You are on page 1of 7

Applied Surface Science 280 (2013) 523529

Contents lists available at SciVerse ScienceDirect

Applied Surface Science


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

Nitrogen doped TiO2 nanotube arrays with high photoelectrochemical


activity for photocatalytic applications
Bao Yuan a,b , Yan Wang a,b, , Haidong Bian a,b , Tiankuo Shen a,b ,
Yucheng Wu a,c, , Zhong Chen a,c
a

Laboratory of Functional Nanomaterials and Devices, School of Materials Science and Engineering, Hefei University of Technology, Hefei, 230009, China
Key Laboratory of Advanced Functional Materials and Devices, Hefei University of Technology, Hefei, 230009, China
c
School of Materials Science & Engineering, Nanyang Technological University, 50 Nanyang Avenue, 639798, Singapore
b

a r t i c l e

i n f o

Article history:
Received 21 December 2012
Received in revised form 7 April 2013
Accepted 6 May 2013
Available online 20 May 2013
Keywords:
Anodization
Nitrogen doped TiO2 nanotube arrays
Methyl orange
Photocurrent
Photocatalytic property

a b s t r a c t
Nitrogen doped TiO2 nanotube arrays (N-TNAs) were prepared by immersing TNAs in 1 M NH3 H2 O solution and then annealing in different temperatures. The morphology, structure and composition of the
N-TNAs were characterized by eld-emission scanning electron microscopy (FESEM), X-ray diffraction
(XRD), X-ray photoelectron spectroscopy (XPS), and UVvis spectroscopy, respectively. Effects of annealing temperatures on structure, photocatalytic properties, and the crystal structure transformation process
of the N-TNAs were discussed. Photocatalytic properties of the N-TNAs were evaluated in term of the
degradation of methyl orange (MO) under UV light and visible light, and the photocurrent of N-TNAs
were tested by electrochemical workstation. The XPS results showed that the N-TNAs were achieved by
interstitial doping and substitutional doping, and the FESEM results showed the morphology was not
changed after doping process. Compared with the pure TNAs, the N-TNAs annealed at 500 C for 2 h with
a mixed phase of anatase and rutile exhibited higher photocatalytic degradation activity to MO. Furthermore, the photocatalytic mechanism of organic pollutants degradation (MO) was discussed based on our
experiments.
2013 Elsevier B.V. All rights reserved.

1. Introduction
Nanostructured materials, especially the highly ordered nanotube materials, have attracted a great deal of attention in various
elds. In recent decades, the preparation of nanomaterials has
diversied with the development of science and technology.
Advances in the nanoscale technology [1] facilitated the fabrication of highly ordered and multidimensional structured materials.
Many efforts have focused on new synthesis methods and photoelectrochemical properties of the tubular structure, large specic
surface area, oriented charge transfer channel and the other distinct
properties. TiO2 nanotube arrays as nanostructure semiconductor
compound have attracted increasing research interests in photocatalysis [25], dye sensitized solar cells [6,7], gas sensors [8,9],
biomedical applications [10] and so on. Particularly, TNAs are
expected to exhibit better photocatalytic properties compared with
nanoparticles or other forms of titanium dioxide [11], due to their
high specic surface area, short diffusion path and high activity in
the band-edge positions, which make it more suitable to be used
as catalyst [12].

Corresponding author. Tel.: +86 551 62901012; fax: +86 551 62904517.
E-mail address: ycwu@hfut.edu.cn (Y. Wu).
0169-4332/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apsusc.2013.05.021

Consequently, the synthesis or modication of TiO2 nanotube arrays have been widely studied [13,14], and considerable
efforts of fabrication TiO2 nanotubes such as hydrothermal
treatment [15], template-deposition [16], sonoelectrochemical
method [17] and anodic oxidation [18,19] have been developed. Gong and co-workers [20] pioneered the synthesis of
vertically ordered TiO2 nanotube arrays up to 500 nm in length
by a potentiostatic electrochemical anodization of titanium in
hydrouoric acid aqueous electrolyte. The experimental process
is very convenient without any complex apparatus. Subsequently, various organic electrolytes including dimethyl sulfoxide
[21], formamide [22] and ethylene glycol [23] have been
adopted to fabricate TiO2 nanotube arrays with greatly extended
length.
Though TNAs as photocatalysts were rstly used in environmental applications [24], many challenges still remain such as
the TNAs could not absorb visible light ( > 387 nm) of the solar
spectrum efciently because of their large band gap (3.2 eV)
as well as the recombination of photogenerated electrons and
holes. In order to overcome these disadvantages, considerable
efforts have been made to modify TNAs in order to reduce
the band gap. In the present case, many transition metal ions
[25,26] and nonmetal ions [2730] have been studied to increase
the visible light absorption or suppress the recombination of

