You are on page 1of 7

Materials Science in Semiconductor Processing 16 (2013) 288–294

Contents lists available at SciVerse ScienceDirect

Materials Science in Semiconductor Processing


journal homepage: www.elsevier.com/locate/mssp

Preparation and characterization of BiFeO3@Ce-doped


TiO2 core-shell structured nanocomposites
Lixiu Gong a, Zhufa Zhou a,b,n, Shumei Wang a,b,n, Ben Wang a
a
College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, PR China
b
National Engineering Laboratory of Modern Silk, Soochow University, Suzhou 215123, PR China

a r t i c l e in f o abstract

Available online 13 November 2012 BiFeO3@TiO2–Ce composite nanoparticles with a BiFeO3 core and a Ce-doped TiO2 shell
Keywords: structure were fabricated via a sol–gel method. The nanoparticles were characterized by
Titania scanning electron microscopy with energy dispersive spectroscopy, X-ray diffraction,
BiFeO3 transmission electron microscopy and UV–vis diffuse reflectance spectroscopy. The
Ce-doped results reveal that core-shell structured BiFeO3@TiO2–Ce nanoparticles show a signifi-
Synergistic effect cant redshift in the UV–vis absorption spectra in comparison with both Ce-TiO2 and
Photocatalysts BiFeO3@TiO2 nanoparticles. The photocatalytic activities of the samples were tested in
the degradation of methyl orange in aqueous solutions under visible light and UV light
irradiations. The core-shell structured BiFeO3@TiO2–Ce sample exhibits higher photo-
catalytic activity, which is attributed to the synergistic effects of BiFeO3 and cerium.
Crown Copyright & 2012 Published by Elsevier Ltd. All rights reserved.

1. Introduction In particular, exploiting a core-shell structured nanoparti-


cles in which TiO2 acted as shell and magnetic materials
Titanium dioxide (TiO2) in different forms has attracted (such as Fe3O4, MnFe2O4, BaFe12O19) introduced as core is
extensive interests in recent years owing to its high photo- also an efficient way [14–16].
catalytic effects on decomposing various organic pollutants As we have known, BiFeO3 has an excellent multiferro-
[1–4], which offers a viable approach to solve the variety of electric properties at room temperature. In addition, It has
environmental problems. But the band gap of TiO2 (3.2 eV been studied widely in the photocatalytic field due to its
for anatase) is so wide that it can only absorb ultraviolet small bandgap. Zhang et al. synthesized TiO2/BiFeO3
(UV) light, which is only 5% in the sun source [5]. In heterostructures, and their results indicated BiFeO3
addition, the photogenerated electron–hole pair recombin- weaken the recombination of electrons generated [17].
ing easily results in poor photocatalytic activity, and hence Shun Li et al. had prepared BiFeO3@TiO2 core-shell struc-
limits its applications. Several efforts have been made, such ture and found that the core-shell nanoparticles had a
as replacing the oxygen position with nonmetallic or metal higher photocatalytic activity in contrast to pure TiO2.
element to overcome the barriers [6–10]. Besides, various BiFeO3 act as electron carriers that can promote interfacial
modifications have been performed to synthetize nanos- charge transfer in the composite systems [18]. Addition-
tructured composites such as Fe3O4/SiO2/TiO2, ZnO/TiO2 ally the magnetic property of the composites helps
and WO3/TiO2, which can efficiently hinder the recombina- recycle the photocatalysts from treated water. The core-
tion of the photogenerated electron–hole pairs [11–13]. shell BiFeO3@TiO2 nanoparticles have a lower energy gap
that enhances visible light absorption. In order to over-
n
come the recombination of photogenerated electron–hole
Corresponding authors at: College of Chemistry, Chemical Engineering
and Materials Science, Soochow University, Suzhou 215123, PR China.
pairs of BiFeO3@TiO2, many groups have been involved in
Tel.: þ 86 512 65880963; fax: þ 86 512 65880089. doping rare earth ions into TiO2 [19,20], among which
E-mail address: zhouzhufa@suda.edu.cn (Z. Zhou). cerium has been received much attentions because of the

