You are on page 1of 5

Materials Characterization 100 (2015) 3135

Contents lists available at ScienceDirect

Materials Characterization
journal homepage: www.elsevier.com/locate/matchar

TEM and XPS studies on the faceted nanocrystals of Ce0.8Zr0.2O2


Debadutta Prusty a, Abhishek Pathak a, Manabendra Mukherjee b,
Bratindranath Mukherjee c, Anirban Chowdhury a,
a
b
c

Research & Development, Tata Steel Limited, Jamshedpur 831001, India


Saha Institute of Nuclear Physics, Kolkata 700064, India
Department of Metallurgical Engineering, Indian Institute of Technology, Banaras Hindu University, Varanasi, 221005, India

a r t i c l e

i n f o

Article history:
Received 23 September 2014
Received in revised form 8 December 2014
Accepted 9 December 2014
Available online 11 December 2014
Keywords:
Electron microscopy
X-ray diffraction
Defects
Chemical synthesis
Differential scanning calorimetry (DSC)

a b s t r a c t
Faceted nanocrystals of Ce0.8Zr0.2O2 synthesised by co-precipitation method were characterised by X-ray diffraction, high-resolution transmission electron microscopy, thermogravimetrydifferential scanning calorimetry and
X-ray photoelectron spectroscopy techniques. The nanocrystals were highly faceted and exhibited a cubic phase.
X-ray photoelectron spectroscopy analyses conrmed the presence of vacancy related defects and revealed the
presence of ~22% of Ce3+ in the nanopowders. High-resolution transmission electron microscopy results conrmed that the nanocrystal sizes are around 31 5 nm and the obtained hexagonal cross-section shape is
bound by hexagonal {111} and square {100} facets. The shape-controlled nanocrystals were synthesised without
using any surfactants or complexing agents and retained their morphology beyond 800 C. This is a simple and
easy method for producing shape-controlled Ce0.8Zr0.2O2 nanoparticles which can be used for catalytic conversion and other related advanced technological areas.
2014 Published by Elsevier Inc.

1. Introduction
Zirconia doped ceria has emerged as the most effective candidate for
catalytic and other technological applications [16]. While the introduction of Zr4+ in small amounts does not alter the uorite crystal structure
of the parent phase, the smaller radius of Zr4+ cation (0.084 nm) compared to that of Ce4+ (0.097 nm) [7] cation results in the distortion of
the lattice. This decreases the activation energy for oxygen ion diffusion,
consequently enhancing its performance in reduction [8]. In fact, a twoorder increase in bulk diffusion coefcient of O2 ions in ceriazirconia
solid solution than in pure ceria has been reported [9]. It also enhances
the formation of defects that plays a crucial role in redox behaviour and
improves the resistance towards ageing at high temperature. Wang
et al. [10] observed that while the catalytic performance of CuO/CeO2
is better than that of CuO/Ce0.8Zr0.2O2 at lower temperatures, above
600 C, the reverse behaviour is seen owing to higher tendency of
CeO2 towards sintering compared to Ce0.8Zr0.2O2. Similar trend has
been reported by other researchers [11].
The traditional method of producing a complex oxide is solid state
reaction. In this process, individual metal oxides, CeO2 and ZrO2 in the
present case, are manually mixed together, ground and pelletized. The
pellet is heated at a high temperature to facilitate solid state diffusion.
Yet, this process suffers from several drawbacks like loss in surface
area due to prolonged exposure to high temperature, compositional
Corresponding author at: Materials Science and Engineering, Indian Institute of
Technology, Patna, Bihar 800013, India.
E-mail address: anirban.chowdhury@gmail.com (A. Chowdhury).

http://dx.doi.org/10.1016/j.matchar.2014.12.009
1044-5803/ 2014 Published by Elsevier Inc.