524

B. Yuan et al. / Applied Surface Science 280 (2013) 523529

Fig. 1. SEM images of the highly ordered TNAs with different magnications.

photogenerated electronholes. Asahi et al. [31] investigated a


visible-light photocatalysis in nitrogen-doped titanium oxides by
sputtering the TiO2 target in a N2 /Ar gas mixture. Tokudome and
co-workers [32] reported nitrogen-doped TiO2 nanotubes by a wet
process. In this paper, nitrogen doped TiO2 nanotube arrays (NTNAs) were fabricated by immersing TNAs in ammonia aqueous
solution following with annealing in air atmosphere. Effects of
annealing temperature on the photocatalytic performance of NTNAs were investigated.

2. Experiments
2.1. Preparation of nitrogen doped TiO2 nanotube arrays
Highly ordered TiO2 nanotube arrays were fabricated by
anodization method. Titanium foil was anodized in ethylene glycol
electrolytes containing 0.3 M ammonium uoride and 2 vol% water
with potential of 60 V for 6 h. The as-prepared TNAs samples were
immersed in 1 M NH3 H2 O solution for 10 h and then annealed in
a tube furnace for 2 h at different temperatures with heating and
cooling rates of 2 C/min.

(a), (b), (c) top views, and (d) cross-sectional views.

2.3. Photoelectrochemical and photocatalytic activity


measurements
The photocatalytic activities of TNAs and N-TNAs were
evaluated by the photocurrent and degradation of MO. Photoelectrochemical measurements were carried out in 0.05 M phospate
buffer solution with pH value of 7 by adjusting the ratio of Na2 HPO4
and NaH2 PO4 . LED lamp with a quartz window was used as UV
source. Photocurrent was measured by an electrochemical workstation with the N-TNAs as working electrode. All the samples were
measured twice to get the average photocurrent value.
In the photocatalytic experiments, the prepared samples were
used as photocatalysts and the MO was chosen as target pollutant.
The experiment was performed in a UV-light reactor with a 300 W
high-pressure mercury lamp (390 nm) and a 250 W metalhalogen
lamp (420 nm). The initial concentration of MO aqueous solution was 20 mg/L. The change of concentration was monitored for
analyzing the photocatalytic activity at different irradiation time
intervals by measuring the absorption at 464 nm using a UV1800
spectrometer.
3. Results and discussion
3.1. The morphology of TNAs and N-TNAs

2.2. Characterization of N-TNAs


The surface morphologies of samples were observed using the
eld-emission scanning electron microscopy (FESEM, FEI Sirion200). Crystal structures of N-TNAs were characterized by X-ray
diffraction spectrometer (XRD, D/Max-rB, Japan). Average crystallite sizes of N-TNAs were determined according to the Scherrer
equation using the full-width half-maximum data of each diffraction peak after correcting the instrumental broadening. Surface
chemical states of samples were analyzed by X-ray photoelectron
spectroscopy (XPS, Thermo ESCALAB-250). All the binding energies were referenced to the C1s peak at 285.0 eV of the surface
adventitious carbon.

SEM morphologies of typical TNAs are shown in Fig. 1. The TNAs


are well-aligned with average diameter of 140 nm, wall thickness
of 10 nm, and length of 30 m. The morphologies of N-TNAs are
shown in Fig. 2. N-TNAs annealed at 500 C are shown in Fig. 3.
Both undoped and the N-doped TiO2 nanotube arrays annealed
at 500 C show similar morphologies with the as-fabricated TNAs.
These results indicate no signicant effect of annealing on surface
morphology and microstructure of the TNAs. However, the nanotube collapsed when the annealing temperature increase to 700 C.
This is ascribed to the rapid grain growth during the phase transition from anatase to rutile at high temperature [33], which is in
good agreement with what reported in literature [34].

