Você está na página 1de 11

Journal of Solid Mechanics Vol. 2, No.

1, 2008

and Materials
Engineering
Retraction: Structural and Photovoltaic
Properties of High Surface Area
One-dimensional Nanostructured TiO2
[Journal of Solid Mechanics and Materials
Engineering, 2007, 1(4), 459–468]*
Sorapong PAVASUPREE**,***, Supachai NGAMSINLAPASATHIAN***,
Yoshikazu SUZUKI*** and Susumu YOSHIKAWA***
**
Department of Materials and Metallurgical Engineering, Faculty of Engineering,
Rajamangala University of Technology Thanyaburi, Klong 6, Pathumthani, 12110, Thailand.
***
Institute of Advanced Energy, Kyoto University, Uji, Kyoto 611-0011, Japan.
E-mail: s-yoshi@iae.kyoto-u.ac.jp, sorapong@iae.kyoto-u.ac.jp, sorapongp@yahoo.com

The article authored by Sorapong Pavasupree, et al., and titled “Structural and
Photovoltaic Properties of High Surface Area One-dimensional Nanostructured
TiO2”, Journal of Solid Mechanics and Materials Engineering, 2007, 1(4),
459–468, has been found to be a duplicate publication of their article on another
journal, and thus retracted by the Editor. We apologize any inconvenience this may
have caused.

*Received 11 Dec., 2007


[DOI: 10.1299/jmmp.2.R01]

R01
Journal of Solid Mechanics Vol. 1, No. 4, 2007

and Materials
Engineering
Structural and Photovoltaic Properties
of High Surface Area One-dimensional
Nanostructured TiO2*
Sorapong PAVASUPREE**,***, Supachai NGAMSINLAPASATHIAN***,
Yoshikazu SUZUKI*** and Susumu YOSHIKAWA***
**
Department of Materials and Metallurgical Engineering, Faculty of Engineering,
Rajamangala University of Technology Thanyaburi, Klong 6, Pathumthani, 12110, Thailand.
***
Institute of Advanced Energy, Kyoto University, Uji, Kyoto 611-0011, Japan.
E-mail: s-yoshi@iae.kyoto-u.ac.jp, sorapong@iae.kyoto-u.ac.jp, sorapongp@yahoo.com

Abstract
One-dimensional nanostructured TiO2 were synthesized by hydrothermal method at
130-170 oC for 12-72 h. The samples characterized by XRD, SEM, TEM, SAED,
HRTEM, and BET surface area. The nanorods/nanoparticles TiO2 with mesoporous
structure showed higher photocatalytic activity (I3- concentration) than the
nanorods TiO2, nanofibers TiO2, mesoporous TiO2, and commercial TiO2 (ST-01,
P-25, JRC-01, and JRC-03). The solar energy conversion efficiency (η) of the cell
using nanorods/nanoparticles TiO2 with mesoporous structure was about 7.12 %
with Jsc of 13.97 mA/cm2, Voc of 0.73 V and ff of 0.70; while η of the cell using
P-25 reached 5.82 % with Jsc of 12.74 mA/cm2, Voc of 0.704 V and ff of 0.649.
The η of the cell using nanorods/nanoparticles TiO2 (from the nanosheet calcined at
450 oC for 2 h) with mesoporous structure was about 7.08 % with Jsc of 16.35
mA/cm2, Voc of 0.703 V and ff of 0.627.

Key words: One-Dimensional, Nanofibers, Nanorods, Nanoparticles, Nanosheets,


TiO2, Photocatalytic Activity, Dye-Sensitized Solar Cell.

