You are on page 1of 6

Article

pubs.acs.org/JPCC

Characterization of Oxygen Vacancy Associates within


Hydrogenated TiO2: A Positron Annihilation Study
Xudong Jiang,† Yupeng Zhang,† Jing Jiang,† Yongsen Rong,† Yancheng Wang,† Yichu Wu,†
and Chunxu Pan*,†,‡

School of Physics and Technology, Hubei Nuclear Solid Physics Key Laboratory and MOE Key Laboratory of Artificial Micro- and
Nano-structures, Wuhan University, Wuhan 430072, China

Center for Electron Microscopy, Wuhan University, Wuhan 430072, China
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: This paper introduces a novel method for


characterizing the oxygen vacancy associates in hydrogenation-
modified TiO2 by using a positron annihilation lifetime
spectroscopy (PALS). It was found that a huge number of
Downloaded via ANADOLU UNIV on April 1, 2021 at 20:28:55 (UTC).

small neutral Ti3+−oxygen vacancy associates, some larger size


vacancy clusters, and a few voids of vacancy associates were
introduced into hydrogenated TiO2. The defects blurred the
atomic lattice high-resolution transmission electron micros-
copy (HRTEM) images and brought about the emergence of
new Raman vibration. X-ray photoelectron spectroscopy
(XPS) measurement indicated that the concentration of
oxygen vacancies was 3% in the TiO2 lattice. The photoluminescence (PL) spectroscopy, photocurrent, and degradation of
methylene blue indicated that the oxygen vacancy associates introduced by hydrogenation retarded the charge recombination and
therefore improved the photocatalytic activity remarkably.

1. INTRODUCTION However, the unclear detailed image of the disorder and the
As one of the promising semiconductors, titania (TiO2) has role of the hydrogenation as well as the mechanism of high
grabbed appreciable attention over the past decades, due to its photoactivity are still ambiguous.
powerful potential applications in photovoltaics, biomedicine, In addition, surface-hydrogenated anatase TiO2 nanowire
and photocatalysis.1−3 However, the photocatalytic efficiency of microspheres were also reported by Zheng and his co-
TiO2 is limited by its wide band gap and the low efficiency of workers.17 They ascribed the enhanced photocatalytic activities
recombination of photogenerated electrons and holes.4 Herein, to the improved optical absorption and efficient photo-
the methods for enhancing the photocatalytic activity of TiO2 generated electron−hole separation induced by the surface
are mainly based on decreasing the band gap and inhibiting the Ti−H bonds as well as the unique structure. Simultaneously,
recombination of photo-generated electrons and holes. Naldoni et al.18 fabricated the black TiO2 nanoparticles with
In general, the nonmetal dopants for TiO2, such as N,5 C,6 crystalline core/disordered shell morphology, and ascribed its
F, and S,8 were widely used for shifting the photoresponse
7 narrowed bandgap to the synergistic presence of oxygen
toward longer wavelengths. Meanwhile, numerous surface vacancies and surface disorder. More examples showed the
modifications have been devoted to reduce the recombination decisive role of defects such as oxygen vacancy in hydrogenated
of photogenerated electrons and holes, for example, TiO2/ TiO2. However, few studies concern the detection of oxygen-
semiconductor composites,9,10 metallic doping,11 heavy metal vacancies induced by hydrogenation in TiO2.
deposition upon TiO2,12 and so on. More current evidence has Positron annihilation lifetime spectroscopy (PALS) provides
demonstrated that oxygen vacancy defects on the TiO2 surface a sensitive method for investigating vacancy-type defects, and it
play an essential part in acting as electron scavengers when has been proven to be useful in determining the intrinsic
electron−hole pair creation occurs by photoexcitation in the defects in semiconductors.19,20 When positrons inject inside
TiO2.13 bulk materials, they get thermalized and annihilated with
Recently, hydrogenation modified TiO2 has triggered electrons, and result in the emission of γ rays, which convey
extensive research interest and is considered to have great information of the lifetime of positrons. Compared to the bulk
potential.14−18 Chen et al.14 reported the preparation of black of the material, positrons preferentially distribute in the regions
hydrogenated TiO2 nanoparticles in a 20 bar hydrogen where electron density is low, e.g., vacancy type defects,
atmosphere at 200 °C for 5 days. It exhibited a greatly
enhanced absorption in the visible light and near-infrared Received: July 31, 2012
region, due to its narrowed band gap around 1.0 eV, which was Revised: September 30, 2012
ascribed to the surface disorder of TiO2 nanoparticles. Published: October 2, 2012