1369-8001/$ - see front matter Crown Copyright & 2012 Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.mssp.2012.10.009
L. Gong et al. / Materials Science in Semiconductor Processing 16 (2013) 288–294 289

following two reasons: (1) the redox couple Ce3 þ /Ce4 þ transparent suspension was stirred for another 6 h and
with the ability of ceria to shift between CeO2 and Ce2O3 aged for 24 h until the transparent gel was obtained. The
under oxidizing and reducing conditions; and (2) the easy gel was dried at 80 1C for 24 h and then ground to form
formation of labile oxygen vacancies (OV) with the semisolid powder before sintering at 500 1C for 2 h in air
relatively high mobility of bulk oxygen species [21]. condition. The resultant powders were BiFeO3@TiO2–Ce
In this paper, a novel TiO2-based BiFeO3@Ce-doped catalysts. Ce-doped TiO2 was obtained in the same experi-
TiO2 core-shell nanocomposite was prepared through a mental conditions just like solution A. Meanwhile,
sol–gel method. The crystal structure, optical property BiFeO3@TiO2 was prepared in the absence of Ce(NO3)3 
and morphologies of the nanocomposite were system- 6H2O.
atically characterized and investigated as well. Finally, the
photocatalytic activity was measured via the degradation 2.3. Characterizations
of MO under both UV light and visible light illuminations.
Due to the synergistic effects of BiFeO3 and Ce, the The phase structure of the as-synthesized particles
photocatalytic activity of these core-shell nanoparticles were measured by X-ray diffraction (XRD) (Rigaku Co,
was dramatically enhanced. Additionally, under an exter- Tokyo, Japan) with a Bragg-Brentano geometry using Cu
nal magnetic field, the nanocomposite could be easily Ka radiation (l ¼1.5405 Å). The morphologies were
separated and recycled from the treated water. observed by a scanning electron microscope (SEM, Hitachi
S-4700, Japan) operated at 25 KV and a transmission
2. Experimental electron microscope (TEM, FEI TecnaiG220, USA) with an
accelerating voltage of 200 kV. An energy dispersive
2.1. Materials spectroscopy (EDS) was employed to examine the com-
position of the resultant nanocomposite photocatalysts.
All the chemicals used in the experiments were analy- The UV–visible diffuse reflectance spectra (DRS) were
tically pure and used as received without further purifica- obtained using a scan UV–vis spectrophotometer (Varian
tion. Tetrabutyl titanate (Ti(OC4H9 n)4, 98%), cerium (III) Cary 500) equipped with an integrating sphere assembly.
nitrate hexahydrate (Ce(NO3)3  6H2O, 99%), bismuth nitrate The photocatalytic performance was evaluated by the
pentahydrate (Bi(NO3)3  5H2O, 99%), and iron nitrate non- degradation of methylene orange (MO) in aqueous solu-
ahydrate (Fe(NO3)3  9H2O, 98.5%) were obtained from tion. The particles were placed in quartz vessels and were
Shanghai Chemical Corporation (Shanghai, China). Anhy- irradiated under UV light (Hg lamp 300 W with a primary
drous ethanol (C2H5OH, 99.7%, Jiuyi Chemical Reagent Co., wavelength of 254 nm, Philips, China) and visible light (Xe
Ltd., China), nitric acid (HNO3, 67%) and deionized water lamp, 500 W; Visible cutoff filter 4400 nm), the concen-
were used in all the experiments. tration of the dye solution was measured by a UV–vis
spectrometer (Perkin-Elmer Lambda 900 UV/vis/NIR,
2.2. Sample preparation Waltham, MA) at wavelength 500 nm. Magnetic measure-
ments were carried out at room temperature using a
The BiFeO3 was synthesized according to the afore- vibrating sample magnetometer (VSM, Lakeshore
published method [22]. The BiFeO3 precursor was pre- 7307, USA).
pared from acetic acid based solution, using bismuth
nitrate pentahydrate (5 mol% excess for the bismuth loss 3. Results and discussion
during the heating process), iron nitrate nonahydrate and
acetic acid as the solutes and 2-methoxyethanol as the 3.1. X-ray diffraction
solution. After the solution became transparent, the pre-
cursor was dried at 80 1C for about 24 h to obtain the XRD measurements are performed to identify the
BiFeO3 xerogel powder. Then the xerogel powder was crystalline phases synthesized by the sol–gel process.
annealed at 500 1C for 2 h and the BiFeO3 nanoparticles The obtained TiO2 and Ce–TiO2 samples consist of anatase
were obtained. as a unique phase are shown in Fig. 1. No characteristic
The coating of Ce–TiO2 on BiFeO3 is as follows: solu- peaks of cerium oxides are observed in Ce–TiO2 samples
tion A consisted of 10 ml tetrabutyl titanate, Ce for the low dosage (Ce 2%) and the extremely high
(NO3)3  6H2O in the required stoichiometry (0.5%, 1.0%, dispersion [21,23]. The BiFeO3@TiO2–Ce 2% sample shows
2.0% and 4.0%, which was the mole percentage of Ce that TiO2 and BiFeO3 phases coexistence at the present
element in the theoretical TiO2 samples) and 2 ml acet- synthesis conditions, and there are no extra peaks, indi-
ylacetone in 100 ml anhydrous ethanol, and the pH value cating that no new phase is formed between TiO2 and
was adjusted to 5–6 by dropwisely adding of HNO3 (70%). BiFeO3. Additionally, it is worth noting that the anatase
Then the mixture was stirred vigorously until the solution (101) crystal plane peak became broad and weak as
became transparent. Solution B contained 4.65 g BiFeO3 presented in the inset of Fig. 1. The average crystallite
(BiFeO3:TiO2 ¼ 1:2, molar ratio) and 150 ml anhydrous sizes of different samples are calculated by applying the
ethanol. In order to make the BiFeO3 dispersed uniformly, Scherrer formula on the anatase (101) diffraction peaks.
the solution B was treated under an ultrasonic bath for The values of pure TiO2, TiO2–Ce and BiFeO3@TiO2–Ce 2%
2 h. The Solution A was added dropwise into solution B were 11.0, 7.6 and 8.9 nm, respectively, suggesting that
under vigorous stirring. Then the mixture solution was cerium doping inhibits crystal growth. The smaller size is
under vigorous stirring at 40 1C for 5 h. The resulting beneficial to improve the photocatalytic activity.
290 L. Gong et al. / Materials Science in Semiconductor Processing 16 (2013) 288–294