B. Yuan et al. / Applied Surface Science 280 (2013) 523529

Fig. 2. SEM images of N-doped TiO2 nanotube arrays.

3.2. Characterization of N-TNAs


Fig. 4 shows the XRD patterns of the TNAs and N-TNAs annealed
at different temperatures. As-synthesized TiO2 nanotube arrays
were reported to be amorphous [35], and well- crystallized anatase
or rutile can be obtained by thermal annealing method. Two diffraction peaks at 25.28 and 48.05 are observed in XRD pattern of
N-TNAs annealed at 300 C for 2 h, which are according to (1 0 1)
and (2 0 0) crystal faces of anatase TiO2 . With the increase of temperature, both anatase and rutile appeared in N-TNAs. Moreover,
the N-TNAs (Fig. 4 d) have more obvious rutile peaks appeared at
27.45 accordance with the (1 1 0) plane than that of TNAs (Fig. 4a)
annealed at 500 C, which indicates that nitrogen doping in TiO2
lattice can decrease temperature of crystal transition from anatase
to rutile. This result can be interpreted that N-doping reduced the
TiO2 crystallite size, as well as increased the specic surface area
and surface energy, leading to the instability of crystal structure.
With temperature increasing (Fig. 4e and f), diffraction peaks of
rutile phase become stronger and diffraction peaks of anatase phase
get weaken and even disappear at 700 C. Crystal grain size can be

525

(a), (b) top views, and (c), (d) cross-sectional views.

Table 1
The phase composition and grain size of the samples annealed at various
temperatures.
Samples

N-TNAs

Annealing temperature ( C)
Weight fraction of anatase (%)
Weight fraction of rutile (%)
Grain size(nm)

300
100
0
28.4

400
100
0
29.2

500
91.7
8.3
32.0

600
51.1
48.9
32.7

700
9.1
90.9
39.0

calculated by Scherrer Eq. (1) at different annealing temperatures.


D=

k
cos 

(1)

where k is Scherrer constant, D is grain size,  is wavelength of Xray,  is diffraction angle, and is the full-width at half-maximum
of the (1 0 1) plane. The results of D are summarized in Table 1. The
anatase and rutile contents in crystalline TiO2 nanotube arrays are
calculated using the formula as follow:
XA = (1 + 1.26IR /IA )

Fig. 3. SEM images of N-doped TiO2 nanotube arrays annealed at 500 C for 2 h.

(a) Top view, and (b) cross-sectional view.

(2)

526

B. Yuan et al. / Applied Surface Science 280 (2013) 523529

T
R

intensity (a.u.)

e
d
c
b
a

20

30

40

50

60

70

80

2 (deg.)
Fig. 4. XRD patterns of TNAs and N-TNAs annealed at temperatures ranging from
300 to 700 C.
a TNAs-500 C, b N-TNAs-300 C, c N-TNAs-400 C, d N-TNAs-500 C, e NTNAs-600 C, f N-TNAs-700 C
A, R, and T represent anatase, rutile and titanium, respectively.

where XA is the weight fraction of anatase in the crystalline TiO2


nanotube arrays, IR and IA is the integrated intensity of the (1 1 0)
reection plane of rutile phase and the (1 0 1) reection plane of
anatase, respectively. The calculation results of different annealing
temperatures are given in Table 1.

