Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Powder Technology 253 (2014) 608–613

Contents lists available at ScienceDirect

Powder Technology
journal homepage: www.elsevier.com/locate/powtec

Preparation and photocatalytic activity of B–N co-doped


mesoporous TiO2
Kui Zhang a, Xiangdong Wang a,⁎, Tianou He a, Xiaoling Guo b, Yaming Feng b
a
School of Science, Xi'an Jiaotong University, Xi'an 710049, Shaanxi, PR China
b
School of Textile and Materials, Xi'an Polytechnic University, Xi'an 710048, Shaanxi, PR China

a r t i c l e i n f o a b s t r a c t

Article history: B-doped and B–N co-doped mesoporous TiO2 photocatalysts were synthesized by a fast sol–gel method. The pre-
Received 18 July 2013 pared samples were characterized by X-ray diffraction (XRD), UV–vis diffuse reflectance spectrum (DRS), trans-
Received in revised form 30 November 2013 mission electron microscopy (TEM), N2 adsorption–desorption, and X-ray photoelectron spectroscopy (XPS). It is
Accepted 13 December 2013
found that the average crystallite size, pore size, and specific surface area of the co-doped TiO2 are 9.1 nm,
Available online 21 December 2013
16.8 nm, and 125.4 m2/g, respectively. Compared with that of the pure mesoporous TiO2, the absorption band
Keywords:
edge of B–N co-doped sample exhibits an evident red-shift and the absorption intensity of visible region increases
Mesoporous TiO2 obviously. The experimental results show that a structure of Ti–N–B–O is formed on the surface of the
B–N co-doping photocatalyst, indicating a synergistic effect of two dopants enhancing the photocatalytic activity, which was
Methyl blue evaluated by the degradation of methyl blue aqueous solution under UV and visible light irradiation.
Photocatalytic activity © 2013 Elsevier B.V. All rights reserved.

1. Introduction which is probably attributed to the synergistic effect between boron and
nitrogen, resulting the narrowing of the band gap. Owing to the
Titanium dioxide (TiO2) is widely investigated because of its high synergistic effect of B and N, B–N co-doped TiO2 is considered as a
photocatalytic activity, non-toxicity, and chemical stability, which much effective photocatalyst under visible light. Xing et al. [21]
make it functional in a wide range of areas such as dye-sensitized prepared the B and N co-doped TiO2 by a novel double hydrothermal
solar cells [1,2], degradation of pollutant in water and air [3–5], and method. They revealed that when nitrogen was introduced into the
hydrogen evolution [6]. Nevertheless, the application of conventional material before boron, the synergistic effect maybe aroused by the for-
TiO2 is very limited, mainly due to two problems: (1) the band gap of mation of Ti–B–N–Ti and Ti–N–B–O compounds on the surface of the
anatase TiO2 is 3.2 eV, which leads it active only in UV region, account- catalyst, while only Ti–N–B–O occurred when boron was introduced
ing for less than 5% of the solar energy; (2) the low surface area of the into the material before nitrogen. Some studies have been performed
conventional TiO2, resulting in a low efficiency when absorbing and to investigate the synergistic effect of boron and nitrogen in the TiO2
reacting. It is, therefore, urgent to develop some modifications of TiO2 catalysts [19–24], but few of these studies have focused on the modify-
to gain the novel photocatalyst, which has high surface area and can ing the crystalline structure and improving the specific surface area.
be activated by visible light. Mesoporous materials have shown unparalleled advantages in pho-
Since the report by Asahi et al. [7] revealed that nitrogen doping tocatalytic application especially due to their high specific surface area,
can significantly improve the photocatalytic activity of TiO2 under which may produce numerous surface reaction sites for the adsorption
visible light, it has been an effective means by doping non-metal ions of the reactant molecular. In addition, for mesoporous TiO2, the small
to narrow the band gap of TiO2, thus expand the light response grain size results in a shorter distance for the electrons and holes to
range to visible light region. Extensive efforts have been made in TiO2 transfer to the reaction sites [25]. Since Antonelli and Ying [26] firstly
doped with non-metal ions, such as B, C, N, S, F [8–12]. In recent years, synthesized mesoporous TiO2 in 1995, many attempts have been
the TiO2 materials co-doped with two kinds of nonmetal atoms such made to modify TiO2 by giving it mesopore structures [27–30]. Howev-
as S–N, C–N, F–N, and B–F [13–18], which are expected to get the syner- er, few studies reported the mesoporous TiO2 doped with other
gistic effect permitting tune the electronic structure and enhance the elements [19,31]. In our previous work, we developed a fast sol–gel
photocatalytic activity, have drawn more and more attention. As the method using polyacrylamide (PAM) and polyethylene (PEG) as the
co-doped elements, however, B–N co-doping has been less studied. templates to synthesis N-doped mesoporous TiO2 with high specific
Liu et al. [19] and In et al. [20] reported that B–N co-doped TiO2 exhibits surface area [32]. In the present work, B doped mesoporous TiO2 and
enhanced photocatalytic activity under UV and visible lights irradiation, B–N co-doped mesoporous TiO2 photocatalysts were prepared by the
fast sol–gel method using the tetrabutyl titanate, urea, and boric acid
⁎ Corresponding author. as Ti precursor, N source, and B source, respectively. The photocatalysts
E-mail address: wang90xd@163.com (X. Wang). with high specific surface area were characterized by X-ray diffraction

