to an increase of the oxygen reduction
kinetics. It has been discussed earlier that the passive film on chromium contains local sites
of a higher electron conductivity due to variations in the film thickness and defectiveness.23
It is then possible that cerium is blocking such weak sites, which otherwise would become
reductively dissolved.
In the case of an untreated passive film the cathodic polarization leads to reductive
dissolution of the oxide film (overall dissolution of iron oxide and local dissolution of
defective sites of chromium oxide) and, as a consequence, the oxygen reduction is less
hindered
on a bare steel surface than on the treated steel surface covered by chromium-rich
oxide and cerium species. It should be pointed out that the electron transfer kinetics on a
bare metal surface are very fast compared with an oxide-covered surface and therefore the
increase of the electron conductivity of the oxide layer by cerium treatment is of minor
significance compared with the enhancement of electron transfer kinetics due to dissolution
of the untreated oxide layer.
The mechanism of cerium incorporation in the oxide layer

The mechanism of cerium incorporation


in the oxide layer is clearly cathodic
precipitation, as indicated by the studies carried out under polarization. The cathodic
potential used in this study [E= - 500 mV (SCE)] 1s
. far above the equilibrium potential of
the Ce(III)+Ce(O)
reaction [Eo= -2.48 V (NHE)], thus the species responsible for the
blocking of the cathodically active sites must be Ce(II1) species. XPS spectra and a
comparison with standards indicate that cerium is indeed present as Ce(III) on the surface
(see Table 1 for binding energy positions for a cerium-treated stainless steel sample and for
standards). According to Pourbaix,24 an increase of pH, which can be expected to take place
at cathodic sites of the surface, will then lead to a precipitation of Ce(OH)3. During cathodic
polarization, the surface pH of the solution will increase and thus insoluble Ce(II1) species
are precipitated. During open-circuit treatments, cerium precipitation will take place at
cathodic sites of the surface. Therefore cerium will be incorporated in the oxide films exactly
at those sites which otherwise would lead to current paths during cathodic polarization. The
finding that the amount of cerium is always relatively low in the oxide film is in good
agreement with the concept of an inhomogeneous cerium distribution in the film.
Furthermore,
since cerium is found distributed throughout the film and its depth
distribution does not show any time dependence, it is more likely that cerium is
incorporated in the film by local destruction of the oxide film followed by a precipitation

S. Virtanen

1912

el al.

Table 1. Binding energies (eV) of Ce 3d spectra


for AISI 304 treated in Ce(NO& and for Ce(II1)
and Ce(IV) standards
AISI 304

3d3.2
3dm

905.13
886.88

Ce(NOh
905.25
886.75

&Ce(SO&
906.4
887.75

of cerium species from the solution than by a diffusion of cerium into an existing film, since
the latter would be expected to lead to a higher surface concentration of cerium.
The results on Fe-Cr alloys with a varying chromium content clearly indicate that
cerium is efficient only if the chromium content in the alloy and subsequently in the oxide
films is sufficiently high. Thus a low chromium-containing film contains too many sites,
which are prone to reductive dissolution, to be blocked with cerium.
Since cerium incorporation takes place by pH-induced precipitation, the influence of
solution pH on the effectiveness of the treatment can be well understood. In the case of
stainless steel, a low pH leads to a high chromium enrichment but, owing to the high
solubility of Ce(OH)s in acidic solutions,24 only a small amount of cerium is found on the
surface. Thus, both the Ce(NO& and NaNOs treatments lead to a similar behavior during
subsequent galvanostatic reduction. On the other hand, in near-neutral solutions (pH 6),
only cerium nitrate is efficient in inhibiting the cathodic kinetics, and in this case the effect is
solely due to the precipitation of cerium species. In the case of pure chromium, cathodic
inhibition efficiency increases with increasing pH, again due to easier cerium precipitation.
The finding of a different galvanostatic reduction behavior for chromium treated in NaNOX
solution of different pH values may be due to formation of oxide films with a pH-dependent
thickness or stoichiometry.
The oxygen reduction inhibition is thus most likely due to the formation of a highly
reduction-resistant oxide film-partially
due to the chromium enrichment and in addition
to the precipitation of insoluble Ce(OH)s at cathodic weak sites of the oxide layer. In order
to achieve the optimum cathodic inhibition, both the chromium enrichment and cerium
precipitation should be maximized. The resistance against reductive dissolution can be of
major importance for localized corrosion resistance. During localized attack such as pitting
or crevice corrosion, the outer surface is under cathodic polarization. Thus an oxide film
which is highly resistant against reductive dissolution prevents high cathodic oxygen
reduction currents and hence suppresses anodic pit growth.
CONCLUSIONS
(1) During exposure of stainless steel to hot (90C) cerium nitrate solution, a gradual
dissolution of iron oxide and an accompanying film growth by chromium passivation take
place, leading to an enrichment of chromium in the oxide layer. The chromium enrichment
increases with exposure time (15 min to 3 h) and with a lower treatment solution pH.
Further, insoluble cerium species are precipitated at the cathodic sites of the surface. The
amount of cerium incorporated does not depend strongly on the treatment time, but
decreases in acidic solutions. The modification of the oxide chemistry by the cerium
treatment leads to an inhibition of oxygen reduction kinetics on the stainless steel surface.

