You are on page 1of 10

J Mater Sci (2014) 49:277286

DOI 10.1007/s10853-013-7702-1

Nanoceria coating imperfections and their effect on the


high-temperature oxidation resistance of a 304 stainless steel
H. F. Lopez H. Zhang

Received: 13 June 2013 / Accepted: 26 August 2013 / Published online: 25 September 2013
Springer Science+Business Media New York 2013

Abstract The microemulsion method was employed in


this work for synthesizing nanocrystalline cerium oxide
particles. Average nanoparticle sizes of 3.45 nm were
produced by these means. XPS determinations indicated
that both Ce3? and Ce4? are present in the synthesized
nanoceria particles, with an average amount of 22.8 % of
Ce3? ions. It was found that the nanocrystalline cerium
oxide coatings lead to significant improvements (of 12
orders of magnitude) in the high-temperature oxidation
resistance of 304 SS. In addition, it was found that by
decreasing the nanoparticle mean size from 10 to 3.45 nm,
the effect of the coating protection was drastically
improved. The experimentally determined parabolic rate
constants kp at 1073 and 1273 K for 304 SS indicated a
reduction of up to two orders of magnitude when nanoceria
coatings with 3.45 nm in mean size were applied. Also, the
scale thickness was reduced from 15 to 5 lm when oxidized at 1073 K for 442 h. Coatings with purchased nanoceria particles (10 nm) were not as effective as the
coatings with synthesized nanoceria. Apparently, at
increasing nanoceria sizes, the oxidation protection is significantly reduced. In addition, it was found that the
method of dipping for coating 304 SS does not provide a
uniform coverage with nanoceria particles. Fe-rich
islands develop during high-temperature oxidation

H. F. Lopez (&)
Department of Materials Science and Engineering, CEAS,
University of Wisconsin-Milwaukee, 3200 N. Cramer Street,
Milwaukee, WI 53211, USA
e-mail: hlopez@uwm.edu
H. Zhang
Civil and Materials Department, University of Illinois-Chicago,
700 S. Halsted St. No. 2029, Chicago, IL 60607, USA
e-mail: haiyin@uic.edu

indicating that some regions do not exhibit the protection


that nanoceria can provide. In contrast, relatively thickcoating regions on the steel substrate exhibit minimal
oxidation, and the resultant scale is fine grained and
uniform.

Introduction
Over the past two decades, there have been extensive
efforts aimed at disclosing the role played by rare earths
(RE), on the reactive element effect (REE) in the oxidation
resistance of high-temperature alloys [13]. In 304 stainless steel (SS), CeO2 nanoparticle coatings have been
highly effective in reducing the oxidation rates by up to
13 orders of magnitude when they are of nanometric sizes
[49]. In contrast, submicrometric or large-nanoparticlesized CeO2 coatings do not seem to provide any significant
oxidation protection when compared to bare 304 SS [8, 9].
Apparently, at relatively large particle nanosizes, the defect
structure of the CeO2 becomes dominated by tetravalent
oxygen anion vacancies. Yet, at reduced nanocrystallite
sizes, there is an increase in the formation of mixed valence
cations, and thus, trivalent oxygen anion vacancies become
increasingly present, probably as a result of surface-tension-driven effects [10].
In particular, the work of Patil et al. [5] and Bauchard
et al. [11] has shown that the presence of mixed valence
oxygen vacancies plays a key role during the establishment
of the initial oxide scale at the substrate/ceria-coating
interface in either chromia-forming substrates [5] or alumina-forming Ni-based alloys [11]. In both cases, it has
been shown that the generation of vacancies driven by
mixed valence cations in the coating significantly contributes to the enhanced oxidation protection. This is