inhomogeneity and sometimes incomplete reaction if the diffusion of


ions in the product layer is slow [12]. Additionally, the particles produced in this process are irregular in shape. In general, to produce
shaped crystals, various solution controlled approaches have been
adopted in which the crystal shape can be manipulated by adding surfactants. Surfactants behave as capping agents, adsorb on specic planes
depending upon their charge density, atom density and dipolar moment, and affect the relative stability and growth of these planes,
resulting in faceted nano-crystals. For example, in ceria system,
tadpole-shaped ceria nanoparticles, consisting of spherical heads and
wire shaped tails, were obtained by adding excess oleic acid in thermal
decomposition of cerium nitrate hexahydrate [13]. Wu et al. [14] prepared single crystalline CeO2 rods, cubes and octahedra by hydrothermal process.
Another requirement for achieving high catalytic activity is that
these faceted crystals retain their shape at higher temperature. Often
on heating at higher temperature, the nanoparticles become of irregular
shape and the advantages of facets cannot be utilised. Sarada et al. [15]
synthesised Ce0.8Zr0.2O2 by co-precipitation method and irregular
shaped nanoparticles were obtained upon calcination at 500 C. Li
et al. [16] prepared CeO2 and Ce0.8Zr0.2O2 by solgel method followed
by calcination at 1000 C and the particles were found to be irregular
and agglomerated. Wu et al. [14] observed that CeO2 nano-rods and
cubes undergo morphological changes above 500 C.
In this work, we report the synthesis of faceted Ce0.8Zr0.2O2
nanocrystals through a simple co-precipitation route without using
any complexing agent or surfactants. The Ce0.8Zr0.2O2 nanoparticles
are of cubic phase and retain the morphology even after being calcined

32

D. Prusty et al. / Materials Characterization 100 (2015) 3135

beyond 800 C. To the best of our knowledge, this is the rst report
for achieving faceted nanocrystals in ceriazirconia system by coprecipitation route without using any surfactant. These nanoparticles
were characterised by X-ray diffraction (XRD), high-resolution transmission electron microscopy (HRTEM), thermogravimetrydifferential
scanning calorimetry (TG-DSC) and X-ray photoelectron spectroscopy
(XPS) techniques. Microstructure and surface properties of the nanoparticles have been discussed along with the defects present in the parent lattice.
2. Experimental
For the synthesis of Ce0.8Zr0.2O2, cerium nitrate (Ce(NO3)36H2O,
99% pure, Loba Chemie) and zirconyl oxychloride (ZrOCl28H2O, 98%
pure, Loba Chemie) were used as precursors. The two salts in the
molar ratio of Ce3+:Zr4+ = 4:1 were dissolved in distilled water and
aqueous ammonia was added drop wise to the mixture under constant
magnetic stirring. The precipitates were collected at pHs 3, 5, 7, and 9.
Subsequently, they were washed with deionised water followed by
four cycles of centrifugation at a speed of 10,000 rpm. The resultant
products were dried in an oven at 110 C for 8 h and crushed in a mortar
to break the agglomerates.
Thermal analysis of the dried powders was performed in air using a
differential thermal calorimeter (DSC) (STA 449 F3 Jupiter NETZSCH)
from room temperature to 1500 C at a heating rate of 10 C/min.
Considering the thermal analysis results, the samples were calcined at
750 C. In order to study the crystallinity of the samples, continuous
scan was performed on the calcined powder samples over the range
of 20100 2 using a diffractometer (PANalytical X'Pert Pro) with Cu
K radiation. The patterns were analysed on PANalytical X'Pert High
Score software.
On thorough analysis of the XRD patterns (not shown here), we
found pH 9 as to be the most suitable pH to obtain stoichiometric
Ce0.8Zr0.2O2 as at lower pHs all cations could not be precipitated.
Hence, we perform all the characterisation on this sample.
The sample synthesised at pH 9 was calcined at 750, 850, 1000, 1200
and 1400 C for 2 h. The morphology of the sample was studied by a
transmission electron microscope (TEM, Tecnai 320 W-lament) operated at 200 kV. A small amount of the powder was dispersed in ethanol
and a small amount of the resultant suspension was dropped on a carbon coated copper grid to prepare the specimen for TEM.
XPS core-level spectra were taken with an Omicron Multiprobe
(Omicron Nano Technology GmbH., UK) spectrometer tted with an
EA125 hemispherical analyser. A monochromated AlK X-ray source
operated at 150 W was used for the experiments. The analyser pass energy was kept xed at 40 eV for all the scans. For quantitative analysis,
background of the data was removed using Shirley method and FHI sensitivity factors were used for the calculations. For insulating samples, as
is the case here, the probing depth of XPS is around 10 nm. Therefore, it
is expected that we probe about two-thirds of the volume of an average
sized (31 5 nm) nanoparticle.