1. Introduction
The synthesis and characterization of one-dimensional (1-D) nanostructured
materials (nanotubes, nanorods, and nanowires) have received considerable attention due to
their unique properties and novel applications [1-4]. Much effort has concentrated on the
important metal oxides such as TiO2, SnO2, VO2, and ZnO [1-9]. Amoung them, TiO2 and
TiO2-derived materials are of importance for utilizing solar energy and environmental
purification. TiO2 has been widely used for various applications such as a semiconductor in
dye-sensitized solar cell, water treatment materials, catalysts, gas sensors, and so on [10-29].
In our previous works, nanofibers TiO2 were synthesized by hydrothermal and post
heat-treatments from natural rutile sand, however, nanofibers TiO2 had rather low surface
area (10-20 m2/g) [20-21]. The author’s group preliminary reported that the
nanowires/nanoparticles composite structure showed better photovoltaic conversion effect
in dye-sensitized solar cell [22].
In this study, nanorods/nanoparticles TiO2 with mesoporous structure (with much
higher surface area, 203 m2/g) has been synthesized, which shows high photocatalytic
activity and high performance in dye-sensitized solar cell. The nanorods/nanoparticles TiO2
from high surface area nanosheets (642 m2/g) also prepared by hydrothermal method for
*Received 9 Nov., 2006
dye-sensitized solar cells applications.
(No. ASMP0661, 0662)
[DOI: 10.1299/jmmp.1.459]

459
Journal of Solid Mechanics Vol. 1, No. 4, 2007
and Materials Engineering

2. Experimental
2.1 Synthesis
2.1.1 Nanorods/nanoparticles TiO2 with mesoporous structure
Titanium (IV) butoxide (Aldrich) was mixed with the same mole of acetyacetone
(ACA, Nacalai Tesque, Inc., Japan) to slowdown the hydrolysis and the condensation
reactions [11-13]. Subsequently, distilled water 40 ml was added in the solution, and the
solution was stirred at room temperature for 5 min. After kept stirring, ammonia aqueous
solution 28 % (Wako Co., Ltd., Japan) 30 ml was added in the solution, then the solution
was put into a Teflon-lined stainless steel autoclave and heated at 150 oC for 20 h with
stirring condition [23]. After the autoclave was naturally cooled to room temperature, the
obtained product was washed with HCl aqueous solution, 2-propanal and distilled water for
several times, followed by drying at 100 oC for 12 h.
2.1.2 Nanorods TiO2
The nanorods TiO2 were synthesized by using the same route of 2.1.1 but
difference in the hydrothermal temperature at 170 oC for 72 h.
2.1.3 Nanofibers TiO2
The nanofibers TiO2 were synthesized by hydrothermal method (150 oC for 72 h)
using natural rutile sand as the starting material and calcined at 700 oC for 4 h (prepared as
ref. 20-21).
2.1.4 Mesoporous TiO2
The mesoporous TiO2 were synthesized by surfactant-assisted sol-gel method at
80 oC for 5 days then calcined at 250 oC for 24 h (washed by 2-propanal) and 400 oC for 4 h
(prepared as ref 15-19).
2.1. 5 Nanosheet TiO2
The nanorods TiO2 were synthesized by using the same route of 2.1.1 but
difference in the hydrothermal temperature at 130 oC for 12 h [24].

2.2 Characterization
The crystalline structure of the samples was evaluated by X-ray diffraction (XRD,
RIGAKU RINT 2100). The microstructure of the prepared materials was analyzed by
scanning electron microscopy (SEM, JEOL JSM-6500FE), transmission electron
microscopy (TEM, JEOL JEM-200CX), and selected-area electron diffraction (SAED). The
Brunauer-Emmett-Teller (BET) specific surface area was determined by the nitrogen
adsorption (BEL Japan, BELSORP-18 Plus).