123

278

J Mater Sci (2014) 49:277286

critical during the initial oxidation stages because it


appears to control the amount of available oxygen to grow
the most stable oxide at the coating/substrate interface.
From the published reports, it is apparent that nanoceria
coatings impart protection in the high-temperature oxidation of 304 SS but not as effective as in 316 or 321 SS [7].
Moreover, thermal cycling seems to be highly efficient in
improving the oxidation resistance of some of these steels,
but not in 304 SS [7]. Although there have been various
coating methods implemented on high-temperature alloys
[1], most of the published reports have been focused on
dipping the steels into a nanoceria solution produced by the
microemulsion method [39]. Under these conditions, it is
expected that a few monolayers of nanoceria particles will
uniformly coat the substrate surface with reported thicknesses between 100 and 800 nm. In addition, the substrate
surfaces are often prepared by either mechanical polishing
and/or electropolishing with or without a short pre-oxidation treatment. This treatment is expected to improve the
wettability of CeO2 after the dipping process [12]. For the
most part, high-temperature exposure of stainless steels has
been carried out for short periods of time (24 h or less) and
a single temperature [47], typically 1273 K.
It has been suggested that the dipping method provides a
highly uniform coating after using sodium bis-(2-ethylhexyl)sulphosuccinate (AOT) as a surfactant [4]. However,
there is evidence [9] that the dipping method can lead to
coating imperfections, which might severely impair the
protection of the steel substrates. When regions of the
substrate are devoid of nanoceria coatings, these regions
can locally oxidize at increasing rates when compared with
nanoceria-coated regions. Moreover, due to the lack of
coating uniformity, there are local regions which exhibit
enhanced protection due to increasingly thick-coating
layers [9]. Hence, in the present work, the oxidation
behavior of nanoceria coatings produced by the dipping
method is investigated by exposure of a 304 SS to dry air at
1073 and 1273 K. In particular, the work is focused on the
oxidation kinetics and possible coating imperfections that
might impact the exhibited oxidation resistance of the
investigated 304 SS.

Experimental
Nanocrystalline cerium oxide was synthesized in this work
following the microemulsion method published by Patil
et al. [4] and Wu et al. [13]. The starting materials

employed were cerium nitrate hexahydrate (Ce(NO3)36H2O), toluene, water, hydrogen peroxide (H2O2),
and sodium bis-(2-ethylhexyl)sulphosuccinate (AOT) as a
surfactant. All of the chemicals and reagents were purchased from Aldrich Chemicals Company, Inc. The synthesis process is described in detail somewhere else, but it
is similar to the one reported by other authors [4]. A
solution containing the nanocrystalline CeO2 was extracted
from the upper layer of the microemulsion system and
collected as a coating solution.
In addition, purchased nano-CeO2 was dispersed in
ethanolwater solutions in order to achieve a uniform
distribution during the coating of the stainless steel. The
nanoparticle sizes were measured using a light scattering
technique based on the Malvern Instrument Zetasizer Nano
ZS. In addition, the top layer solution containing the
nanocrystalline ceria particles was deposited on a polished
silicon substrate for further examination. Chemical analyses were carried out using energy dispersive spectroscopy
(EDS) in a TOPCON, ABT-32 scanning electron microscope (SEM) at an accelerating voltage of 20 kV. In
addition, the surface chemistry of the synthesized cerium
oxide particles was analyzed using a 5950A HP X-ray
photoelectron spectroscopy (XPS) spectrometer. For this
purpose, monochromatic Al-Ka X-radiation at a power of
600 W was employed. A binding energy of 284.5 eV for
C(1s) was assumed as a reference in order to remove any
charging shift effects caused by the samples.
AISI 304 SS coupons with 20 9 20 mm dimensions
were cut from 1.0-mm-thick steel sheets. The samples were
polished using 1200 grit SiC paper and ultrasonically
cleaned. Table 1 gives the composition of the 304 SS. All
coupons were ultrasonically cleaned in iso-propanol and
air-dried, and then pre-oxidized at 973 K for 2 min under
dry air conditions. Pre-oxidation improves surface wetting
and helps to form a well-adherent coating. The coupons
were manually dip coated in the prepared solutions. The
coating process was repeated several times with intermediate drying at 373 K. The coating times were the same for
all the coupons in order to obtain similar coating thicknesses. Discontinuous isothermal oxidation tests were
carried out on both bare and coated coupons in dry air at
1073 and 1273 K for up to 24 and 400 h, respectively. A
highly accurate Sartorium balance (seven digits) was used
for weight measurements of the oxidized samples with
time, and the exhibited oxidation rates were determined. In
addition, the oxide scale morphology was characterized by
EDX and SEM including EDS means.