N-TNAs were characterized by XPS to analyze the surface composition and chemical states of the as-prepared nanotubes. Fig. 5
shows XPS spectra of N-TNAs. The individual peaks of Ti 2p1 at
464.98 eV, Ti 2p3 at 459.29 eV, respectively, can be clearly seen
in Fig. 5b. And the distance of peaks between Ti 2p1 and Ti 2p3
is 5.69 eV, which means that the chemical state of the titanium
is Ti4+ (TiO2 ) [36]. Compared with standard binding energy, there
is a small shift of Ti 2p peak due to the change of local chemical
state of Ti ions inuenced by N incorporation and TiNO bond
formation. As shown in Fig. 5c, peaks of O 1s locate at 530.51 eV
and 532.6 eV, respectively. The former corresponds to O 1s electrons of TiO bonding, and the latter is from OH bonding. These
results conrm the existance of hydroxyl (OH ) on the TiO2 surface
which can facilitate the photocatalytic properties of TiO2 nanotube
arrays [37,38]. The peak of C 1s can be ascribed to the contaminant of organic carbon from the process of preparing N-TNAs and
annealing treatment.
Fig. 6 shows the high-resolution XPS N 1s core level spectra of NTNAs. The nitrogen element present on the N-TNAs and its relative
content is identied. As shown in Fig. 6a, peak of N 1s is at about
400.28 eV without sputtering treatment. Jagadale and co-workers
[39] research indicated that the peak of N 1s located ranging from
398.8 to 400.8 eV which belonged to electron binding energy of
TiNO. To eliminate the effect of surface contamination layer, the
XPS of the sample after sputtering treatment is shown in Fig. 6b.
Two peaks of N 1s locate at 400.37 eV and 397.0 eV, respectively.
This result can be assigned to the formation of TiN bond [40].
Therefore, N-doping was achieved through substitutional doping
and interstitial doping.

O 1s

Ti 2p1
Ti 2p3

N 1s

1200

1000

800

600

C 1s

400

200

Binding energy / eV

Ti 2p3

O 1s

Ti 2p1

475

470

465

460

455

Binding energy / eV
Fig. 5. XPS spectrum of N-TNAs samples annealed at 500 C for 2 h.

450

545

540

535

530

525

Binding energy / eV
a XPS survey spectrum of N-TNAs.

b, c high-resolution XPS spectra of Ti 2p and O 1s.

B. Yuan et al. / Applied Surface Science 280 (2013) 523529

400.28 eV

527

397.0 eV

N 1s

N1s

400.37 eV

410

408

406

404

402

400

398

396

394

410

408

406

404

Binding energy / eV
Fig. 6. XPS spectrums of N 1s in N-TNAs samples.

402

400

398

396

394

392

Binding energy / eV
a surface layer of N 1s; b depth prole of N 1s.

3.3. Photoelectrochemical properties

0.12

UV on

20

30

40

Time/min
Fig. 8. Photocatalytic degradation of MO for TNAs and N-TNAs with different annealing temperature under high-pressure mercury lamp illumination.

photocatalytic activity of TiO2 mainly depends on the photocurrent


values and electronhole pairs transfer ability. So, we can deduce
that the N-TNAs annealed at 500 C will exhibit higher photocatalytic activity than the other N-TNAs or TNAs.

a
200

300

400

500

Degradation Rate/%

UV off

c
100

10

100

b
g

0.00

25

UV off

50

0.08

0.04

75

3.3.2. The photocatalytic properties of N-doped TiO2 nanotube


arrays
The photocatalytic activity of TNAs and N-TNAs were studied by
degradation of methyl orange solution, and the results were shown

Current / mA

Degradation Rate/%

100
3.3.1. The photocurrent of N-doped TiO2 nanotube arrays
Fig. 7 shows the photocurrent versus time plots of N-TNAs with
different annealing temperatures, which performed in a buffer
solution containing 0.05 M Na2 HPO4 and 0.05 M NaH2 PO4 (pH 7)
with bias potential of 0.2 V.
The rise and fall of the photocurrent corresponded well to the
UV illumination being switched on and off for all the samples,
and the dark current is very small. Compared to the pure TNAs
without annealing, the photocurrents increase signicantly with
the annealing of the N-TNAs or TNAs, and the maximum value
(0.116 mA) can be observed on the N-TNAs annealed at 500 C.
Higher photocurrent means more photo-induced charges were
motivated from the samples. And it also means lower recombination of electronhole.
As shown in Fig. 4, with the increase of annealing temperature,
N-TNAs changes from amorphous to anatase phase. The increase
of anatase phase ratio may enhance the separation and transfer efciency of the photogenerated carriers, resulting in higher
photocurrent [41]. Peak intensity of anatase phase weakens when
annealing temperature is higher than 500 C. N-TNAs are gradually
transformed to rutile phase, leading to a decline of photocurrent.
For instance, the photocurrent (0.051 mA) of N-TNAs annealed at
700 C is less than that of the sample annealed at 500 C. The

80

60

Time / s
40
Fig. 7. Photocurrent spectra of as-prepared TiO2 nanotube arrays and N-doped TiO2
nanotube arrays.
a pure TiO2 nanotube arrays without annealing, b pure TiO2 nanotube arrays
annealed at 500 C; N-doped TiO2 nanotube arrays annealed at different temperatures, c 300 C, d 400 C, e 500 C, f 600 C, g 700 C.
The inset shows the photocurrent versus annealing temperature of N-doped TiO2
nanotube arrays.