0032-5910/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.powtec.2013.12.024
K. Zhang et al. / Powder Technology 253 (2014) 608–613 609

(XRD), transmission electron microscopy (TEM), X-ray photoelectron


spectroscopy (XPS), UV–vis spectroscopy, and N2 adsorption–desorp- (c)
tion. The photocatalytic activity of the catalysts is evaluated by the pho-
tocatalytic degradation of methyl blue (MB) in water under UV and
visible light irradiation.

Intensity (a.u.)
2. Experimental
(b)

2.1. Preparation of photocatalysts

A fast sol–gel preparation of the B and B-N co-doped mesoporous (a)


TiO2 was performed as follows. 10 mL tetrabutyl titanate was dissolved
into 20 mL anhydrous ethanol with stirring to obtain solution A. 10 mL
nitric acid (5 wt.%), H3BO3 (0.055, 0.091, 0.182 g corresponding to 0.03, 10 20 30 40 50 60 70
0.05, 0.10 of B/Ti molar ratios), and urea (0, 1.76 g corresponding to 0, 1 2θ (degree)
of N/Ti molar ratios) were added to the solution of 10 mL deionized
water and 280 mL anhydrous ethanol with stirring to obtain solution Fig. 1. XRD patterns of (a) pure TiO2, (b) (0.03B)-TiO2, and (c) (0.03B, 1.0N)-TiO2 .

B. 0.05 g PAM and 8.8 g PEG was added to 20 mL anhydrous ethanol


with stirring to obtain solution C. Solution C was added to solution B 2.3. Evaluation of photocatalytic activity
while stirring to get the mixture. Then solution A was added dropwise
to the mixture above under stirring. The resultant mixture was stirred The photocatalytic activity of as-prepared samples was evaluated
at room temperature for 1 h until a white gel was obtained. The gel by degradation of methyl blue in water (MB, 10 mg/L) under UV and
was dried at 30 °C for 16 h to gain the xerogel. The resultant xerogel visible light irradiation. The catalyst (0.05 g) was added into a quartz re-
was crushed to obtain fine powder and further calcined in air at actor (100 mL), which contained 50 mL of MB solution. Prior to irradia-
400 °C for 4 h to obtain the catalyst. The samples were labeled as tion, the suspension was magnetically stirred for 40 min in the dark to
(mB, nN)-TiO2, where m and n corresponded to the initial molar reach an adsorption-desorption equilibrium. A 300 W high-pressure
ratios of B to Ti and N to Ti, respectively. For the comparison, undoped mercury lamp was used as UV radiation with a peak of 365 nm and a
TiO2 was prepared according to the above procedure in the absence of 300 W xenon lamp through a UV-cutoff filter (≤400 nm) was used as
solution B. the visible light source. The lamp was cooled by flowing water in the
quartz jacket around the lamp during the photocatalytic reaction, thus
the ambient temperature would stay constant. The analytic suspension
2.2. Characterizations (5 mL) was taken out of the reactor at regular intervals, which would be
centrifuged immediately and then filtered to separate TiO2 from the
X-ray diffraction (XRD) patterns of photocatalysts were collected at solution. The residual concentration of the MB in the remaining clear
room temperature in the range 10–75° (2θ) using a Rigaku D/MAX 2550 liquid was analyzed by a spectrophotometer (UV-7220, Beifenruili,
diffractometer (Cu Kα radiation, λ = 1.5406 Å), operated at 40 kV and China) at 662 nm. The degradation ratio of MB can be calculated by
100 mA. The crystalline size was estimated by applying the Scherrer (A0 − A) / A0 × 100%.
equation. N2 absorption isotherms were collected on an AUTOSORB-1
nitrogen adsorption apparatus at − 196 °C and all samples were 3. Results and discussion
degassed at 120 °C for 2 h. The instruments employed for XPS analysis
was a PHI 5300 ESCA with an Mg K X-ray source at a power of 250 W. 3.1. XRD and BET analysis
The shift of binding energy due to relative surface charging was
corrected using the C1s lever at 284.6 eV as an internal standard. The The XRD patterns of the as-synthesized pure TiO2, (0.03B)–TiO2, and
UV–vis absorbance spectra were obtained using a Shimadzu-2501 spec- (0.03B, 1.0N)–TiO2 samples are presented in Fig. 1. All the diffraction
trophotometer. BaSO4 was the reference sample and the spectra were peaks can be ascribed to pure anatase phase TiO2 after calcination.
recorded at room temperature in air within the range 200–900 nm. These test results indicate that there has been virtually no phase transi-
TEM images were recorded on a JEM-2100F made in Japan. tion in TiO2 in the process of B-doping or B–N co-doping. The diffraction

