You are on page 1of 7

Solid State Sciences 15 (2013) 29e35

Contents lists available at SciVerse ScienceDirect

Solid State Sciences


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

Construction of 3D hierarchical SnO2 microspheres from porous nanosheets


towards NO decomposition
Thi Hang Le a, Quang Duc Truong b, Takeshi Kimura b, Huihui Li b, Chongsen Guo b, Shu Yin b, Tsugio Sato b,
Yong-Chien Ling a, *
a
b

Department of Chemistry, National Tsing Hua University, Hsinchu 30013, Taiwan


Institute of Multidisciplinary Research for Advanced Materials, Tohoku University, Sendai 980-8577, Japan

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 13 January 2012
Received in revised form
3 September 2012
Accepted 10 September 2012
Available online 18 September 2012

Three-dimensional (3D) hierarchical architectures are currently attracting worldwide interest owing to
their fascinating morphology-dependent properties and potential applications. Herein we constructed
SnO2 microspheres with 3D hierarchical ower-like architectures self-assembled with porous SnS2
nanosheets by a facile hydrothermal method with subsequent calcination. The chemical and physical
properties as well as photocatalytic application of SnO2 microspheres were investigated. The size and
morphology were examined with scanning electron microscopy and transmission electron microscopy.
The phase and crystalline structure were determined with powder X-ray diffraction. The UVeVis
absorption property was determined with UVeVis diffuse reectance. The photocatalytic activities
were evaluated with nitrogen monoxide (NO) decomposition under UVeVis light irradiation. The effects
of calcination temperature on morphology and NO decomposition were also studied.
Crown Copyright 2012 Published by Elsevier Masson SAS. All rights reserved.

Keywords:
SnO2
Morphology
Hierarchical structure
Hydrothermal method
NO decomposition

1. Introduction
The control of the crystalline structure, size, shape, morphology,
and surface chemistry plays an important role in material
science and engineering eld. Particularly, the morphological
control of metal oxides has drawn considerable attentions owing to
their versatile ability, efciency, and long-term durability in a wide
range of applications [1e4]. The three-dimensional (3D) hierarchical nanostructures formed by self-assembly of building blocks
may provide novel or optimal functional properties that are
not found in the building blocks. Therefore, efforts have been
devoted to the construction of 3D nanostructures based on the
various building blocks such as nanoparticles, nanorods, nanowires, nanobers, nanotubes, nanoparticles, and nanosheets. The
nanosheet-assembled nanostructures are of particular importance
owing to their unique chemical and physical properties as well as
potential applications in catalysis, gas sensing, and lithium-ion
batteries [1e4]. For example, complex 3D dendritic architectures
of TiO2 consisted of ultrathin nanosheets exhibited fast and
reversible lithium insertion/extraction for batteries [1]. Zhong and

* Corresponding author. Tel.: 886 3 5721484; fax: 886 3 5711082.


E-mail address: ycling@mx.nthu.edu.tw (Y.-C. Ling).

co-workers synthesized self-assembled 3D ower-like iron oxide


nanostructures showing an excellent ability to remove various
water pollutants [2]. Polshettiwar et al. have reported the synthesis
of silica nanospheres with high specic surface area, which served
as a potential absorbent or catalysis support [3]. Recently, 3D
ower-like hierarchical Fe3O4/Bi2O3 coreeshell architectures
exhibiting powerful visible-light photocatalytic activity towards
the degradation of rhodamine B have been synthesized [4].
Generally, the material physical and chemical properties are heavily
dependent on their structural features, triggering the development
of controllable synthetic methods towards functional applications.
Tin dioxide (SnO2) semiconductor is one of the most promising
candidates for molecular sensing and energy applications [5e12].
In particular, two dimensional (2D) SnO2 nanosheets have attracted
much attention owing to their remarkable receptivity variation in
gaseous environment, excellent lithium storage capacity and cycle
performance. The SnO2 nanosheets have been widely applied as
gas-sensing materials and anode materials in lithium-ion batteries
[5e12]. Despite of the successful growth of nanosheets, the
synthesis of 3D hierarchical architectures consisting of nanosheets,
which may exhibit excellent performance in gas sensing, energy
storage, and catalysis, is still challenging.
In this report, we demonstrate the synthesis of 3D hierarchical
ower-like SnO2 microspheres self-assembled by porous SnO2