© 2012 American Chemical Society 22619 dx.doi.org/10.1021/jp307573c | J. Phys. Chem. C 2012, 116, 22619−22624
The Journal of Physical Chemistry C Article

vacancy clusters, and microvoids. Herein, the electron-positron (160 W), which generates light with maximum intensity at 365
annihilation photons allow measurement of the lifetime of nm, was used as a light source away from the working electrode
positrons, which provides information regarding various with a distance of 10 cm. The intensity of photocurrent was
defects.21,22 measured by an electrochemical workstation (CHI660C,
In this research, commercial P25 was chosen as the pristine Chenhua, China).
TiO2 photocatalyst and annealed at 400 °C in H2 under 2.4. Measurement of Photocatalytic Activities. The
atmospheric pressure for 10 h. The positron annihilation photocatalytic activity was evaluated by detecting the
characterization revealed the existence of a large number of degradation rate of methylene blue (MB). In brief, a 250 W
various oxygen vacancy associates in the hydrogenated TiO2, high pressure mercury lamp was used as a light source for
which would play the key role of inhibiting the recombination photocatalytic reaction, 100 mg of photocatalyst was dispersed
of holes and electrons during photocatalyzing. into 100 mL of 10 mg/L MB solution, and then the mixture
was stirred constantly under the light at a distance of 250 mm.
2. EXPERIMETAL SECTION The concentration of MB solution was measured by the UV−
2.1. Preparation of Hydrogenated TiO2. Commercial vis spectrophotometer (UV-2550, Shimadzu, Japan) every 30
pure white P25 TiO2 nanoparticles purchased from Degussa min.
were used for the pristine sample. The hydrogenation was
carried out in a tube furnace filled with ultrahigh purity 3. RESULTS AND DISCUSSION
hydrogen gas (99.99%) under atmospheric pressure, and the Figure 1 illustrates the XRD patterns of the pristine and
pristine P25 was annealed at 400 °C for 10 h, as described in hydrogenated P25 TiO2 nanoparticles. The strong XRD
earlier reports.15,18 The flow rate of hydrogen is 10 sccm. The
obtained taupe powders were denoted as H-P25, distinguished
from the pristine white P25.
2.2. Characterization. The morphologies of P25 and H-
P25 nanoparticles were observed by using high-resolution
transmission electron microscopy (HRTEM, JEM 2010FEF
HRTEM, JEOL, Japan). An X-ray diffraction spectrometer
(XRD) (D8 Advanced XRD, Bruker AXS, Germany) with a Cu
Kα source was employed to analyze their crystalline structures.
The scanning speed was 4°/m in the range of 20−80°. X-ray
photoelectron spectroscopy (XPS) measurements were per-
formed in a VG Multilab2000 spectrometer to obtain
information on the chemical binding energy of the samples.
The ultraviolet−visible diffuse reflectance spectra (UV−vis
DRS) of the samples were measured by using a diffuse
reflectance accessory of a UV−vis spectrophotometer (UV-
2550, Shimadzu, Japan). The Raman spectra were obtained
using a LabRAM HR spectrometer (HORIBA Jobin Yvon,