140 0.025
a b
Adsorbed volumn (cm3/g,STP)

120
0.020
dV/dD (cm3/g-nm)

100
0.015
80

60 0.010

40 0.005

20
0.000
0
0.0 0.2 0.4 0.6 0.8 1.0 0 10 20 30 40 50 60 70 80
Relative pressure (p/p0) Pore diameter(nm)

Fig. 2. N2 adsorption-desorption isotherms (a) and the BJH corresponding pore size distribution curve (b) of (0.03B, 1.0N)-TiO2.
610 K. Zhang et al. / Powder Technology 253 (2014) 608–613

Table 1 sample were determined to be 125.4 m2/g and 0.1776 cm3/g. Other
The characterization results of different samples. samples have the similar nitrogen adsorption–desorption isotherms
Sample Crystallite sizea SBETb Pore sizec Pore volumed and BJH pore size distributions to the (0.03B, 1.0N)-TiO2 sample. Their
(nm) (m2/g) (nm) (cm3/g) textural properties are listed in Table 1. Considering the crystallite size
TiO2 12.9 121.6 16.9 0.1754 of the sample, the mesopores are believed to be formed by the agglom-
(0.03B)-TiO2 9.5 130.6 15.6 0.1804 eration and connection of adjacent nanoparticles in the sample [34].
(0.03B, 1.0N)-TiO2 9.1 125.4 16.8 0.1776 This network nanostructure offers more efficient transportation for
a
Evaluated by the Scherrer's equation. the reactant molecules to the active sites, which are expected to en-
b
Measured by the BET method. hance the photocatalytic activity [26].
c
Estimated from the Barrett-Joyner-Halenda (BJH) formula.
d
Estimated by the cryolac number of N2 adsorption at relative pressure of 0.99.
3.2. TEM analysis