Table 1 Chemical composition of 304 stainless steel (wt%)


Grade

Mn

Si

Cr

Ni

Mo

Ti

304

B0.08

B2.0

B0.045

B0.03

B0.75

1820

810.5

123

J Mater Sci (2014) 49:277286

Results and discussion


Nanoparticle characterization
Figure 1a, b shows the statistical distributions of the
effective particle sizes of nanocrystalline cerium oxide
agglomerates. In Fig. 1a, it can be observed that the
effective size of synthesized nanoparticles is 24 nm, and
the average size is 3.45 nm. Notice in particular that the
synthesized nanocrystalline cerium oxide particles show a
highly uniform distribution. In contrast, the size range in
purchased nanoceria particles varies from 6 to 100 nm, and
the average size is close to 10 nm (Fig. 1b).
Figure 2a shows the distribution of nanoceria clusters on
the 304 SS substrate after dipping and drying, whereas
Fig. 2b shows the EDS peak intensities of the synthesized
nanoceria particles. Notice from the EDS spectrum the
dominant presence of cerium ions. The silicon peak corresponds to the substrate, while the sodium and sulfur
peaks correspond to the AOT surfactant. From Fig. 2a, it
can be observed that the ceria nanoparticles tend to form
agglomerates on the substrate surface resulting in a nonhomogeneous distribution.
Figure 3 shows the Ce(3d) spectra obtained in this
study. An interpretation of the exhibited Ce(3d) spectrum is
not straightforward. Hence, Gaussian functions are used in
this work to model the peaks. Although it is fairly simple to
obtain a good match between the simulation and the
experimental spectrum, it is difficult to get consistent
results. The fitted spectrum consists of 10 overlapped peaks
resulting from the presence of both Ce3? (4 peaks) and
Ce4? (6 peaks).
The concentrations of Ce3? and Ce4? in cerium oxide
were determined from their integrated peak areas. The
results gave an average concentration of Ce3? ions of
22.8 5 % and Ce4? concentration of 77.2 %. These

279

values are within the range of concentration values reported by other workers [6, 13]. From these results, it is evident that the synthesized nanocrystalline oxide is mainly
tetravalent cerium oxide. This agrees with the work of
Deshpande et al. [14], which shows that with a reduction in
CeO2 particle sizes below 5 nm, the concentration of Ce3?

Fig. 2 a Nanoceria particle agglomerates on the surface of a stainless


steel substrate and b EDS spectrum of synthesized cerium oxide
nanoparticles
1

Experimental
Fitted

Normalized Intensity

0.9
0.8

Ce4+

Ce

4+

3+

0.7

Ce

Ce4+

0.6

Ce4+
Ce3+

0.5

0.3

Ce4+

Ce3+

0.4
Ce

Ce3+

4+

0.2
0.1
0
880

885

890

895

900

905

910

915

920

Binding Energy (eV)

Fig. 1 Nanocrystalline cerium oxide particle size distributions.


a Particles synthesized in this work and b purchased nanoparticles

Fig. 3 Ce(3d) XPS spectrum for cerium oxide nanoparticles synthesized by the microemulsion method (experimental and fitted peaks)

