Vous êtes sur la page 1sur 8

Structure and high photocatalytic activity of (N, Ta)-doped TiO2 nanoparticles

N. T. H. Le, T. D. Thanh, V.-T. Pham, T. L. Phan, V. D. Lam, D. H. Manh, T. X. Anh, T. K. C. Le, N. Thammajak,
L. V. Hong, and S. C. Yu
Citation: Journal of Applied Physics 120, 142110 (2016); doi: 10.1063/1.4961718
View online: http://dx.doi.org/10.1063/1.4961718
View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/120/14?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
Bandgap tailoring of in-situ nitrogen-doped TiO2 sputtered films intended for electrophotocatalytic applications
under solar light
J. Appl. Phys. 116, 153510 (2014); 10.1063/1.4898589
Cationic (V, Y)-codoped TiO2 with enhanced visible light induced photocatalytic activity: A combined
experimental and theoretical study
J. Appl. Phys. 114, 183514 (2013); 10.1063/1.4831658
Preparation and characterization of Eu 3+ doped In 2 O 3 nanoparticles
AIP Conf. Proc. 1512, 448 (2013); 10.1063/1.4791104
Visible light photocatalytic activity in nitrogen-doped TiO 2 nanobelts
Appl. Phys. Lett. 94, 093101 (2009); 10.1063/1.3093820
Hydrothermal synthesis and visible light photocatalysis of metal-doped titania nanoparticles
J. Vac. Sci. Technol. B 25, 430 (2007); 10.1116/1.2714959

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 120.188.94.247 On: Fri, 09 Sep 2016
16:55:59

JOURNAL OF APPLIED PHYSICS 120, 142110 (2016)

Structure and high photocatalytic activity of (N, Ta)-doped TiO2 nanoparticles


N. T. H. Le,1 T. D. Thanh,1,2,a) V.-T. Pham,3,4 T. L. Phan,2 V. D. Lam,1 D. H. Manh,1
T. X. Anh,3 T. K. C. Le,5 N. Thammajak,6 L. V. Hong,1 and S. C. Yu2,a)

1
Institute of Materials Science, Vietnam Academy of Science and Technology, 18-Hoang Quoc Viet, Hanoi,
Vietnam
2
Department of Physics, Chungbuk National University, Cheongju 28644, South Korea
3
Center for Quantum Electronics, Institute of Physics, Vietnam Academy of Science and Technology, Hanoi,
Vietnam
4
Synchrotron SOLEIL, LOrme des Merisiers, Bo^te Postale, 48, 91192 Gif-sur-Yvette Cedex, France
5
Institut des Sciences Moleculaires dOrsay, CNRS, Univ Paris-Sud, 91405 Orsay Cedex, France
6
Synchrotron Light Research Institute, 111 University Avenue, Muang, Nakhon Ratchasima 30000, Thailand

(Received 8 April 2016; accepted 31 July 2016; published online 30 August 2016)
A hydrothermal method was used to prepare three nano-crystalline samples of TiO2 (S1), N-doped
TiO2 (S2), and (N, Ta)-codoped TiO2 (S3) with average crystallite sizes (D) of 1325 nm. X-ray
diffraction studies confirmed a single phase of the samples with a tetragonal/anatase structure. A
slight increase in the lattice parameters was observed when N and/or Ta dopants were doped into
the TiO2 host lattice. Detailed analyses of extended X-ray absorption spectra indicated that N- and/
or Ta-doping into TiO2 nanoparticles influenced the co-ordination number and radial distance (R)
of Ti ions in the anatase structure. Concerning their absorption spectra, (N, Ta)-doping narrowed
the band gap (Eg) of TiO2 from 3.03 eV for S1 through 2.94 eV for S2 to 2.85 eV for S3. Such
results revealed the applicability of these nanoparticles in the photocatalytic field working in the
ultraviolet (UV)-visible region. Among these, photocatalytic activity of S3 was the strongest. By
using S3 as a catalyst powder, the degradation efficiency of methylene blue solution was about
99% and 93% after irradiation of UV-visible light for 75 min and visible-light for 180 min, respectively. Published by AIP Publishing. [http://dx.doi.org/10.1063/1.4961718]
I. INTRODUCTION