Role of cerium in treatment

of stainless steels

1913

(2) The effect of the cerium treatment cannot be solely attributed to the formation of a
chromium-rich oxide layer, since cathodic reactions are more strongly inhibited on the
cerium-treated stainless steel surface than on passivated pure chromium. Moreover, cerium
treatment of pure chromium leads to a retardation of the oxygen reduction reaction on the
chromium passive film.
(3) The inhibition of the oxygen reduction reaction is not due to a lower electron
conductivity of the oxide layer. The effect can be attributed to the high resistance of the
cerium-treated oxide film against reductive dissolution. This is partially due to a higher
chromium content of the passive film and, in addition, to the precipitation of cerium at weak
sites in the oxide layer which otherwise under cathodic polarization would lead to reductive
dissolution thus providing current paths for the oxygen reduction reaction. Factors leading
to a maximum chromium
enrichment and/or cerium precipitation and incorporation
increase the cathodic inhibition efficiency.
Acknowledgements-The
authors would like to thank MRCO and Long Manufacturing
Ltd for financial support
of this work, and Dr Mark Kozdras and Dr Brian Chiedle (Long Manufacturing
Ltd) for helpful discussions and
comments.

REFERENCES
1.
2.
3.
4.
5.
6.
7.

8.
9.

IO.
I I.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.

D.R. Arnott, B.R.W. Hinton and N.E. Ryan, Corrosion 45, 12 (1989).
F. Mansfeld, S. Lin, S. Kim and H. Shih, Corrosion 45, 615 (1989).
A.J. Davenport,
H.S. Isaacs and M.W. Kendig, Corros. Sci. 32, 653 (1991).
A.J. Aldykiewicz,
H .S. [saacs and A.J. Davenport,
J. Elecrrochem. Sot. 142, 3342 (1995).
Y.C. Lu and M.B. Ives, Corros. Sci. 34, 1773 (1993).
Y.C. Lu and M.B. Ives, Corros. Sci. 3-7, I45 (1995).
Y.C. Lu and M.B. Ives, in Proc. Critical Factors in Localized Corrosion II, ed. P.M. Natishan,
R.G. Kelly,
G.S. Frankel and R.C. Newman. The Electrochemical
Society, Inc., Pennington, NJ, Proc. Vol. 95-15, p. 253,
1996.
C.B. Breslin, C. Chen, F. Mansfeld, E. Otero and F.J. Perez, Extended Abstracts of the 189th Meeting of the
Electrochemical
Society, Paper No. 113, Los Angeles, CA (1996).
Y.C. Lu, M.B. Ives, G.I. Sproule and M.J. Graham, in Proc. H.H. Uhlig Memorial Symp., ed. A. Asphahari,
H. Biihni and R.M. Latanision.
The Electrochemical
Society, Inc., Pennington,
NJ, Proc. Vol. 94-26, p. 151,
1995.
L. Wegrelius and 1. Olefjord, Mater. Sci. Forum 185-188, 347 (1995).
A. Rossi and B. Elsener, Mater. Sci. Forum 185-188, 337 (1995).
J. Chastain,
Handbook of X-ray Photoelectron Spectroscopy. Perkin-Elmer,
Eden Prairie, 1992.
Z.H. Lu, B. Bryskieicz, J. McCaffrey,
Z. Wasilewski and M.J. Graham, J. Vat. Sci. Technol. Bll, 2033
(1993).
D.F. Mitchell, G.I. Sproule and M.J. Graham, Appl. Surf Sci. 21, 199 (1985).
D.F. Mitchell, J.S. Arlow, J.R. Phillips and G.I. Sproule, Surf. Interface Anal. 14, 302 (1989).
S. Virtanen and H. Biihni, Mater. Sci. Forum 185-188, 965 (1995).
K.J. Vetter, Elektrochemische Kinetik. Springer-Verlag,
Berlin, 1961.
J.W. Schultze, in Passivity of Metals,ed. R.P. Frankenthal
and J. Kruger. The Electrochemical
Society, Inc.,
Pennington,
NJ, p. 82, 1978.
Z. Szklarska-Smialowska,
K.-C. Chou and Z. Xia, Corros. Sci. 32, 609 (1991).
A.M.P. Simoes, M.G.S. Ferreira, G. Lorang and M. da Cunha Belo, Electrochim. Acta 36, 315 (1991).
P. Schmuki, S. Virtanen, A.J. Davenport
and C.M. Vitus, J. Electrochem. Sot. 143, 574 (1996).
P. Schmuki, S. Virtanen, H. BGhni, H.S. Isaacs, A.J. Davenport
and T. Stenberg, in Proc. Surface Oxide
Films, ed. J. Bardwell. The Electrochemical
Society, Inc., Pennington,
NJ, Proc. Vol. 96-18, p. 234, 1996.
T.P. Moffat, H. Yang, F.-R.F. Fan and A.J. Bard, J. Electrochem. Sot. 139, 3158 (1992).
M. Pourbaix, Atlas dEqu&bres Electrochimiques. Gautiers-Villars
and Vie, Paris, 1963.

Potrebbero piacerti anche