3.2. SEM-EDS and TEM core-shell structure is well-supported by the results of


EDS analysis (Fig. 4(b)).
Scanning electron microscopy (SEM) images of the
products, Fig. 2, reveal that the pure BiFeO3 particles 3.3. UV–vis diffuse reflectance spectra
exhibits an agglomeration and irregular shapes (Fig. 2a),
compared to the uniform spherical particles with smooth The UV–vis diffuse reflectance spectra (DRS) of TiO2,
surface for BiFeO3@TiO2–Ce 1% (Fig. 2b). Both samples BiFeO3@TiO2, Ce–TiO2 and BiFeO3@TiO2–Ce 1% catalysts
have a narrow size distribution range (50–100 nm). The are shown in Fig. 5. The results demonstrate that the
size observed with SEM is extremely different from the broad band gap of pure TiO2 shows narrow absorption
XRD calculated size, indicating that the particles observed edges below 385 nm. Additionally, Ce–TiO2 samples exhi-
with SEM is not a single crystallite but the agglomerates bit a significant light absorption in the range from 400 to
of many single crystallites. The EDS patterns of bare 500 nm. Cerium ion has two possible oxidation states,
BiFeO3 and BiFeO3@TiO2–Ce are presented in Fig. 3. It is Ce3 þ and Ce4 þ . The absorption range from 400 to 500 nm
clearly seen that BiFeO3 mainly contains Bi, Fe and O cannot be ascribed to Ce4 þ because the 4f state of Ce4 þ
elements (Fig. 3(a)). However, except for Bi, Fe and O, does not contain any electron. Although Ce3 þ ion pos-
there are other elements such as Ti and Ce co-exist in sesses a single optically-active electron in the 4f1 orbital,
BiFeO3@TiO2–Ce sample (Fig. 3(b)). which has an excited state 2F7/2 and a ground state of 2F5/2,
The TEM images of typically BiFeO3@TiO2–Ce sample the light absorption in visible region also cannot be
are shown in Fig. 4. At higher magnification, the inter- attributed to 4f–4f transitions, because the characteristic
planar spacing of 0.354 nm and 0.397 nm in Fig. 4(a) 4f–4f transitions only be observed in the infrared spectral
corresponding to the (101) plane of anatase TiO2 and region [26]. Therefore, the absorption range from 400 to
(012) plane of perovskite BiFeO3, respectively, which 500 nm is due to 4f–5d transition of Ce3 þ [21,27]. The
agrees with the previously reported values [24,25]. The band gap energy of the sample could be calculated by the
inset of Fig. 4(a) displays the TEM image of the sample. It equation (ahu)n ¼ k(hu Eg), where a is the absorption
can be seen that BiFeO3 particles are surrounded by coefficient, k is the parameter related to the effective
anatase TiO2–Ce nanocrystallines. The existence of the masses associated with the valence and conduction bands,
n is 1/2 for a direct transition, hu is the absorption energy,
and Eg is the band gap energy [28]. Plotting (ahu)1/2; versus
hu based on the pattern response, as shown in Fig. 4(a),
gives the extrapolated intercept corresponding to the Eg
value as shown in Fig. 4(b). The band gaps of TiO2, Ce–TiO2,
BiFeO3@TiO2 and BiFeO3@TiO2–Ce are about 3.13 eV,
2.76 eV, 2.58 eV and 2.41 eV, respectively. The Ce–TiO2
sample has a lower Eg than pure TiO2. This is ascribed to