1293-2558/$ e see front matter Crown Copyright 2012 Published by Elsevier Masson SAS. All rights reserved.
http://dx.doi.org/10.1016/j.solidstatesciences.2012.09.004

30

T.H. Le et al. / Solid State Sciences 15 (2013) 29e35

nanosheets using a facile hydrothermal method with subsequent


calcination of hydrothermally prepared SnS2 precursor. The
synthesis was performed by hydrothermal treatment of SnCl4 in
cysteine solution and calcinating the thus-obtained SnS2 precursor.
Up to date, SnO2 was investigated mainly as a component in
composite photocatalysts such as SnO2/TiO2 [13], SnO2/ZnO [14],
and SnO2/Pt/C [15]. The study of SnO2 alone as self-standing photocatalyst has not been considered appreciably. For the rst time,
the photocatalytic activities of 3D hierarchical ower-like SnO2
microspheres were evaluated by means of oxidative decomposition
of nitrogen monoxide (NO). The effects of calcination temperature
on morphology and NO decomposition were investigated.
2. Experimental section
2.1. Synthetic method
The synthesis was performed rst by dissolving 4 mmol of
cysteine (99%, Kanto Chemicals Co., Inc.) in 25 ml of deionized
water under continuous stirring, followed by adding 2 mmol of
SnCl4$5H2O (98%, Kanto Chemicals Co., Inc.), allowed for reaction
for 15 min. The resultant transparent solution was transferred into
a 50 ml Teon-lined autoclave heated at 190  C for 15 h in an
electric oven. The obtained SnS2 precipitate was allowed to cool
down to room temperature, collected by centrifugation, washed
with a mixture of deionized water ethanol several times to remove
the residual chloride ion and organic compounds, and nally dried
in air at 60  C for 1 day. The obtained SnS2 precursors were further
calcinated at 450  C (or 550  C) for 5 h to yield the SnO2 microspheres. For clear expression, all SnO2 microspheres referred
thereafter are calcinated at 450  C (denoted as SnO2 or SnO2-450)
unless otherwise specied.

(20  15  0.5 mm3) placed at the centre of a sealed plastic reactor


(373 cm3 volume). A testing gas containing 1 ppm NO in air
(balanced N2) was allowed to ow into the reactor at a ow rate of
200 cm3 min1. A 450 W high-pressure mercury lamp was used as
a light source. A Pyrex jacket cooled by running water was used to
cut-off l < 290 nm light. Additional removable lters used to cutoff l < 400 nm: and l < 510 nm light, respectively, were placed
between the lamp and the reactor if needed. The remained NO
concentration was determined using a NOx analyzer (Yanaco, ECL88A). The NO concentration was assured constant at w1 ppm in
dark before the photocatalytic decomposition experiment. The NO
concentration was monitored every 10 s during the irradiation. The
510 and 400 nm lters were respectively removed every 10 min to
stimulate irradiation with light of desirable wavelength range. The
photocatalysts studied include P25 (Degussa, used as reference),
SnS2 and SnO2 microspheres calcinated at 450 (and 550)  C. Details
about the experimental setup for NO decomposition have been
reported elsewhere [16].
3. Results and discussion
3.1. Characterization of SnO2 microspheres
Fig. 1A presents the XRD pattern of the SnO2 microspheres. All
indexed diffraction peaks can be indexed as (110), (101), (200) and
(211) from the known rutile-structured tetragonal phase single
crystalline SnO2 and matched well to those reported for rutile
tetragonal SnO2 (Joint committee on powder diffraction standards
JCDPS No. 77-0452). The absence of clear characteristic peaks from