So far, titanium dioxide (TiO2) is regarded as one of the


most fascinating materials for environmental applications. It
is widely investigated as a photocatalyst due to its high
photo-activity. Besides, it presents numerous advantages,
such as low production cost, its environment friendly nature,
and its high chemical stability. We know that, TiO2 has three
different polymorphs, which are anatase, rutile, and brookite.1 Herein, Ti4 ions are co-ordinated to six oxygen (O2)
ions, forming octahedral TiO6.2 Among these, anatase TiO2
is an n-type semiconductor with the band-gap energy
Eg 3.2 eV,3 which can be used in the split-water application to make hydrogen and oxygen without an external bias
under ultraviolet (UV) irradiation.4
For solar-cell applications, it is well known that 43% of
the whole solar energy belongs to the visible-light range
while only 4% belongs to UV light. TiO2 is thus not appropriate for the solar-light-energy conversion. Because of this
reason, many studies have been carried out to improve its
optical properties upon impurity doping and intrinsic
defects.515 Among these, N-doping into TiO2 indicates a
great promise in achieving visible light active photocatalysis.7,8,16 With a small ionization energy and high stability,
N can be easily introduced in the TiO2 structure due to its comparable atomic size with oxygen. Recently, Obata and coThis paper is part of the Special Topic Cutting Edge Physics in Functional
Materials published in J. Appl. Phys. 120, 14 (2016).
a)
Authors to whom correspondence should be addressed. Electronic addresses:
thanhxraylab@yahoo.com and scyu@cbnu.ac.kr
0021-8979/2016/120(14)/142110/7/$30.00

workers5 reported that (N, Ta)-codoping into TiO2 can narrow


the band gap, making absorption edges shifted towards longer
wavelengths under increasing doping concentration.
In fact, many efforts have been deployed to dope TiO2
with non-metals (including N, C, and S),510,16,17 with metals
(such as Cr, Fe, Ta, Ag, and Pt),5,11,12,18,19 and multi-phases
as well as composites.1315,20 With this route, photoresponse from the UV to the visible light region would be
efficiently extended. Furthermore, some theoretical calculations have also been performed and suggested that TiO2
doped with non-metals has considerable effects on the bandgap energy (Eg).21 Photocatalytic activity of TiO2 in the
visible-light region can be increased by N-doping. This provides good opportunities for extensive applications related to
oxidation of CO, ethanol, gaseous 2-propanol, acetaldehyde,
NOx, and the decomposition of dyes such as methylene blue
(MB). To the best of our knowledge, very few reports have
been published on (N, Ta)-codoped TiO2 based photocatalytic nanoparticles for photo-oxidative degradation of
organic pollutants.5,22 In reference to this problem, we prepared TiO2 nanoparticles doped with N and/or Ta. The influence of doping on their crystal structure was then studied by
means of X-ray diffraction and absorption. Photocatalytic
and surface adsorption properties for MB decomposition
were also taken into account.
II. EXPERIMENTAL

Three samples of pure TiO2, N-doped TiO2, and (N,


Ta)-codoped TiO2 nanoparticles (denoted as S1, S2, and S3,

120, 142110-1

Published by AIP Publishing.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 120.188.94.247 On: Fri, 09 Sep 2016
16:55:59

142110-2

Le et al.

J. Appl. Phys. 120, 142110 (2016)

FIG. 1. FE-SEM images for the nanocrystalline samples of (a) S1, (b) S2, and (c) S3.

respectively) were prepared by a hydrothermal method at


200  C for 24 h. Tetraisopropyl orthotitanate, isopropanol,
NH4OH, Ta, HF, and HNO3 were used as initial materials.
NH4OH and homogeneous solutions of Ta in HNO3 and HF
acids were used as nitrogen and Ta sources. The obtained
products from the hydrothermal process were dried in an
oven for 1 h at 70  C, and then annealed at 300  C for 1 h in
air with a heating rate of 10  C/min. X-ray diffraction (XRD)
patterns of final nanoparticles at room temperature were
recorded in the range 2h 20 80 by using a SIEMENS ). The
D5000 diffractometer (Cu-Ka radiation, k 1.5406 A
surface morphology of the nanoparticles has been checked by
using a field-emission scanning electron microscope (FE-SEM
S4800, Hitachi). The surface area of the nanoparticles has
been measured on a Quantachrome instrument (AutosorbiQ-MP) based on the BET (Brunauer-Emmett-Teller) method
with a Nitrogen adsorption. Room temperature extended X-ray
absorption fine structure (EXAFS) spectra in the transmission
mode for the Ti K-edge (with E0 4966.4 eV) were collected
at the BL8 beam line (Siam Photon Source, Synchrotron Light
Research Institute, Thailand). EXAFS analyses were performed by using a software package Ifeffit.23
UV-visible absorption spectra of nanoparticles in powder of S1, S2, and S3 samples were obtained by using a
UV-visible spectrophotometer (Cary Varian 5000). The photocatalytic activity of these nanoparticles (herein S1, S2, and
S3 as the catalyst powers) was evaluated by measuring the
decomposition of the distilled water solution of MB (with a
concentration of 5 mol/l) under UV-visible light irradiations.
UV-visible irradiations were carried out by using a 400 W
Hg-Xe lamp with a power density of 15 mW/cm2 with and
without an optical filter (the UV lights with wavelengths