France) with a laser excitation of 488 nm. Photoluminescence Figure 1. XRD patterns of P25 before (a) and after (b) hydrogenation.
(PL) spectra were measured at room temperature on a
fluorescence spectrophotometer (F-4600, Hitachi, Japan) with
an excitation wavelength of 300 nm, a scanning speed of 240 diffraction peaks indicated that the pristine P25 was highly
nm/min, and a photomultiplier tube (PMT) voltage of 700 V. crystallized with a mixture of anatase (70−80%) and rutile
PALS was measured using a conventional ORTEC-265 fast− (20−30%), which has been reported elsewhere,23 and no
fast coincident system at room temperature. The coincidence obvious change was observed for the H-P25 TiO2. The anatase
spectrometer used had a prompt time resolution of 150 ps grain sizes of the P25 TiO2 nanoparticles were around 16 nm
(fwhm) for the γ-rays from a 60Co source selected under the calculated from the Scherrer. By contrast, Figure 2 shows the
experimental conditions. Three hundred milligrams of sample absorption spectra of the pristine and hydrogenated P25 TiO2
powder was pressed into a disk (diameter: 13 mm, thickness: 1 nanoparticles. The hydrogenation had little effect on the
mm). The source (∼1.3MBq) of 22Na was sandwiched between absorption of P25 in the UV region, while H-P25 exhibited a
two identical sample disks. The positron lifetime spectrum remarkable absorption in the visible light region, which was also
containing 106 counts was analyzed by the PATFIT program to similar to the results of Chen et al.14 Nevertheless, the gray
decompose several lifetime components. coloration of H-P25 in our experiment was due to the use of
2.3. Photocurrent Tests. Five milligrams of photocatalysts slow cooling rate after hydrogenation.18
was dispersed in 5 mL of ethanol and ultrasonically vibrated for HRTEM observations are shown in Figure 3 and Figure 4,
30 min. The 0.25 mL resultant slurry was then dip-coated onto respectively, with related fast Fourier transform (FFT) analysis
a 10 × 20 mm indium−tin oxide (ITO) glass electrode and images in the inset. It revealed clear lattice fringes and a sharp
dried under high-pressure mercury lamp irradiation to eliminate FFT image in pristine P25 TiO2, which demonstrated a high
ethanol. The prepared TiO2/ITO electrode, platinum elec- level of crystallization. However, after hydrogenation, the lattice
trode, and saturation calomel electrode (SCE) were used as the images and the FFT image of H-P25 became blurred, which
working electrode, counter electrode, and reference electrode, indicated the distortion of the TiO2 atomic lattice structures
respectively. The electrolyte was 0.5 mol/L Na2SO4 aqueous during hydrogenation.
solution. The working electrode was activated in the electrolyte Raman spectroscopy is a spectroscopic technique to measure
for 0.5 h before measurement. A high-pressure mercury lamp molecular vibrations. It can be used for detecting the structural
22620 dx.doi.org/10.1021/jp307573c | J. Phys. Chem. C 2012, 116, 22619−22624
The Journal of Physical Chemistry C Article