peak at 25.4°, corresponding to characteristic peak of crystal plane (101) The microstructure of the (0.03B, 1.0N)-TiO2 sample was investigat-
of anatase phase TiO2, became broader and the relative intensity de- ed with transmission electron microscopy (TEM) and high-resolution
creased after boron and nitrogen are introduced into TiO2, suggesting transmission electron microscopy (HRTEM). TEM images of the sample
the decrease of the crystal size. According to Scherrer equation, the av- are shown in Fig. 3(a) and (b). As estimated from the TEM images,
erage crystallite sizes of pure TiO2, (0.03B)–TiO2, and (0.03B, 1.0N)–TiO2 grains are all round-shaped with a uniform size of about 10 nm,
samples are estimated to be about 10.9, 9.5, and 9.1 nm, respectively, which are in agreement with the XRD analysis. In addition, it is clearly
corresponding to (a), (b), and (c) in Fig. 1. It confirms that the boron seen that the sample has a typical mesopore structure with a
and nitrogen doping can decrease the crystal size mainly due to the de- wormhole-like framework. As observed in Fig. 3(b), the pore size is
formation of lattice and oxygen vacancies left by the substitution of O about 15–20 nm, which is consistent with N2 sorption analysis. The
atoms for B or N atoms [19]. The size decreases from (b) to (c) maybe Debye rings showed in Fig. 3(d) suggest a sequence of diffraction rings
aroused by the new structure Ti–N–B–O formed at the internal surface consistent with what is expected for anatase TiO2 as the nanocrystal of
of TiO2 [21]. the mesoporous sample. The average inter planar space is determined
Fig. 2 shows the nitrogen adsorption–desorption isotherms and pore to be 0.35 nm (Fig. 3(c)), which is assigned to the anatase TiO2 facet
size distribution curves of the (0.03B, 1.0N)-TiO2 sample. As shown in (101) from the corresponding wide-angle XRD pattern [35,36].
Fig. 2(a), the N2 sorption isotherm of the sample is consistent with
type IV of hysteresis loops, which is a typical mesoporous structure of 3.3. UV–vis DRS of catalysts
the material according to the IUPAC classification. There is a hysteresis
loop appearing at the relatively low pressure (0.4 b P/P0 b 0.8), indicat- The UV–vis diffraction reflectance spectra of (mB)-TiO2 samples are
ing a small range of pore size distribution, which is centered at 16.8 nm shown in Fig. 4. It is found that the amount of boron doped into TiO2
estimated by the Barrett–Joyner–Halenda (BJH) approach (Fig. 2(b)). affects the optical absorption of the sample both in the UV and visible
The mesopores allow light to scatter inside their pore channel, thus en- regions. The (0.1B)-TiO2 sample exhibits the strong absorption in visible
hancing the harvesting of light [33]. According to the liner part of the region, and the (0.03B)-TiO2 exhibits the strongest response to light in
sorption isotherm, the specific surface area and pore volume of the wavelength N 550 nm. Furthermore, it can be clearly observed that the

Fig. 3. TEM characterization: (a, b) TEM images, (c) HRTEM images, (d) SAED pattern.
K. Zhang et al. / Powder Technology 253 (2014) 608–613 611

and C1s are 459.0, 531.0, 400.5, 192.1, and 284.6 eV, respectively,
(0.1B)-TiO2
which are approximately in agreement with the previous data by
(0.03B)-TiO2 other researchers [20–23]. The C1s peak (284.6 eV) is usually associated
(0.05B)-TiO2 with the residual carbon from precursor solution and the adventitious
Absorbance (a.u.)