below 420 nm were removed by this filter). The UV-visible


absorption spectra of the distilled water solution of MB with
a suspension of the catalyst powers before and after irradiation were recorded on the UV-visible spectrophotometer
UV-VIS (JASCO V-550).
III. RESULTS AND DISCUSSION

The FE-SEM images of nanoparticles after fabrication


are shown in Fig. 1. It appears that the surface morphology
of nanoparticles is quite uniform. Average sizes of nanoparticles are estimated to be less than 20 nm for S1 and S3, and
larger than 20 nm for S2. With these nanosizes, it is expected
to have a large surface-area-to-volume ratio, which increases
photocatalytic activity of materials.
Together with morphology investigations, the crystal
structure of the as-prepared samples S1S3 was also checked
by recording their XRD patterns at room temperature. XRD
patterns for the samples are shown in Fig. 2(a). The Millerindexed peaks indicate that the samples are in a single phase
of TiO2 in the anatase/tetragonal structure, belonging to
space group I41/amd. If more attention is given to XRD
peaks, one can see a slight shift in their diffraction peaks
towards smaller angles. This confirms a modification in the
anatase structure when TiO2 nanoparticles are doped with N
and/or Ta. Based on these XRD data, we calculated the lattice parameters (a, b, and c) and the unit-cell volume (V) for
the samples, as shown in Table I. It points out that N and/or
Ta doping into TiO2 nanoparticles increased slightly the lattice parameters, as well as the unit-cell volume. This is
explained to be due to N3 and Ta4 ions with larger ionic
, rTa4 0.68 A
) substituted for O2 and/
radii (rN3 1.46 A
4

) in the TiO2 host


or Ti ions (rO2 1.40 A, rTi4 0.65 A

FIG. 2. (a) Room-temperature XRD


patterns and (b) the W-H analyses
for nanocrystalline samples S1S3.
The solid lines in (b) are the fitting
curves to the function bS cos h (Kk/D)
2e sin h.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 120.188.94.247 On: Fri, 09 Sep 2016
16:55:59

142110-3

Le et al.

J. Appl. Phys. 120, 142110 (2016)

TABLE I. Average crystallite size, lattice parameters, and band-gap energy


of the samples.
Samples
S1
S2
S3

D (nm)

)
a (A

)
c (A

c/a

3)
V (A

Eg (eV)