Figure 2. UV−vis DRS of P25 and H-P25.

Figure 3. HRTEM micrographs of P25 nanoparticles.

Figure 5. Raman spectra of P25 before and after hydrogenation: (a)


survey; (b) magnified.

cm−1 might result from the hydrogenation-induced structural


changes at the surface.
XPS analysis of the Ti 2p region evidences a slight shift to
low binding energy. Figure 6 shows the XPS spectra of the Ti
2p doublet region after peak deconvolution and background
removal. Comparing with pristine P25 with symmetrical Gauss
distribution (Figure 6a), the 2p doublet peaks of the
hydrogenated P25 exhibited a tail in the region of lower
Figure 4. HRTEM micrographs of the hydrogenated P25 nano-
binding energy, which indicated the presence of lower valence
particles. states of Ti, as shown in Figure 6b. The small peaks at 457.6 eV
for Ti 2p3/2 and 463.3 eV for Ti 2p1/2 clearly indicate the
presence of Ti3+. In addition, the main sharp peaks of Ti 2p3/2
changes on the surface of P25 TiO2 before and after located at 458.3 eV and Ti 2p3/2 located at 464.0 eV were
hydrogenation. As shown in Figure 5, pristine P25 TiO2 assigned to Ti4+ in TiO2. The Ti3+/Ti4+ ratio was calculated to
exhibited Eg modes (143 cm−1, 198 cm−1 and 639 cm−1), a be approximately 14% from the peak areas. In general, the
B 1g mode (399 cm −1 ) and an A 1g mode (515 cm −1 existence of Ti3+ in TiO2 indicates the oxygen vacancies will be
superimposed by 519 cm−1), in good agreement with a former generated to maintain electrostatic balance according to the
report.24 However, besides the peaks appearing in pristine P25, following chemical equation:
the Raman peaks of the hydrogenated P25 TiO2 became
fluctuated more, and two weak peaks emerged at 316 cm−1 and 4Ti4 + + O2 − → 4Ti4 + + 2e−/□ + 0.5O2
810 cm−1, as shown in Figure 5b. The band centered at about → 2Ti4 + + 2Ti 3 + + □ + 0.5O2
320 cm−1 should be assigned to second-order scattering, which
always showed an A1g symmetry component.25 In addition, the The □ represents an empty position originating from the
band at 810 cm−1 still could not be assigned to any type of removal of O2− in the lattice. It can be deduced that an oxygen
anatase Raman-active modes. A new band at 321.1 cm−1 was vacancy generated companied with two Ti3+ ions from the
also observed in the surface-hydrogenated anatase TiO 2 equation. Therefore, the concentration of oxygen vacancy in
nanowire microspheres.17 Therefore, the phenomena of the the TiO2 lattice is calculated to be 3% in H-P25. Naldoni et
emerging of the second-order scattering and new band at 810 al.18 reported a 5% concentration of oxygen vacancy, which
22621 dx.doi.org/10.1021/jp307573c | J. Phys. Chem. C 2012, 116, 22619−22624
The Journal of Physical Chemistry C Article

Figure 7. Positron lifetime spectra with fitting lines of P25 (blue) and
H-P25 (red).

Table 1. Positron Lifetimes and Relative Intensities with


Their Deviations of Pristine P25 and H-P25
τ3 I3
sample τ1 (ps) τ2 (ps) (ns) I1 (%) I2 (%) (%)
P25 value 140.5 376.9 NA 11.22 88.78 NA
deviation 14.1 2.1 NA 1.08 1.08 NA
H-P25 value 188.1 378.7 19.3 14.30 85.45 0.25
deviation 19.9 3.8 10.7 2.74 2.75 0.07

a straight line, because they closed to the time resolution. The


slop of the straight line was according to λ = Δ ln D(t) /Δt
strictly. The λ represents annihilation rate, which is the
reciprocal of the τ.
The shorter component (τ1) is generally due to the free
Figure 6. Ti 2p XPS spectra of P25 (a) and H-P25 (b).