300

400

500

600

700

Annealing Temperature/
Fig. 9. Relationships between annealing temperature and degradation rate of NTNAs under high-pressure mercury lamp illumination for 30 min.

528

B. Yuan et al. / Applied Surface Science 280 (2013) 523529


1.2

b
0 min

1.0

40 min

0.5

0.0
200

400

600

Wavelength (nm)

Absorbance (a.u.)

Absorbance (a.u.)

1.5

0h

0.6
4h

0.0
200

400

600

Wavelength (nm)

Fig. 10. UVvis absorption spectra of photocatalytic degradation MO recorded at different time intervals by the N-TNAs annealed at 500 C.
a under high-pressure mercury lamp illumination.
b under metalhalogen lamp illumination.

in Fig. 8. For comparison, the photocatalytic performance of pure


TiO2 nanotube arrays was also studied under UV illumination. The
concentration of MO was monitored by measuring the absorbance
of MO solution at the wavelength of 464 nm.
Lai and co-workers [42] reported the photocatalytic degradation with the presence of TiO2 nanotube arrays photocatalyst was
signicantly higher than self-degradation under UV illumination.
This result indicated the TNAs played an important role in the
photocatalysis process and the effect of self-degradation can be
neglected. As shown in Fig. 8, 98.5% of the MO is degraded by the
N-TNAs annealed at 500 C, while only 6.6% for as-prepared TiO2
nanotube arrays without annealing. These results can be attributed
to crystallization via annealing treatment, which can increase their
adsorption abilities and promote the transfer efciency of the photogenerated charges.
The variation tendency of degradation rate with annealing temperature is shown in Fig. 9. The photocatalytic degradation rate of
MO increases rst and then falls with increasing annealing temperature. N-TNAs annealed at 500 C gives the best photocatalytic
performance, due to the transformation of phase composition,
which is in accordance with the results of Fig. 7. These results could
be ascribed to the transformation of phase composition. N-TNAs
were annealed to form anatase and rutile phases by annealing at
500 C (according to the XRD results). When the temperature is
above 600 C, the diffraction peaks of rutile phase become stronger
and peaks of anatase phase weaken and even disappear. Therefore,
the rapid TiO2 grain growth at high temperatures results in thick
tube walls and large grain size (as shown in Table 1), which decrease
the surface area of N-TNAs. Hence, the photocatalytic activity of
anatase is better than rutile.
Fig. 10a shows the plot of wavelength vs. absorbance under different light source irradiation. With increasing irradiation time,
it can be seen that the characteristic absorption peak at 464 nm
decreases sharply due to the degradation of MO on the N-TNAs
photocatalyst. This absorption peak disappears at about 40 min,
which indicates that MO is almost completely removed. However,
the degradation ratio of MO is about 19% under metal-halogen lamp
illumination after 4 h (Fig. 10b), indicating that the N-doped TiO2
nanotube arrays have certain visible light photocatalytic activity.
This is due to the fact that visible light can excite the valence band
electrons in the N-TNAs. And N dopants could be assumed to play a
vital role in promoting the amount of surface hydroxyl group, which
will increase the photoresponse in the visible light range. Moreover,
the highly ordered tubular structure of N-TNAs can be benecial for
promoting the diffusion access of reactants. Therefore, the visible
light photocatalysis has benecial effect for the enhancement of
photocatalytic activity. Moreover, the highly ordered tubular structure of N-TNAs can be benecial to promote the diffusion access of
reactants.

Fig. 11. Schematic diagram of photocatalysis process of N-doped TiO2 nanotube


arrays photocatalyst under UV light irradiation.

