You are on page 1of 16

Oxidation of Metals, Vol. 64, Nos.

3/4, October 2005 ( 2005)


DOI: 10.1007/s11085-005-6560-x

Oxygen Transport during the High Temperature


Oxidation of Pure Nickel
S. Chevalier, F. Desserrey, and J. P. Larpin
Received March 11, 2003; revised April 21, 2005

The high temperature oxidation of nickel has been investigated in air under
atmospheric pressure in the temperature range 600900 C. The oxidation
kinetic curves deviate from the parabolic law for temperatures over 800 C.
The observation of scale morphologies and the use of two stage oxidation experiments under 16 O2 /18 O2 atmospheres showed that oxygen transport through the NiO scale had to be taken into consideration during the
oxidation process. Despite the main outward diffusion of Ni species through
the oxide scale, the inward oxygen diffusion at lower temperatures (<800 C)
or the oxygen transport, probably as molecular species, via pores or microcracks were found to play a major role in the formation of duplex oxide
scales, made of small equiaxed oxide grains at the metal/oxide interface
overgrown by larger columnar grains at the gas/oxide interface. Oxygen
diffusion coefcients into thermally grown NiO scales were determined and
compared to the values of Ni diffusion coefcients from the literature.
KEY WORDS: pure nickel oxidation; two stage oxidation under
logies; duplex scale; oxygen diffusion coefcients.

16 O /18 O ;
2
2

SIMS; morpho-

INTRODUCTION
The high temperature (HT) oxidation of nickel has been studied for many
years and is commonly considered as a model system, because of the
growth of a single NiO scale assumed to be controlled by outward cation diffusion.13 Nevertheless, oxygen penetration and formation of NiO
Laboratoire

de Recherches sur la Reactivite des Solides, UMR 56 13 CNRS, Universite de


Bourgogne, 9 avenue Alain Savary, BP 47870, 21078 Dijon cedex, France.
To whom correspondence should be sent. e-mail: sebastien.chevalier@u-bourgogne.fr
219
0030-770X/05/10000219/0 2005 Springer Science+Business Media, Inc.

220

Chevalier, Desserrey, and Larpin

duplex scales, generally composed of small equiaxed grains at the metal


-oxide interface overgrown by larger columnar grains in the external part,
was reported for scales grown in the temperature range under 1000 C.
Atkinson et al.1 indicated that oxygen penetrated down to the metal
oxide interface and that short-circuit diffusion of oxygen contributed to
the growth of the inner layer of the duplex scale. Moon et al.4 observed,
after sequential oxidation in 16 O2 and 18 O2 , that a layer of Ni18 O formed
on top of a Ni16 O layer but also that 18 O had penetrated into the Ni16 O
layer. They proposed that gaseous oxygen penetrated the scale along pathways such as through-scale micro-channels or closely spaced transient
ssures. More recently, Kyung et al.5 observed that inert markers were
located along the interface between the predominantly columnar outer
NiO layer and the very ne-grained inner layer of NiO, indicating that the
formation of the inner layer is controlled by inward oxygen transport.
The mechanisms by which the duplex structure formation is enhanced
is still disputed, but several hypotheses have been claimed. Two recent
papers5, 6 listed most of them. Mrowec7 proposed that the growth stresses
led to oxide separation from the metal due to insufcient oxide plasticity.
Gibbs and Hales8 rened this model, assuming that oxygen could penetrate
the scale via pores, which were formed by vacancy condensation and dissociation of the oxide. The dissociation of the oxide was also proposed by Yurek
and Schmalzried.9, 10 Evans et al.11 suggested that the stresses produced during the oxidation process were a probable source of porosity which acted as
an available space for the inner layer growth. To Kofstads opinion,12 the
internal oxidation was also due to the porosity formation because creep
occurred in the oxide layer during its growth. Smeltzer et al.13, 14 supposed
that transport of both anions and cations signicantly contributed to scale
growth, and that the possibility of the formation of a new oxide within
the existing layer might exist. Atkinson et al.1517 advanced that short-circuit diffusion of oxygen appeared to occur to a greater or lesser extend in
practically all oxide lms. The inward oxygen gas transport was responsible for the growth of new oxide within the lms and for the growth of the
inner layer of the duplex oxide layers. Robertson and Manning18 estimated
that single layers grew when dislocation climb eliminated space created at
the metaloxide interface by outward metal diffusion. Oxidation proceeded
until the space was lled with new oxide which formed the more inner layer
oxide. Recently, Peraldi et al.19 observed that the growth of duplex NiO
scales did not proceed only by the outward diffusion of cationic species. It
seemed that the inner and outer scale formation was also associated to the
inward transport of oxygen.
The reasons for the formation of duplex NiO microstructure seem
to be related to inward diffusion of oxygen. Nevertheless, the required