the formation of Ce 4f and oxygen defect levels. Ce 4f and
oxygen defect states are expected to hybridize and to form
an impurity band, leading to the decrease of bandgap.
Compared with the peaks of pure TiO2 sample, the
BiFeO3@TiO2 displays the Eg of 2.58 eV, which is exactly the
same as that of the pure BiFeO3 [29]. The spectral response in
the visible region of the BiFeO3@TiO2 is mainly attributed to
the photosensitization of BiFeO3. BiFeO3, as a semiconductor,
Fig. 1. XRD patterns of samples: (a) pure TiO2, (b) Ce–TiO2, (c) BiFe
has photocatalytical active in visible light for narrow band
O3@TiO2–Ce 2.0% and (d) BiFeO3. The inset shows the enlarged (101) gap. When the two semiconductors are in contact, electronic
peak of samples (a), (b) and (c). interaction occurs at their contact point. The spontaneous

Fig. 2. SEM images of: (a) pure BiFeO3 and (b) BiFeO3@TiO2–Ce.
L. Gong et al. / Materials Science in Semiconductor Processing 16 (2013) 288–294 291

Fig. 3. EDX analysis images of: (a) the pure BiFeO3 powder and (b) the core-shell structured BiFeO3@ TiO2–Ce nanoparticles.

Fig. 4. (a) A HRTEM micrograph of the BiFeO3@TiO2–Ce composite nanoparticles. the inset: a TEM micrograph of the BiFeO3@TiO2–Ce composite
nanoparticles and (b) the corresponding EDS spectra.

polarization in the ferroelectric domains of BiFeO3 leads to the samples in MO solution are kept in dark for 30 min to
band bending that transports photoinduced electrons and reach adsorption/desorption equilibrium. As shown in
holes in opposite directions. Because the photogenerated Fig. 6(a) and (b), no evident difference of the adsorption
carriers are separated by the internal polarization, they are percentage is found, suggesting that the adsorption ability
less likely to recombine, and this may enhance photocatalytic of different samples is not the key factor to determine the
efficiency [17]. The absorption of BiFeO3@TiO2–Ce samples in photodegradation rate. MO solution does not have any
the visible region is stronger than the undoped and doped self-degradation under the radiation of UV or visible light.
sample with Ce4 þ or modified with BiFeO3 alone. The It can be seen that the photocatalytic activity of BiFeO3@-
enhanced visible light absorption may be attributed to the TiO2 is much better than pure TiO2, which is owing to the
synergistic effect of Ce3 þ /Ce4 þ and BiFeO3. interfacial effect between BiFeO3 and TiO2. The photo-
electrons in BiFeO3@TiO2 could easily transfer from the
3.4. Photocatalytic activity analysis conduct band (CB) of BiFeO3 to the neighboring CB of TiO2.
Thus, the recombination between photoelectrons and holes
The photocatalytic performance of the samples is could be effectively inhibited, leading to a higher decom-
tested in the degradation of MO in aqueous suspensions position rate. The results also demonstrate that all cerium
under UV and visible light irradiation. Before the reaction, doped samples achieve a higher degradation than pure
292 L. Gong et al. / Materials Science in Semiconductor Processing 16 (2013) 288–294