(110)
(101)

SnO2 (PDF#77-0452)
(211)
(200)

20

30

40

50

60

2 Theta (degree)

B
Sn

Intensity / a.u

The phase and crystalline structure of SnS2 precursors and SnO2


microspheres was characterized using powder X-ray diffraction
(XRD; Rigaku RINT V-2200, 40 kV and 30 mA) with Cu Ka radiation
(l 1.5406 
A). Data were collected with 2qeq scanning mode at
a scan speed of 4 min1 and a step size of 0.02 . The surface
morphology of SnS2 precursors and SnO2 microspheres was
examined using eld-emission scanning electron microscopy (FESEM; Hitachi S-4800) at an accelerating voltage of 5 kV. The
composition of SnO2 was determined by an energy dispersive X-ray
(EDX) spectrometer (Hitachi S-4800). The SnO2 microspheres were
further examined using a transmission electron microscopy (TEM;
Hitachi H-7650, 100 kV, HRTEM HF-2000). The microspheres were
prepared by dispersing the microspheres in ethanol, dropping onto
Cu microgrids coated with a holey carbon lm, and nally evaporating the ethanol. The UVeVis diffuse reectance spectra (DRS) in
the range of 200e800 nm were recorded with a UVeVis 2450
(Shimadzu). The thermogravimetric analysis (TGA) and differential
thermal analysis (DTA) was measured with a DTG-60H (Shimazdu)
heated at a rate of 10 K min1 under N2 ow of 50 ml min1 up to
1000  C with Al2O3 as a reference. The adsorption and desorption
isotherms were measured with N2 at 77 K (Micromeritics ASAP
2010). The specic surface area was measured by Brunauere
EmmetteTeller (BET) method and BarretteJoynereHalenda (BJH)
pore size distribution (Micromeritics ASAP 2010).

Intensity / a. u.

2.2. Characterization

Sn

Sn
S
Sn Sn

2.3. Oxidative decomposition of NO

1.5
The photocatalytic activity of SnO2 microspheres was evaluated
with NO decomposition reaction. The experimental setup consisted
of a glass holder to attach photocatalyst in a working space

2.5

3.5

4.5

Energy (KeV)
Fig. 1. (A) XRD pattern and (B) EDX spectrum of SnO2 microspheres.

Fig. 2. SEM (AeC) and TEM (D, E) images of SnO2 microspheres. The inset in E shows the SEAD of the corresponding sample.

dV/dD/cm-3g-1nm-1

0.014

60

Adsorbed volume / cm-3g-1

SnS2 (JCDPS No. 89-2538) reveals rutiled-structure tetragonal


phase SnO2 as the main product. The peaks are relatively broadened
compared to those in JCDPS No. 77-0452, indicating the small
crystallite size nature of SnO2 microspheres. However, the EDX
spectrum in Fig. 1B shows signal of S, although the atomic ratio of S
is small. Nevertheless, the presence of sulphur indicates that the
nal product is rutiled-structure tetragonal phase SnO2 with small
quantity of SnS2.
Fig. 2AeC shows typical SEM images of the SnO2 microspheres
with different magnication. The low-magnication SEM image of
individual microsphere in Fig. 2A exhibits monodispersed microsphere with w7e10 mm in diameter consisted of 3D ower-like
nanostructures. The high-magnication SEM images in Fig. 2BeC
further reveal that the ower-like nanostructures are constructed
from numerous nanosheets. The TEM images (Fig. 2D) clearly show
the presence of nanosheets as building blocks for self-assembly of
SnO2 microspheres. The nanosheets are consisted of densely
assembled nanoparticles (NPs) with porous frameworks. The
selected area electron diffraction pattern (SAED) derived from small
NPs shown as inset in Fig. 2E demonstrates the polycrystalline
nature of the SnO2 nanosheets.