hydrocarbon from XPS instrument itself. The relative atomic concentra-


(0.02B)-TiO2
tions of boron and nitrogen in the sample, estimated from the XPS data,
were determined to be about 1.23 and 1.78 atom%, respectively.
Fig. 6(b) shows the high-resolution XPS spectra for B1s. Four XPS
peaks at 187.3, 189.2, 192.1, and 193.7 eV were observed. According
to the previous literature, the peak at 193.7 eV is attributed to the
bond of B–O [37]. The peaks at 187.3 and 192.1 eV are ascribed to the
structures of Ti–B and Ti–O–B [20,38], respectively. The peak at
189.2 eV is attributed to the bond of B\N, which is supposed to exist
in forms of Ti–B–N or Ti–N–B [19]. To make sure of the structural
300 400 500 600 700 800 forms of B and N existing in the sample, the N1s XPS spectra of the sam-
Wavelength (nm) ple are measured and the high-resolution XPS spectra of N1s are shown
in Fig. 6(c). As seen from Fig. 6(c), three peaks at 399.1, 400.5, and
Fig. 4. UV–vis absorption spectra of (mB)-TiO2 samples.
401.5 eV appeared on the spectra. The peak at 400.5 eV is attributed
to the structure of Ti–O–N [20]. The peak at 401.5 eV may be attributed
absorption edge, which is decided by the intercept on the wavelength to the bond of Ti–N–O [25,39]. The peak at 399.1 eV indicates a structure
axis for a tangent drawn on absorption spectra, of all samples doped of Ti–N–B [21] rather than Ti–B–N, in which the binding energies of B1s
with boron extend into visible light from about 420 to 480 nm. Taking and N1s are 190.5 and 398.1 eV, respectively [19]. Taking the spectra of
into account of the absorption intensities in UV and visible regions, B1s and N1s into account in this work, we may conclude that the dop-
the amount of boron doped of the (0.03B)-TiO2 sample is appropriate. ants of boron and nitrogen form a structure of Ti–N–B–O in the sample.
The relative atomic concentration of boron in the sample is about 1.23 The high-resolution XPS spectrum of Ti2p, shown as Fig. 6(d), can
atom % determined by the XPS data. This relative atomic concentration further verify the existence of this Ti–N–B–O structure. As seen from
of boron is approximately the maximized amount of boron doped Fig. 6(d), the characterization peak shifts from 458.2 eV, which is the
(about 1.13 atom %), which was demonstrated to be the very significant normal binding energy usually for pure TiO2, to 459 eV, the binding en-
activity for B doped sample [20]. ergy for TiO2 co-doped with boron and nitrogen [19–21]. This change of
Fig. 5 shows a description about the UV–vis diffraction reflectance the binding energy should be attributed to the electronegativity of O in
spectra of pure TiO2, (0.03B)-TiO2, and (0.03B, 1.0N)-TiO2. Compared Ti–N–B–O structure, which is a good agreement with the results of other
with the pure TiO2, the absorption of B-doped sample extends some- literature [21].
what into the visible region. Further, it is clear that the absorption
edge of B-N co-doped sample exhibits an evident red-shift, indicating 3.5. Photocatalytic activity
a possible synergistic effect of the two dopants. The plots of transformed
Kubelka-Munk function versus the energy of absorbed light give the The UV-light and visible light photocatalytic activities of pure meso-
band gap of the samples [25]. The band gap energies are estimated to porous TiO2, (0.03B)-mesoporous TiO2, and (0.03B, 1.0N)-mesoporous
be 3.18, 3.02, and 2.86 eV for pure TiO2, (0.03B)-TiO2, and (0.03B, TiO2 were evaluated by photocatalytic degradation of MB aqueous solu-
1.0N)-TiO2, respectively. The narrowed band gap arises from the contri- tion (shown in Fig. 7). Fig. 7(a) shows a comparison of photocatalytic
butions of B–N co-doping atoms, which formed the Ti–N–B–O structure activity of the samples under UV-light irradiation. As seen from the fig-
demonstrated by the XPS tests. The results confirm that both dopants ure, although all the samples exhibit high photocatalytic activity under
will make the band gap of TiO2 narrow and therefore improve the visi- UV irradiation, it has somewhat difference for the samples. It can be
ble light photocatalytic activity. seen that the doped TiO2 exhibit a higher photocatalytic activity, espe-
cially for (0.03B, 1.0N)-TiO2, which may be because the doped samples
3.4. XPS analysis have higher absorption intensity than pure TiO2, as shown in Fig. 5(a).
The photocatalytic activity of samples under visible light was shown
The XPS spectra of (0.03B, 1.0N)-TiO2 sample is shown in Fig. 6. The in Fig. 7(b). Compared with pure mesoporous TiO2, doped mesoporous
obvious peaks of titanium, oxygen, nitrogen, boron, and carbon can be TiO2 photocatalysts showed obviously enhancing effect on the photo-
detected in Fig. 6(a), and the binding energies of Ti2p, O1s, N1s, B1s, catalytic activity, especially for the B–N co-doped mesoporous TiO2. As

(0.03B, 1.0N)-TiO2 a (0.03B, 1.0N)-TiO2


b
(0.03B)-TiO2 (0.03B)-TiO2
TiO 2 TiO2
Absorbance (a.u.)

(αEphoto)2 (a.u.)

200 300 400 500 600 700 800 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0
Wavelength (nm) Ephoto (eV)

Fig. 5. UV–vis absorption spectra of pure TiO2, (0.03B)-TiO2, and (0.03B, 1.0N)-TiO2.
612 K. Zhang et al. / Powder Technology 253 (2014) 608–613

a O1s b 192.1eV
B1s
Ti2p

Intensity (a.u.)

Intensity (a.u.)
C1s 193.7eV 187.3eV
N1s B1s 189.2eV

600 500 400 300 200 100 195 194 193 192 191 190 189 188 187 186
Binding energy (eV) Binding energy (eV)

c 400.5eV d 459.0eV
N1s Ti2p
Intensity (a.u.)