2.3 Photocatalytic activity measurement


The photocatalytic activity was measured through the formation rate of I3- due to
the oxidation photo-reaction of I- to I2 in excess I- conditions [11-13, 19]. A reaction system
was set up by adding 50 mg of a sample powder into 10 ml of 0.2 M of potassium iodide
aqueous solution then stirred and irradiated with UV light (15 W, Vilber Lourmat VL-115L,
with a maximum emission at about 365 nm) at room temperature. After the irradiation of 15,
30, 45, and 60 min, the suspension was withdrawn and centrifuged. After the clear
supernatant was diluted ten times, the concentration of liberated I3- ions was monitored by
the absorbance at 288 nm, using an UV–Vis spectrophotometer (Shimadzu UV 2450). The
molar extinction coefficient was determined to be 4.0 x 104 (cm mol/l)-1. No I3- formation
was observed when the experiments were conducted in the dark or in the absence of the
TiO2 samples. For reference, 4 difference commercially available nanoparticles TiO2
powders, ST-01 (Ishihara Sangyo Kaisha, Ltd., Japan), P-25 (Nippon Aerosil Co., Ltd.,
Japan), JRC-01 and JRC-03 (The Catalysis Society of Japan) were tested.

460
Journal of Solid Mechanics Vol. 1, No. 4, 2007
and Materials Engineering

2.4 Dye-sensitized solar cell measurement


TiO2 electrodes were prepared as follows; 1 g of TiO2 powder was mixed with 0.1
mL of ACA, and was ground mechanically. During vigorous stirring, 5 ml of mixture of
water and ethanol (1:1, in vol %) was added and 0.4 ml of polyoxethylene (10) octylphenyl
ether (Triton x-100) was added to facilitate spreading of the paste on the substrate. The
obtained colloidal paste was coated on fluorine-doped SnO2 conducting glass (FTO, sheet
resistance 15 Ω/□, Asahi glass Co., Ltd.) by squeegee technique. After coating, each layer
was dried at room temperature and then annealed at 400 oC for 5 min. The coating process
was repeated to obtain thick films. The resulting films were sintered at 450 oC for 2 h in air.
Sintered TiO2 electrodes were soaked in 0.3 mM of ruthenium (II) dye (known as N719,
Solaronix) in a t-butanol/ acetonitrile (1:1, in vol %) solution. The electrodes were washed
with acetonitrile, dried, and immediately used for measuring photovoltaic properties. The
electrolyte was composed of 0.6 M dimethylpropylimidazolium iodide, 0.1 M lithium
iodide (LiI), 0.05 M iodide (I2), and 0.5 M 4-tert-butylpyridine in acetonitrile.
The thickness of the TiO2 films was measured with a Tencor Alpha-step Profiler.
Photocurrent-voltage curve was measured under simulated solar light (CEP-2000,
Bunkoh-Keiki, AM 1.5, 100 mW/cm2). The light intensity of the illumination source was
calibrated by using a standard silicon photodiode (BS520, Bunkoh-Keiki).
The amount of adsorbed dye was determined by desorbing the dye from the
titania surface into a mixed solution of 0.1 M NaOH and ethanol (1:1 in volume fraction)
and measuring its absorption spectrum. The concentration of adsorbed dye was analyzed by
UV-vis spectrophotometer (UV-2450 SHIMADZU).

3. Results and discussion


3.1 Characterization results
Fig. 1 (a) shows the X-ray diffraction pattern of the as-synthesized
nanorods/nanoparticles TiO2, The peaks were rather sharp, which indicated the obtained
TiO2 had relatively high crystallinity, and attributable to the anatase phase.
Fig. 1 (b) gives the nitrogen adsorption isotherm and the pore size distribution of
the as-synthesized nanorods/nanoparticles TiO2. The isotherm shows a typical IUPAC type
IV pattern with sharp inflection of nitrogen adsorbed volume at P/P0 about 0.80 (type H2
hysteresis loop), indicating the existence of mesopores. The pore size distribution of the
sample, as shown in the inset of Fig. 1 (b), showed that the nanorods/nanoparticles TiO2 had
average pore diameter about 7-12 nm. The BET surface area and pore volume of sample are
about 203 m2/g and 0.655 cm3/g, respectively.
(101)

450
(a) 250
(b)
Adsorbed amount / ml(STP) g-1

400
-1

200
g
-1

350
dv/dR / mm nm
Intensity (a.u.)