Fig. 5. UV–vis absorption spectra (a) and the optical absorption edges Fig. 6. Photodegradation of MO with different catalysts under (a) UV
(b) of TiO2, Ce–TiO2, BiFeO3@TiO2 and BiFeO3@TiO2–Ce. light and (b) visible light irradiation, (a): blank; (b): TiO2; (c): TiO2–Ce;
(d):BiFeO3@TiO2; (e): BiFeO3@TiO2–Ce 0.5%; (f): BiFeO3@TiO2–Ce 1.0%;
(g): BiFeO3@TiO2–Ce 2.0%; (h): BiFeO3@TiO2–Ce 4.0%. The inset is the
MO color change of BiFeO3@TiO2–Ce 2.0%.
TiO2, and the MO degradation rate increase with the
increase of the concentration of cerium initially. However,
the photocatalytic activity decrease, as well as the concen- 2.5 eV like reported in the previous work [29], TiO2 is
tration of cerium, achieve a higher level. Under UV light considered as n-type semiconductor with a band gap of
irradiation, the optimal doping levels of cerium is 2%, and 3.2 eV, the details are shown in Fig. 7. Fe3 þ existence at
the MO color changes dramatically as shown in the inset of the interface of BiFeO3@TiO2 sample is reported pre-
Fig. 6(a); While under visible light illumination, the sample viously [18]. Therefore, electrons can be excited from
doped with 1% cerium shows the best performance. The the partially filled d orbitals of Fe3 þ to the Ti 3d orbitals
best degradation rate under UV light irradiation and visible in TiO2 under lower energy visible light condition. It is
light irradiation is 93.5% and 74.8%, respectively. These reported that the excitation of one semiconductor results
results are closely related to the smaller particle size, larger in an electron injection into the lower lying CB of the
surface area and higher concentration of hydroxy groups on second semiconductor [32]. The core-shell structured
the surface of TiO2. On the contrary, when the doping BiFeO3@TiO2–Ce systems consist of two semiconductors.
concentration exceeds the doping levels (2% under UV light BiFeO3 has relatively higher conduction band (CB), hence
and 1% under visible light), the photoinduced electrons will the photoinduced electrons on the BiFeO3 nanoparticle
be trapped by Ce4 þ in the interstitial lattice of TiO2 rather can transfer to the lower lying CB of TiO2 via the interface.
than transferred to the surface [23]. Meanwhile, many Subsequently, the electrons are promoted to the surface
introduced holes can be also trapped by Ce3 þ , the trap of TiO2 and then trapped by Ce4 þ /Ce3 þ . The cerium has
center become the recombination center, resulting in the better electron scavenging capacity than oxygen vacancies.
decrease of photoactivity of Ce doped TiO2 [30]. Besides, the Its unoccupied 4f orbital can accept electron and undergo
excessive cerium induce the agglomeration, which will strong interaction with the chromophore of dye molecule.
weaken the incident light, hence reduce the photoquantum Due to the presence of Ce3 þ , the labile oxygen vacancies
efficiency [31]. and mobility of bulk oxygen species are increased [33]. On
Clearly, BiFeO3@TiO2–Ce exhibits the highest photo- the other hand, the particle size of samples decreased
catalytic activity for the degradation of MO. In order to significantly. In this aspect, the larger surface area is
explain the enhanced visible light activity of BiFeO3@- efficient for absorbing O2 and H2O. The trapped electrons
TiO2–Ce, possible mechanisms are proposed. The bandgap absorb O2 molecules, and the holes accumulate in the
value of p-type semiconductor (BiFeO3) is assume as valence band of BiFeO3 react with the absorbed H2O which
L. Gong et al. / Materials Science in Semiconductor Processing 16 (2013) 288–294 293

which is mainly attributed to the high volume concentra-


tion of the non-magnetic coating layer of TiO2–Ce [12].
In addition, the magnetic separation ability of the
BiFeO3@TiO2–Ce (2%) nanoparticles is tested in water by
placing a magnet near the glass bottle. As displayed in
the inset of Fig. 8, under a magnetic field, the nanoparticles
can be easily attracted towards the magnet, demonstrating
that the resultant magnetic photocatalyst gives it excellent
performance on magnetic recycling.