Intensity (a.u.)
401.5eV
399.1eV

398 399 400 401 402 403 450 455 460 465 470 475
Binding energy (eV) Binding energy(eV)

Fig. 6. (a) XPS spectra of (0.03B, 1.0N)-TiO2, (b) high-resolution XPS spectra of B1s, (c) high-resolution XPS spectra of N1s, (d) high-resolution XPS spectra of Ti2p.

we can see from the data shown in Fig. 7(b), (0.03B)-doped and (0.03B, produce a lot of electrons and holes, which will be recombined by the
1.0N) co-doped TiO2 photocatalysts can improve the degradation rate of excess amount of B–N species on the surface of the sample [21,40]
MB under visible light. It is obvious that the photocatalytic activity of B– and therefore affect the photocatalytic activity.
N co-doped TiO2 is better than that of B-doped TiO2, which is largely due
to the outstanding synergistic effect of boron and nitrogen co-doping by 4. Conclusions
forming the structure of Ti–N–B–O [21]. The result is in great agreement
with the UV–vis DRS analysis. In summary, mesoporous TiO2 photocatalyts co-doped with boron
Compared with Fig. 7(a) and Fig. 7(b), it is evident that the photocat- and nitrogen were successfully prepared by using the fast sol–gel
alytic activities of all samples in UV region are higher than in visible method. The B–N co-doped catalyst has a specific surface area of
region. This is because the absorption in UV region is stronger than in 125.4 m2/g and the average pore size is decided to be 16.8 nm, which
the visible region, which has been shown in Fig. 5. It is also found that implies a typical mesopore structure. The XPS analysis shows that a
the enhancement of photocatalytic activity by boron and nitrogen in structure of Ti–N–B–O was formed on the surface of the co-doped sam-
UV region seems to be negligible compared with that in visible region. ple. The special structure greatly improves the photocatalytic activity
This may be ascribed to the structure of Ti–N–B–O formed on the sur- under visible light, suggesting a synergistic effect of boron and nitrogen.
face of the sample. To our best knowledge, the UV light irradiation will The absorption band edge of the B–N co-doped mesoporous TiO2

1.0 a TiO2
1.0 b
(0.03B, 1.0N)-TiO 2
0.8 (0.03B)-TiO2 0.8

0.6 0.6
C/C0

C/C0

0.4 0.4 (0.03B)-TiO2

(0.03B, 1.0N)-TiO2
0.2 0.2
TiO2

0.0 0.0
0 20 40 60 80 100 0 60 120 180 240
Time (min) Time (min)

Fig. 7. Photocatalytic degradation kinetics of MB solutions under irradiation of (a) UV light and (b) visible light.
K. Zhang et al. / Powder Technology 253 (2014) 608–613 613