150
3

300
100
(004)

(200)
(105)

250
(201)

50
(204)

200
(116)
(220)
(215)

0
1 10 100
150 Pore radius / nm

100
10 20 30 40 50 60 70 80 adsorption
50
C u -K α 2 θ (degree) desorption

0
0 0.5 1
Relative Pressure P/P 0

Fig. 1. (a) X-ray diffraction pattern of the as-synthesized nanorods/nanoparticles TiO2. (b)
Nitrogen adsorption isotherm pattern of the as-synthesized nanorods/nanoparticles TiO2,
and the pore size distribution of sample with pore diameter about 7-12 nm (inset).

461
Journal of Solid Mechanics Vol. 1, No. 4, 2007
and Materials Engineering

Fig. 2 (a-b) show SEM images of the as-synthesized nanorods/nanoparticles TiO2,


indicating the rods-like morphology. In addition, it’s TEM (Fig. 2 (c)) clearly shows not
only nanorods but also nanoparticles. The nanorods in the composite powder had diameter
about 10-20 nm and the lengths of 100-200 nm, the nanoparticles had diameter about 5-10
nm. The electron diffraction pattern shown in the inset of (Fig. 2(c)) supported that the
nanorods/nanoparticles composite was anatase-type TiO2. The lattice fringes of the
nanorods and the nanoparticles appearing in the image (d = 0.35 nm) also allowed for the
identification of the anatase phase (Fig. 2 (d)). HRTEM images of nanorods/nanoparticles
with clear lattice fringes, again confirming its high crystallinity.
Fig. 3 show the summarized results (SEM, TEM, SAED, HRTEM, and BET
surface area) of the prepared TiO2 (nanorods TiO2, nanofibers TiO2, and mesoporous TiO2).
The BET surface area of the nanorods TiO2 (Fig. 3 (a), anatase structure with
diameter 60-80 nm and the lengths of 300-600 nm), the nanofibers TiO2 (Fig. 3 (b), anatase
structure with diameter 20-100 nm and the lengths of several ten micrometers), and
mesoporous TiO2 (Fig. 3 (c), anatase structure with particles diameter 7-15 nm) were about
19, 10, and 80 m2/g, respectively. The BET surface area of the commercial TiO2, ST-01,
(anatase structure with particles diameter 4-5 nm), P-25 (anatase 70 % and rutile 30 %
structure with particles diameter 30-50 nm), JRC-01 (anatase structure with particles
diameter 15-30 nm), JRC-03 (rutile structure with particles diameter 20-50 nm) were about
284, 56, 71, and 48 m2/g, respectively (Fig. 3 (d-f)).

(a) (b)

(c) (d)

Fig. 2. (a-b) SEM, (c) TEM, SAED, and (d) HRTEM images of the as-synthesized
nanorods/nanoparticles TiO2.

462
Journal of Solid Mechanics Vol. 1, No. 4, 2007
and Materials Engineering

SEM images TEM, SAED images

(a) Nanorods TiO2 (19 m2/g) (d) P-25 (57 m2/g)

(b) Nanofibers TiO2 (10 m2/g) (e) JRC-01 (71 m2/g)

(c) Mesoporous TiO2 (80 m2/g) (f) JRC-03 (48 m2/g)

Fig. 3. SEM, TEM, SAED, and BET surface area results of the prepared and commercial
TiO2. (a) nanorods TiO2 prepared by hydrothermal at 170 oC for 72 h. (b) nanofibers TiO2
prepared by hydrothermal at 150 oC for 72 h as ref. 20-21. (c) mesoporous TiO2 prepared by
sol-gel at 80 oC for 7 days as ref. 15-16. (d-f) the commercial nanoparticles TiO2 (P-25,
JRC-01, and JRC-03).
3.2 Photocatalytic activity results
The I3- concentration at 60 min of the irradiation period of the
nanorods/nanoparticles TiO2 were about 3.02 x 10-4 M (Fig. 4 (a)), which is higher than that
of other synthesized powders (nanorods TiO2, nanofibers TiO2, mesoporous TiO2) and also
that of four commercially available titania nanomaterials, ST-01, P-25, JRC-01, and JRC-03
which exhibit I3- concentration about 2.68 x 10-4 M, 1.50 x 10-4 M, 0.66 x 10-4 M, and 0.25
x 10-4 M, respectively. The photocatalytic activity was almost proportional to the BET
surface area (Fig. 4 (b)), indicating that adsorption of I– on titania surface was rate
determining step in the initial stage of the reaction (before 15 min) [11-13]. In our previous
works, it is obviously revealed that the introduction of mesopore into titania photocatalyst
substantially improved the photocatalytic performance [17, 19].