123

280

ions continually increases from 17 to 44 %. In addition, a


lattice expansion is expected at these particle sizes.
Tsunekawa et al. [15] reported anomalous lattice expansions in cerium oxide nanoparticles for sizes ranging from
2 to 8 nm. Apparently, at these nanoparticle sizes, the
formation of oxygen vacancies is favored resulting in lattice expansions. In turn, the development of oxygen
vacancies can be related to the CeO2Ce2O3 reduction as
suggested by Zhou and Huebner [16].
Oxidation resistance and imperfections
Figure 4a, b is the plot of mass gain per unit area (DW/A)
versus time, t, for bare and coated 304 stainless steels
coated with nanocrystalline CeO2 after exposure to 1073
and 1273 K in dry air. The oxidation behavior was
assumed to be rate limited by diffusion and of parabolic

Fig. 4 Mass gain per unit area


(DW/A) versus time in coated
and uncoated 304 SS after
exposure to dry air at a 1073 K
and b 1273 K

123

J Mater Sci (2014) 49:277286

nature. Nevertheless, it was found that at 1073 K, there is a


significant deviation from parabolic behavior, particularly
in the bare and coated (using purchased nanoceria particles) 304 SS (see Fig. 4a). In this case, mass gain with time
is almost linear during the first 20 h. This is followed by a
second stage where the mass gain associated with scale
thickening becomes relatively slow as expected for diffusion controlled scale growth.
The anomalous oxidation behavior can be explained by
considering the kinetic aspects associated with the formation of a continuous protective scale on the substrate.
Apparently, in this steel, a continuous adherent and protective oxide layer is initially unable to develop during the
first 20 h. In turn, direct contact of the oxidant from the
atmosphere with the substrate is not avoided until a continuous scale forms. This can happen whenever the
nucleation of scale grains is hindered by either limited

J Mater Sci (2014) 49:277286

281

d(DW=A)=dt kp =2(DW=A)

(a)

0.5

(W/A)2 (mg 4/cm2)

0.4

kpu

0.3

T = 1073K
0.2

0.1

kppc

100

200

300

400

Time (h)

(b)

0.9
0.8
0.7

kpu

0.6

T = 1273 K
0.5

kppc

0.4
0.3
0.2

kpsc

0.1
0

10

20

30

Time (h)

Figure 5 shows the exhibited slopes (parabolic


constants, kp) for the various oxidation conditions used in
this work. Notice in particular that at 1073 K, there are two
slopes associated with the transition from linear to
parabolic behavior in the uncoated and coated 304 SS
using purchased nanoceria. The average experimental kp
values are given in Table 2. From this Table, it is apparent
that nanocrystalline CeO2 coatings significantly improve
the oxidation resistance of the 304 SS in agreement with
the published reports [37]. The exhibited rate constants
indicate that oxidation rates can be reduced by up to two
orders of magnitude, with the strongest effect achieved by
employing synthesized nanoceria as a coating. Moreover,
long-term exposure at 1073 K indicated that after only
168 h, serious damage through scale spallation occurred in
bare 304 SS samples (see Fig. 6b). Conversely, the coated
sample remained undamaged even after 350 h of dry air
exposure (see Fig. 6a).
Assuming that there are no appreciable differences in
the uniformity and thickness of the nanoceria surface
coatings, the exhibited variations in the experimental kp

kpsc

(W/A)2 (mg2/cm 4 )

nucleation on the substrate or when scale growth is relatively slow due to doping effects on the grain boundary
diffusivity of the metal ion species. During this lapse of
time (20 h), there is significant oxidation due to direct
contact of the dry air with the substrate.
From the experimental outcome, the benefit of
employing nanoceria coatings on providing oxidation
protection is clearly evident. Notice that the introduction of
nanoceria particles leads to minimizing or even avoiding
this anomalous oxidation behavior in the 304 steel. When
commercial nanoceria was used as a coating, coverage of
the substrate with a protective scale occurred in approximately 10 h (see Fig. 4a, b). Apparently, the presence of
nanoceria either provided additional nucleation sites for the
oxide scale and/or improved substrate protection, but it is
not effective enough to totally avoid direct contact with air
during the early stages of oxidation.
In addition, when synthesized nanoceria coatings were
implemented, it is found that a protective scale is immediately developed upon exposure to the high-temperature
(1073 K) oxidation environment. Accordingly, the positive
influence of nanoceria coatings on the protection of this
steel which exhibits an anomalous early oxidation behavior
is clearly evident.
Although there are some deviations from parabolic
behavior, particularly at 1073 K in the uncoated 304 SS
and to some extent in the one coated with purchased nanoceria, average parabolic rate constants for oxidation, kp
were estimated for comparison purposes using the
expression