Oxygen Transport during the High Temperature Oxidation of Pure Nickel

221

experimental conditions to obtain the duplex structure (oxide thickness, oxidation conditions, Ni purity, . . . ) and the diffusion pathways are yet unclear.
Only a few studies have been made on the determination of oxygen
diffusion coefcient in NiO.2024 As far as we know, only two studies concern the oxygen diffusion in polycrystalline NiO scales.23, 24
The aim of this work was to study the oxidation of Ni in air under
atmospheric pressure over the temperature range 6001000 C, and to rely
kinetics and morphologies of the NiO scales to 16 O2 /18 O2 oxidation tests
followed by secondary ion mass spectrometry (SIMS) analyses. Oxygen
diffusion coefcients (effective, bulk and grain boundary diffusion coefcients) in thermally grown NiO scales were calculated and compared to the
oxygen as well as the nickel diffusion coefcients from the literature.

EXPERIMENTAL PROCEDURE
Ni (99.98% from Aldrich) samples were oxidized in laboratory air
under atmospheric pressure in a thermobalance over the temperature
range 6001000 C. Prior to oxidation, 10 10 1 mm samples were polished down to 1 m with diamond paste and washed in alcohol in an
ultrasonic bath.
Two stage oxidation experiments were processed by oxidizing rstly Ni
specimen under 200 mbar 16 O2 pressure and secondly under 200 mbar of
18 O pressure (isotopic enrichment was about 90%). The samples were not
2
cooled between the two steps of the oxidation test to avoid thermal shocks,
which could provoke cracks and/or spallation of the oxide scales. After the
second oxidation stage, the oxidizing atmosphere was evacuated to reach a
low pressure and the samples were quickly removed outside the furnace to
be cooled down to room temperature. Two tests were performed in order,
on one hand, to understand the NiO scale growth mechanism (Table Ia)
and, on the other hand, to determine the oxygen diffusion coefcients in
the thermally grown NiO scales (Table Ib). SIMS proles were made with a
MIQ 256 Riber analyzer using 13 keV Ar+ beam as primary ion beam.
The effective, bulk and grain boundary oxygen diffusion coefcients
werecalculated assuming that oxygen diffusion occurs in a B regime2527 ,
Db t /2, where is the grain boundary width, Db is the bulk diffusion coefcient, t is the diffusion time and the average grain size.
The relative concentration of 18 O was calculated from the measured
ion intensities I(18 O) and I(16 O) using the expression c(18 O) = I(18 O)/
[I(18 O)+I(16 O)]. A typical prole, ln[c(18 O)] = f(x), is shown in Fig. 1, and
clearly exhibits two domains.28, 29 The rst domain is related to effective diffusion (Deff ) and the second domain to grain boundary diffusion

222

Chevalier, Desserrey, and Larpin

Table I. Experimental Conditions for the Two Stage Oxidation Experiments under
16 O /18 O , a) NiO Growth Mechanism and (b) Oxygen Coefcients
2
2
T( C)

Duration under

16 O
2

(min)

Duration under

(a) Nio growth Mechanism


600
130
650
75
700
30
800
30
900
25
(b) Oxygen diffusion co-efcient
600
300
650
300
700
120
800
120
900
60

18 O
2

(min)

540
120
45
45
35
60
50
45
45
30

Bulk and grain boundary diffusion(Deff)


4

Grain boundary diffusion (Dgb)

18

18

16

ln ( O/ O+ O)

-1

-2
0.0

0.1

0.2

0.3

0.4

0.5

x (m)
Fig. 1. Plot of ln(18 O/18 O+16 O) = f(x) indicating the domain where effective diffusion predominates and the domain where grain boundary diffusion predominates.