13
25
18

3.780
3.790
3.793

9.474
9.509
9.509

2.506
2.509
2.507

135.34
136.57
136.79

3.03
2.94
2.85

lattice. By using the Williamson-Hall (W-H) method,


detailed descriptions were shown elsewhere,24 the average
values of the crystallite size (D) and the lattice strain (e)
could be obtained from the intercept and the slope, respectively, of the relation bS cos h (Kk/D) 2e sin h, where bS
is the full width at half maximum of the XRD peak, h is the
Bragg angle, and K 0.9 is the shape factor. W-H analyses
for the samples are plotted in Fig. 2(b). As a result, D values
are about 13, 25, and 18 nm, and e values are about
7.6  104, 7.2  105, and 2.1  104 for S1, S2, and S3
samples, respectively. These D values are in good agreement
with the results obtained from FE-SEM observations, as
mentioned above. The increase of D in doped-TiO2 anatase
could be related to a distortion of the lattice. N3 and Ta4
ions with larger ionic radii substituted for O2 and Ti4 ions,
respectively, distorted the TiO2 anatase crystalline lattice.
This distortion is a positive process in the formation and the
growing up of the nanoparticles in the anatase structure.25,26
Additionally, it is known that the specific surface area of the
particles is the summation of the areas of the exposed surfaces of the particles per unit mass. The BET method is the
most common method for determining the surface area of
the powders or porous materials, where Nitrogen gas is generally used as an analysis gas. In this work, the surface area
for S1, S2, and S3 samples are found to be 145.5, 59.0, and
109.5 m2/g, respectively. Clearly, there is an inverse relationship between the surface area and particle size in the
nanoparticles. It means that the surface area of S2 is the lowest, corresponding to the largest average size.
To further understand the local crystal structure of the
samples, we have studied their EXAFS spectra for the Ti
K-edge. The EXAFS signals for the Ta K-edge are very
weak, and thus are out of our attention. In this work, all spectra were recorded in the transmission mode at room temperature using a double monochromator equipped with Ge (2 2 0)
crystal. Energy calibration for the monochromator before the
measurements was performed using Ti foil at the K-edge. We
have also used the commercial TiO2 anatase powder as a
comparative sample. The EXAFS spectra of the nanoparticles
and the commercial TiO2 anatase powder are shown in Fig.
3(a). One can see that there is not very much difference in the
features of the absorption edge of nanoparticles in comparison with the commercial TiO2 anatase powder. This indicates
that the anatase structure of nanoparticles is not significantly
changed when N and/or Ta are doped into the TiO2 host lattice. By performing the Fourier transform (FT) of k3v(k)
EXAFS spectra, we can see a small difference in the peak
positions associated with bond distances of Ti-O and Ti-Ti,
and the peaks intensities of the Ti-O scattering path, see Fig.
3(b). Comparing with sample S1 and commercial anatase

FIG. 3. (a) EXAFS spectra recorded for the Ti K-edge and (b) their Fouriertransformed EXAFS spectra in the R space. The spectra of commercial anatase TiO2 powders are also shown for reference (denoted as Anatase).

TiO2 powder, these distances become slightly shorter when


TiO2 nanoparticles are doped with N and/or Ta (for the samples S2 and S3). To determine quantitatively these distances,
we have performed an EXAFS refinement of the commercial
anatase and S1S3 samples based on the anatase-structural
model. Scattering amplitude and phase shift for the anatase
structure were calculated by using ab initio (FEFF9).27 The
amplitude reduction S02 was fixed to be 0.8 for the Ti K-edge
EXAFS.28 Three first shells about the Ti ion were included in
the fit. According to the FEFF calculation, the first and the
second peaks in these FT spectra are from single scattering
paths of Ti-O and Ti-Ti, respectively, see Fig. 3(b). The third
FT peak contains single- and multiple-scattering contributions from a variety of paths such as Ti-Ti and Ti-O-Ti,
respectively. The EXAFS refinement is performed in the
for the FT of
r-space with a fitting range of r 13.8 A
1.
k3-weighted EXAFS with k ranging from 3.612.85 A
EXAFS fits in the k-space and the r-space are displayed in
Fig. 4. An overall good fit is obtained for all samples as
reflected by the fit goodness R-factors. Best fit parameters
such as the distances (R), the coordination number (CN), and
the Debye-Waller factor (r2) of the above scattering paths are
summarized in Table II. The changes of these parameters
related to the doped samples (S2 and S3) suggest us the incorporation of N and Ta dopants into the TiO2 host lattice.
Notably, the coordination number of Ti ions in the octahedral
TiO6 of the anatase structure was lower than 6 (S1 and S3)
suggesting a coexistence of Ti3 and Ti4 in the nanoparticles. It means that there is an existence of the high oxygen

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 120.188.94.247 On: Fri, 09 Sep 2016
16:55:59

142110-4

Le et al.

J. Appl. Phys. 120, 142110 (2016)

FIG. 4. EXAFS refinement of commercial anatase TiO2 and nanocrystalline


samples S1S3 at the Ti K-edge. (a) k3weighted EXAFS spectra in k-space and
(b) the corresponding Fourier transform
in r-space.