annihilation of positrons in a defect-free crystal, denoted as the
first lifetime of positrons.27 In a disordered system, smaller
vacancies (such as monovacancies, etc.) or shallow positron
indicated the fast cooling rate could be a plus to keep the traps (such as oxygen vacancies associated in CeO2) may
concentration of oxygen vacancy in hydrogenation. reduce the surrounding electron density,19 which increases the
PALS provides information including size, type, and relative lifetime of τ1. In pristine P25, some monovacancies exist
concentration of various vacancies according to the lifetime of naturally in the crystal lattice. Herein, the τ1 for pristine P25
the positrons.21,22 Generally, the positions would undergo rapid was 140.5 ps, which could be ascribed to the inherent free
thermalization by dissipation of their energy to the surrounding positron lifetime in pristine TiO2. However, the τ1 for H-P25
medium and then diffuse through the medium with an average was remarkably longer at 188.1 ps, 33% greater than in pristine
diffusion length after entering into solid. For anatase TiO2, the P25. Therefore, the prolonged τ1 of H-P25 demonstrates that a
position diffusion length is approximately 45 nm referring to huge number of small neutral Ti3+−oxygen vacancy associates
the previous report,26 which is much longer than the grain size were introduced into the TiO2 lattice by hydrogenation.
of our sample (about 16 nm of anatase). Herein, the positron Correspondingly, the relative intensity (I1) of H-P25 was also
can diffuse throughout the lattice and reach the grain surface to increased dramatically, i.e., from 11.22% in pristine P25 to
annihilation. 14.3% in H-P25.
The time-dependent positron decay spectrum D(t) = ∑iIi The longer lifetime component (τ2) is attributed to positrons
exp(−t/τi) was analyzed as the sum of exponential decay captured by defects of larger size.27 The average electron
components convoluted with the Gaussian resolution function density in larger-sized defects is lower than that in small size
of the spectrometer. Figure 7 illustrates the typical positron defects, which decreases the annihilation rate and increases the
lifetime spectra from the samples P25 and H-P25. The positron positron lifetime correspondingly. Therefore, the values of τ2
lifetime spectrum N(t) = ∑iIi/τi exp(−t/τi) and the absolute were much larger than the values of τ1. In our samples, both
value of the time were the deconvolution of the positron decay P25 and H-P25, the positions could diffuse throughout the
spectrum D(t) by the PATFIT program. The fitting results are lattice and reach the grain boundaries, where there existed a
listed in Table 1 with three positron lifetime components, τ1, τ2, huge number of boundary-like defects. So the longer lifetime
and τ3, and the relative intensities, I1, I2, and I3, for pristine P25 was generated with larger proportion. For pristine P25, the τ2
and H-P25. The obtained lifetime components τ2 for both P25 was 376.9 ps with the intensity (I2) of 88.78%. Obviously, the
and H-P25 and τ3 for H-P25 were added as straight lines in the boundary-like defect accounts for more proportion than that of
plots for illustration. The τ1 components were not indicated as natural monovacancies in pristine P25 due to I2 > I1. For H-
22622 dx.doi.org/10.1021/jp307573c | J. Phys. Chem. C 2012, 116, 22619−22624
The Journal of Physical Chemistry C Article

