0% found this document useful (0 votes)
16 views8 pages

Go Harshad I 2011

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
16 views8 pages

Go Harshad I 2011

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 8

Journal of Colloid and Interface Science 356 (2011) 473–480

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Fabrication of cerium oxide nanoparticles: Characterization and optical properties


Elaheh K. Goharshadi a,b,⇑, Sara Samiee a, Paul Nancarrow c
a
Dept. of Chemistry, Ferdowsi University of Mashhad, Mashhad 91779, Iran
b
Center of Nano Research, Ferdowsi University of Mashhad, Iran
c
School of Chemistry and Chemical Engineering, Queen’s University Belfast, UK

a r t i c l e i n f o a b s t r a c t

Article history: Ceria (CeO2) is a technologically important rare earth material because of its unique properties and var-
Received 23 October 2010 ious engineering and biological applications. A facile and rapid method has been developed to prepare
Accepted 20 January 2011 ceria nanoparticles using microwave with the average size 7 nm in the presence of a set of ionic liquids
Available online 25 January 2011
based on the bis (trifluoromethylsulfonyl) imide anion and different cations of 1-alkyl-3-methyl-imi-
dazolium. The structural features and optical properties of the nanoparticles were determined in depth
Keywords: with X-ray powder diffraction, transmission electron microscope, N2 adsorption–desorption technique,
Ceria nanoparticle
dynamic light scattering (DLS) analysis, FTIR spectroscopy, Raman spectroscopy, UV–vis absorption spec-
Ionic liquid
Microwave
troscopy, and Diffuse reflectance spectroscopy. The energy band gap measurements of nanoparticles of
Optical property ceria have been carried out by UV–visible absorption spectroscopy and diffuse reflectance spectroscopy.
The surface charge properties of colloidal ceria dispersions in ethylene glycol have been also studied. To
the best of our knowledge, this is the first report on using this type of ionic liquids in ceria nanoparticle
synthesis.
Ó 2011 Elsevier Inc. All rights reserved.

1. Introduction photocatalytic oxidation of water [27], high temperature oxidation


safe guards [3], as an anode material for lithium ion battery system
Ceria (cerium oxide, CeO2) is a cubic fluorite-type oxide in [28], buffer layers with silicon wafer [29], gates for metal-oxide
which each cerium site is surrounded by eight oxygen sites in fcc semiconductor device [30], solar cells [31], low-temperature water
arrangement and each oxygen site has a tetrahedron cerium site. gas shift catalyst [32], free-radical scavenger [33], removal soot
Recently, vacancy-engineered ceria nanoparticles have emerged from diesel engine exhaust [34], removal organics from wastewa-
as a fascinating material due to its wide applications and unique ter [35], water splitting for the generation of hydrogen gas
properties such as UV absorbing ability [1], high thermal stability [27,36], photodegradation of toluene in the gas phase [37], photo-
[2], facile electrical conductivity and diffusivity [3], high hardness, catalytic behavior under sunlight irradiation to degrade dyes
specific chemical reactivity [4], ability to store and transport oxy- [38,39], protective of primary cells from the detrimental effects
gen as large oxygen storage capacity [5], high refractive index, of radiation therapy [40], neuroprotection to spinal cord neurons
and the quick and expedient mutation of the oxidation state of cer- [41], prevention of retinal degeneration induced by intercellular
ium between Ce(III) and Ce(IV) [3]. These unique properties are yet peroxide [41], potent antioxidant in cell culture model [42], and
to be tapped for a large number of advanced applications including gas sensor [43].
glass polishing material [6], UV-blockers and filters [1,7–9], auto- A variety of methods have been applied to fabricate ceria nano-
motive exhaust promoter [10,11], As a coating for corrosion pro- particles such as hydrothermal synthesis [18,44–46], reversed mi-
tection for metals and alloys [12,13], as an additive to glass to celles route [47], coprecipitation [16,4,48], forced hydrolysis [49],
protect light-sensitive material [14], as an oxidation catalyst [15], an electrochemical method [50], solvothermal synthesis [51],
as an oxygen ion conductor in solid oxide full cells [16], electrolyz- sol–gel process [52,53], pyrolysis [54], sonochemical [55,56], and
ers [17], oxygen pumps and amperometric monitors[18], an effec- microwave [57–59].
tive high temperature oxidation resistant coating [19], sunscreens Among these preparation methods, microwave has the advanta-
[20], solid electrolytes [21], additives in ceramics [22,23], abrasive ges of rapid volumetric heating and the consequent dramatic in-
of the CMP slurry in semiconductor fabrication [24], selective crease in reaction rates, high-efficiency, high purity, and rapid
hydrogenation catalysis of unsaturated compounds [25,26], formation of nanoparticles with a narrow size distribution, and
no serious agglomeration.
Ionic liquids (ILs) represent a big success for industrial and engi-
⇑ Corresponding author at: Dept. of Chemistry, Ferdowsi University of Mashhad,
Mashhad 91779, Iran. Fax: +98 511 8796416.
neering chemistry at the beginning of the 21st century due to their
E-mail address: [email protected] (E.K. Goharshadi). large liquid state range, favorable solvation behavior, low melting