vacancies in these nanoparticles. A lower Ti-O CN of S1 and


S3 compared to that of S2 is consistent with the smaller particle sizes found by XRD. Interestingly, reductions in the bond
lengths of Ti-O and Ti-Ti are found in the doped samples
with respect to the undoped one. This fact that is not contradictory to the increasing lattice parameters resulted from the
XRD study since a direct comparison between EXAFS and
XRD results are not possible, as EXAFS is sensitive to the
local environment whereas XRD is sensitive to the long range
order. Since EXAFS probes the local environment, its signal
can be contributed not only by the crystalline phase but also
by the amorphous phase while the information of the nonBragg diffracting amorphous phase cannot be achieved from
XRD data. The decreasing Ti-O bond lengths in doped
TABLE II. Best-fit parameters to the Ti K-edge EXAFS for commercial anatase TiO2 powder, and the nanocrystalline samples S1S3.
Samples

CN

r2

R-factor

Commercial anatase
TiO2 powder

Ti-O
Ti-Ti1a
Ti-Ti2
Ti-O-Ti

6.4
2.8
5.9
11.8

1.971
3.055
3.803
3.872

0.0063
0.0019
0.0107
0.0107

0.0079

S1

Ti-O
Ti-Ti1
Ti-Ti2
Ti-O-Ti

5.3
3.1
6.2
12.4

1.961
3.047
3.794
3.858

0.0057
0.0029
0.0098
0.0098

0.0176

S2

Ti-O
Ti-Ti1
Ti-Ti2
Ti-O-Ti

6.0
3.01
5.81
11.6

1.949
3.040
3.784
3.841

0.0056
0.0021
0.0103
0.0103

0.0064

Ti-O
Ti-Ti1
Ti-Ti2
Ti-O-Ti

5.1
2.5
5.4
10.8

1.954
3.044
3.790
3.849

0.0048
0.0022
0.0104
0.0104

0.0084

S3

In the main text, the annotation of Ti-Ti often refers to Ti-Ti1.

samples is explained by a rather high doping concentration of


10% wt. N (0.57 N/Ti ratio), that contains not only substitutional but also interstitial N atoms in the host lattice.29 A
slight reduction of Ti-Ti bond lengths in the doped sample
compared to the undoped TiO2 is believed to be related to a
small structural distortion when dopants are introduced to the
host lattice. This bond length reduction was also observed in
the EXAFS study of TiO2 crystallized at high pressure.30
Following the structural and the morphology analyses,
the photocatalytic ability of the nanoparticles has been
studied by using absorption spectroscopy operating in the
wavelength range k 200800 nm. UV-visible spectra of the
nanoparticles are shown in Fig. 5(a). There is a shift of
the absorption edge and distinct shoulders towards longer
wavelengths when TiO2 nanoparticles are doped with N and/
or Ta. There is an absorption tail in the long wavelength
region, which is enhanced in S2 and S3 samples. These
results could be related to local lattice distortions and defects
in the anatase structure caused by N and Ta dopants, which
are responsible for the shift of the onset of their absorption
edge near 400 nm and for the absorption in the visible
region.
It is known that the relation between the absorption
coefficient (a) and incident photo energy (E h with h is
the Plancks constant and  is the frequency of the incident
light) can be written as follows:31
ah A1 h  Eg 1=2 ;

(1)

ah A2 h  Eg 2 ;

(2)

where Eq. (1) for the direct transition, Eq. (2) for the indirect
transition, A1 and A2 are constants, and Eg is the band-gap
energy. Assuming that N- and/or Ta-doped TiO2 is an indirect semiconductor (similar to TiO2), a plot of the square
root of the absorption coefficient a versus the energy E is

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 120.188.94.247 On: Fri, 09 Sep 2016
16:55:59

142110-5

Le et al.

J. Appl. Phys. 120, 142110 (2016)

FIG. 5. (a) UV-visible spectra and (b)


the square root of absorption coefficient versus the photon energy for the
nanocrystalline samples S1S3. The
solid lines in (b) are linear fitting to
determine the value of Eg.