463
Journal of Solid Mechanics Vol. 1, No. 4, 2007
and Materials Engineering

(a)

(b)
Concentration of I 3 x 10 (M)

3
-4
-

1
Prepared TiO2
Commercial TiO2

0
0 50 100 150 200 250 300
2
BET surface area (m /g)

Fig. 4. (a) Photocatalytic activity (I3- concentration) of the nanorods/nanoparticles TiO2 with
mesoporous structure (prepared by this study), the mesoporous TiO2 calcined at 250 oC for
24 h and 400 oC for 4 h (prepared as ref. 15 -19), the nanorods TiO2 (prepared by this study),
the nanofibers TiO2 (prepared as ref. 20-21), and commercial TiO2 (ST-01, P-25, JRC-01,
and JRC-03). (b) The comparison between I3- concentration-BET surface area of the
prepared TiO2 (nanorods/nanoparticles, mesoporous, nanorods, and nanofibers) and the
commercial TiO2 (ST-01, P-25, JRC-01, and JRC-03).

3.3 Dye-sensitized solar cell results


Fig. 5 shows comparison between photocurrent–voltage characteristics of the cell
using nanorods/nanoparticles TiO2 with mesoporous structure (thinckness = 8.3 µm) and
P-25 (thinckness = 13.8 µm). The solar energy conversion efficiency of the cell using
nanorods/nanoparticles TiO2 with mesoporous structure was about 7.12 % with Jsc of 13.97
mA/cm2, Voc of 0.73 V and ff of 0.70; while η of the cell using P-25 reached 5.82 % with
Jsc of 12.74 mA/cm2, Voc of 0.704 V and ff of 0.649. High current density and efficiency
might be attributed to higher amount of adsorbed dye (11.44 x 10-8 mol/cm2 for
nanorods/nanoparticles TiO2 and 5.68 x 10-8 mol/cm2 for P-25), owing to larger surface area
of the nanorods/nanoparticles TiO2 (Table 1). The η of the cell using nanorods/nanoparticles
TiO2 (from the nanosheet calcined at 450 oC for 2 h) with mesoporous structure was about
7.08 % with Jsc of 16.35 mA/cm2, Voc of 0.703 V and ff of 0.627 (Fig. 6-7). The η of

464
Journal of Solid Mechanics Vol. 1, No. 4, 2007
and Materials Engineering

dye-sensitized solar cell fabricated from the high surface area one-dimensional
nanostructured TiO2 (prepared by this study) higher than the η of dye-sensitized solar cell
fabricated from the one-dimensional nanostructured TiO2 using the TiO2 (ST-01)-derived
nanotubes (η = 3.1 %) nanowires (η = 6.01 %) [22, 25-27], from the TiO2 (P-25)-derived
nanotubes (η = 2.9 %) [28]. The Jsc and Voc also higher than recently reported of the TiO2
nanorods as additive to TiO2 film for improvement in the performance of dye-sensitized
solar cells [29].
The nanorods/nanoparticles TiO2 and nanosheets TiO2 were found to be superior
solar energy conversion efficiency to commercial nanoparticles P-25. The optimal thickness
of this electrode should be further investigated. This material possesses high potential for
further development in the application of dye-sensitized solar cells.