0021-9797/$ - see front matter Ó 2011 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2011.01.063
474 E.K. Goharshadi et al. / Journal of Colloid and Interface Science 356 (2011) 473–480

temperature, non-volatility, non-flammability, stability in air, high represents the different conditions for preparing ceria nanoparti-
ionic conductivity, high thermal stability in a wide temperature cles used in this work.
range, relatively low viscosity, low vapor pressure, and selectivity
in chemical reactions [60]. ILs are synthesized by combining bulky 2.3. Characterization techniques
organic cations, e.g., imidazolium or pyridinium, with a wide vari-
ety of anions. The large range of organic cation and anion pairs The powder phases were determined by means of a Bruker/D8
makes it possible to design solvents with specific properties to suit Advanced diffractometer in the 2° range from 20° to 80°, by step
specific applications such as separation and extraction processes of 0.04°, with graphite monochromatic Cu Ka radiation
[61–63], and synthesis of nanoparticles [64–67]. (k = 1.541 Å). The TEM analyses of the samples were done using a
In the present study, we employed the microwave method to LEO 912 AB, instrument and the electron beam accelerating voltage
prepare the nanometer-sized cerium oxide ultrafine particles using was 120 kV.
a set of ILs based on the bis (trifluoromethylsulfonyl)imide anion The surface area, pore volume, and average pore radius of sam-
and different cations of 1-alkyl-3-methyl-imidazolium. The struc- ples were measured by an ASAP-2010 system from Micromeritics.
tural features and optical properties of the ceria nanoparticles were The samples were degassed to remove moisture and impurities at
determined in depth with X-ray powder diffraction (XRD), trans- 200 °C for 4 h under 50 mTorr vacuum and their BET area, pore vol-
mission electron microscope (TEM), FTIR spectroscopy, Raman ume, and average pore radius were determined.
spectroscopy, UV–vis absorption spectroscopy, and Diffuse reflec- The DLS experiment and the zeta potential of the particles were
tance spectroscopy (DRS), N2 adsorption–desorption technique, measured by the Zetasizer (Nano-ZS) from Malvern Instrument.
and DLS analysis. Also, in this study, the dispersion behavior of cer- The Fourier transform infrared spectra of the fabricated ceria
ia nanoparticles in ethylene glycol was investigated under different nanoparticles were recorded at room temperature with a KBr pellet
pH values by the method of zeta potential. on a Shimadzu 4300 Spectrometer ranging from 450 to 3600 cm1.
The FT-Raman spectra in the region 3200–200 cm1 were recorded
employing a 180° back-scattering geometry and a Bomem MB-154
2. Materials and methods Fourier transform Raman spectrometer operating at the 1064 nm
excitation line of a Nd:YAG laser. It was equipped with a ZnSe
2.1. Materials beam splitter and a TE cooled InGaAs detector. Rayleigh filtration
was afforded by a set of two holographic technology filters. The
All ILs, namely 1-ethyl-3-methyl-imidazolium bis (trifluoro- spectra were accumulated for 1500 scans with a resolution of
methylsulfonyl) imide ([C2mim][NTf2]), 1-butyl-3-methyl-imi- 2 cm1. The laser power at the sample was 500 MW.
dazolium bis (trifluoromethylsulfonyl) imide ([C4mim][NTf2]), The UV–vis absorbance spectra were obtained for the samples
and 1-hexyl-3-methyl-imidazolium bis (trifluoromethylsulfonyl) using an Agilent photodiode-array Model 8453 equipped with
imide ([C6mim][NTf2]) used in this work were synthesized accord- glass of 1 cm path length. The spectra were recorded at room tem-
ing to the literature [68]. All ionic liquids were analyzed by NMR, perature in air within the range 300–550 nm. The diffuse reflec-
Karl-Fischer titration for water content, and ion chromatography tance spectra were recorded by a Scinco apparatus.
for chloride content. In all cases, the water mass fraction was found
to be less than 0.001 and chloride mass fraction was less than
3. Results and discussion
5  106.
All other chemicals used were of analytical grade and used as
Fig. 1 shows the XRD patterns of the samples. All the diffraction
received without further purification.
peaks are labeled and can be indexed as the face-centered cubic
phase. The strong and sharp diffraction peaks indicate the good
crystallization of the samples. No additional peaks in the XRD were
2.2. Experimental
observed, revealing the high purity of the prepared ceria
nanoparticles.
In a typical synthesis of ultrafine CeO2, 4.585 mmol
The average Crystallite size, D, can be calculated using the well-
Ce(NO3)36H2O was dissolved in 2 ml ionic liquid and proper
known Scherrer’s equation:
amounts of NaOH solution (0.18 M) were rapidly added. As soon
as the NaOH solution was added, the reaction started and white kk
Ce(OH)3 precipitate was formed. The solution was kept in ambient Dhkl ¼ ð1Þ
bhkl  cos hhkl
environment with a stirring rate of 500 rpm for 30 min. The pH va-
lue of the reaction solution during the process maintained at 11.74. where Dhkl is the Crystallite size perpendicular to the normal line of
It should be noted that NaOH solution generates OH and in such a (hkl) plane, k is a constant (0.9), bhkl is the full width at half maxi-
basic environment, the oxidation transfer of Ce(III) ? Ce(IV) takes mum of the (hkl) diffraction peak, hhkl is the Bragg angle of (hkl)
place because of the dissolved oxygen from air. Gradually, the color peak and k is the wavelength of X-ray. The peak position and the
of the solution changed to purple. The reaction was carried on for FWHM were obtained by fitting the measured peaks with two
30 min in air in a domestic microwave oven. The microwave oven Gaussian curves in order to find the true peak position and width
(1000 W) followed a working cycle of 10 s on and 5 s off (30% corresponding to monochromatic Cu Ka radiation. The characteris-
power) and a yellowish precipitate started to appear. The temper- tics of the nanoceria samples from XRD patterns are summarized in
ature increase after the microwave treatment was 65 °C. After
cooling to room temperature, the resulting precipitate was centri-
Table 1
fuged (10 min with 13,000 rpm) and washed with ethanol and
Synthesis conditions used in this work.
deionized water several times to remove excess ILs and any possi-
ble ionic remnants. At last, the products were dried in a vacuum Sample ILs