Fig. 5 (DW/A)2 versus time plots showing the exhibited parabolic


constants kp at a 1073 K and b 1273 K on the 304 SS exposed to dry
air. kup = uncoated, kpc
p = coated with purchased nano-CeO2,
kpc
p = coated with synthesized nano-CeO2

Table 2 Experimental parabolic rate constants in 304 stainless steel


Temperature
(K)

Parabolic rate constant, kp (mg2 cm-4 h-1)


Uncoated

Coated with
synthesized oxide

Coated with
purchased oxide

1073

8.7 9 10-3

9.6 9 10-5

4.1 9 10-4

1273

-2

-3

9.3 9 10-3

4.0 9 10

4.4 9 10

values can be attributed to the differences in nanoparticle


sizes. From the work of Roure et al. [17], it has been found
that CeO2 nanoparticle size plays a strong role on the
exhibited protection behavior of 304 SS. In their work, they
found that when coatings are made of nanometric CeO2
particles with sizes above 10 nm, the oxidation resistance
is significantly reduced. The particle size effect on
enhancing the oxidation protection is closely related to

123

282

Fig. 6 a Synthesized nanocrystalline CeO2 coated and b uncoated


304 SS after exposure to dry air at 1273 K. Uncoated 304 SS was
exposed to oxidant for 168 h and coated 304 SS for 350 h

possible nanoceria particle dissolution [17, 18]. It can also


be influenced by the development of oxygen ion vacancies
when nanoparticles reach sizes below 8 nm [16], which are
expected to favor inward oxygen diffusion through the
nanoceria coating.
According to the literature [14, 15], nanoceria coatings
are highly effective in promoting heterogeneous nucleation
of chromium oxide grains. Therefore, a fine and uniform
distribution of nanoceria particles will favor the development of a fine-grained chromia scale resulting in full substrate coverage in relatively short times. In addition, some
nanoceria particles will necessarily have to dissolve when
in contact with the newly formed chromium oxide in order
to account for the experimental observations [16] of preferential segregation of cerium ions to the scale grain
boundaries (gbs). In turn, the presence of Ce?4 ions at gbs
has been considered as the main factor in hindering outward diffusion of the Cr?3 or Fe?3 ions. Under these
conditions, oxygen inward diffusion is expected to be the
dominant transport mechanism in the oxidation process. As
a result, the oxidation rates drop by 13 orders of magnitude depending on the steel and the environment.
Coating imperfections
Figure 6a, b shows the effect of nanocrystalline CeO2
coatings on the scale appearance of 304 SS exposed to
1273 K dry air. At this temperature, there is significant
spallation in the uncoated sample (Fig. 6b) indicating that
it does not possess good adherence to the alloy substrate.
The developed scale exhibits a coarse-grained structure and
has a strong tendency to promote cracking and spallation.