sample exhibits an evident red-shift and its absorption intensity of vis- [17] Y. Wu, M. Xing, B. Tian, J. Zhang, F. Chen, Preparation of nitrogen and fluorine
co-doped mesoporous TiO2.microsphere and photodegradation of acid orange 7
ible region is higher than that of the undoped sample, which is respon- under visible light, Chem. Eng. J. 162 (2010) 710–717.
sible for B–N co-doping that make the energy gap of TiO2 narrower. [18] E.A. Reyes-Garcia, Y. Sun, D. Raftery, Solid-state characterization of the nuclear and
Compared with the un-doped sample, B–N co-doped mesoporous TiO2 electronic environments in a boron-fluoride co-doped TiO2 visible-light
photocatalyst, J. Phys. Chem. C 111 (2007) 17146–17154.
appears to have higher photocatalytic activity under irradiation of visi- [19] G. Liu, Y. Zhao, C. Sun, F. Li, G. Lu, H. Cheng, Synergistic effects of B/N doping on the
ble light. visible-light photocatalytic activity of mesoporous TiO2, Angew. Chem. Int. Ed. 47
(2008) 4510–4516.
[20] S. In, A. Orlov, R. Berg, F. García, S. Pedrosa-Jimenez, M.S. Tikhov, D.S. Wright, R.M.
Acknowledgment Lambert, Effective visible light-activated B-doped and B, N-codoped TiO2
photocatalysts, J. Am. Chem. Soc. 129 (2007) 13790–13791.
[21] M. Xing, W. Li, Y. Wu, J. Zhang, X. Gong, Formation of new structures and their
This work has been supported by the Science and Technology synergistic effects in boron and nitrogen codoped TiO2 for enhancement of photo-
Planning Project of Shaanxi Province (2013K09-04) and the Scientific catalytic performance, J. Phys. Chem. C 115 (2011) 7858–7865.
[22] N.O. Gopal, H.H. Lo, S.C. Ke, Chemical state and environment of boron dopant in B,
Research Project of the Provincial College Key Laboratory of Shaanxi N-codoped anatase TiO2 nanoparticles: an avenue for probing diamagnetic dopants
Province (2010JS007). in TiO2 by electron paramagnetic resonance spectroscopy, J. Am. Chem. Soc. 130
(2008) 2670–2671.
[23] X. Zhou, F. Peng, H. Wang, H. Yu, Boron and nitrogen-codoped TiO2 nanorods: syn-
References thesis, characterization, and photoelectrochemical properties, J. Solid State Chem.
184 (2011) 3002–3007.
[1] P. Docampo, S. Guldin, U. Steiner, H.J. Snaith, Charge transport limitations in [24] X. Ding, X. Song, P. Li, Z. Ai, L. Zhang, Efficient visible light driven photocatalytic re-
self-assembled TiO2 photoanodes for dye-sensitized solar cells, J. Phys. Chem. Lett. moval of NO with aerosol flow synthesized B, N-codoped TiO2 hollow spheres,
4 (2013) 698–703. J. Hazard. Mater. 190 (2011) 604–612.
[2] H. Nishikiori, Y. Uesugi, R.A. Setiawan, T. Fujii, W. Qian, M.A. El-Sayed, Photoelectric [25] D. Wang, L. Jia, X. Wu, L. Lu, A. Xu, One-step hydrothermal synthesis of N-doped
conversion properties of dye-sensitized solar cells using dye-dispersing titania, TiO2/C nanocomposites with high visible light photocatalytic activity, Nanoscale 4
J. Phys. Chem. C 116 (2012) 4848–4854. (2012) 576–584.
[3] D. Wang, X. Li, J. Chen, X. Tao, Enhanced visible-light photoelectrocatalytic degrada- [26] D.M. Antonelli, J.Y. Ying, Synthesis of hexagonally packed mesoporous TiO2 by a
tion of organic contaminants at iodine-doped titanium dioxide film electrode, Ind. modified sol–gel method, Angew. Chem. Int. Ed. 34 (1995) 2014–2017.
Eng. Chem. Res. 51 (2012) 218–224. [27] L. Chen, B. Yao, Y. Cao, K. Fan, Synthesis of well-ordered mesoporous titania with
[4] J. He, G.K. Reddy, S.W. Thiel, P.G. Smirniotis, N.G. Pinto, Ceria-modified manganese tunable phase content and high photoactivity, J. Phys. Chem. C 111 (2007)
oxide/titania materials for removal of elemental and oxidized mercury from flue 11849–11853.
gas, J. Phys. Chem. C 115 (2011) 24300–24309. [28] W. Shao, F. Gu, C. Li, M. Lu, Interfacial confined formation of mesoporous spherical
[5] P. Wang, Y. Ao, C. Wang, J. Hou, J. Qian, Enhanced photoelectrocatalytic activity for TiO2 nanostructures with improved photoelectric conversion efficiency, Inorg.
dye degradation by graphene–titania composite film electrodes, J. Hazard. Mater. Chem. 49 (2010) 5453–5459.
223–224 (2012) 79–83. [29] T. Coquil, C. Reitz, T. Brezesinski, E.J. Nemanick, S.H. Tolbert, L. Pilon, Thermal con-
[6] C.G. Silva, R. Juárez, T. Marino, R. Molinari, H. García, Influence of excitation wave- ductivity of ordered mesoporous titania films made from nanocrystalline building