The photocatalytic mechanism for degradation of organic pollutants (MO) can be explained. Under the UV light or visible light
illumination, electrons (e ) migrate from the valence band to the
conduction band. Meanwhile, holes (h+ ) will be left at the valence
band. Then the electrons and holes move to TiO2 nanotube arrays
surface under the electric eld, and react with the H2 O, O2 and OH
etc. on the TNAs surface. Several highly oxidability species (OH ,
O2 ) are generated (H2 O + h+ OH , O2 + e O2 ). These radicals and peroxo ions are able to oxidize organic pollutant to CO2
and H2 O. The whole photocatalysis process could be described as
Fig. 11. The N 2p energy level of N-TNAs located above the valence
band of pure TiO2 . Electrons can be excited from valence band to
the N 2p energy level and it will reduce the band gap of TiO2 compared to the pure TiO2 . As a result, photogenerated carriers could
be effectively separated to take part in the photocatalytic process,
which leading to a higher photocatalytic activity than that of pure
TiO2 nanotube arrays.
4. Conclusions
In this study, TiO2 nanotube arrays were successfully prepared by anodization. N-doped TiO2 nanotube arrays were then
synthesized by immersing the TNAs into the ammonia aqueous
solution. The XPS characterization results showed that the NTNAs were mainly achieved through substitutional doping and
interstitial doping. And the morphology was not changed after

B. Yuan et al. / Applied Surface Science 280 (2013) 523529

annealing and doping process. Photoelectrochemical experiment


showed that photocurrent and photocatalytic activity of N-TNAs
strongly depended on the annealing temperature, and the N-TNAs
annealed at 500 C have the best activity to degrade organic pollutants like methyl orange.
Acknowledgements
This work was nancially supported by the National Natural Science Foundation of China (Nos. 91023030, 51272062, 51128201
and 51202052), Natural Science Foundation of Anhui Province (No.
1308085QE74), the Specialized Research Fund for the Doctoral
Program of Higher Education (No. 20100111110012) and the International Scientic and Technological Cooperation Project of Anhui
Province (No. 10080703017).
References
[1] H.C. Liang, X.Z. Li, Effects of structure of anodic TiO2 nanotube arrays on photocatalytic activity for the degradation of 2,3-dichlorophenol in aqueous solution,
Journal of Hazardous Materials 162 (2009) 14151422.
[2] J.M. Macak, M. Zlamal, J. Krysa, P. Schmuki, Self-organized TiO2 nanotube layers
as highly efcient photocatalysts, Small 3 (2007) 300304.
[3] G.K. Mor, K. Shankar, M. Paulose, O.K. Varghese, C.A. Grimes, Enhanced photocleavage of water using titania nanotube arrays, Nano Letters 5 (2005) 191195.
[4] S. Sakthivel, H. Kisch, Daylight photocatalysis by carbon-modied titanium