123

J Mater Sci (2014) 49:277286

In contrast, no spallation is observed in the coated sample


(see Fig. 6a). Moreover, white spots are observed on the
surface of the coated sample. The white appearance indicates an enhancement of oxidation protection as a result of
locally increased coating thicknesses at these locations,
particularly on the left bottom edge of the sample (Fig. 6a).
Although thick coatings are not desirable for protection [1],
it is not clear what the critical thickness is beyond which no
effective protection can be achieved. In this work, these
relatively thick regions resulted after the drying process of
the dipped coatings. Yet, these locally coated regions were
sound as no flaws nor spallation was found at these sites.
Figure 7a, b is SEM micrograph of the exhibited oxide
scale in the 304 SS after 10 h exposure at 1273 K. Notice
that in the bare 304 SS, the scale developed is relatively
coarse and contains numerous scale flaws (Fig. 7a). In
contrast, the scale developed in the steel coated with synthesized nanoceria is highly adherent with an extremely
fine grain size structure. In addition, EDS intensity peaks
for the scales developed in both the bare and coated 304 SS
are included in Fig. 7a, b. It can be observed that in the
bare steel, Fe is dominant indicating that haematite is the
dominant oxide (Fig. 7a). In the coated steel, the resultant
scale is chromium rich as inferred from the EDS intensity
peaks (Fig. 7b). Ceria is also found to be present on the
scale surface supporting the idea of nanoceria particles
acting as inert markers [16].
A closer look to the black oxide regions of the coated
304 SS using the SEM indicated the development of
isolated hills where rapid oxide growth occurs (see
Fig. 8). EDX analyses show that the developed oxide scale
on these hills is iron rich, most likely haematite. A
comparison of the oxide morphology with that found in the
oxidized bare 304 SS indicates that both scales are rather
similar. In contrast, between hills, chromia seems to be
the dominant oxide. Apparently, some regions on the surface of the substrate are not fully covered with nanoceria
by the dipping process to provide a protective coating.
Hence, upon exposure to hot dry air, an oxide scale similar
to that found in the uncoated 304 SS is developed (see
Fig. 7a).
In contrast, the white regions on the scale surface (see
Fig. 9) indicate that the dominant oxide scale is Cr-rich
suggesting the development of a fine-grained chromia
scale. Moreover, the scale developed is extremely fine with
no evidence of any disruption of the original coating as a
result of oxidation. This is confirmed by EDX analyses of
the oxidized surface which show high intensity Ce peaks
beyond the ones found at other locations (see Figs. 7, 8).
Accordingly, it is evident that the development of a uniform coating thickness by the dipping method does not
necessarily occur in the 304 SS as suggested in the literature [4].

J Mater Sci (2014) 49:277286

283

Fig. 7 SEM micrographs


including EDX spectra of 304
SS oxidized at 1273 K for 10 h
in dry air. a Uncoated steel and
b CeO2-coated steel

Scale development
From the published literature [13], it is evident that oxide
scales formed on nano-CeO2-coated 304 SS exhibit a
superior resistance to cracking and spallation. Yet, the
mechanisms involved are not fully understood. Among the
potential mechanisms are (a) the development of a finegrained oxide scale, which creeps at relatively fast rates at
high temperatures, thus releasing any stress built-ups on the
scale in agreement with the observations of Fig. 3b, or
(b) inward oxygen diffusion along gbs prevents excess

cation vacancy condensation in the form of pores [1]. EDS


analyses on the scale surface of the 304 SS oxidized at
1273 K for 10 h in dry air indicated that in the bare steel,
the top of the scale was Fe-rich, whereas in the CeO2coated steel, it was Ce-rich (see Fig. 3). In turn, this indicates that in the absence of a protective CeO2 coating,
outward Fe cation diffusion becomes dominant during
scale growth.
In contrast, the scale surface of the coated sample is Crrich suggesting that outward Fe cation diffusion along the
gbs is drastically hindered in the presence of

123

284

J Mater Sci (2014) 49:277286

Fig. 8 SEM micrograph of a dip-coated 304 SS including EDS intensity peaks showing the development of islands where rapid oxidation
occurs. The EDS peaks indicate that the islands are Fe-rich, whereas the regions surrounding the islands Cr-rich