Fig. 1. TGA plot of the precipitate dried at 110 C.

reveal that cubic ceria can accommodate Zr4+ ions in its lattice up to
25%. The patterns are indexed as per the standard pattern of JCPDS le
28-0271. The average crystallite sizes of the sample at 750 and 850 C
from the Scherrer method were calculated to be 16 and 17 nm, respectively. A small peak at 29.86 2 starts emerging as a shoulder from the
(111) peak after 850 C and separates at higher temperature. This is attributed to a tetragonal phase rich in zirconia. However, it is difcult to
assign the exact phase as the XRD patterns of the phases (JCPDS 381437 for tetragonal Ce0.16Zr0.84O2, 42-1164 for tetragonal ZrO2, 01080-0785 for tetragonal Ce0.18Zr0.82O2) that match the experimental
pattern are very similar. Similar behaviour has also been reported in
previous works [18,19]. Chuang et al. [19] prepared CexZr1 xO2
(0.16 b x b 1) by co-precipitation method and observed that upon ageing at 1200 C, all the solid solutions having CeO2 content between 16
and 83% decompose into two separate phases; one rich in zirconium
and the other in cerium. Also, there are two smaller extra peaks at
25.86 and 42.98 2; their evidence can also be traced in the 750 C pattern. However, we are not sure about their origin.
Fig. 3(a) depicts the Ce 3d XPS spectra of Ce0.8Zr0.2O2 calcined at
850 C. The Ce 3d spectrum comprises two multiplets (labelled as u
and v) arising from spin orbit splitting of 3d3/2 and 3d5/2 states. For
the Ce4+ state, each spin-orbit component is composed of two sets of
three features (u + u + u, v + v + v), while two sets of two features (u0 + u, v0 + v) are found in each spin-orbit component for
Ce3+. The appearance of four pairs of doublets in our result conrms
the presence of cerium in both oxidation states on the surface.

3. Results and discussions


The typical TGA plot carried out in air of the precipitate (Fig. 1) illustrates four distinct steps of mass loss up to a temperature of 440 C. The
rst step is attributed to the evaporation of molecular water physically
adsorbed on the surface and/or trapped within the pores. The remaining
steps relate to the removal of water of crystallisation and decomposition
of metal hydroxides. The mass loss becomes negligible after 440 C, implying a complete decomposition of the precipitates.
The XRD patterns of the sample calcined at 750, 850, 1000, 1200 and
1400 C are shown in Fig. 2. The sample is found to be in fully crystalline
state at 750 C; corroborating well with the DSC results. All the samples
exist in cubic uorite structure with a lattice parameter of 5.374
0.006 . This is in well accord with the earlier studies [10,17] which

Fig. 2. XRD patterns of the Ce0.8Zr0.2O2 sample calcined at different temperatures. The *
sign corresponds to extra peaks in the sample. The patterns have been indexed as per
the standard pattern of JCPDS le 28-271.

D. Prusty et al. / Materials Characterization 100 (2015) 3135

Fig. 3. (a) XPS spectra of Ce 3d region of Ce0.8Zr0.2O2 calcined at 850 C in air. The individual peaks are shown by deconvoluted curves. (b) XPS spectra of Zr 3d region of Ce0.8Zr0.2O2
calcined in air at 850 C Fig. (c) O1s spectra of Ce0.8Zr0.2O2 calcined sample. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.)

Two approaches have primarily been used for quantifying the


amount of a particular oxidation state: the factor group analysis [20,
21] and the deconvolution of the spectrum into several peaks [22,23].
In our case, the latter is used. The fraction of Ce in 3 + state is determined from the ratio (Iu + Iv)/(Iu + Iv + Iu + Iv) according to a previous report [24]. Calculated concentrations of Ce3+ and Ce4+ are 22%
and 78%, respectively. This is extremely useful data for the application
of these powders in catalytic applications.
The existence of Ce3+ in ceriazirconia has also been observed by
many researchers. Reddy et al. [25] observed 15% of Ce3 + in