(Dgb ). The rst part of the oxygen prole allows to determine Deff
according to Ficks law solution for a constant surface concentration:


x
C(x) Cs
(1)
= erf
C0 C s
2 Deff t
where C0 is the natural tracer concentration in the sample (0.2 at. %30 )
and Cs the constant supercial tracer concentration.

Oxygen Transport during the High Temperature Oxidation of Pure Nickel

223

In the second part of the proles, Dgb can be determined from the
Whipple-Leclaire equation:31



4Db dInC 5/3
Dgb = 0.661
(2)
t
dx6/5
where ddxln6/5C is the slope of the curve ln C=f(x6/5 ) and t is the diffusion time.
Hart32 proposed that Deff corresponded to a balance between dislocation and lattice diffusion. The concept of Hart can be applied to transport
through oxide lms as done before by Smeltzer et al.33 and by Atkinson.1
The relation between Dgb and Db could be written as:
Deff = (1 f )Db + f Dgb

(3)

where f the fraction of sites associated to grain boundaries, which is generally expressed as:
3
(4)
f=

Assuming that = 107 cm and combining equations (2) and (3), Db is


obtained via the following second order equation:



0.661f 4Db
d ln C 5/3
Deff = 0
(5)
6/5
(1 f ) Db +

t
dx


s
= f (x)
Deff is calculated from the slope of the curve arg erf C(x)C
C0 Cs
in the rst part of the diffusion prole, Db with Eq. (5) and Dgb from Eq.
(3) using the second part of the prole.
The top surface morphologies and fracture cross-sections of the oxidized samples were analyzed by scanning electron microscopy (SEM),
using a eld-emission gun (FEG), coupled with an energy dispersive X-ray
analysis (EDX).
RESULTS
Oxidation Kinetics
Figure 2a exhibits the plots m/A vs. time corresponding to isothermal oxidation tests under atmospheric pressure in the temperature
range 600900 C. The weight gains increase with temperature. The plots
of m/A vs. time1/2 (Fig. 2b) indicate a deviation from parabolic law for
the higher temperature experiments. The oxidation performed at 900 C
exhibits the largest deviation to parabolic law. The kp values estimated
from the linear parts of the plots (Table II) lead to an activation energy
of 220 5 kJ.mol1 , which is comparable to values previously reported.6

224

Chevalier, Desserrey, and Larpin

(a) 2,5

600C
650C

m/A (mg.cm-2)

2,0

700C
800C

1,5

900C

1,0

0,5

0,0
0

10

15

20

25

t (h)
(b) 2,5

600 C

m/A (mg.cm-2)

2,0

650C
700C

1,5

800C
900C

1,0

0,5

0,0
0

3
1/2

1/2

(h )

Fig. 2. Kinetic curves for pure nickel oxidation in air under atmospheric pressure for
24 hr in the temperature range 600900 C.

Oxygen Transport during the High Temperature Oxidation of Pure Nickel

225

SEM Observations
The oxidized surfaces were observed using a FEG-SEM. The top
surface morphologies after 24 hr oxidation tests at 600, 650, 700, 800
and 900 C are shown in Fig. 3(ae). They exhibit the classical cellular
morphology with oxide ridges and open porosities. The oxide grain size
increases with temperature as well as the numbers of pores and a very few
platelets are visible on some specimens. These observations are in agreement with a recent work of Peraldi et al.34 concerning the morphology
and microstructure of NiO scales grown at high temperature on highpurity nickel. The observation of platelets in NiO and the explanation of
their formation were proposed by Dufour et al.35, 36 They demonstrated
that the presence of impurities within the metal and the stresses accumulated in the twin primary oxide layer was responsible for the platelet
growth. The very limited number of platelets observed on our samples is
probably due to the purity of the nickel tested in this study.
Figure 4 shows the fracture cross-section of a nickel specimen oxidized at 800 C for 12 hr. A duplex microstructure composed of small equiaxed oxide grains close to the metaloxide interface and columnar oxide
grains in the external part of the scale, is clearly visible. The fracture
cross-section was obtained by quenching the sample in liquid nitrogen and
deforming the sample with pliers. As a result, the sample has a curved
shape. The duplex microstructure could not be observed on samples oxidized at lower temperature.
Oxide Scale Growth Mechanism
Two-stage oxidation experiments under 16 O2 and 18 O2 provide some
insights on the mechanisms of the scale formation. The location of the
second oxidant, 18 O2 , leads to know if metal and/or oxygen transport
through the NiO scale are responsible for the oxide scale growth under
high temperature conditions. For the three tested temperatures, 600 C