shown Fig. 5(b). The band-gap energy (Eg) values of nanoparticles thus can be evaluated by extrapolating the linear
part of the curves to zero. For the undoped TiO2 (S1) sample,
Eg 3.03 eV is a little bit smaller than that of pure TiO2
(Eg 3.2 eV).3 This can be due to more effects generated by
nano-sizes of the undoped TiO2 particles. However, Eg of
the N-doped (S2) and (N, Ta)-codoped (S2) samples are
slightly decreased; those are found to be 2.94 eV and 2.85 eV
for S2 and S3, respectively (see Table I). Clearly, the (N,
Ta)-codoping into TiO2 nanoparticles narrows significantly
the band-gap width of 0.18 eV. This value is higher than
that in the case of N-doped TiO2 (0.09 eV) because the substitution of N for O in TiO2 narrows the band gap, which can
be related to the mixing of N 2p states with O 2p states.
Basically, N dopants in the TiO2 host lattice form donor
states with energy levels just above the top of the valence
band. This leads to the narrowing of the band gap of TiO2.5
Additionally, the oxygen vacancies would be created when
TiO2 nanoparticles are doped with N and/or Ta as mentioned
above on the EXAFS results, which may possibly create
shallow trap states that lead to the narrowing of band-gap.
Such defects enhance visible light absorbance.32 According
to Obata and co-workers,5 calculations of the electronic
structure of (Ta, N)-codoped TiO2 also reveal the lower
energy (positive) shift of the N 2p narrow band. It is also
suggested that complete Ta and N neighboring can lead to
the hybridization of N 2p and O 2p, constituting the valence
band of TiO2.5
In an attempt to evaluate photocatalytic activity, we
have examined the decomposition of the MB solution by
using our nanoparticles (S1, S2, and S3) as the catalyst powers under UV-visible light irradiation with different irradiation times. For each typical photocatalytic experiment, a
total of 0.15 g catalyst powders were put into 50 ml of the
MB solution and mixed well in a Pyrex glass. Prior to irradiation, suspensions were magnetically stirred in the dark for
60 min to ensure the establishment of an adsorption/desorption equilibrium. And then these suspensions were irradiated
with different times (with and without an optical filter
k  420 nm) under constant air-equilibrated conditions. The
concentrations of the remnant MB in the solutions have been
assessed based on the intensity of absorption peaks before
(C0) and after (C) irradiation. The degradation efficiency

was calculated by: Degradation efficiency (%) [(C0  C)/


C0]  100%.
The absorption spectra of the MB solution before and
after irradiation by UV-visible light for a representative sample S3 are shown in Fig. 6. One can see that the characteristic
absorption band of the MB solution is about 660 nm. After
15 min of irradiation, the intensity of the absorption peak
was significantly decreased. Besides, the absorption peak
was also shifted toward shorter wavelengths. After 75 min of
irradiation by UV-visible light, the intensity of the absorption peak decreases to zero indicating that the MB solution
has been completely decolorized. This result suggests that
our nanoparticles can be used for photocatalytic activity
under UV-visible light irradiation.
The photocatalytic degradation of MB solution over our
nanoparticles under UV-visible light irradiation is shown in
Fig. 7(a). One can see that the nanoparticles are essentially
active since they could be activated by UV-visible light due
to their small energy band-gap. The degradation efficiency
varies as a function of irradiation time. The solid lines in Fig.
7(a) are only a guide to the eye, which are performed following a power law, [(C0  C)/C0]  100% a  tn (a is a constant, t is the irradiation time, and n is an exponent). Clearly,

FIG. 6. UV-visible absorption spectra of MB solution before and after UVvisible light irradiation for different exposure times in the presence of sample S3.

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 120.188.94.247 On: Fri, 09 Sep 2016
16:55:59

142110-6

Le et al.

J. Appl. Phys. 120, 142110 (2016)

FIG. 7. Degradation efficiency (a) and


ln(C/C0) (b) versus time under UVvisible light irradiation, degradation
efficiency (c) and ln(C/C0) (d) versus
time under visible-light irradiation for
samples S1S3. The solid lines in (a)
and (c) are only a guide to the eye,
they are performed following a power
law of a  tn, herein (a) n 0.103,
0.135, and 0.045, and (c) n 0.471,
0.675, and 0.488 for S1, S2, and S3,
respectively. The solid lines in (b) and
(d) are linear fitted to determine the
degradation reaction rate.