33

Ce0.75Zr0.25O2 synthesised by co-precipitation method and 23% in


Ce0.5Zr0.5O2 by a microwave combustion method and attributed this to
the incorporation of more Zr4+ ions. A Ce3 +/(Ce3 + + Ce4+) ratio as
high as 0.32 was found in the as-synthesised Ce0.2Zr0.8O2 prepared by
urea based hydrothermal method and decreased to 0.20 on calcination
at 500 C for 4 h [22]. Sarada et al. [26] prepared Ce0.8Zr0.2O2 by coprecipitating cerium nitrate and zirconyl nitrate using NaOH and
found 33% Ce3+ in the sample calcined at 500 C. Liu et al. [17] synthesised CeO2 (111) and Ce0.8Zr0.2O2 (111) epitaxial thin lms on yttria
stabilised zirconia (YSZ) surface by oxygen plasma assisted molecular
beam epitaxy and found Ce3 + ions to be stable in Ce0.8Zr0.2O2 even
after prolonged annealing in oxygen atmosphere while no features of
Ce3+ ions were revealed in oxygen treated CeO2.
Now, we discuss the possible causes of existence of Ce3+ in our sample. The strongest reason so far, as proposed in previous similar works,
has been existence of oxygen vacancies on the surface as a result of
which some Ce4+ ions undergo reduction to maintain charge neutrality.
By performing neutron diffraction, Mamintov and Egami [27] demonstrated existence of Frenkel-type anion defects in uorite structures
such as ceria where some of the oxygen ions in the tetrahedral sites migrate to octahedral voids, creating some vacancies in tetrahedral sites.
However, the concentrations of interstitial ions and oxygen vacancies
are same within the limits of error and annihilation of the above pairs
occurs on heating at higher temperature by recombination of vacancies
and interstitials. On the contrary, in ceriazirconia (containing 31% Zr)
prepared by the same processing method, a vacancy concentration of
~ 17% and an interstitial oxygen concentration of 8% were obtained
[28]. Therefore, apart from vacancies originating due to interstitial
oxygen ions, there are excess intrinsic vacancies existing in ceriazirconia. It has been suggested that introduction of Zr4+ ions facilitates reduction as its smaller size reduces the lattice strain associated with
Ce4 + Ce3 + due to an increase of ionic size [29], thus resulting in
more Ce3+ ions.
In Fig. 3(b), the blue line shows the core level spectrum of Zr 3d at
binding energies, 182.6 eV for Zr 3d5/2 and 184.7 eV for Zr 3d3/2. These
peak values have been reported for a typical characteristic of the Zr4+
in ZrO2 [30,31]. The Zr 3d5/2 peak at 181.4 eV in the spectra marked by
red line in Fig. 3(b) can be assigned to the non-stoichiometric oxides
of zirconium as observed by others [32,33].
The O1s spectrum is presented in Fig. 3(c). The broad and asymmetric O1s spectrum indicates that there are several oxygen containing
species present on the surface. After deconvolution, three distinct
peaks were observed at 528.9, 530.3 and 532.2 eV. According to
published data on X-ray photoelectron spectroscopy [34], the binding
energy of oxygen in CeO2 is 529.2 eV. Therefore we assign the O1s
peak at 528.9 eV to oxygen bound to Ce4 + sites. According to the
reference data [34], peak position of 530.3 eV corresponds to O1s
in Ce2O3, Therefore we assign this peak to oxygen bound to Ce3+ sites.
It may be noted that binding energy for O1s in ZrO2 also lie in a
range of 530.2 to 530.9 eV [34,35]. Since Ce and Zr both are present in
our sample and the observed peak at 530.3 eV is relatively larger and
broader compared to those reported in the literature [36,37], we believe
that contribution from ZrO2 may be present here. The peak at 532.2 eV is
attributed to the adsorbed oxygen [10]/carbon tape support to the
sample.
Fig. 4 shows the TEM micrograph of Ce0.8Zr0.2O2 sample calcined at
850 C. The corresponding particle size distribution is depicted as a histogram in the inset plot. The average particle size was calculated to be
31 5 nm. We observe that most of the particles are of hexagonal
shape with well-dened sharp corners.
Fig. 5a depicts the selected area diffraction pattern of the sample calcined at 850 C. The pattern consists of well-dened rings, which indicates the sample retains its nano-crystalline nature even at such a
high temperature. The average lattice parameter calculated using the
radii of the rings is 5.493 0.071 . The rst four planes have been
indexed on the basis of face-centred cubic phase of Ce0.8Zr0.2O2.

34

D. Prusty et al. / Materials Characterization 100 (2015) 3135

Fig. 4. TEM micrograph of Ce0.8Zr0.2O2 nanoparticles calcined at 850 C. The inset shows
the corresponding particle size distribution.