oven at 60 °C overnight. A similar procedure was used for the fab- 1 No ILs was used
rication of ceria nanoparticles in the absence of ILs. The time for 2 [C2mim][NTF2]
3 [C4mim][NTF2]
the reaction, appearing a yellowish precipitate, was 30 min
4 [C6mim][NTF2]
whereas in the presence of ILs this time was 10 min. Table 1
E.K. Goharshadi et al. / Journal of Colloid and Interface Science 356 (2011) 473–480 475

(220)
(111)

(311)

(420)
(331)
(400)
(222)
(200)
(d)

Intensity (a. u.)


(c)

(b)

(a)
20 30 40 50 60 70 80 90

2θ (degree)
Fig. 1. The XRD patterns of nanoceria for: (a) sample 1, (b) sample 2 (c) sample 3, (d) sample 4.

Table 2 in the lattice parameter was observed with a decrease in the crys-
The characteristics of the nanoceria samples from XRD patterns. tallite size. As a general rule, nanoparticles of oxides exhibit a lat-
Sample Dhkl (nm) 2h Lattice parameter (A°) Cell volume (A°3) tice expansion with reduction in particle size while metal
1 6.45 56.345 5.4111 158.44
nanoparticles exhibit a lattice contraction [70]. The reciprocal of
2 5.85 56.339 5.4116 158.48 the diameter (D1) is proportional to the surface to volume ratio
3 7.54 56.378 5.4082 158.18 (S/V) and consequently, the increase of the lattice parameter can
4 7.27 56.343 5.4113 158.45 be related to the higher surface to volume ratio in the smaller par-
ticles, resulting in a higher contribution from the surface layer. In
oxide particles, the bonds have a directional character and at the
Table 2. Since the systematic error in the lattice parameter (a) de- outer surface of each particle, there would be unpaired electronic
creases as the Bragg angle increases [69], the values of the average orbitals, which would repel each other [71]. This contribution from
particle size and the lattice parameter were calculated for the the surface layer increases with decreasing particle size and leads
highest-angle line (3 1 1) with a reasonable intensity. The average to larger values of the lattice parameter than in the bulk. However,
Crystallite size of ceria nanoparticles is smallest when [C2mim] some authors [72], have attributed the lattice expansion with
[NTf2] is used as an ionic liquid. Among three ILs used, this ionic decreasing particle size to an increase in vacancy or impurity con-
liquid reduces the particle size as a result of its adsorption on the centrations. Hence, the lattice expansion may be assigned to the
surface of CeO2 nanoparticles which inhibits the aggregation of presence of Ce3+ and or oxygen vacancies. The high reactivity of
ceria nanoparticles [63,65]. In other words, 1-ethyl-3-methyl- the nanoparticles would have caused the reduction of Ce4+
imidazolium bis(trifluoromethylsulfonyl) imide is the strongest (0.97 A°) ion and consequent formation of Ce3+ ion with a higher
adsorbed ILs on the surface of ceria nanoparticles because of its ionic radii (1.14 A°) [73].
small alkyl group and small steric inhibition. The morphology of the products was characterized by TEM.
Table 2 shows the dependence of the calculated lattice param- Typical TEM images of samples 1 and 3 are shown in Fig. 2. The
eter, a, to the observed crystallite size of the samples. An increase average particle size of about 5 nm for sample 1 and 10 nm for

Fig. 2. The TEM images of nanoceria for: (a) sample 1, (b) sample 3.
476 E.K. Goharshadi et al. / Journal of Colloid and Interface Science 356 (2011) 473–480

[C 2 mim][NTf 2 ]
[C 4 mim][NTf 2 ]
100
[C 6 mim][NTf 2 ]

80

Transmittance
60

40
/ 0
0

20

0
1000 1500 2000 2500 3000 3500
Wavenumber (cm-1)
Fig. 3. FT-IR spectra of used ILs.