50
40

0.012
0.01
0.008
0.006
0.004
0.002
0
0

20

40

60

80

Pore diameter / nm

30
20
10
0
0

0.2

0.4

0.6

0.8

Relative Pressure (P/Po)


Fig. 3. Adsorption and desorption curves of SnO2 microspheres. The inset shows the
pore size distribution.

32

T.H. Le et al. / Solid State Sciences 15 (2013) 29e35

(101 )
(001)

(100)

Intensity / a.u.

(102)

(112)

SnS2 (PDF#89 -2358)

10

20

30

40

50

60

2 Theta (degree)
Fig. 4. XRD pattern of SnS2 precursor.

The porous nature of SnO2 nanostructures were further studied


with N2 gas adsorption/desorption analysis. The isotherm curve in
Fig. 3 can be categorized as IUPAC type III with a hysteresis loop in
the range of 0.4e1 P/P0. The BET specic surface area was found as
65.8 m2 g1. The pore size distribution demonstrates a maximum
at w5 nm and another maximum at >80 nm.
3.2. Characterization of SnS2 precursor
In order to understand the formation of hierarchical SnO2
microspheres from ower-like nanostructures, we also characterized

the SnS2 precursor obtained by hydrothermal treatment of SnCl4


with cysteine.
Fig. 4 presents the XRD pattern of the SnS2 precursor. All
indexed diffraction peaks can be indexed as (001), (100), (101),
(102) and (112) from the hexagonal phase single crystalline SnS2
and matched well to those reported for hexagonal SnS2 (Joint
committee on powder diffraction standards JCDPS#89-2358, cell
constants of a 3.638 
A and c 5.880 
A), revealing hexagonal
phase SnS2 was the major product.
Fig. 5AeC shows typical SEM images of the SnS2 precursor with
different magnication. The low-magnication SEM image of individual microsphere in Fig. 5A and B exhibits monodispersed
microsphere with w7e10 mm in diameter consisted of 3D ower-like
nanostructures. The high-magnication SEM image in Fig. 5C further
reveals that the ower-like nanostructures are constructed from
numerous nanosheets, which is similar to the SnO2 microspheres
geometry. The appearance of similar ower-like nanostructures in
SnO2 and SnS2 nanosheets suggests that the formation of SnO2
nanosheets directly from SnS2 nanosheets is highly plausible.
We therefore studied the conversion of SnS2 to SnO2 as well as the
effect of calcination temperature on the morphology of SnO2 nanosheets. The TGA curve in Fig. 6 reveals that the SnS2 undergoes two
stages of the weight-loss process. At the rst stage from room
temperature to 350  C, w5% of the total weight was lost, which is
attributed to the evaporation of adsorbed water and surface organic
residues. The second stage of weight-loss process occurred at 400  C
can be assigned to the conversion of SnS2 to SnO2. The DTA curve in
Fig. 6 reveals this calcination process was highly exothermal. The 18%
weight-loss was drastic (decreased from 5% to 23%) as shown in the
corresponding TGA curve. The numerical result is consistent with the
calculated value based on the assumption that the oxidation of SnS2
to SnO2 took place in this step. The corresponding weight-loss was
calculated using the formula of [M(SnS2)  M(SnO2)]/M(SnS2)

Fig. 5. SEM images of SnS2 precursor with different magnications.

T.H. Le et al. / Solid State Sciences 15 (2013) 29e35

Fig. 6. TGA and DTA curves of SnS2 microspheres.

(where M denotes the molecular weight) and yielded w0.175.