Fig. 9 a SEM micrograph of a dip-coated 304 SS showing a white


region in the 304 steel and b EDS intensity peaks indicating that the
scale in the white region is Cr- and Fe-rich and that the Ce peaks
are relatively intense

123

nanocrystalline CeO2. In turn, this can account for the


exhibited reduction in the scale growth rates. When CeO2
nanoparticles cover the FeCr surfaces, they provide extra
nucleation sites for the formation of Cr2O3 resulting in an
immediate and complete surface coverage by Cr2O3 [17,
18]. In general, nano-sized particles are increasingly
unstable when compared to micro-size particles due to their
increasing high surface-to-volume ratios which seem to
favor their dissolution [19].
It is well known that scale growth in high-temperature
alloys at temperatures below 1273 K is limited to mass
transport along short-circuit diffusion paths. Hence,
blocking of these paths by Ce?4 ions has been suggested to
account for the reduced oxidation rates in nanoceria-coated
stainless steels [46]. However, for Ce?4 ions to effectively
saturate the chromia gbs, dissolution of nanoceria particles
is a necessary condition. Determinations using SIMS have
shown that Ce?4 ions are present in NiO, both in the bulk
as well as at the gbs [1, 17]. Similar results have been
observed in chromia scales [2, 18].
It is well known that during the oxidation process, oxide
scales grow by cation outward diffusion. Inward lattice
diffusion of oxygen anions is too slow to be dominant
during scale growth, even though gb inward oxygen diffusion can always be active. Hence, when a fine-grained
oxide scale forms as a result of the nano-CeO2 coating, the
diffusional mechanisms for mass transport across the scale
shift from Fe or Cr outward diffusion to O inward diffusion. In the coatings containing nano-sized CeO2 particles,

J Mater Sci (2014) 49:277286

a scale with an extremely large density of gbs is expected


to form which can act as high-diffusivity paths for inward
oxygen diffusion. This as long as sufficient Ce ions are
segregated at the gbs to hinder cation vacancy diffusion by
either (a) reducing the number of vacancies or (b) by pair
defect association between Ce ions and cation vacancies
[13].
Alternatively, a shift in the diffusional mechanisms can
be explained through an increase in the activation energies
for cation diffusion [20]. A reduction in oxidation rates is
expected whenever the activation energies for either iron or
chromium ion diffusion through the scale thickness along
the gbs is higher than that for oxygen inward diffusion. In
turn, this would explain the reduced high-temperature
oxidation rates found in this work. Zhang [8] simulated the
changes in activation energies using molecular dynamics.
In her work, she considered the role of nanoceria as a
dopant in the chromium oxide scale. In this case, the presence of Ce ions in the chromium oxide scale resulted in an

285

increase in activation energies for Fe diffusion beyond the


one for oxygen inward diffusion. From these results, it is
expected that oxygen inward diffusion becomes the dominant mass transport mechanism during scale growth, thus
accounting for the exhibited reduction in oxidation rates in
the nanoceria-coated 304 SS.
Further support for the development of a fine-grained
structure is given by the developed scale cross sections.
Figure 10a, b is SEM micrograph of the scale cross sections for both bare and nanocrystalline CeO2-coated 304
samples oxidized at 1073 K for 442 h in dry air. Notice
that the scale thickness of the coated steel is less than
5 lm, whereas in the uncoated sample, it is approximately
15 lm (b). Hence, it is evident that nanocrystalline CeO2
coatings significantly hinder the scale growth rates. In the
uncoated sample, voids and cracks are observed, but they
do not provide a conclusive evidence for the adherence of
the scale as these defects can be induced through sample
preparation. Nevertheless, a comparison of both scale cross
sections indicates that the thick scale develops a relatively
coarse grain structure during oxidation at this temperature.

Conclusions
1.

2.

3.

4.

5.

6.