Fig. 4. FT-IR spectra for: (a) sample 1, (b) sample 2 (c) sample 3, (d) sample 4.

sample 3 which is in agreement with the results deduced from the bulk ceria, which is assigned to triply degenerate F2g mode as a
XRD. symmetric breathing mode of the oxygen atoms around the cerium
Figs. 3 and 4 show the FT-IR spectra of the ILs and our sam- ions [76]. This peak is shifted to 463 cm1 and 461 cm1, for sam-
ples, respectively. The strong peaks around 2851 and 2916 cm1 ple 1 and samples 2–4, respectively. The curve becomes broad for
in Fig. 3, attributed to the alkyl chain of the two characteristic ceria nanoparticles. The broad peaks indicate that the size of the
bands of the imidazolium ring around 3100 cm1 correspond to particles is small. The red shift of this main Raman band is attrib-
the symmetric and asymmetric stretching of the CAH bond in uted to the lattice relaxation revealed by XRD while several other
positions four and five of the imidazolium ring [74]. All charac- factors, such as phonon confinement, increased lattice strain, non-
teristic peaks of ionic liquids can not be found in the spectra of uniform strain, and variations in phonon relaxation may account
ceria samples in Fig. 4. The CeAO stretching band around for the broadening of the F2g band at decreased crystallite particle
450 cm1 shown in Fig. 4 confirms the formation of CeO2 [75]. sizes [76]. The position of Raman peak varies with the change of
The peak at ca. 3400 cm1 corresponds to the physically ad- the interatomic force which is characterized by the change of bond
sorbed water on the samples. length, as well as the change of the lattice spacing. The relationship
Fig. 5 exhibits the Raman spectra of the ceria nanoparticles. It is between the Raman shift and the change of the lattice parameter
known that only one sharp Raman peak at 464 cm1 is observed for can be expressed by the following equation [76]:
E.K. Goharshadi et al. / Journal of Colloid and Interface Science 356 (2011) 473–480 477

6
dBET ¼ ð3Þ
qasBET
The size distribution of the samples 1 and 3 was examined by a
DLS analyzer (Fig. 7). The DLS data suggest that the samples 1 and 3
mainly consist of particles of about 38 nm and 59, respectively.
Clearly, the particle size by DLS analysis is larger than of TEM anal-
ysis since the hydrodynamic radius is probed with DLS.
The zeta potential, which is the overall charge a particle ac-
quires in a specific medium, is a good indicator of the stability of
the colloidal system. The zeta potential of colloidal dispersions of
samples 1 (Fig. 8) and 3 in ethylene glycol has been studied. The
zeta potential of sample 3 is much less than that of sample 1.
The pH was determined using the method recommended by IUPAC
[77]. The pH measured in ethylene glycol was found to differ by up
to 1 unit from that of water. The isoelectric point (IEP) was found
to be at pH = 4.12. The colloidal system is the least stable at this
point due to the absence of particle surface charges. Patila et al.
Fig. 5. The Raman spectra of nanoceria for sample 1. Inset: (a) sample 2 (b) sample [78] measured the zeta potential of ceria nanoparticles in water
3, (c) sample 4. prepared by microemulsion and hydrothermal methods. In the
case of the microemulsion nanoceria, the IEP was approximately
4.5 and is about 9.5 for hydrothermal nanoceria. The IEP measured
3cxo Da
Dx ¼  ð2Þ by Patsalas et al. [79] was 3. As pH increases beyond the IEP, the
ao absolute value of zeta potential of the particle surface increase
where xo is Raman frequency of bulk CeO2, ao is the bulk ceria lat- and hence the electrostatic repulsion force between particles be-
tice constant (0.5411 nm), Da is the change in lattice constant, and c comes sufficient to prevent attraction. The zeta potential of the
is the Gruneisen constant with a value of 1.24 for ceria [76]. In our suspensions of ceria nanoparticles reaches as high as 59 mV at
experiment, the experimental frequency shifts, Dx, are 1 cm1 pH = 10.
and 3 cm1 for sample 1 and samples 2–4, respectively. Hence, Optical responses of the ceria nanoparticles have been investi-
the change in the lattice constant is 0.0003 nm and 0.0009 nm for gated via UV–vis absorption spectroscopy, and the results are pre-
sample 1 and samples 2–4, respectively which are in agreement sented in Fig. 9 for a typical sample. All the samples distinctly
with the results deduced from XRD.
Fig. 6 shows the nitrogen adsorption–desorption isotherms at
76.09 K for samples 1 and 3. The isotherms for both samples are
Table 3
apparently of type IV, typical of mesoporous materials. The desorp- The specific surface area and porosity of samples.
tion branch of the isotherm follows a different path to the adsorp-
tion branch, although the curve closes as the relative pressure Sample aSBET aSBJH V PBJH dBET dpore
(m2 g1)a (m2 g1)b (cm3 g1)c (nm)d (nm)e
approaches 0.4. The BET surface areas and porosity of these two
samples are summarized in Table 3. It should be noted that the 1 91.77 120.8876 0.132306 8.98 4.3778
adsorption capacity of sample 1 is less than that of sample 3. This 3 87.31 117.9597 0.197642 9.44 6.7020