Values
After
Table II. kp
Oxidation Tests in air Under
Atmospheric Pressure in the
Temperature Range 600900 C
T ( C)
600
650
700
800
900

kp g2 cm4 . s1
2.4 1014
8.4 1014
4.2 1012
7.0 1012
5.8 1011

226

Chevalier, Desserrey, and Larpin

Fig. 3. SEM oxide surface morphologies after 24 hr in air under atmospheric pressure,
(a) 600 C, (b) 650 C, (c) 700 C, (d) 800 C and (e) 900 C.

(Fig. 5a), 650 C (Fig. 5b) and 900 C (Fig. 5c), a large peak of 18 O is
observed in the external part of the NiO scale, indicating a major diffusion
of Ni ions. A slight peak of 18 O is present at the metal-oxide interface at
600 C. This peak intensies at 650 C and become stronger at 900 C. This
clearly indicates that oxygen transport increases as the oxidation temperature increases.

Oxygen Transport during the High Temperature Oxidation of Pure Nickel

227

Fig. 4. SEM observation of a fracture cross-section of NiO scale formed during 12 hr at


800 C in air under atmospheric pressure.

Oxygen Diffusion Coefcients


Diffusion proles of 18 O were obtained on pre-grew Ni16 O scales.
O ), oxygen grain boundary diffuOxygen effective diffusion coefcients (Deff
O
sion coefcients (Dgb ) and oxygen bulk diffusion coefcients (DbO ) were
determined for temperatures of 600, 650, 700 , 800 and 900 C, following
the pre-described method (Fig. 1). The results are summarized in Table III
and compared to results of Atkinson et al.37, 38 who determined the Ni
diffusion coefcients through NiO scales. For the three tested temperatures
oxygen diffusion coefcients at the NiO grain boundaries are over the oxygen diffusion coefcients in the bulk of NiO. The values calculated by
Atkinson et al.37, 38 indicate that for these temperatures the nickel diffusion coefcients at the NiO grain boundaries are over the nickel diffusion
coefcients in the bulk of NiO.

228

Chevalier, Desserrey, and Larpin

counts

(a)
3.0x10

2.5x10

2.0x10

1.5x10

1.0x10

5.0x10

600C
O16
O18

0.0
0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

thickness (m)

(b)
3.0x10

2.5x10

2.0x10

15x10

1.0x10

5.0x10

counts

650C
O16
O18

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

thickness (m)

counts

(c)
5x10

4x10

3x10

2x10

1x10

900C
O16
O18

0.0

0.5

1.0

1.5

2.0

thickness (m)

Fig. 5. Oxygen isotope SIMS proles after twostage oxidation experiments under 16 O2 and 18 O2
atmospheres, (a) 600 C, (b) 650 C and (c) 900 C.

DISCUSSION
The thermogravimetric study of the oxidation of pure nickel in the
temperature range 600900 C indicates that the kinetic curves deviate
from the parabolic rate law over 800 C. Below this temperature, the specic weight gain curves follow a parabolic law. According to Wagners

f = 3/

O
Deff
cm2 .s1

600
650
700
800
900

0.2
0.2
0.25
0.4
0.5

0.015
8.3 1015
0.015
6.2 1014
0.012
2.1 1013
0.0075
1.8 1011
0.006
4.2 1011


3
Ni
(a) Dgb
= 3 108 exp 171.7 10
after Atkinson et al.38
RT

3
(b) DbNi = 2.2 102 exp 24710
after Atkinson et al.37
RT
Ni
Ni
Ni
(c) Deff = (1 f ) Db + f Dgb after Hart32 .