The HRTEM image of a nanocrystal of Ce0.8Zr0.2O2 calcined at 850 C


is presented in Fig. 5b. The interplanar spacings calculated from the lattice fringes are in well in agreement with the d-spacings reported in
JCPDS le no. 28-0271. The gure depicts well-dened edges with
sharp corners for the Ce0.8Zr0.2O2 nanocrystals. The planes representing
the faces have been identied from their lattice parameters and angles
between them. The obtained hexagonal cross-section is often observed
in truncated octahedron consisting of eight hexagonal {111} facets and
six square {100} facets [38]. However, conventional TEM does not
guarantee about the 3D nanocrystal morphology. To reveal the actual
morphology techniques such as electron tomography are usually
employed and further works will be carried out in future to conrm
the same.
Ceria particles of different shapes such as nano-rod, nano-cubes exposing more reactive planes such as {110} and {100} planes have previously been synthesised [14]; however the requirement of addition of
surfactants and a strict process control makes them less preferable for
large scale production. In fact, ceria nanoparticles synthesised by
shape controlled methods are not thermodynamically stable and
become of irregular shape upon annealing at elevated temperatures
[14]. Sarada et al. [26] prepared Ce0.8Zr0.2O2 nanoparticles using coprecipitation method and obtained irregular shaped particles on calcination at a temperature as low as 500 C. However, in our case, the particles retain their shape even after annealing at 850 C.
As a general discussion, an important synthesis aspect of coprecipitation method needs to be highlighted also in this case. The coprecipitation process moves through a complex sequence of reactions
nally leading to a formulation of metal hydroxide compounds. Therefore, the hydroxide forming ability of the metal cation is crucial in this
regard. It has also been found that their reaction kinetics is strongly dependent on the process parameters (pH, temperature etc.), which will
also nally inuence the rate of completion of such reactions. This is
why it is observed that some constituent precursor retains their identity
by not taking part in the reaction. Previously, we have observed that for
the case of lanthanum zirconate made by the same method and an additional step of ultrasonication was needed to control the stoichiometry
of the system [39]. The main cause for this was correlated to the slow reactivity of lanthanum precursors, a fraction of which did not participate
in the co-precipitation reaction. In the present case, no such
ultrasonication step was required. Both cerium and zirconium are

Fig. 5. (a) Selected area diffraction pattern of the sample calcined at 850 C and (b) HRTEM
image of the Ce0.8Zr0.2O2 nanocrystal depicting 2D hexagonal morphology.

excellent hydroxide formers and hence the desired reaction product


was suitably achieved without any additional ultrasonication step.
4. Conclusion
Microstructure and surface properties of the Ce0.8Zr0.2O2 nanocrystals were characterised by HRTEM, XPS and XRD techniques. The
well-developed faceted Ce0.8Zr0.2O2 nanocrystals are about 2530 nm
even after carrying out calcinations at 850 C. This is an important result
since the synthesis process does not use any surfactant to control the
shape and size of the Ce0.8Zr0.2O2 crystals. XPS analyses revealed the
presence of 22% of Ce3+ in the nanopowders, an extremely useful fact
for catalytic applications. The underlying reason has been correlated
with the presence of oxygen vacancies on the surface. The results also
indicate the importance of selecting precursors for co-precipitation
and their propensity of hydroxide formation for obtaining faceted
nanocrystals without using any surfactants. This economical method