all these experiment data can be depicted well by the power


laws. For (N, Ta)-codoped TiO2 nanoparticles (S3), we found
that their photocatalytic activities were enhanced under UVvisible light irradiation. Among the samples S1S3, sample
S3 has the highest photodegradation activity, with a MB solution conversion of 99% after 75 min irradiation by UVvisible light. To compare the degradation reaction rate of
samples, a first order kinetic rate was assumed, ln(C/
C0) k  t, where k is the degradation reaction rate. Fig. 7(b)
shows ln(C/Co) versus irradiation time for samples, herein the
straight lines are plotted for the determination of the rate of
constant. The degradation reaction rate under UV-visible light
irradiation is found to be k1 (102 min1) 4.597, 3.771,
and 5.005 for samples S1, S2, and S3, respectively.
As mentioned above, our nanoparticles have the bandgap energy Eg smaller than 3.2 eV and the absorption tails
take place in the visible-light region. These suggest that our
nanoparticles can be used for photocatalytic activity under
the visible-light irradiation. In this content, an optical filter
has been used to remove the UV-lights of k  420 nm. Fig.
7(c) shows the visible-light photocatalytic degradation of the
MB solution under the visible-light irradiation of the samples
S1S3. Clearly, the degradation rate obtained by the visiblelight irradiation is slower than that under UV-visible light
irradiation, see Fig. 7(d). Thus, the degradation reaction rate
under the visible-light irradiation is found to be k2 (102
min1) 0.424, 0.357, and 1.310 for samples S1, S2, and

S3, respectively. Among these, the degradation activity of


sample S3 was two times higher than that obtained for samples S1 and S2. After 200 min irradiation by visible-light, the
degradation of the MB solution was only about 50% for S1
and S2. Meanwhile, sample S3 showed apparent visible-light
photocatalytic activity with the degradation of 93% after
180 min irradiation.
We know that the catalyst efficiency is dependent on the
surface area of the catalyst powers. In our case, the surface area
for samples S1, S2, and S3 is 145.5, 59.0, and 109.5 m2/g,
respectively. Meanwhile, the degradation reaction ordering rate
of nanoparticles on irradiation both under UV-visible light and
visible-light is S3  S1  S2. Clearly, the degradation of MB
solution over S2 is the slowest, relating to the smallest value of
the surface area (59.0 m2/g). However, it is surprising that in
spite of the surface area of S1 (145.5 m2/g) is higher than that
obtained for S3 (109.5 m2/g), the degradation reaction rates of
S3 (k1 5.005  102 min1 and k2 1.310  102 min1)
are higher than those obtained for S1 (k1 4.597  102
min1 and k2 0.424  102 min1). It means that S3 is
more active than S1 in both cases of the irradiation. This
could be also related to another reason that is a smaller
band-gap energy of the S3, which can be attributed to the
positive shift of the N 2p narrow band caused by donor (Ta)
and acceptor (N) interactions, and the charge compensation
as well.5 Their hybridization could produce the delocalized
and dispersed valence band, generating holes with a high

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 120.188.94.247 On: Fri, 09 Sep 2016
16:55:59

142110-7

Le et al.

mobility, which is favorable for visible-light sensitive


photocatalysts.
IV. CONCLUSIONS

We have used a hydrothermal method for the preparation of high quality nanoparticles of TiO2, N-doped TiO2,
and (N, Ta)-codoped TiO2 with the pure anatase structure.
The results show that Ta and/or N doped TiO2 influences the
anatase structure and narrows significantly the band-gap
width. The examinations of the decomposition of MB solution in the presence of doped/undoped TiO2 nanoparticles
have been done with different irradiation times. Under UVvisible light irradiation in 75 min, the MB solution conversion was about 99%. However, under visible-light (with
k > 420 nm), this degradation was only about 50% after
200 min irradiation for N-doped/undoped TiO2 nanoparticles. Meanwhile, this degradation efficiency was about
93% after 180 min irradiation if we use (N, Ta)-codoped
TiO2 nanoparticles as a catalyst powder. Our results suggest
that the (N, Ta)-codoped TiO2 nanoparticles are powerful for
photocatalytic degradation of MB solution under UV or visible-light.
ACKNOWLEDGMENTS

Financial and technical supports were respectively


provided by the National Foundation for Science and
Technology Development (NAFOSTED) under Grant No.
103.01-2011.34 and National Key Laboratory for Electronic,
Institute of Materials Science-VAST in Vietnam, and the
Converging Research Center Program through the Ministry of
Science, ICT and Future Planning, Korea (2015055808).
1

N. T. Nolan, M. K. Seery, and S. C. Pillai, J. Phys. Chem. C 113, 16151


(2009).
2
D. Nicholls, Complexes and First-Row Transition Elements (MacMillan
Education, Hong Kong, 1974).
3
R. Asahi, Y. Taga, W. Mannstadt, and A. J. Freeman, Phys. Rev. B 61,
7459 (2000).