T ( C)

Grain
size
(m)
5.5 1013
4.1 1012
1.8 1011
2.4 1010
7.0 109

O
Dgb
cm2 .s1

1.7 1017
7.7 1017
1.4 1016
2.1 1016
9.4 1015

DbO
cm2 .s1
1.6 1011
5.7 1011
1.8 1010
1.3 109
7.0 109

Ni
Dgb
cm2 .s1 (a)

3.6 1017
2.3 1016
1.2 1015
2.0 1014
2.2 1013

DbNi
cm2 .s1 (b)
2.4 1013
8.5 1013
2.2 1012
9.8 1012
4.1 1011

Ni
Deff
cm2 .s1 (c)

Table III. Oxygen Diffusion Coefcients Calculated in the Temperature Range from 600 to 900 C and Comparison with Nickel Diffusion
Coefcients Determined by Atkinson et al.37,38

Oxygen Transport during the High Temperature Oxidation of Pure Nickel


229

230

Chevalier, Desserrey, and Larpin

theory39 , to get a parabolic law implies that diffusion of species through


the growing NiO scale is dominating the oxidation process. The deviation
from the parabolic oxidation kinetics can point out that diffusion of species through the scale is not more the determining step of the oxide scale
growth.40 The contribution of local stresses in the scale, the grain growth
in the scale, the diffusion in the substrate or the depletion effects (in the
case of alloys) can be responsible for the deviation of the curves.41 Formation of pores, cracks or ssures within the scale16, 42 can also explain that
the diffusion of species is not more the determining step of the oxidation
process, but that interfacial reaction steps can become predominant.
The use of two-stage oxidation under 16 O2 and 18 O2 is a technique
permitting to analyze the oxide scale growth mechanisms. For the tested
temperatures (600, 650 and 900 C), the SIMS proles (Fig. 5) show that
the 18 O isotope is incorporated mainly into the oxide scale at the gas/oxide
interface, as expected for an oxide growing by cation diffusion.16 The presence of a second 18 O peak close to the metal/oxide interface suggests that
oxygen transport participates also to the NiO scale growth. A recent paper
of Huntz et al.43 relates that the presence of impurities in nickel leads
to inward diffusion of oxygen with formation of an inner equiaxed oxide
lm, thus to a duplex oxide scale. The oxide scale growth stresses can also
play a key role in the growth of NiO scales.44, 45 . In particular, the stress
relief in the oxide seems to occur by microcrack formation which allows
transport of molecular oxygen through the oxide scale at 800 and 900 C.46
The rst study concerns diffusion of oxygen in the temperature range
11001600 C which is far away from the temperatures tested in this work.
As a consequence, we can not compare the values of oxygen diffusion
coefcients. The second study determines that the value of effective diffusion coefcient of oxygen in NiO at 550 C is 31016 cm2 . s1 for Ni
(111).24 This value is comparable to the values we determined in the
temperature range 600700 C. Figure 6 exhibits the Arrheniuss plot of
our effective diffusion coefcients associated with the value determined
by Berger et al.24 The so-calculated activation energy is 230 kJ . mol1 .
The determination of the oxygen diffusion coefcients shows that oxygen diffuses much more rapidly at the NiO grain boundaries than in the
O ) to the values calcuNiO bulk. The comparison of these values (Dgb
lated by Atkinson et al.37, 38 for Ni diffusion at NiO grain boundaries
Ni ) indicates that Ni diffusion at NiO grain boundaries is between one
(Dgb
and two orders of magnitude faster than O diffusion via the same paths
when T<800 C (Table III). Whatever the diffusive species, O or Ni, the
oxide grain boundaries appear to be the main diffusion paths, at least
over the temperature range 500900 C. At 800 and 900 C, the O diffu-

Oxygen Transport during the High Temperature Oxidation of Pure Nickel

231

-28
-29
-30

lnDeff

-31

eff

(this work)

eff

(Berger at al. )

44

-32
-33
-34
-35
-36
1,0x 10

-3

1,1x1 0

-3

1,2x 10

-3

1/T (K-1)
Fig. 6. Arrhenius plot of oxygen effective diffusion coefcients and comparison with the
value determined by Berger et al.44

sion coefcients were found to be larger than the nickel diffusion coefcients estimated from Atkinsons results.38 The formation of open pores
or micro-channels, which lead to the fast penetration of oxygen through
the oxide scale, may induce major errors and artefacts in the determination of oxygen diffusion coefcients. The formation of pores or microchannels were already observed in NiO scales.5 It can explain, in our
case, the large 18 O peak observed close to the metaloxide interface at
900 C.
The rapid transport of oxygen can also be related to the formation
of the duplex oxide scales, composed of equiaxed NiO grains close to
the metal/oxide interface, overgrown by columnar NiO grains. The formation of the inner oxide layer corresponds to the apparition of a large
18 O peak at the inner interface, as already mentioned by Monceau et al.47
This suggests that over 800 C, oxygen penetrates through the oxide scale
via pores or micro-channels and form new oxide at the metal/oxide interface16, 17 . The formation mechanism of the pores or the micro-channels is
still unclear, but stress relief in the oxide layer during its growth seems to
promote the microcrack formation.46