16

Photocurrent density (mA/cm 2 )


14

12
10 Nanorods/nanoparticles TiO2
2
Jsc = 13.97 mA/cm
8 Voc = 0.73 V, FF = 0.70
η = 7.12 %, thinkness = 8.3 µm
6
P-25
4 2
Jsc = 12.74 mA/cm
2 Voc = 0.704 V, FF = 0.649
η = 5.82 %, thinkness = 13.8 µm
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Voltage (V)

Fig. 5. The comparison between photocurrent-voltage characteristic of a typical dye


sensitized solar cells fabricated by the nanorods/nanoparticles TiO2 with mesoporous
structure and commercial nanoparticles TiO2 (P-25).

Table 1 BET surface area after calcined at 450 oC for 2 h and amount of adsorbed dye of
nanorods/nanoparticles TiO2, P-25, and nanosheets TiO2.

Samples BET surface area after calcined Amount of adsorbed


at 450 oC for 2 h (m2/g) dye (mol/cm2)
Nanorods/nanoparticles TiO2 100 11.44 x 10-8
P-25 54 5.68 x 10-8
Nanosheets TiO2 134 13.94 x 10-8

(a) (b)

Fig. 6. TEM images of (a) the as-synthesized nanosheets TiO2 and (b) the
nanorods/nanoparticles TiO2 from the nanosheets calcined at 450 oC for 2 h.

465
Journal of Solid Mechanics Vol. 1, No. 4, 2007
and Materials Engineering

18

Photocurrent density (mA/cm 2 )


16
14
12
Nanosheet-like TiO2
10 Jsc = 16.35 mA/cm
2

Voc = 0.703 V, FF = 0.627


8
η = 7.08 %, thinkness = 8.4 µm
6
P-25
4 Jsc = 12.74 mA/cm
2

2 Voc = 0.704 V, FF = 0.649


η = 5.82 %, thinkness = 13.8 µm
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Voltage (V)

Fig. 7. The comparison between photocurrent-voltage characteristic of a typical dye


sensitized solar cells fabricated by the nanosheet TiO2 and commercials nanoparticles TiO2
(P-25).

4. Conclusion
In summary, the nanorods had diameter about 10-20 nm and the lengths of
100-200 nm, the nanoparticles had diameter about 5-10 nm. Nanorods/nanoparticles TiO2
with mesoporous structure showed higher photocatalytic activity (I3- concentration) than the
nanorods TiO2, nanofibers TiO2, mesoporous TiO2, and commercial TiO2 (ST-01, P-25,
JRC-01, and JRC-03). The η of the cell using nanorods/nanoparticles TiO2 was about
7.12 % while η of the cell using P-25 reached 5.82 %. The η of the cell using
nanorods/nanoparticles TiO2 (from the nanosheet calcined at 450 oC for 2 h) was about
7.08 %. These materials are promising for chemical and energy-related applications such as
photocatalytic activity materials, a semiconductor in dye-sensitized solar cell.

Acknowledgements
The authors would like to express gratitude to Prof. S. Isoda and Prof. H. Kurata,
Institute for Chemical Research, Kyoto University for the use of TEM apparatus Prof. T.
Yoko, Institute for Chemical Research, Kyoto University for the use of XRD equipment.
Prof. Mochizuki at AIST for the kind supply of Pt counter electrode. We are also grateful to
theGeomatec Co. Ltd. for providing a part of conducting glass. This work was supported by
grant-in-aids from the Ministry of Education, Science Sports, and Culture of Japan under
the 21 COE program, the Nanotechnology Support Project, NEDO under high-performance
dye-sensitized solar cell project, and from JSPS research fellow.