P25, the τ2 was increased to 378.7 ps because of hydrogenation,


which demonstrated that some larger size vacancy clusters were
induced inside the hydrogenated TiO2, which resulted from
interaction between the small neutral Ti3+−oxygen vacancy
associates. However, the increase of τ2 was only 0.4% for the
monovacancies introduced predominately by hydrogenation.
The decrease of I2 in H-P25 was due to the dramatic increase of
I1, because of the relationship of I1 + I2 + I3 = 100%.
The longest lifetime component (τ3) was probably assigned
to the annihilation of orthopositronium atoms formed in the
large voids (nanoscale) in materials.27 For the present work, τ3
= 19.3 ns was detected only in H-P25, and no τ3 single in
pristine P25. However, the intensity of τ3 (I3) was also so slight
that it was only 0.25% in the hydrogenated TiO2. That is to say,
few voids of oxygen vacancy associates in nanoscale were
formed in the hydrogenated TiO2 crystalline grains.
Hydrogen gas (H2) is generally used as a reducing agent. By
heating TiO2 in a hydrogen atmosphere, numerous oxygen Figure 8. PL spectra of P25 and H-P25.
vacancies, associated with Ti3+ were introduced in H-P25. Thus
the shorter lifetime component (τ1) of positrons was prolonged TiO2 samples.28 For pristine P25, the intensities of the
by the oxygen vacancy associates, where the surrounding secondary peaks (at 451 and 469 nm) are lower than their
electron density decreased. Some of the oxygen vacancy main peaks (at 400 and 412 nm). However, for H-P25, the
associates tended to flock to larger size vacancy clusters and relative intensity of the secondary peaks was enhanced
few voids in nanoscale, which increased the τ2 and generated τ3 remarkably, when compared with the main peaks. Particularly,
accordingly. Herein, the concentration of the small neutral the intensity of peak at 469 nm reached the summit of the
Ti3+−oxygen vacancy associates, larger size vacancy clusters, survey spectrum, which demonstrated the formation of oxygen
and voids of vacancy associates decreased subsequently. The defects in the TiO2 lattice by hydrogenation. However, for the
vacancy associates also disordered the surface structure of lower intensity of background, the absolute intensity of the
modified TiO2 and thus blurred the atomic lattice HRTEM secondary PL peaks of H-P25 was lower than that of pristine
images, as shown in Figure 4. Moreover, the surface structural P25.
variations were caused by vacancies in the modification of It has been well-known that there is a positive correlation
hydrogenation, thereby the emergence of the second-order between e−−h+ separation efficiency, photocurrent, and
scattering and new band at 810 cm−1 of Raman vibration were photocatalytic activity. Figure 9 illustrates the photocurrent
brought about in the hydrogenated TiO2. Oxygen vacancy
defects on the surface of TiO2 play an essential part in
governing the adsorption of O2 molecules, which interact with
Ti3+ sites or act as electron scavengers.13 Additionally, the Ti3+
species from the removal of oxygen atoms would act as holes
scavengers. These behaviors could retard the charge recombi-
nation and therefore improve the photocatalytic activity.
The measurement of PL emission was performed to reveal
the efficiency of charge carrier trapping, migration, transfer, and
separation, and to understand the fate of photogenerated
electrons and holes in semiconductors, since PL emission
results from the recombination of free carriers.28 Figure 8
shows the comparison of PL spectra of pristine and
hydrogenated P25 TiO2 in the wavelength range of 350−550
nm with the excitation at 300 nm. The main emission peaks of
pristine P25 and H-P25 appeared at 400 and 412 nm, which
were attributed to the emission of bandgap transition with the
energy of emission corresponding to the bandgap energy of
anatase (3.2 eV) and rutile (3.0 eV), respectively. The PL Figure 9. Photocurrents of P25 and H-P25.
emission intensity of H-P25 at 400 and 412 nm are only one-
third of the pristine P25, which indicated that the profiles of pristine P25 and H-P25. It was found that value of
recombination rate of photogenerated electrons and holes the H−P25 photocurrent was 3 times higher than that of
had been inhibited considerably in H-P25, because of the pristine P25, in good agreement with the PL test.
formation of oxygen vacancies during the hydrogenation. The Figure 10 shows the time profiles of ln(C0/C) under UV−vis
oxygen vacancies actually served as electron capture traps, and irradiation for degradation of MB aqueous solution, where C is
hence separated the charge carriers and reduced the the concentration of MB at irradiation time t, and C0 is the
recombination significantly. concentration in absorption equilibrium of photocatalysts
The secondary PL emission peaks at 451 and 469 nm were before irradiation. The linear relationship between the ln(C0/
equivalent to 2.75 and 2.65 eV, respectively. These PL signals C) and irradiation time confirmed the stability of the TiO2
were attributed to surface oxygen vacancies and defects of the before and after hydrogenation. The line slope of H-P25 was 2
22623 dx.doi.org/10.1021/jp307573c | J. Phys. Chem. C 2012, 116, 22619−22624
The Journal of Physical Chemistry C Article

(12) Wang, C.; Pagel, R.; Bahnemann, D. W.; Dohrmann, J. K. J.