4. Conclusions

In summary, novel BiFeO3@TiO2–Ce core-shell nanopho-


tocatalysts have been successfully prepared by a sol–gel
method. Photocatalytic experiments show that BiFeO3@
Fig. 7. Schematic diagram of proposed mechanism for the synergistic
TiO2–Ce has a good photocatalytic activity which may be
effect of Ce and BiFeO3 under visible light irradiation. In the figure, the attributed to the synergistic effect of Ce and BiFeO3. BiFeO3
Fermi level, conduction band edge and valence band edge are given by plays an important role in extending light absorption into
the symbols Ef, CB and VB, respectively. the visible region and Ce dopant can inhibit the recombina-
tion between photoinduced electron and hole pair. Because
of the ferromagnetism of BiFeO3 core, these core-shell
nanoparticles can be conveniently separated by using an
external magnetic field, which greatly extend their poten-
tial applications.

Acknowledgments

The authors would like to acknowledge the support of the


project funded by the Priority Academic Program Develop-
ment (PAPD) of Jiangsu Higher Education Institutions.

References

Fig. 8. M–H hysteresis loops of (a) pure BiFeO3 and (b) BiFeO3@TiO2–Ce [1] L.M. Jiang, G. Zhou, Materials Science in Semiconductor Processing
15 (2012) 108–111.
2.0% nanocomposites. The inset is a image of BiFeO3@TiO2–Ce 2.0%
[2] O. Vázquez-Cuchillo, A. Cruz-López, L.M. Bautista-Carrillo,
suspension under a magnetic field.
A. Bautista-Hernández, L.M. Torres Martı́nez, S. Wohn Lee, Research
on Chemical Intermediates 36 (2010) 103–113.
can be expressed in the Eqs. (1)–(3). The absorption spectra [3] X. Zhang, X.L. Zhao, H.J. Su, Journal of Chemical Engineering 28
(2001) 1241–1246.
and the photo utilization of TiO2 were broadened and
[4] T.C. Cheng, K.S. Yao, Y.H. Hsieh, L.L. Hsieh, C.Y. Chang, Materials and
improved by the BiFeO3 core. While Ce-doping retard the Design 31 (2010) 1749–1751.
recombination of the photoinduced electron–hole pair and [5] J.L. Gole, J.D. Stout, C. Burda, Y.B. Lou, X.B. Chen, Journal of Physical
increase the absorption of MO [34]. Therefore, the photo- Chemistry B 108 (2004) 1230–1240.
[6] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293
catalytic activity of co-modified TiO2 is efficiently (2001) 269–271.
improved. [7] H.X. Li, X.Y. Zhang, Y.N. Huo, J. Zhu, Environmental Science and
Technology 41 (2007) 4410–4414.
[8] G.S. Wu, J.P. Wang, D.F. Thomas, A. Chen, Langmuir 24 (2008)
Ce4 þ þe  -Ce3 þ (1) 3503–3509.
3þ 4þ [9] K. Matsubara, T. Tatsuma, Advanced Materials 19 (2007)
Ce þ O2 -O
2 þCe ð2Þ 2802–2806.
[10] M.S. Nahar, J. Zhang, K. Hasegawa, S. Kagaya, S. Kuroda, Materials
H2Oþ h -dOHþ H
þ þ
(3) Science in Semiconductor Processing 12 (2009) 168–174.
[11] D.L. Liao, C.A. Badour, B.Q. Liao, Journal of Photochemistry and
3.5. Analysis of magnetism of BiFeO3@TiO2–Ce Photobiology A: Chemistry 194 (2008) 11–19.
[12] Y.H. Fan, C.H. Ma, W. Li, Y.H. Yin, Materials Science in Semicon-
nanoparticles ductor Processing 15 (2012) 582–585.
[13] H.M. Yang, R.R. Shi, K. Zhang, Y.H. Hu, A.D. Tang, X.W. Li, Journal of
The magnetic properties of pure BiFeO3 and BiFeO3@ Alloys and Compounds 398 (2005) 200–202.
[14] M.S. Islam, Y. Kusumoto, M. Abdulla-Al-Mamun, Y.J. Horie,
TiO2–Ce (2%) nanoparticles are shown in Fig. 8. In case of
A.C. Photocatalytic, Catalysis Communications 16 (2011) 39–44.
BiFeO3 nanoparticles, an obvious ferromagnetic (FM) beha- [15] W.Y. Fu, H.B. Yang, M.H. Li, L.X. Chang, Q.J. Yu, J. Xu, et al., Materials
vior is observed and the saturated magnetization (Ms) is Letters 60 (2006) 2723–2727.
estimated to be 0.8 emu/g. A similar behavior for the [16] H.M. Xiao, X.M. Liu, S.Y. Fu, Composites Science and Technology 66
(2006) 2003–2008.
BiFeO3@TiO2–Ce composites is also observed. In contrast, [17] Y.L. Zhang, A.M. Schultz, P.A. Salvador, G.S. Rohrer, Journal of
Ms of the TiO2–Ce coated BiFeO3 decrease to 0.5 emu/g, Materials Chemistry 21 (2011) 4168–4174.
294 L. Gong et al. / Materials Science in Semiconductor Processing 16 (2013) 288–294