decrease should be related to the better crystallization of sample 3 a


Calculated from BET equation.
b
as confirmed by the XRD and TEM. BJH desorption cumulative surface area of pores between 17 and 3000 A°
The particle size, dBET, can be calculated from the specific sur- diameter.
c
BJH desorption cumulative pore volume of pores between 17 and 3000 A°
face area, aSBET , data by assuming spherical particles using the value
diameter.
of the density for particles (Table 3): d
Calculated using Eq. (3).
e
BJH desorption average pore diameter.

120
Volume Adsorbed (cm 3/g) STP

Sample 1
Sample 3
100

80

60

40

20

0
0.0 0.2 0.4 0.6 0.8 1.0
Relative Pressure (P/Po)

Fig. 6. Nitrogen adsorption–desorption isotherms for samples 1 and 3.


478 E.K. Goharshadi et al. / Journal of Colloid and Interface Science 356 (2011) 473–480

(a) 25
(b) 25

20
20
Frequency

15

Frequency
15

10 10

5 5

0 0
27 30 33 36 39 42 45 48 51 42 45 48 51 54 57 60 63 66 69 72 75 78 81 84 87 90 93
Diameter (nm) Diameter (nm)
Fig. 7. The DLS data of size distribution of (a) sample (1) and (b) sample 3.

exhibit a strong absorption band (below 400 nm in wavelength) at of the absorption edge [82], see kmax,UV in Table 4. This blue shift
the UV region due to the charge-transfer transitions from O 2p to in absorption spectra is attributed to the quantum size effect [55].
Ce 4f, which overruns the well-known f–f spin–orbit splitting of Since the size-related band gap shift of semiconductor nano-
the Ce 4f state [8,79,80]. The spectrum of each sample shows most crystals can be quantified, it is possible to calculate an optical par-
of the UV light (200–350 nm) is blocked allowing ceria nanoparti- ticle size with the band gap shift measured from absorption
cles to be used as a UV blocker. spectra [83]. The relation between the particle size and effective
An estimate of the optical band gap, Eg, can be determined by band gap of a nanomaterial can be written as:
the following equation for a semiconductor:  
ph2 1 1 1:8e2
Eg;n ¼ Eg;b þ þ  ð5Þ
ðahtÞn ¼ Bðht  Eg Þ ð4Þ 2R2 me mh eR
where Egb is the bulk band gap (3.19 eV), R is the particle radius, me
where ht is the photon energy, a is the absorption coefficient, B is a and mh are the effective masses of the electron and hole, respec-
constant relative to the material, and n is either 2 for a direct tran- tively and it considered as me ¼ mh ¼ 0:4mo , where mo is the mass
sition or 1/2 for an indirect transition. The energy intercept of a plot of a free electron, h is Planck´ s constant, e is the bulk optical dielec-
of (aht)n versus ht yields Eg. Since ceria is a direct band gap semi- tric constant. When R is very small (comparable to the Bohr radius),
conductor, the value of n is 2 [81]. The values of direct band gap en- the 1/R2 term is dominant and the band gap increases with decreas-
ergy, Eg, of the samples are summarized in Table 4. The results ing size, see Table 4.
reveal that the band gap energy increases when the particle size Diffuse reflectance spectroscopy is a more convenient technique
is reduced. Since ceria is a direct band gap semiconductor, a de- to characterize unsupported nanomaterials than UV–vis absorp-
crease in particle size is exhibited to be manifested by a blue shift tion spectroscopy, since it takes advantage of the enhanced

60

40

20

-20

-40

-60

0 1 2 3 4 5 6 7 8 9 10 11

Fig. 8. The zeta potential of ceria naoparticles dispersed in ethylene glycol of sample (1) under various pH values.
E.K. Goharshadi et al. / Journal of Colloid and Interface Science 356 (2011) 473–480 479

Fig. 9. UV–vis spectrum of sample 1dispersed in ethanol. Plot of (aht)2 versus photon energy (inset).

Table 4 5. Conclusions
The optical properties of ceria nanoparticles.