232

Chevalier, Desserrey, and Larpin

Relating the two-stage oxidation experiments with the kinetics and the
SEM morphologies of the scale, we assume that the nickel oxide scale
growth is controlled by Ni diffusion via NiO grain boundaries at temperature lower than 800 C, with a slight diffusion of oxygen through the same
pathways. Above 800 C, the formation of cracks or micro-channels within
the scale can induce a faster penetration of oxygen to the metal/oxide
interface; this phenomena explains the deviation of the kinetic curves from
the parabolic law, because interfacial reaction steps become predominant
in the oxidation process. As a consequence, a duplex oxide layer is formed
with an external columnar structure corresponding to outward diffusion of
nickel and an inner part composed of equiaxed oxide grains, corresponding to inward penetration of oxygen. Therefore, the duplex scale microstructure obeys complex growth kinetics, as proposed by Peraldi et al.48
who suggest that no simple kinetics law is appropriate to analyze NiO
scale growth for intermediate temperatures (<900 C).
CONCLUSION
The oxidation of Ni was carried in air under atmospheric pressure
over the temperature range 600900 C. The kinetic curves deviate from the
parabolic law above 800 C. The use of two stage oxidation experiments
denitely proves that oxygen transport through the oxide scale has to be
taken into consideration. At temperatures lower than 800 C, oxygen diffusion via NiO grain boundaries is lower than nickel diffusion via the same
pathways. Above 800 C, inward transport of oxygen occurs via microcracks or micro-channels leading to deviation from the parabolic law and
formation of duplex oxide scales, made of columnar NiO grains close to
the metal/gas interface and equiaxed grains close to the metal/oxide interface.
ACKNOWLEDGMENT
A part of the present work was made thanks to Procope nancial
support (YM 00356); the authors are thankful to the French and the
German governments. The authors would like to acknowledge Pr. Borchardt (Technische Universitat Clausthal, Germany) for fruitful discussions
and Dr. G. Strehl (Technische Universitat Clausthal, Germany) for the
two-stage oxidation experiments. A special thank to Dr. O. Heintz (LRRS,
University of Burgundy, France) for the realization of the SIMS proles
and to Dr. T. Montesin (LRRS, University of Burgundy, France) for his
scientic contribution.

Oxygen Transport during the High Temperature Oxidation of Pure Nickel

233

REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.