[18] S. Li, Y.H. Lin, B.P. Zhang, J.F. Li, C.W. Nan, Journal of Applied Physics [26] W.M. Yen, M. Raukas, S.A. Basun, W.V. Schaik, U. Happek, Journal of
105 (2009) 054310. Luminescence 69 (1996) 287–294.
[19] Q.G. Zeng, Z.M. Zhang, Z.J. Ding, Y. Wang, Y.Q. Sheng, Scripta [27] F.B. Li, Z.X. Li, M.F. Hou, K.W. Cheah, W.C.H. Choy, Applied Catalysis
Materialia 57 (2007) 897–900. A: General 285 (2005) 181–189.
[20] Y.N. Huo, J. Zhu, J.X. Li, G.S. Li, H.X. Li, Journal of Molecular Catalysis [28] M.J. Yoon, M.J. Seo, C.J. Jeong, J.H. Jang, K.S. Jeon, Chemistry of
A: Chemical 278 (2007) 237–243. Materials 17 (2005) 6069–6079.
[21] T.Z. Tong, J.L. Zhang, B.Z. Tian, F. Chen, D.N. He, M. Anpo, Journal of [29] X. Wang, Y. Lin, C. Zhang, J.Y. Bian, Journal of Sol–Gel Science and
Colloid and Interface Science 315 (2007) 382–388. Technology 60 (2011) 1–5.
[22] F. Gao, X.Y. Chen, K.B. Yin, S. Dong, Z.F. Ren, F. Yuan, et al., Advanced [30] S.I. Shah, W. Li, C.P. Huang, O. jung, C. Ni, National Academy of
Materials 19 (2007) 2889–2892. Sciences 99 (2002) 6482–6486.
[23] H. Liu, M.Y. Wang, Y. Wang, Y.G. Liang, W.R. Cao, Y. Su, Journal of [31] Z.F. Bian, J. Zhu, S.H. Wang, Y. Cao, X.F. Qian, H.X. Li, Journal of
Photochemistry and Photobiology A: Chemistry 223 (2011) Physical Chemistry C 112 (2008) 6258–6262.
157–164. [32] M.A. Fox, M.T. Dulay, Chemical Reviews 83 (1995) 341–357.
[24] C.L. Zhu, M.L. Zhang, Y.J. Qiao, G. Xiao, F. Zhang, Y.J. Chen, Journal of [33] B.M. Reddy, A. Khan, Y. Yamada, T. Kobayashi, S. Loridant, J.C. Volta,
Physical Chemistry C 14 (2010) 16229–16235. Journal of Physical Chemistry B 107 (2003) 5162–5167.
[25] U.A. Joshi, J.S. Jang, P.H. Borse, J.S. Lee, Applied Physics Letters 92 [34] Q.Q. Wang, S.H. Xu, F.L. Shen, Applied Surface Science 257 (2011)
(2008) 242106. 7671–7677.

You might also like