Moreover, the nal product, SnO2 free of S, can only be obtained at
850  C calcination temperature.
The biomolecular cystenine-assisted synthesis recently has been
used to prepare various sulde nanostructures [17e19]. In this
study, cysteine plays dual roles as sulphur source and structuredirecting agent. In particular, cysteine possesses carboxyl group
(COOH) and thiol group (SH), which may simultaneously coordinate with tin (IV) ion in aqueous solution. For example, tin (IV) ion
has two donoreacceptor bonds with sulphur atoms in thiol groups
and two covalent bonds with oxygen atoms in carboxyl groups,
facilitating the formation of a complex. At the elevated temperature
of hydrothermal reaction, the organic components decomposed,
resulting in the formation of SnS2. The SnS2 structure was stabilized
by the organic ligands selectively adsorbed onto its surface,
promoting the growth of 2D nanosheets. Because of their ultrathin
feature, these nanosheets are highly exible and can readily be self-

33

assembled into hierarchical architectures. The SnS2 nanosheet


structures can be preserved and easily converted into SnO2 nanosheets during the calcination without being collapsed.
The calcination temperature plays a critical role in determining
the crystalline structure as well as morphology of the SnO2 microspheres. Fig. 7 shows the typical SEM and TEM images of SnO2-550
microspheres prepared by calcinating the SnS2 precursor at 550  C.
The morphology appears similar to the SnO2-450 microspheres as
both are constructed from porous nanosheets. However, detailed
inspection of the morphology of nanosheets (Fig. 7A and B) reveals
there is slightly different, i.e., the ower-like nanostructures being
less uniform and unordered agglomeration as well as smaller in
size and numbers of nanosheets on each sphere, compared to that of
SnO2-450 (Fig. 2A and B). These factors might explain the lower
specic surface area for SnO2-550 in Table 1. The same conclusion
also holds when comparing the TEM images (Fig. 7C and D vs.
Fig. 2D and E). This may be due to the fact that the SnS2
nanoparticles have strong tendency to aggregate to each other in
order to minimize their total surface energy under higher calcination temperature.
3.3. Photocatalytic activity
Nitrogen monoxide (NO) is a well-known air pollutant produced
by incomplete combustion in air ambient such as in automobile
engines and fossil fuel power plants. NO is gradually oxidized in air
to form nitrogen dioxide (NO2), which is a toxic air pollutant [20e
27]. The study of NO decomposition by photocatalyst is therefore
becoming highly desirable. To the best of our knowledge, there has
been no report on photocatalytic decomposition of NO with SnO2
catalyst. In this report, the photocatalytic activity of the 3D owerlike SnO2 nanostructures towards NO decomposition was evaluated. This experiment was carried out by oxidative destruction of
NO under irradiation using a 450 W high-pressure mercury lamp

Fig. 7. SEM (A, B) and (C, D) TEM images of SnO2 microspheres prepared by calcination of SnS2 precursor at 550  C.

34

T.H. Le et al. / Solid State Sciences 15 (2013) 29e35

Table 1
BET specic surface area, NO decomposition (%) and ratio of various photocatalyst
irradiated with light of different wavelength range.
BET (m2 g1)

NO decomposition (%) and ratio (over P25)

l > 290 nm

l > 400 nm

l > 510 nm

P25
SnS2
SnO2-450a
SnO2-550a

52.6
50.3
65.8
61.4

59.5
52.0
57.4
57.2

30.2
15.3
11.2
11.5

8.0
3.1
4.0
4.2

(1.00)
(0.87)
(0.96)
(0.96)

(1.00)
(0.51)
(0.37)
(0.38)

(1.00)
(0.38)
(0.50)
(0.52)

The number represents the calcination temperature in  C.