A. Atkinson, R. I. Taylor, and A. Hughes, Philosophical Magazine A 45, 823 (1982).


S. Mitra, Material Science Engineering A 174, 103 (1994).
M. J. Graham and R.J. Hussey, Oxidation Metals 44, 339 (1995).
D. P. Moon, A. W. Harris, P. R. Chalker, and S. Mountoford, Materials Science Technology 4, 1101 (1988).
H. Kyung and C. K. Kim, Materials Science Engineering B 76, 173 (2000).
R. Haugsrud, Corrosion Science 45, 211 (2003).
S. Mrowec, Corrosion Science 7, 563 (1967).
G. B. Gibbs and R. Hales, Corrosion Science 17, 487 (1977).
G. J. Yurek, H. Schmalzried, Berichte Der Bunsengesellschaft Physical Chemistry 78,
1379 (1974).
G. J. Yurek, H. Schmalzried, Berichete Der Bunsengesellschaft Physical Chemistry 79,
255 (1975).
A. G. Evans, D. Rajdev, and D. L. Douglass, Oxidation of Metals 4, 151 (1972).
P. Kofstad, Oxidation of Metals 24, 265 (1985).
W. W. Smeltzer, R. R. Haering, and J. S. Kirkaldy, Acta Metallurgica 9, 880 (1961).
J. M. Perrow, W. W. Smeltzer, and J. D. Embury, Acta Metallurgica 16, 1209 (1968).
A. Atkinson, R. I. Taylor, and P. D. Goode, Oxidation of Metals 13, 543 (1979).
A. Atkinson and D. W. Smart, Journal of the Electrochemical Society 135, 2886 (1988).
A. W. Harris and A. Atkinson, Oxidation of Metals 34, 229 (1990).
J. Robertson and M. I. Manning, Materials Science and Technology 4, 1064 (1988).
R. Peraldi, D. Monceau, and B. Pieraggi, Materials Science Forum 369372, 189 (2001).
C. Dubois and C. Monty, J. Philibert, Philosophical Magazine A3, 419 (1982).
J. Cabrera-Cano, A. Dominguez-Rodriguez, and R. Marquez, J. Castaing, and
J. Philibert, Philosophical Magazine A 46, 397 (1982).
C. Dubois, C. Monty, and J. Philibert, Solid State Ionics 12, 75 (1984).
A. Atkinson and F. C. W. Pummery, in Transport in Nonstoichiometric Compounds, eds,
G. Simkovic and V. S. Stubican, (Plenum Press, New York and London, 1985) p. 359.
P. Berger, L. Gaillet, R. El Tahhann, G. Moulin, and M. Viennot, Nuclear Instrument
and Methods in Physics Research B 181, 382 (2001).
Y. M. Mishin and G. Borchardt, Journal of Physics III France 3, 863 (1993).
Y. M. Mishin and G. Borchardt, Journal of Physics III France 3, 945 (1993).
Y. M. Mishin and G. Borchardt, Journal of Physics III France 7, 1797 (1997).
S. C. Tsa, A. M. Huntz, and C. Dolin, Oxidation of Metals 43, 581 (1995).
S. C. Tsa, A. M. Huntz, and C. Dolin, Materials Science and Engineering A212, 6 (1996).
C. M. Lederer, J. M. Hollander, and I. Perlman, Table of Isotopes, 6th edn, (J. Wiley and
sons, INC., New-York, 1967) p. 6.
R. T. P. Whipple, Philosophical Magazine 45, 1225 (1954).
E. W. Hart, Acta. Metallurgia 5, 597 (1957).
W. W. Smeltzer, R.R. Haering, and J. S. Kirkaldy, Acta Metallurgia 9, 880 (1961).
R. Peraldi, D. Monceau, and B. Pieraggi, Oxidation of Metals 58, 249 (2002).
F. Morin, L. C. Dufour, and G. Trudel, Oxidation of Metals 37, 39 (1992).
L. C. Dufour and F. Morin, Oxidation of Metals 39, 137 (1993).
A. Atkinson and R. I. Taylor, Philosophical Magazine A39, 581 (1979).
A. Atkinson and R. I. Taylor, Philosophical Magazine A43, 979 (1981).
C. Wagner, Journal of Electrochemical Society 99, 369 (1952).
G. Borchardt and G. Strehl, Materials Aspects in Automotive Catalytic Converters
(Wiley-VCH, Munich, October 2001), 106116.
G. Borchardt and G. Strehl, Proc. of the International Conf. On Materials Aspects in
Automotive Catalytic Converters (MACC), Munich, Germany (34 October 2001).
K. P. Lillerud and P. Kofstad, Journal of Electrochemical Society 127, 2397 (1980).
A. M. Huntz, A. Lefeuvre, M. Andrieux, C. Severac, G. Moulin, R. Molins, and
F. Jomard, Proc. of the 5th Int. Conf. on Microscopy of Oxidation, Limerick, Ireland,
2628 August 2002 (to be published).

234

Chevalier, Desserrey, and Larpin

44. P. Y. Hou and R. M. Cannon, Materials Science Forum 251254, 325 (1997).
45. A. M. Huntz, G. Calvarin Amiri, H. E. Evans, and G. Cailletaud, Oxidation of Metals
57, 499 (2002).

46. W. Przybilla and M. Schutze,


Oxidation of Metals 58, 103 (2002).
47. D. Monceau, R. Peraldi, and B. Pieraggi, Defect and Diffusion Forum 194199, 1675
(2001).
48. R. Peraldi, D. Monceau, and B. Pieraggi, Oxidation of Metals 58, 275 (2002).

You might also like