7.
Fig. 10 SEM micrographs of the scale cross sections developed on
a nanoceria dip-coated 304 SS and b uncoated 304 SS oxidized at
1073 K for 442 h

Nanocrystalline cerium oxide particles with average


particle sizes of 3.45 nm were produced by the microemulsion method.
Characterization of the synthesized particles by XPS
means indicated that both Ce3? and Ce4? were present
and that the average amount of Ce3? ions was
22.8 5 %.
The high-temperature oxidation resistance of the 304
SS was found to improve by up to two orders of
magnitude when nanoceria coatings with mean particle
sizes of 3.45 nm were implemented.
Parabolic rate constants kp were found to be reduced
from 8.7 9 10-3 mg2 cm-4 h-1 (uncoated sample) to
9.6 9 10-5 mg2 cm-4 h-1 (synthesized ceria-coated
samples) at 1073 K.
In the synthesized ceria-coated 304 SS, the scale
thickness was barely 5 lm when oxidized at 1073 K
for 442 h. In contrast, in the uncoated steel under
similar exposure conditions, the scale thickness was
15 lm.
In the coated steels, the developed scale was highly
adherent, and no spalling was evident. This was not the
case in the uncoated steel which exhibited significant
spalling and scale detachment.
The method of coating by dipping did not provide a
uniform coverage of nanoceria particles on the 304 SS.
Fe-rich islands were found to develop during hightemperature oxidation which indicated that some

123

286

J Mater Sci (2014) 49:277286

substrate regions are not fully protected due to the


inhomogeneous thickness of the protective coating. In
contrast, relatively thick-coating regions on the steel
substrate exhibited minimal oxidation.

References
1. Moon DP, Bennett MJ (1989) Mater Sci Forum 43:269
2. Polman EA, Fransen T, Gellings PJ (1989) J Phys Condens
Matter 1:4497
3. Stringer J (1989) Mater Sci Eng A 120:129
4. Patil S, Kuiry SC, Seal S, Vanfleet R (2002) J Nanopart Res
4:433
5. Patil S, Kuiry SC, Seal S (2004) Proc R Soc Lond Ser A 460:3569
6. Thanneeru R, Patil S, Deshpande S, Seal S (2007) Acta Mater
55:3457
7. Seal S, Roy SK, Bose SK, Kuiry SC (2000) JOM-e 52, http://
www.tms.org/pubs/jurnals/JOM/0001/Seal/Seal-0001.html

123

8. Zhang H (2007) Role of nanocrystalline cerium oxide coatings on


austenitic stainless steels. University of Wisconsin-Milwaukee,
Milwaukee
9. Lopez HF (2009) High-temperature oxidation resistant nanocoatings on austenitic stainless steels. In: MRS conference proceedings, symposium 5 advanced materials, XVIII international
congress on Mats. Res., August 1620, MRS, Cancun, 1243:21
10. Morris VN, Farrell RA, Sexton AM, Morris MA (2006) J Phys
Conf Ser 26:119
11. Bouchaud B, Balmain J, Bonnet G, Pedraza F (2013) Appl Surf
Sci 268:218
12. Czerwinski F, Szpunar JA (1996) Thin Solid Films 289:213
13. Wu Z, Gu L, Li H, Benfield RE, Yang Q, Grandjean D, Li Q, Zhu
H (2001) J Phys Condens Matter 13:52695284
14. Deshpande S, Patil S, Kuchibhatla S VNT, Seal S (2005) Appl
Phys Lett 87:133113
15. Tsunekawa SK, Ishikawa K, Li ZQ, Kawazoe Y, Kasuya A
(2000) Phys Rev Lett 85:3440
16. Zhou XD, Huebner W (2001) Appl Phys Lett 79:3512
17. Roure S, Czerwinski F, Petric A (1994) Oxid Met 42:75
18. Czerwinski F, Szpunar JA (1997) J SolGel Sci Technol 9:103
19. Campbell CT, Parker SC, Starr DR (2002) Science 298:811
20. Duffy DM, Tasker PW (1986) Phil Mag 54:759

You might also like