Sample kmax,UV (nm) Eg,UV (eV)a dUV (nm)b kmax,DRS Eg,DRS (eV)c Microwave-assisted heating method has been successfully
established for the preparation of nanocrystalline CeO2 particles
1 317 3.48 7.95 351 3.53
2 280 3.70 6.33 350 3.54 in the presence of ILs. The method is found to be convenient, rapid,
3 318 3.47 8.00 352 3.52 and efficient.
4 306 3.11 7.27 356 3.48 One interesting point of the present work is that the average
a
Calculated from UV–vis spectra. particle size of samples calculated with the help of XRD pattern,
b
Calculated using Eq. (5). TEM images, BET surface area, and the shift of the band gap absorp-
c
Calculated from DRS spectra. tion in the UV–vis spectrum agree closely.
The optical properties of nanoparticles of ceria have been inves-
tigated by UV–vis absorption spectroscopy and Diffuse reflectance
scattering phenomenon in powder materials. Effects of light scat- spectroscopy. The band gap energy values of both methods are
tering in the absorption spectra of powder samples dispersed in li- nearly identical.
quid media can be avoided using DRS. The DRS technique does not The zeta potential of the suspensions of ceria nanoparticles
require a powder sample to be dispersed in any liquid medium, so reaches as high as 59 mV at pH = 10 so the dispersion stability
the material is not contaminated or consumed [84]. Diffuse reflec- of ceria nanoparticles in ethylene glycol is the best.
tance spectra of samples are shown in Fig. 10. Table 4 presents the
Eg values obtained from two methods: UV–vis absorption spectros- Acknowledgments
copy and DRS. The Eg values of both methods are nearly identical.
The authors express their gratitude to Ferdowsi University of
Mashhad for support of this project (P986). The authors would also
like to acknowledge Professor Sayyed Faramarz Tayyari for taking
the Raman spectra, Mrs. Somayeh Laleh for taking the FTIR spectra,
and Mrs. Roksana Pesian for taking TEM images.

References

[1] S. Tsunekawa, R. Sahara, Y. Kawazoe, A. Kasuya, Mater. Trans. JIM 41 (2000)


1104–1107.
[2] A. Trovarelli, C. de Leitenburg, M. Boaro, G. Dolcetti, Catal. Today 50 (1999)
353–367.
[3] F. Zhou, X. Zhao, H. Xu, C. Yuan, J. Phys. Chem. C 111 (2007) 1651–1657.
[4] H.I. Chen, H.Y. Chang, Solid State Commun. 133 (2005) 593–598.
[5] A.M. Shahin, F.G. Grandjean, J. Long, T.P. Schuman, Chem. Mater. 17 (2005)
315–321.
[6] E. Bekyarova, P. Fornasiero, J. Kaspar, M. Graziani, Catal. Today 45 (1998) 179–
183.
[7] T. Morimoto, H. Tomonaga, A. Mitani, Thin Solid Films 351 (1999) 61–65.
[8] S. Tsunekawa, T. Fukuda, A. Kasuya, J. Appl. Phys. 87 (2000) 1318–1321.
[9] M. Yamashita, K. Kameyama, S. Yabe, S. Yoshida, Y. Fujishiro, T. Kawai, T. Sato,
J. Mater. Sci. 37 (2002) 683–687.
Fig. 10. Diffuse reflectance spectra for: (a) sample 1, (b) sample 2 (c) sample 3, (d) [10] G.R. Rao, P. Fornasiero, R. Di Monte, J. Kaspar, G. Vlaic, G. Balducci, S. Meriani,
sample 4. G. Gubitosa, A. Cremona, M. Graziani, J. Catal. 162 (1996) 1–9.
480 E.K. Goharshadi et al. / Journal of Colloid and Interface Science 356 (2011) 473–480