References
[1] Rao, C. N. R. and Nath, M., Inorganic nanotubes, Dalton Transactions, Vol. 1, 2003, pp. 1-24.
[2] Patzke, G. R., Krumeich, F. and Nesper, R., Oxidic nanotubes and nanorods - anisotropic
modules for a future nanotechnology, Angewandte Chemie International Edition, Vol. 41,
2002, pp. 2446-2461.
[3] Huang, M., Mao, S., Feick, H., Yan, H., Wu, Y., Kind, H., Weber, E., Russo, R. and Yang, P.,
Room-temperature ultraviolet nanowire nanolasers, Science, Vol. 292, 2001, pp. 1897-1899.
[4] Pan, Z. W., Dai, Z. R. and Wang, Z. L., Nanobelts of semiconducting oxides, Science, Vol.

466
Journal of Solid Mechanics Vol. 1, No. 4, 2007
and Materials Engineering

291, 2001, pp. 1947-1949.


[5] Miao, L., Tanemura, S., Toh, S., Kaneko, K. and Tanemura, M., Heating-sol–gel template
process for the growth of TiO2 nanorods with rutile and anatase structure, Applied Surface
Science, Vol. 238, 2004, pp. 175-179.
[6] Xu, C., Zhan, Y., Hong, K. and Wang, G., Growth and mechanism of titania nanowires, Solid
State Communications, Vol. 126, 2003, pp. 545-549.
[7] Cheng, B., Russell, J. M., Shi, W., Zhang, L. and Samulski, E. T., Large-Scale, solution-phase
growth of single-crystalline SnO2 nanorods, Journal of the American Chemical Society, Vol.
126, 2004, pp. 5972-5973.
[8] Pavasupree, S., Suzuki, Y., Kitiyanan, A., Pivsa-Art, S. and Yoshikawa, S., Synthesis and
characterization of vanadium oxides nanorods, Journal of Solid State Chemistry, Vol.
178, 2005, pp. 2152-2158.
[9] Pavasupree, S., Suzuki, Y., Pivsa-Art, S. and Yoshikawa, S., Synthesis and characterization of
nanoporous, nanorods, nanowires metal oxides, Science and Technology of Advanced
Materials, Vol. 6, 2005, pp. 224-229.
[10] Fujishima, A., Rao, T. N. and Tryk, D. A., Titanium dioxide photocatalysis, Journal of
Photochemistry and Photobiology C: Photochemistry Reviews, Vol. 1, 2000, pp. 1-21.
[11] Adachi, M., Murata, Y., Harada, M. and Yoshikawa, S., Formation of titania nanotubes with
high photo-catalytic activity, Chemistry Letters, Vol. 8, 2000, pp. 942-943.
[12] Adachi, M., Okada, I., Ngamsinlapasathian, S., Murata, Y. and Yoshikawa, S.,
Dye-sensitized solar cell using semiconductor thin film composed of titania nanotubes,
Electrochemistry, Vol. 70, 2002, pp. 449-452.
[13] Adachi, M., Murata, Y., Okada, I. and Yoshikawa, S., Formation of titania nanotubes and
applications for dye-sensitized solar cells, Journal of The Electrochemical Society, Vol
150, (8), 2003, pp. G488-G493.
[14] Ngamsinlapasathian, S., Sreethawong, T., Suzuki, Y. and Yoshikawa, S., Single and
double-layered mesoporous TiO2/P25 TiO2 electrode for dye-sensitized solar cell, Solar
Energy Materials and Solar Cells, Vol. 86, 2005, pp. 269-282.
[15] Pavasupree, S., Suzuki, Y., Pivsa-Art, S. and Yoshikawa, S., Preparation and
characterization of mesoporous MO2 (M=Ti, Ce, Zr, and Hf) nanopowders by a modified
sol-gel method, Ceramics International, Vol. 31, 2005, pp. 959-963.
[16] Pavasupree, S., Suzuki, Y., Pivsa-Art, S. and Yoshikawa, S., Preparation and characterization
of mesoporous TiO2–CeO2 nanopowders respond to visible wavelength, Journal of Solid State
Chemistry, Vol. 178, 2005, pp. 128-134.
[17] Sreethawong, T., Suzuki, Y. and Yoshikawa, S., Synthesis, characterization, and photocatalytic
activity for hydrogen evolution of nanocrystalline mesoporous titania prepared by
surfactant-assisted templating sol-gel process, Journal of Solid State Chemistry, Vol. 178,
2005, pp. 329-338.
[18] Kitiyanan, A., Ngamsinlapasathian, S., Pavasupree, S. and Yoshikawa, S., The preparation
and characterization of nanostructured TiO2–ZrO2 mixed oxide electrode for efficient
dye-sensitized solar cells, Journal of Solid State Chemistry, Vol. 178, 2005, pp. 1044-1048.
[19] Sakulkhaemaruethai, S., Pavasupree, S., Suzuki, Y., and Yoshikawa, S., Photocatalytic
activity of titania nanocrystals prepared by surfactant-assisted templating method—effect of
calcination conditions, Materials Letters, Vol. 59, 2005, pp. 2965-2968.
[20] Suzuki, Y., Pavasupree, S., Yoshikawa, S. and Kawahata, R., Natural rutile-derived titanate
nanofibers prepared by direct hydrothermal processing, Journal of Materials Research, Vol.
20, 2005, pp. 1063-1070.
[21] Pavasupree, S., Suzuki, Y., Yoshikawa, S. and Kawahata, R., Synthesis of titanate, TiO2 (B),
and anatase TiO2 nanofibers from natural rutile sand, Journal of Solid State Chemistry, Vol.
178, 2005, pp. 3110-3116.
[22] Suzuki, Y., Ngamsinlapasathian, S., Yoshida, R. and Yoshikawa, S., Partially