dioxide, Angewandte Chemie International Edition 42 (2003) 49084911.
[5] Q. Li, Y.W. Li, P. Wu, R. Xie, J.K. Shang, Palladium oxide nanoparticles on
nitrogen-doped titanium oxide: accelerated photocatalytic disinfection and
post-illumination catalytic memory, Advanced Materials 20 (2008) 37173723.
[6] G.K. Mor, K. Shankar, M. Paulose, O.K. Varghese, C.A. Grimes, Use of highlyordered TiO2 nanotube arrays in dye-sensitized solar cells, Nano Letters 6
(2006) 215218.
[7] J.R. Jennings, A. Ghicov, L.M. Peter, P. Schmuki, A.B. Walker, Dye-sensitized
solar cells based on oriented TiO2 nanotube arrays: transport, trapping, and
transfer of electrons, Journal of the American Chemical Society 130 (2008)
1336413372.
[8] O.K. Varghese, D. Gong, M. Paulose, K.G. Ong, C.A. Grimes, Extreme changes
in the electrical resistance of titania nanotubes with hydrogen exposure,
Advanced Materials 15 (2003) 624627.
[9] S. Singh, H. Kaur, V.N. Singh, K. Jain, T.D. Senguttuvan, Highly sensitive and
pulse-like response toward ethanol of Nb doped TiO2 nanorods based gas sensors, Sensors and Actuators B 171 (172) (2012) 899906.
[10] S.C. Roy, M. Paulose, C.A. Grimes, The effect of TiO2 nanotubes in the enhancement of blood clotting for the control of hemorrhage, Biomaterials 28 (2007)
46674672.
[11] A.L. Castro, M.R. Nunes, A.P. Carvalho, F.M. Costa, M.H. Florencio, Synthesis of
anatase TiO2 nanoparticles with high temperature stability and photocatalytic
activity, Solid State Science 10 (2008) 602606.
[12] H. Li, B.L. Zhu, Y.F. Feng, S.R. Wang, S.M. Zhang, W.P. Huang, Synthesis, characterization of TiO2 nanotubes-supported MS (TiO2 NTS @MS, M = Cd, Zn) and
their photocatalytic activity, Journal of Solid State Chemistry 180 (2007)
21362142.
[13] M.S. Sander, M.J. Cote, W. Gu, B.M. Kile, Template-assisted fabrication of dense,
aligned arrays of titania nanotubes with well-controlled dimensions on substrates, Advanced Materials 16 (2004) 20522057.
[14] B.G. Lee, S.Y. Hong, J.E. Yoo, J. Choi, Electropolishing for the formation of anodic
nanotubular TiO2 with uniform length and density, Applied Surface Science
257 (2011) 71907194.
[15] D.L. Morgan, H.Y. Zhu, R.L. Frost, E.R. Waclawik, Determination of a morphological phase diagram of titania/titanate nanostructures from alkaline
hydrothermal treatment of Degussa P25, Chemistry of Materials 20 (2008)
38003802.
[16] S.I. Na, S.S. Kim, W.K. Hong, J.W. Park, J. Jo, Y.C. Nah, T. Lee, D.Y. Kim, Fabrication
of TiO2 nanotubes by using electrodeposited ZnO nanorod template and their
application to hybrid solar cells, Electrochimica Acta 53 (2008) 25602566.
[17] S.K. Mohapatra, M. Misra, V.K. Mahajan, K.S. Raja, A novel method for the
synthesis of titania nanotubes using sonoelectrochemical method and its application for photoelectrochemical splitting of water, Journal of Catalysis 246
(2007) 362369.