[11] P. Fornasiero, G. Balducci, R. Di Monte, J. Kašpar, V. Sergo, G. Gubitosa, A. [48] T.J. Kirk, J. Winnick, J. Electrochem. Soc. 140 (1993) 3494–3496.
Ferrero, M. Graziani, J. Catal. 164 (1996) 173–183. [49] W.P. Hsu, L. Ronnquist, E. Matijevic, Langmuir 4 (1988) 31–37.
[12] A.S. Hamdy, Mater. Lett. 60 (2006) 2633–2637. [50] Y. Zhow, R.J. Philips, J.A. Switzer, J. Am. Ceram. Soc. 78 (1995) 981–985.
[13] X. Zhong, Q. Li, J. Hu, Y. Lu, Corros. Sci. 50 (2008) 2304. [51] E. Verdon, M. Devalette, G. Damazeau, Mater. Lett. 25 (1995) 127–131.
[14] C. Lin, J.D.C. Basil, R.M. Hunia, P.P. Bihuniak, US Patent 1994, 5316854. [52] A. Verma1, N. Karar, A.K. Bakhshi, H. Chander, S.M. Shivaprasad, S.A. Agnihotry,
[15] M.S. Yakimova, V.K. Ivanov, O.S. Polezhaeva, A.A. Trushin, A.S. Lermontov, Y.D. J. Nanoparticle Res. 9 (2007) 317–322.
Tretyakov, Dokl. Chem. 427 (2009) 186–189. [53] X. Chu, W. Chung, L.D. Schmidt, J. Am. Ceram. Soc. 76 (1993) 2115–2118.
[16] H. Yahiro, Y. Baba, K. Eguchi, H. Arai, J. Electrochem. Soc. 135 (1988) 2077– [54] H. Xu, L. Gao, H. Gu, J. Guo, D. Yan, J. Am. Ceram. Soc. 85 (2002) 139–144.
2081. [55] L. Yin, Y. Wang, G. Pang, Y. Koltypin, A. Gedanken, J. Colloid Interface Sci. 246
[17] O.A. Marina, L.R. Pederson, US Patent, 2008, 7468218. (2002) 78–84.
[18] M. Hirano, E. Kato, J. Am. Ceram. Soc. 79 (1996) 777–780. [56] D. Zhang, H. Fu, L. Shi, C. Pan, Q. Li, Y. Chu, W. Yu, Inorg. Chem. 46 (2007) 2446–
[19] S. Patil, S.C. Kuiry, S. Seal, R. Vanfleet, J. Nanoparticles Res. 4 (2002) 433–438. 2451.
[20] T. Masut, M. Yamamoyo, T. Sakata, H. Mori, G. Adachi, J. Mater. Chem. 10 [57] H. Yang, C. Huang, A. Tang, X. Zhang, W. Yang, Mater. Res. Bull. 40 (2005)
(2000) 353–357. 1690–1695.
[21] H. Inaba, H. Tagawa, Solid State Ionics 83 (1996) 1–16. [58] C.S. Riccardi, R.C. Lima, M.L. dos Santos, P.R. Bueno, J.A. Varela, E. Longo, Solid
[22] S.B. Bhadun, A. Chakraborty, R.M. Rao, J. Am. Ceram. Soc. 71 (1988) C410– State Ionics 180 (2009) 288–291.
C411. [59] Y. Tao, F.H. Gong, H. Wang, H.P. Wu, G.L. Tao, Mater. Chem. Phys. 112 (2008)
[23] G.L. Messing, S.C. Zhang, G.V. Jayanthi, J. Am. Ceram. Soc. 76 (1993) 2707– 973–976.
2726. [60] A.E. Visser, R.P. Swatloski, W.M. Reichert, S.T. Griffin, R.D. Rogers, Ind. Eng.
[24] M. Jiang, N.D. Wood, R. Komanduri, J. Eng. Mater. Technol. Trans. ASME 120 Chem. Res. 39 (2000) 3596–3604.
(1998) 304–312. [61] A.E. Visser, R.P. Swatloski, R.D. Rogers, Green Chem. 2 (2000) 1–4.
[25] J.L.G. Fierro, J. Soria, J. Sanz, M.J. Rojo, J. Solid State Chem. 66 (1987) 154–162. [62] A.E. Visser, R.P. Swatloski, S.T. Griffin, D.H. Hartman, R.D. Rogers, Sep. Sci.
[26] K.S. Sim, L. Hilaire, F. Le Normand, R. Touroude, V. Paul-Boncour, A. Percheron- Technol. 36 (2001) 785–804.
Guegan, J. Chem. Soc., Faraday Trans. 87 (1991) 1453–1460. [63] E.K. Goharshadi, Y. Ding, P. Nancarrow, J. Phys. Chem. Solids 69 (2008) 2057–
[27] G.R. Bamwenda, H. Arakawa, J. Mol. Catal. A 161 (2000) 105–113. 2060.
[28] F. Zhou, X. Ni, Y. Zhang, H. Zheng, J. Colloid, J. Colloid Interface Sci. 307 (2007) [64] E. Redel, R. Thomann, C. Janiak, Inorg. Chem. 47 (2008) 14–16.
135–138. [65] E.K. Goharshadi, Y. Ding, M.J. Namayandeh, P. Nancarrow, Ultrason. Sonochem.
[29] J. Tashiro, A. Sasaki, S. Akiba, S. Satoh, T. Watanabe, H. Funakubo, M. 16 (2009) 120–123.