length (UV or visible light) on the photocatalytic activity of titania containing gold blocks and sol–gel reagents, J. Phys. Chem. C 114 (2010) 12451–12458.
nanoparticles for the generation of hydrogen or oxygen from water, J. Am. Chem. [30] D. Lee, S. Park, S. Ihm, K. Lee, Synthesis of bimodal mesoporous titania with high
Soc. 133 (2011) 595–602. thermal stability via replication of citric acid-templated mesoporous silica, Chem.
[7] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Visible-light photocatalysis in Mater. 19 (2007) 937–941.
nitrogen-doped titanium oxides, Science 293 (2001) 269–271. [31] Y. Wu, M. Xing, J. Zhang, Gel-hydrothermal synthesis of carbon and boron co-doped
[8] N. Lu, H.M. Zhao, J.Y. Li, X. Quan, S. Chen, Characterization of boron-doped TiO2 TiO2 and evaluating its photocatalytic activity, J. Hazard. Mater. 192 (2011)
nanotube arrays prepared by electrochemical method and its visible light activity, 368–373.
Sep. Purif. Technol. 62 (2008) 668–673. [32] E. Liu, X. Guo, L. Qin, G. Shen, X. Wang, Fabrication and photocatalytic activity
[9] P. Xu, J. Lu, T. Xu, S. Gao, B. Huang, Y. Dai, I2-hydrosol-seeded growth of (I2)n-C-codoped of highly crystalline nitrogen doped mesoporous TiO2, Chin. J. Catal. 33 (2012)
meso/nanoporous TiO2 for visible-light-driven photocatalysis, J. Phys. Chem. C 114 1665–1671.
(2010) 9510–9517. [33] D. Huang, S. Liao, J. Liu, Z. Dang, L. Petrik, Preparation of visible-light responsive N–
[10] Q. Xiang, J. Yu, W. Wang, M. Jaroniec, Nitrogen self-doped nanosized TiO2 nano- F-codoped TiO2 photocatalyst by a sol–gel-solvothermal method, J. Photochem.
sheets with exposed 001 facets for enhanced visible-light photocatalytic activity, Photobiol. A Chem. 184 (2006) 282–288.
Chem. Commun. 47 (2011) 6906–6908. [34] G. Tian, H. Fu, L. Jing, B. Xin, K. Pan, Preparation and characterization of stable
[11] M.R. Bayati, A.Z. Moshfegh, F. Golestani-Fard, On the photocatalytic activity of the biphase TiO2 photocatalyst with high crystallinity, large surface area, and enhanced
sulfur doped titania nano-porous films derived via micro-arc oxidation, Appl. photoactivity, J. Phys. Chem. C 112 (2008) 3083–3089.
Catal. A Gen. 389 (2010) 60–67. [35] G. Liu, H. Yang, X. Wang, L. Cheng, J. Pan, G. Lu, H. Cheng, Visible light responsive
[12] S. Tosoni, O. Lamiel-Garcia, D. Fernandez Hevia, J.M. Doña, F. Illas, Electronic struc- nitrogen doped anatase TiO2 sheets with dominant 001 facets derived from TiN,
ture of F-doped bulk rutile, anatase, and brookite polymorphs of TiO2, J. Phys. J. Am. Chem. Soc. 131 (2009) 12868–12869.
Chem. C 116 (2012) 12738–12746. [36] X. Xue, Y. Wang, H. Yang, Preparation and characterization of boron-doped titania
[13] J.F. Ju, X. Chen, Y.J. Shi, J.W. Miao, D.H. Wu, Hydrothermal preparation and photocat- nano-materials with antibacterial activity, Appl. Surf. Sci. 264 (2013) 94–99.
alytic performance of N, S-doped nanometer TiO2 under sunshine irradiation, [37] D.M. Chen, D. Yang, Q. Wang, Z.Y. Jiang, Effects of boron doping on photocatalytic ac-
Powder Technol. 237 (2013) 616–622. tivity and microstructure of titanium dioxide nanoparticles, Ind. Eng. Chem. Res. 45
[14] P. Zhou, J. Yu, Y. Wang, The new understanding on photocatalytic mechanism of (2006) 4110–4116.
visible-light response N–S codoped anatase TiO2 by first-principles, Appl. Catal. B [38] Q. Ling, J. Sun, Q. Zhou, Preparation and characterization of visible-light-driven tita-
Environ. 142–143 (2013) 45–53. nia photocatalyst co-doped with boron and nitrogen, Appl. Surf. Sci. 254 (2008)
[15] Q. Xiang, J. Yu, M. Jaroniec, Nitrogen and sulfur co-doped TiO2 nanosheets with ex- 3236–3241.
posed 001 facets: synthesis, characterization and visible-light photocatalytic activi- [39] H. Irie, Y. Watanabe, K. Hashimoto, Nitrogen-concentration dependence on photo-
ty, Phys. Chem. Chem. Phys. 13 (2011) 4853–4861. catalytic activity of TiO2−xNx powders, J. Phys. Chem. B 107 (2003) 5483–5486.
[16] J. Xu, B. Yang, M. Wu, Z. Fu, Y. Lv, Y. Zhao, Novel C–F-codoped TiO2 inverse opal with [40] G. Liu, C. Sun, L. Cheng, Y. Jin, H. Lu, L. Wang, S.C. Smith, G. Lu, H. Cheng, Efficient
a hierarchical meso-/macroporous structure: Synthesis, characterization, and promotion of anatase TiO2 photocatalysis via bifunctional surface-terminating Ti–
photocatalysis, J. Phys. Chem. C 114 (2010) 15251–15259. O–B–N structures, J. Phys. Chem. C 113 (2009) 12317–12324.

You might also like