Phys. Chem. B 2004, 108, 14082−14092.
(13) Thompson, T. L.; Yates, J. T. Top. Catal. 2005, 35, 197−210.
(14) Chen, X.; Liu, L.; Yu, P. Y.; Mao, S. S. Science 2011, 331, 746−
450.
(15) Wang, G.; Wang, H.; Ling, Y.; Tang, Y.; Yang, X.; Fitzmorris, R.
C.; Wang, C.; Zhang, J. Z.; Li, Y. Nano Lett. 2011, 11, 3026−3033.
(16) Zhang, Z.; Bai, M.; Guo, D.; Hou, S.; Zhang, G. Chem. Commun.
2011, 47, 8439−8441.
(17) Zheng, Z.; Huang, B.; Lu, J.; Wang, Z.; Qin, X.; Zhang, X.; Dai,
Y.; Whangbo, M. Chem. Commun. 2012, 48, 5733−5735.
(18) Naldoni, A.; Allieta, M.; Santangelo, S.; Marcello., Marelli;
Fabbri, F.; Cappelli, S.; Bianchi, C. L.; Psaro, R.; Santo, V. D. J. Am.
Chem. Soc. 2012, 134, 7600−7603.
(19) Liu, X.; Zhou, K.; Wang, L.; Wang, B.; Li, Y. J. Am. Chem. Soc.
2009, 131, 3140−3141.
(20) Kong, M.; Li, Y.; Chen, X.; Tian, T.; Fang, P.; Zheng, F.; Zhao,
X. J. Am. Chem. Soc. 2011, 133, 16414−16417.
(21) Cruz, R. M.; Pareja, R.; Gonzalez, R. Phys. Rev. B 1992, 45,
Figure 10. Photocatalytic degradation of MB by P25 and H-P25. 6581−6586.
(22) Dutta, S.; Chattopadhyay, S.; Jana, D.; Banerjee, A.; Manik, S.;
times higher than that of pristine P25, which demonstrated the Prahan, S. K.; Sutradar, M.; Sarkar, A. J. Appl. Phys. 2006, 100,
114328−114333.
superior photocatalytic property of H-P25.
(23) Zhang, J.; Zhang, Y.; Lei, Y.; Pan, C. Catal. Sci. Technol. 2011, 1,
273−278.
4. CONCLUSIONS (24) Zhang, W. F.; He, Y. L.; Zhang, M. S.; Yin, Z.; Chen, Q. J. Phys.
As a sensitive approach for measuring vacancy-type defects, D 2000, 33, 912−916.
PALS provides an effective route to qualitatively characterize (25) Giarola, M.; Sanson, A.; Monti, F.; Mariotto, G. Phys. Rev. B
the oxygen-vacancy-induced defects in TiO2. It will be of 2010, 81, 174305−174311.
significance to evaluate and understand the importance of (26) Zhang, Y.; Ma, X.; Chen, P.; Li, D.; Pi, X.; Yang, D.; Coleman,
oxygen vacancies, the associated with Ti3+ sites in TiO2 as P. G. Appl. Phys. Lett. 2009, 95, 252102−252104.
electron and hole scavengers, which thus separated the charge (27) Sanyal, D.; Banerjee, D. Phys. Rev. B 1998, 58, 15226−15230.
(28) Xiang, Q.; Lv, K.; Yu, J. Appl. Catal., B 2010, 96, 557−564.
carriers and remarkably improved the photocatalytic activity of
TiO2.

■ AUTHOR INFORMATION
Corresponding Author
*Tel.: +86-27-68752481 ext. 5201; Fax.: +86-27-68752003; e-
mail: cxpan@whu.edu.cn.
Notes
The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS
This research was financially supported by the National Basic
Research Program of China (973 Program) (No.
2009CB939705 and No. 2009CB939704).

■ REFERENCES
(1) Fujishima, A.; Honda, K. Nature 1972, 238, 37−38.
(2) Zhang, J.; Xi, J.; Ji, Z. J. Mater. Chem. 2012, 22, 17700−17708.
(3) Jiang, X.; Shi, A.; Wang, Y.; Li, Y; Pan, C. Nanoscale 2011, 3,
3573−3577.
(4) Chen, X.; Mao, S. S. Chem. Rev. 2007, 107, 2891−2959.
(5) Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Science
2001, 293, 269−271.
(6) Irie, H.; Watanabe, Y.; Hashimoto, K. Chem. Lett. 2003, 32, 772−
773.
(7) Yu, J. C.; Yu, J.; Ho, W.; Jiang, Z.; Zhang, L. Chem. Mater. 2002,
14, 3808−3816.
(8) Ohno, T.; Akiyoshi, M.; Umebayashi, T.; Asai, K.; Mitsui, T.;
Matsumura, M. Appl. Catal., A 2004, 265, 115−121.
(9) Jiang, X.; Wang, Y.; Pan, C. J. Alloys Compd. 2011, 509, L137−
L141.
(10) Zhang, Y.; Fei, L.; Jiang, X.; Pan, C.; Wang, Y. J. Am. Ceram. Soc.
2011, 94, 4157−4161.
(11) Zhao, W.; Chen, C.; Li, X.; Zhao, J. J. Phys. Chem. B 2002, 106,
5022−5028.

22624 dx.doi.org/10.1021/jp307573c | J. Phys. Chem. C 2012, 116, 22619−22624

You might also like