529

[18] X.Y. Hu, T.C. Zhang, Z. Jin, S.Z. Huang, M. Fang, Y.C. Wu, L.D. Zhang, Singlecrystalline anatase TiO2 dous assembled micro-sphere and their photocatalytic
activity, Crystal Growth and Design 9 (2009) 23242328.
[19] Y. Wang, Y.C. Wu, Y.Q. Qin, G.B. Xu, X.Y. Hu, J.W. Cui, H.M. Zheng, Y. Hong, X.Y.
Zhang, Rapid anodic oxidation of highly ordered TiO2 nanotube arrays, Journal
of Alloys and Compounds 509 (2011) 157160.
[20] D. Gong, C.A. Grimes, O.K. Varghese, W.C. Hu, R.S. Singh, Z. Chen, E.C. Dickey,
Titanium oxide nanotube arrays prepared by anodic oxidation, Journal of Materials Research 16 (2001) 33313334.
[21] M. Paulose, K. Shankar, S. Yoriya, H.E. Prakasam, O.K. Varghese, G.K. Mor, Anodic
growth of highly ordered TiO2 nanotube arrays to 134 m in Length, Journal of
Physical Chemistry B 110 (2006) 1617916184.
[22] K. Shankar, G.K. Mor, A. Fitzgerald, C.A. Grimes, Cation effect on the electrochemical formation of very high aspect ratio TiO2 nanotube arrays in
formamide-water mixtures, Journal of Physical Chemistry C 111 (2007)
2126.
[23] H.E. Prakasam, K. Shankar, M. Paulose, O.K. Varghese, C.A. Grimes, A New
benchmark for TiO2 nanotube array growth by anodization, Journal of Physical
Chemistry C 111 (2007) 72357241.
[24] X. Quan, S.G. Yang, X.L. Ruan, H.M. Zhao, Preparation of titania nanotubes and
their environmental applications as electrode, Environmental Science & Technology 39 (2005) 37703775.
[25] R.X. Cai, Y. Kubota, A. Fujishima, Effect of copper ions on the formation of
hydrogen peroxide from photocatalytic titanium dioxide particles, Journal of
Catalysis 219 (2003) 214218.
[26] N. Serpone, Is the band gap of pristine TiO2 narrowed by anion- and cationdoping of titanium dioxide in second-generation photocatalysts? Journal of
Physical Chemistry B 110 (2006) 2428724293.
[27] T. Umebayashi, T. Yamaki, H. Itoh, Band gap narrowing of titanium dioxide by
sulfur doping, Applied Physics Letters 81 (2002) 454456.
[28] T. Ohno, T. Mitsui, M. Matsumura, Photocatalytic activity of S-doped TiO2 photocatalyst under visible light, Chemistry Letters 32 (2003) 364365.
[29] G. Liu, F. Li, D.W. Wang, D.M. Tang, C. Liu, X.L. Ma, G.Q. Lu, H.M. Cheng, Electron
eld emission of a nitrogen-doped TiO2 nanotube array, Nanotechnology 19
(2008) 025606025611.
[30] Y.W. Wang, Y. Huang, W.K. Ho, L.Z. Zhang, Z.G. Zou, S.C. Lee, Biomoleculecontrolled hydrothermal synthesis of CNS-tridoped TiO2 nanocrystalline
photocatalysts for NO removal under simulated solar light irradiation, Journal
of Hazardous Materials 169 (2009) 7787.
[31] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Visible-light photocatalysis
in nitrogen-doped titanium oxides, Science 293 (2001) 269271.
[32] H. Tokudome, M. Miyauchi, N-doped TiO2 nanotube with visible light activity,
Chemistry Letters 33 (2004) 11081109.
[33] S.P. Albu, A. Ghicov, S. Aldabergenova, P. Drechsel, D. Leclere, G.E. Thompson,
J.M. Macak, P. Schmuki, Formation of double-walled TiO2 nanotubes and robust
anatase membranes, Advanced Materials 20 (2008) 41354139.
[34] P. Xiao, Y.H. Zhang, G.Z. Cao, Effect of surface defects on biosensing
properties of TiO2 nanotube arrays, Sensors and Actuators B 155 (2011)
159164.
[35] C.L. Zhu, H.L. Yu, Y. Zhang, T.S. Wang, Q.Y. Ouyang, L.H. Qi, Y.J. Chen, X.Y. Xue,
Fe2 O3 /TiO2 tube-like nanostructures: synthesis, structural transformation and
the enhanced sensing properties, ACS Applied Materials & Interfaces 4 (2012)
665671.
[36] A. Fujishima, K. Honda, Electrochemical photolysis of water at a semiconductor
electrode, Nature 238 (1972) 3738.
[37] H. Irie, Y. Watanabe, K. Hashimoto, Nitrogen-concentration dependence on
photocatalytic activity of TiO2x Nx powders, The Journal of Physical Chemistry
B 107 (2003) 54835486.
[38] O. Diwald, T.L. Thompson, E.G. Coralski, J.T. Yates Jr., The effect of nitrogen ion
implantation on the photoactivity of TiO2 rutile single crystals, The Journal of
Physical Chemistry B 108 (2004) 5257.
[39] T.C. Jagadale, S.P. Takale, R.S. Sonawane, H.M. Joshi, S.I. Patil, B.B. Kale, S.B.
Ogale, N-doped TiO2 nanoparticle based visible light photocatalyst by modied peroxide sol-gel method, The Journal of Physical Chemistry C 112 (2008)
1459514602.
[40] Y.P. Yu, X.J. Xing, L.M. Xu, S.X. Wu, S.W. Li, N-derived signals in the x-ray photoelectron spectra of N-doped anatase TiO2 , Journal of Applied Physics 105 (2009)
123535123539.
[41] S.D. Mo, W.Y. Ching, Electronic and optical properties of three phases of
titanium dioxide: rutile, anatase, and brookite, Physical Review B 51 (1995)
1302313032.
[42] Y.K. Lai, J.Y. Huang, H.F. Zhang, Y.P. Subramaniam, Y.X. Tang, D.G. Gong, L. Sundar, L. Sun, Z. Chen, C.J. Lin, Nitrogen-doped TiO2 nanotube array lms with
enhanced photocatalytic activity under various light sources, Journal of Hazardous Materials 184 (2010) 855863.

You might also like