J. Appl. Phys. 120, 142110 (2016)


4

A. Kudo and Y. Miseki, Chem. Soc. Rev. 38, 253 (2009).


K. Obata, H. Irie, and K. Hashimoto, Chem. Phys. 339, 124 (2007).
6
J. Zhang, Y. Wu, M. Xing, S. A. K. Leghari, and S. Sajjad, Energy
Environ. Sci. 3, 715 (2010).
7
C. W. Dunnill, Z. Ansari, A. Kafizas, S. Perni, D. J. Morgan, M. Wilson,
and I. P. Parkin, J. Mater. Chem. 21, 11854 (2011).
8
N. T. Nolan, D. W. Synnott, M. K. Seery, S. J. Hinder, A. V.
Wassenhoven, and S. C. Pillaia, J. Hazard. Mater. 211212, 88 (2012).
9
W. J. Zhou, Y. H. Leng, D. M. Hou, H. D. Li, L. G. Li, G. Q. Li, H. Liu,
and S. W. Chen, Nanoscale 6, 4698 (2014).
10
R. G. Chaudhuri and S. Paria, Dalton Trans. 43, 5526 (2014).
11
J. S. Zhong, Q. Y. Wang, and Y. F. Yu, J. Alloys Compd. 620, 168 (2015).
12
B. L. An, Y. H. Fu, F. Z. Dai, and J. Q. Xu, J. Alloys Compd. 622, 426
(2015).
13
W. Wang, Y. Ni, and Z. Xu, J. Alloys Compd. 622, 303 (2015).
14
H. Wu, J. Fan, E. Liu, X. Hu, Y. Ma, X. Fan, Y. Li, and C. Tang, J. Alloys
Compd. 623, 298 (2015).
15
C. Liu, D. Meng, Y. Li, L. Wang, Y. Liu, and S. Luo, J. Alloys Compd.
624, 44 (2015).
16
A. Fujishima, X. Zhang, and D. A. Tryk, Surf. Sci. Rep. 63, 515 (2008).
17
H. Irie, S. Washizuka, and K. Hashimoto, Thin Solid Films 510, 21
(2006).
18
M. E. Kurtoglu, T. Longenbach, K. Sohlberg, and Y. Gogotsi, J. Phys.
Chem. C 115, 17392 (2011).
19
B. Naik and K. M. Parida, Ind. Eng. Chem. Res. 49, 8339 (2010).
20
R. Verma and S. K. Samdarshi, J. Alloys Compd. 629, 105 (2015).
21
T. Umebayashi, T. Yamaki, S. Yamamoto, A. Miyashita, S. Tanaka, T.
Sumita, and K. Asai, J. Appl. Phys. 93, 5156 (2003).
22
H. Li, S. Yin, Y. Wang, and T. Sato, Appl. Catal., B 132133, 487 (2013).
23
M. Newville, J. Synchrotron Radiat. 8, 322 (2001).
24
G. K. Williamson and W. H. Hall, Acta Metall. 1, 22 (1953).
25
H. Sun, Y. Bai, W. Jin, and N. Xu, Sol. Energy Mater. Sol. Cells 92, 76
(2008).
26
G. Yang, Z. Jiang, H. Shi, T. Xiao, and Z. Yan, J. Mater. Chem. 20, 5301
(2010).
27
J. Rehr, J. Kas, M. Prange, A. Sorini, Y. Takimoto, and F. Vila, C. R.
Phys. 10, 548 (2009).
28
K. Fukuda, I. Nakai, C. Oishi, M. Nomura, M. Harada, Y. Ebina, and T.
Sasaki, J. Phys. Chem. B 108, 13088 (2004).
29
M. Ceotto, L. L. Presti, G. Cappelletti, D. Meroni, F. Spadavecchia, R.
Zecca, M. Leoni, P. Scardi, C. L. Bianchi, and S. Ardizzone, J. Phys.
Chem. C 116, 1764 (2012).
30
P.-T. Hsiao, M.-D. Lu, Y.-L. Tung, and H. Teng, J. Phys. Chem. C 114,
15625 (2010).
31
N. Serpone, D. Lawless, and R. Khairutdinov, J. Phys. Chem. 99, 16646
(1995).
32
J. A. Rengifo-Herrera, K. Pierzchala, A. Sienkiewicz, L. Forro, J. Kiwi,
and C. Pulgarin, Appl. Catal., B 88, 398 (2009).
5

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 120.188.94.247 On: Fri, 09 Sep 2016
16:55:59

Vous aimerez peut-être aussi