D. Prusty et al. / Materials Characterization 100 (2015) 3135

of synthesising Ce0.8Zr0.2O2 nanopowders at multigram scale can be easily scaled up further at pilot scale for catalytic and other allied technological applications.
References
[1] G. Balducci, P. Fornasiero, R. Di Monte, J. Kaspar, S. Meriani, M. Graziani, An unusual
promotion of the redox behaviour of CeO2ZrO2 solid solutions upon sintering at
high temperatures, Catal. Lett. 33 (12) (1995) 193200.
[2] G. Balducci, J. Kapar, P. Fornasiero, M. Graziani, M.S. Islam, Surface and reduction
energetics of the CeO2ZrO2 catalysts, J. Phys. Chem. B 102 (3) (1998) 557561.
[3] P. Fornasiero, R. Dimonte, G.R. Rao, J. Kaspar, S. Meriani, A. Trovarelli, M. Graziani,
Rh-loaded CeO2ZrO2 solid-solutions as highly efcient oxygen exchangers: dependence of the reduction behavior and the oxygen storage capacity on the structuralproperties, J. Catal. 151 (1) (1995) 168177.
[4] K. Tomishige, Y. Furusawa, Y. Ikeda, M. Asadullah, K. Fujimoto, CeO2ZrO2 solid solution catalyst for selective synthesis of dimethyl carbonate from methanol and carbon dioxide, Catal. Lett. 76 (12) (2001) 7174.
[5] K. Tsukuma, M. Shimada, Strength, fracture toughness and Vickers hardness of
CeO2-stabilized tetragonal ZrO2 polycrystals (CeTZP), J. Mater. Sci. 20 (4) (1985)
11781184.
[6] R. Di Monte, J. Kapar, Nanostructured CeO2ZrO2 mixed oxides, J. Mater. Chem. 15
(6) (2005) 633648.
[7] R. Shannon, C. Prewitt, Effective ionic radii in oxides and uorides, Acta Crystallogr.
B 25 (5) (1969) 925946.
[8] J. Yu, L. Zhang, J. Lin, Direct sonochemical preparation of high-surface-area
nanoporous ceria and ceriazirconia solid solutions, J. Colloid Interface Sci. 260
(1) (2003) 240243.
[9] M. Boaro, C. de Leitenburg, G. Dolcetti, A. Trovarelli, The dynamics of oxygen storage
in ceriazirconia model catalysts measured by CO oxidation under stationary and
cycling feedstream compositions, J. Catal. 193 (2) (2000) 338347.
[10] S. Wang, X. Zheng, X. Wang, S. Wang, S. Zhang, L. Yu, W. Huang, S. Wu, Comparison
of CuO/Ce0.8Zr0.2O2 and CuO/CeO2 catalysts for low-temperature CO oxidation, Catal.
Lett. 105 (3) (2005) 163168.
[11] S. Wang, X. Wang, J. Huang, S. Zhang, S. Wang, S. Wu, The catalytic activity for CO
oxidation of CuO supported on Ce0.8Zr0.2O2 prepared via citrate method, Catal.
Commun. 8 (3) (2007) 231236.
[12] A. West, Solid State Chemistry and its Applications, Wiley, 2007.
[13] T. Yu, J. Joo, Y.I. Park, T. Hyeon, Large scale nonhydrolytic solgel synthesis of uniform sized ceria nanocrystals with spherical, wire, and tadpole shapes, Angew.
Chem. 117 (45) (2005) 75777580.
[14] Z. Wu, M. Li, J. Howe, H. Meyer III, S. Overbury, Probing defect sites on CeO2
nanocrystals with well-dened surface planes by Raman spectroscopy and O2 adsorption, Langmuir 26 (21) (2010) 1659516606.
[15] B.Y. Sarada, K. Dhathathreyan, M. Rama Krishna, Meliorated oxygen reduction reaction of polymer electrolyte membrane fuel cell in the presence of ceriumzirconium
oxide, Int. J. Hydrogen Energy 36 (18) (2011) 1188611894.
[16] L. Li, F. Chen, J. Lu, M. Luo, Study of defect sites in Ce1 xMxO2 (x = 0.2) solid
solutions using Raman spectroscopy, J. Phys. Chem. A 115 (27) (2011) 79727977.
[17] G. Liu, J. Rodriguez, J. Hrbek, J. Dvorak, C. Peden, Electronic and chemical properties
of Ce0.8Zr0.2O2 (111) surfaces: photoemission, XANES, density-functional, and NO2
adsorption studies, J. Phys. Chem. B 105 (32) (2001) 77627770.
[18] C. Bozo, F. Gaillard, N. Guilhaume, Characterisation of ceriazirconia solid solutions
after hydrothermal ageing, Appl. Catal. A 220 (1) (2001) 6977.