providing visible light with l > 510 nm and l > 400 nm as well as
UV light with l > 290 nm, respectively. For comparison purpose,
a popular commercially available photocatalyst, P25 titania
(Degussa), was used as a reference photocatalyst.
Fig. 9 (and Table 1) illustrates that P25 exhibited higher NO
decomposition ability than SnS2 and SnO2 irrespective of the irradiation light wavelength. It can be seen that P25 exhibited high
photocatalytic activity under the irradiation of visible light
l > 510 nm and l > 400 nm. The higher activity of P25 titania,
presumably owing to the visible light induced activity of P25 being
related to the residual impurity level such as Cl leading to the
formation of Cl-doped TiO2 [25]. The SnS2 showed higher NO
decomposition ratio than that of SnO2 with irradiation light of
l > 400 nm. This improvement is expected owing to the narrow
band gap (2.35 eV) inherent in SnS2 leading to strong absorption in
the visible light region (Fig. 8). Surprisingly, all SnO2 exhibited the
high NO decomposition ability with irradiation light of l > 290 nm,
which was comparable to P25 and higher than SnS2. The
enhancement of the photocatalytic activity of SnO2 comparing to
SnS2 may be beneted from their larger BET specic surface area
and higher crystalline structure. In summary, the better photocatalytic activity of the present nanosheet-based microspheres may
be attributed to their unique physical and structural properties
with efcient and fast charge transfer ability, resulting from the
continuous conducting path of nanosheets and their hierarchical
assembly structures. The electron-hole pair separation became
easier and facilitated the reaction with surface adsorbed species,
resulting in increased photocatalytic activity. Moreover, the hierarchical porous structures can also improve the absorption of light
reections, increasing light-harvesting ability and consequently the
photocatalytic activity.
The mechanism of photocatalytic deNOx has been discussed
previously [21]. Under the irradiation of light, the electron-hole
pairs can be generated on the catalyst and form active oxygen
species such as oxygen radical (O
2 ) and oxohydroxyl radical

0.9
0.8
NO (ppm)

Photocatalyst

> 510 nm

0.7
0.6
0.5
0.4

0.3
0.2
0.1
0

> 400 nm

c
> 290 nm
d

10

15

20

25

30

Time (min)
Fig. 9. The remained NO concentration as a function of irradiation time with light of
different wavelength range using photocatalyst (a) P25 TiO2 (Degussa), (b) SnS2
microspheres, SnO2 microspheres prepared by calcination at (c) 450  C, and (d) 550  C.

(OOH). The NO can react with the reactive oxygen radicals,


molecular oxygen, and water in the air to produce HNO2 and/or
HNO3. About 20% of NO directly decomposed to nitrogen and
oxygen [21]. The decomposition of NO with various photocatalysts
such as TiO2 [16,20e22], SrTiO3 [23e26], and ZnO [27] has been
reported. In this report, the deNOx abilities by SnS2 and SnO2 were
evaluated for the rst time, which might open a new direction to
study the effect of dopants and morphology on the photocatalytic
activity of SnS2 and SnO2 materials.
4. Conclusion
The porous SnO2 nanosheets with hierarchical ower-like
nanostructures have been successfully synthesized by a facile
hydrothermal method followed by calcination treatment. The SnO2
microspheres with w7e10 mm in diameter constructed from
numerous porous nanosheets. The 3D hierarchical ower-like SnO2
microspheres exhibited comparable photocatalytic activity towards
NO decomposition with respect to that of commercial P25 TiO2.
Acknowledgements
Financial support by the National Science Council (NSC98-2113M-007-016-MY3 and NSC99-2627-M-007-01) of Taiwan is gratefully acknowledged.

Reflectance / a.u.

References

a
b
c
200

300

400

500

600

700

800

900

Wavelength (nm)
Fig. 8. UVeVis diffuse reectance spectra of (a) SnS2 precursor, and SnO2 microspheres
calcinated at (b) 450  C and (c) 550  C.