Yoshimoto, Thin Solid Films 415 (2002) 272–275. [66] R. Jalal, E.K. Goharshadi, M. Abareshi, M. Moosavi, A. Yousefi, P. Nancarrow,
[30] S.F. Galata, E.K. Evangelou, Y. Panayiotatos, A. Sotiropoulos, A. Dimoulas, Mater. Chem. Phys. 121 (2010) 198–201.
Microelectron. Reliab. 47 (2007) 532–535. [67] M.J. Obliosca, J.I.H. Arellano, H.M. Huang, D.S. Arco, Mater. Lett. 64 (2010)
[31] A. Corma, P. Atienzar, H. Gara, J.Y. Chane-Ching, Nat. Mater. 3 (2004) 394– 1109–1112.
397. [68] P. Bonhôte, A. Dias, N. Papageorgiou, K. Kalyanasundaram, M. Gra1tzel, Inorg.
[32] S. Hilaire, X. Wang, T. Luo, R.J. Gorte, J. Wagner, Appl. Catal. A: Gen. 215 (2001) Chem. 5 (1996) 1168–1178.
271–278. [69] B.D. Cullity, Elements of X-ray Diffraction, second ed., Adison-Wesley
[33] S. Babu, A. Velez, K. Wozniak, J. Szydlowska, S. Seal, Chem. Phys. Lett. 442 Publishing Co., USA, 1978.
(2007) 405–408. [70] P. Vaqueiro, M.A. López-Quintela, J. Mater. Chem. 8 (1998) 161–163.
[34] J. Lahaye, S. Boehm, P.H. Chambrion, P. Ehrburger, Combust. Flame 104 (1996) [71] P. Ayyub, V.R. Palkar, S. Chattopadhyay, M. Multani, Phys. Rev. B 51 (1995)
199–207. 6135–6138.
[35] Y.I. Matatov-Meytal, M. Sheintuch, Ind. Eng. Chem. Res. 37 (1998) 309–326. [72] X.D. Liu, H.Y. Zhang, K. Lu, Z.Q. Hu, J. Phys.: Condens. Matter 6 (1994) L497–
[36] K.H. Chung, D.C. Park, Catal. Today 30 (1996) 157–162. L501.
[37] M.D. Hernández-Alonso, A.B. Hungrı́a, A. Martı́nez-Arias, M. Fernández-Garcı́a, [73] S. Banerjee, P.S. Devi, J. Nanoparticle Res. 9 (2007) 1097–1107.
J.M. Coronado, J.C. Conesa, J. Soria, J. Appl. Catal. B: Environ. 50 (2004) 167– [74] B.D. Fitchett, J.C. Conboy, J. Phys. Chem. B 108 (2004) 20255–20262.
175. [75] C. Ho, J.C. Yu, T.A. Kwong, C. Mak, S. Lai, Chem. Mater. 17 (2005) 4514–4522.
[38] Y.Q. Zhai, S.Y. Zhang, H. Pang, Mater. Lett. 61 (2007) 1863–1867. [76] J.R. McBride, K.C. Hass, B.D. Poindexter, W.H. Weber, J. Appl. Phys. 76 (1994)
[39] P. Borker, A.V. Salker, Mater. Chem. Phys. 103 (2007) 366–370. 2435–2442.
[40] R.W. Tarnuzzer, J. Colon, S. Patil, S. Seal, Nano Lett. 5 (2005) 2573–2577. [77] K. Izutsu, Electrochemistry in Nonaqueous Solutions, Wiley-VCH, Weinheim,
[41] M. Das, S. Patil, N. Bhargava, J.-F. Kang, L.M. Riedel, S. Seal, et al., Biomaterials 2002.
28 (2007) 1918–1925. [78] S. Patila, A. Sandberg, E. Heckertc, W. Selfc, S. Seal, Biomaterials 28 (2007)
[42] S. Patil, A. Sandberg, E. Heckert, W. Self, S. Seal, Biomaterials 28 (2007) 4600– 4600–4607.
4607. [79] P. Patsalas, S. Logothetidis, Phys. Rev. B 68 (2003) 035104–135117.
[43] T.S. Stefanik, H.L. Tuller, J. Eur. Ceram. Soc. 21 (2001) 1967–1970. [80] M.D. Hernández-Alonso, A.B. Hungria, A.J. Martı́nez-Arias, M. Coronado, Phys.
[44] Z. Yang, Y. Yang, H. Liang, L. Liu, Mater. Lett. 63 (2009) 1774–1777. Chem. Chem. Phys. 6 (2004) 3524–3529.
[45] R. Yu, L. Yan, P. Zheng, J. Chen, X. Xing, J. Phys. Chem. C 112 (2008) 19896– [81] H. Gu, M.D. Soucek, Chem. Mater. 19 (2007) 1103–1110.
19900. [82] L.E. Brus, J. Phys. Chem. 90 (1986) 2555–2560.
[46] M. Hirano, E. Kato, J. Mater. Chem. 10 (2000) 473–477. [83] Y. Dieckmann, H. Colfen, H. Hofmann, A. Petri-Fink, Anal. Chem. 81 (2009)
[47] T. Masui, K. Fujiwara, K. Machida, G. Adachi, T. Sakata, H. Mori, Chem. Mater. 9 3889.
(1997) 2197–2204. [84] A.E. Morales, E.S. Mora, U. Pal, Rev. Mex. Fis. S 53 (2007) 18–22.

You might also like