467
Journal of Solid Mechanics Vol. 1, No. 4, 2007
and Materials Engineering

nanowires-structured porous TiO2 electrode for dye-sensitized solar cell, Central European
Journal of Chemistry, Vol. 4, 2006, pp. 476-488.
[23] Pavasupree, S., Ngamsinlapasathian, S., Nakajima, M., Suzuki, Y. and Yoshikawa, S.,
Synthesis, characterization, photocatalytic activity and dye-sensitized solar cell performance
of nanorods/nanoparticles TiO2 with mesoporous structure, Journal of Photochemistry and
Photobiology A: Chemistry, Vol. 184, 2006, pp. 163-169.
[24] Pavasupree, S., Ngamsinlapasathian, S., Suzuki, Y. and Yoshikawa, S., Synthesis and
dye-sensitized solar cell performance of nanorods/nanoparticles TiO2 from high surface area
nanosheet TiO2, Journal of Nanoscience and Nanotechnology , Vol. 6, 2006, pp. 3585-3692.
[25] Suzuki, Y. and Yoshikawa, S., Synthesis and thermal analyses of TiO2-derived nanotubes
prepared by the hydrothermal method, Journal of Materials Research, Vol. 19, 2004, pp.
982-985.
[26] Yoshida, R., Suzuki, Y. and Yoshikawa, S., Effects of synthetic conditions and heat treatment
on the structure of partially ion-exchanged titanate nanotubes, Materials Chemistry and
Physics, Vol. 91, 2005, pp. 409-416.
[27] Yoshida, R., Suzuki, Y. and Yoshikawa, S., Synthesis of TiO2(B) nanowires and TiO2 anatase
nanowires by hydrothermal and post-heat treatments, Journal of Solid State Chemistry, Vol.
178, 2005, pp. 2179-2185.
[28] Uchida, S., Chiba, R., Tomiha, M., Masaki, N. and Shirai, M., Application of titania
nanotubes to a dye-sensitized solar cell, Electrochemistry, Vol. 70, 2002, pp. 418-420.
[29] Yoon, J. –H., Jang, S. –R., Vittal, R., Lee, J. and Kim, K. –J., TiO2 nanorods as additive to
TiO2 film for improvement in the performance of dye-sensitized solar cells, Journal of
Photochemistry and Photobiology A Chemistry, Vol. 180, 2006, pp. 184-188.

468

Você também pode gostar