[1] J.S. Chen, Y.L. Tan, C.M. Li, Y.L. Cheah, D. Luan, S. Madhavi, F.Y.C. Boey,
L.A. Archer, X.W. Lou, J. Am. Chem. Soc. 132 (2010) 6124.
[2] L.-S. Zhong, J.-S. Hu, H.-P. Liang, A.-M. Cao, W.-G. Song, L.-J. Wan, Adv. Mater.
18 (2006) 2426.
[3] V. Polshettiwar, D. Cha, X. Zhang, J.M. Basset, Angew. Chem. Int. Ed. 122
(2010) 9846.
[4] Y. Wang, S. Li, X. Xing, F. Huang, Y. Shen, A. Xie, X. Wang, J. Zhang, Chem. Eur. J.
17 (2011) 4802.
[5] H.B. Wu, J.S. Chen, X.W. Lou, H.H. Hng, J. Phys. Chem. C 115 (2011) 24605.
[6] Y. Masuda, K. Kato, J. Cryst. Growth 311 (2009) 593.
[7] K. Sakaushi, Y. Oaki, H. Uchiyama, E. Hosono, H.S. Zhou, H. Imai, Small 6 (2010) 776.
[8] S.J. Ding, X.W. Lou, Nanoscale 3 (2011) 3586.
[9] X.M. Yin, C.C. Li, M. Zhang, Q.Y. Hao, S. Liu, L.B. Chen, T.H. Wang, J. Phys. Chem.
C 114 (2010) 8084.
[10] Y. Li, Y.Q. Guo, R.Q. Tan, P. Cui, W.J. Song, Mater. Lett. 63 (2009) 2085.
[11] C. Wang, Y. Zhou, M. Ge, X. Xu, Z. Zhang, J.Z. Jiang, J. Am. Chem. Soc. 132
(2010) 46.
[12] S.J. Ding, D.Y. Luan, F.Y.C. Boey, J.S. Chen, X.W. Lou, Chem. Commun. 47
(2011) 7155.
[13] L.R. Hou, C.Z. Yuan, Y. Peng, J. Hazard. Mater. 139 (2007) 310.

T.H. Le et al. / Solid State Sciences 15 (2013) 29e35


[14] W.W. Wang, Y.J. Zhu, L.X. Yang, Adv. Funct. Mater. 17 (2007) 59.
[15] A. Ignaszak, C. Teo, S. Ye, E. Gyenge, J. Phys. Chem. C 114 (2010) 16488.
[16] S. Yin, H. Hasegawa, D. Maeda, M. Ishitsuka, T. Sato, J. Photochem. Photobiol.
A: Chem. 163 (2004) 1.
[17] B. Zhang, X.C. Ye, W.Y. Hou, Y. Zhao, Y. Xie, J. Phys. Chem. B 110 (2006)
8978.
[18] F. Zuo, S. Yan, B. Zhang, Y. Zhao, Y. Xie, J. Phys. Chem. C 112 (2008) 2831.
[19] H. Tong, Y.J. Zhu, L.X. Yang, L. Li, L. Zhang, J. Chang, L.Q. An, S.W. Wang, J. Phys.
Chem. C 111 (2007) 3893.
[20] Y. Aita, M. Komatsu, S. Yin, T. Sato, J. Solid State Chem. 177 (2004) 3235.

35

[21] S. Yin, Y. Aita, M. Komatsu, J. Wang, Q. Tang, T. Sato, J. Mater. Chem. 15


(2005) 674.
[22] H.H. Li, S. Yin, T. Sato, Appl. Catal. B: Environ. 106 (2011) 586.
[23] J. Wang, S. Yin, Q. Zhang, F. Saito, T. Sato, J. Mater. Chem. 13 (2003) 2348.
[24] J. Wang, S. Yin, M. Komatsu, Q. Zhang, F. Saito, T. Sato, Appl. Catal. B: Environ.
52 (2004) 11.
[25] U. Sulaeman, S. Yin, T. Sato, Appl. Phys. Lett. 97 (2010) 103102.
[26] U. Sulaeman, S. Yin, T. Sato, Appl. Catal. B: Environ. 102 (2011) 286.
[27] T. Long, X. Dong, X. Liu, J. Liu, S. Yin, T. Sato, Res. Chem. Intermed. 36
(2010) 61.

You might also like