35

[19] C. Chuang, H. Hsiang, F. Yen, C. Chen, S. Yang, Phase evolution and reduction behavior of Ce0.6Zr0.4O2 powders prepared using the chemical co-precipitation method,
Ceram. Int. 39 (2) (2013) 17171722.
[20] A. Gonzalez-Elipe, J. Holgado, R. Alvarez, G. Munuera, Use of factor analysis and XPS
to study defective nickel oxide, J. Phys. Chem. 96 (7) (1992) 30803086.
[21] M. Trudeau, A. Tschpe, J. Ying, XPS investigation of surface oxidation and reduction
in nanocrystalline CexLa1 xO2 y, Surf. Interface Anal. 23 (4) (1995) 219226.
[22] R. Si, Y. Zhang, C. Xiao, S. Li, B. Lin, Y. Kou, C. Yan, Non-template hydrothermal route
derived mesoporous Ce0.2Zr0.8O2 nanosized powders with blue-shifted UV absorption and high CO conversion activity, Phys. Chem. Chem. Phys. 6 (5) (2004)
10561063.
[23] E. Beche, G. Peraudeau, V. Flaud, D. Perarnau, An XPS investigation of (La2O3)1 x
(CeO2)2x(ZrO2)2 compounds, Surf. Interface Anal. 44 (8) (2012) 10451050.
[24] A. Nelson, K. Schulz, Surface chemistry and microstructural analysis of CexZr1 xO2 y
model catalyst surfaces, Appl. Surf. Sci. 210 (3) (2003) 206221.
[25] B. Reddy, G. Reddy, L. Reddy, I. Ganesh, Synthesis of nanosized ceriazirconia solid
solutions by a rapid microwave-assisted combustion method, Open Phys. Chem. J.
3 (2009) 2429.
[26] B. Sarada, K. Dhathathreyan, M. Rama Krishna, Meliorated oxygen reduction reaction of polymer electrolyte membrane fuel cell in the presence of ceriumzirconium
oxide, Int. J. Hydrogen Energy 36 (18) (2011) 1188611894.
[27] E. Mamontov, T. Egami, Structural defects in a nano-scale powder of CeO2 studied by
pulsed neutron diffraction, J. Phys. Chem. Solids 61 (8) (2000) 13451356.
[28] E. Mamontov, T. Egami, R. Brezny, M. Koranne, S. Tyagi, Lattice defects and oxygen
storage capacity of nanocrystalline ceria and ceriazirconia, J. Phys. Chem. B 104
(47) (2000) 1111011116.
[29] G. Balducci, J. Kapar, P. Fornasiero, M. Graziani, M. Islam, J. Gale, Computer simulation studies of bulk reduction and oxygen migration in CeO2ZrO2 solid solutions, J.
Phys. Chem. B 101 (10) (1997) 17501753.
[30] M. Benjaram, B. Chowdhury, G. Panagiotis, An XPS study of the dispersion of MoO3
on TiO2ZrO2, TiO2SiO2, TiO2Al2O3, SiO2ZrO2, and SiO2TiO2ZrO2 mixed oxides,
Appl. Catal. A 211 (2001) 1930.
[31] S. Nandi, S. Chakraborty, M. Bera, C. Maiti, Structural and optical properties of ZnO
lms grown on silicon and their applications in MOS devices in conjunction with
ZrO2 as a gate dielectric, Bull. Mater. Sci. 30 (3) (2007) 247254.
[32] L. Kumar, D. Sarma, S. Krummacher, XPS study of the room temperature surface oxidation of zirconium and its binary alloys with tin, chromium and iron, Appl. Surf.
Sci. 32 (3) (1988) 309319.
[33] J. Anderson, J. Fierro, Bulk and surface properties of copper-containing oxides of the
general formula LaZr1 xCuxO3, J. Solid State Chem. 108 (2) (1994) 305313.
[34] J.F. Moulder, W.F. Stickle, P.E. Sobol, K.D. Bomben, Handbook of X-ray Photoelectron
Spectroscopy, vol. 40, Perkin Elmer, Eden Prairie, MN, 1992.
[35] A. Galtayries, R. Sporken, J. Riga, G. Blanchard, R. Caudano, XPS comparative study of
ceria/zirconia mixed oxides: powders and thin lm characterisation, J. Electron
Spectrosc. Relat. Phenom. 88 (1998) 951956.
[36] J. Fang, X. Bi, D. Si, Z. Jiang, W. Huang, Spectroscopic studies of interfacial structures
of CeO2TiO2 mixed oxides, Appl. Surf. Sci. 253 (22) (2007) 89528961.
[37] A.E. Nelson, K.H. Schulz, Surface chemistry and microstructural analysis of
Ce x Zr 1 x O 2 y model catalyst surfaces, Appl. Surf. Sci. 210 (3) (2003)
206221.
[38] J.P.Y. Tan, H.R. Tan, C. Boothroyd, Y.L. Foo, C.B. He, M. Lin, Three-dimensional structure of CeO2 nanocrystals, J. Phys. Chem. C 115 (9) (2011) 35443551.
[39] D. Prusty, A. Pathak, A. Chintha, B. Mukherjee, A. Chowdhury, Structural investigations on the compositional anomalies in lanthanum zirconate system synthesized
by coprecipitation method, J. Am. Ceram. Soc. 97 (3) (2014) 718724.

You might also like