You are on page 1of 7

Journal of Alloys and Compounds 548 (2013) 70–76

Contents lists available at SciVerse ScienceDirect

Journal of Alloys and Compounds


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

Characterization and photocatalytic properties of N-doped BiVO4 synthesized


via a sol–gel method
Min Wang ⇑, Qiong Liu, Yinsheng Che, Lifang Zhang, Dong Zhang
College of Environmental and Chemical Engineering, Shenyang Ligong University, Shenyang 110165, China

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

Article history: A N-doped BiVO4 photocatalyst with high visible light activity was synthesized by the complexing sol–gel
Received 23 July 2012 method using citric acid as a chelate and hexamethylene tetramine (C6H12N4) as a nitrogen source. The
Received in revised form 20 August 2012 as-prepared N-doped BiVO4 samples were characterized by X-ray photoelectron spectroscopy (XPS),
Accepted 22 August 2012
X-ray diffraction (XRD), scanning electron microscopy (SEM), specific surface area (BET) and UV–Vis dif-
Available online 10 September 2012
fuse reflectance spectroscopy (DRS). The photocatalytic activity was evaluated by photocatalytic degra-
dation of methyl orange (MO) solution under visible light. This technique revealed that pure BiVO4
Keywords:
and all the N-doped samples were in a monoclinic phase; no peaks of any other phases or impurities were
BiVO4
Photocatalytic activity
detected. Nitrogen atoms were doped into the BiVO4 lattice and filled the atomic sites of oxygen to form
N-doped O–Bi–N–V–O bonds, which contributed to the appearance of the more active species V4+ and oxygen
Methyl orange solution vacancies. The doped nitrogen resulted in a red shift in the absorption edge. However, the N-doping only
slightly changed the morphologies and BET special surface areas of the samples. The photocatalytic activ-
ity of BiVO4 significantly depended on the N-doping content and the calcination temperature. The max-
imum activity was observed for the catalyst obtained via calcination at 500 °C, for which the molar ratio
of N to Bi was 0.20. Excess N-doping decreased the light absorption.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction structure exhibits high photocatalytic activities due to its relatively


narrow band gap energy (2.4 eV) compared to BiVO4 with a tetrag-
Currently, dye pollution is a major source of environmental onal phase (3.1 eV) [18]. However, the photocatalytic activity of
pollution in waste fields that is difficult to eliminate with conven- the monoclinic scheelite BiVO4 is not as strong because of its poor
tional water treatment technologies [1–4]. In recent decades, a adsorptive performance and low quantum yield caused by the dif-
great deal of interest has been devoted to photocatalytic degrada- ficult migration of photo-generated electron–hole pairs.
tion of organic water pollutants by semiconductor particles [5–7]. To remove such obstacles to high photocatalytic activity, doping
Among these semiconductors, titanium oxide (TiO2) is the prefera- of nonmetals, metals and metal oxides is used to accelerate the sep-
ble material for its chemical stability, lack of toxicity and low cost. aration of electron–hole pairs, as well as to efficiently extend the
However, the TiO2 photocatalyst has not been applied widely in usable region of visible light. Using the impregnation technique,
the field of environmental pollution control, given that its large Long et al. [19] successfully synthesized Co3O4/BiVO4 composite
band-gap energy (Eg = 3.2 eV) considerably limits the use of natu- photocatalysts that exhibited high photocatalytic activity on the
ral solar light or artificial visible light. Therefore, the development degradation of phenol under visible-light irradiation. Zhang et al.
of high activity photocatalysts driven by visible light is now of [15] investigated the silicon-doped BiVO4 film via the modified
great importance, especially in view of serious energy shortages. metalorganic decomposition (MOD) method, for which the phenol
Recently, a large number of novel semiconductor photocatalysts elimination rate on the Si-doped BiVO4 film electrode in the photo-
with visible light response have been developed. Bismuth vanadate electrocatalytic process was 1.84 times greater than that on the
(BiVO4), one of the visible-light-driven semiconductor photocata- BiVO4 film electrode. Suo et al. [20] found that the Cu-doped BiVO4
lysts [8–14], has recently attracted considerable attention for its catalyst exhibited improved photocatalytic activity for the degrada-
photocatalytic activity in the degradation of organic compounds tion of toluene in the gas phase, which reached over 90% and
[15] and organic dyes [16] and in water splitting for hydrogen extended a red-shift. Chatchai et al. [21] studied the photoelectro-
and oxygen evolution [17]. The photocatalytic properties strongly chemical properties of a FTO/WO3/BiVO4 composite electrode con-
depend on the crystal structure. BiVO4 with a monoclinic scheelite sisting of monoclinic WO3 and crystalline monoclinic BiVO4. WO3/
BiVO4 was observed to exhibit elevated photocurrent efficiencies,
⇑ Corresponding author. Tel./fax: +86 24 24680345. reaching values ten times those of a bare BiVO4 electrode under
E-mail address: minwang62@msn.com (M. Wang). both UV and visible light irradiation. Ag–BiVO4 composites were

0925-8388/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jallcom.2012.08.140
M. Wang et al. / Journal of Alloys and Compounds 548 (2013) 70–76 71

prepared by Zhang et al. [22] through a photoreduction technique; intervals, samples were collected from the mixtures, immediately centrifuged and
then filtered through a 0.45 lm Millipore filter to remove catalyst particles. The
these composites showed visible-light activities superior to those of
filtrate concentration was determined by recording the absorbance at 464 nm using
pure BiVO4 in the decomposition of phenol. Yu et al. [23] fabricated a UV-1800 UV–Vis spectrophotometer (Puxi, China). The reproducibility of the
Co3O4 and CuO/BiVO4 by intense ultrasound irradiation at room results, which was determined by repeating the measurements at least three times,
temperature. The composite photocatalysts exhibited much higher was within an acceptable limit (±5%).
photocatalytic activity than that of pure BiVO4 due to the high crys-
tallinity and the formed p–n heterojunction of Co3O4/BiVO4 or CuO/ 3. Results and discussion
BiVO4. These two factors can effectively suppress the recombina-
tion of photogenerated hole–electron pairs. 3.1. XPS analysis
Although there have been many reports on nonmetal-, metal-
and noble metal-doping in BiVO4, the effect of N doping on the pho- The overall XPS spectra of BVO and BVON-20 samples are
tocatalytic activity of BiVO4 for the photodegradation of organic shown in Fig. 1. Seven peaks corresponding Bi 5d, Bi 4f, Bi 4d5/2,
contaminants has seldom been reported, to the best of our knowl- Bi 4d3/2, V 2p, O 1s, and C 1s appeared for both BVO and BVON-
edge. Herein, we synthesized pure BiVO4 and N-doped BiVO4 phot- 20 powders. The N 1s peak at approximately 399.8 eV can be
ocatalysts through the sol–gel method and characterized them by detected obscurely for the BVON-20 compared with the pure
XPS, XRD, SEM, BET and DRS. The effects of N-doping concentration BiVO4, which was due to the presence of nitrogen from the sol–
and calcination temperature on the photocatalytic degradation of gel precursor. The C 1s peak at 285 eV most likely originated from
methyl orange were investigated under visible-light irradiation. the calcinated residue of the organic materials used and is used for
Based on the results of photocatalyst characterization and activity calibration. The total nitrogen concentration calculated from the
testing, the mechanism of the enhanced activity of N-doped BiVO4 XPS spectrum was 2.87%. It indicated that the actual N-doping con-
composite catalysts was also discussed. centration in sample was only about 1/10 of the original doping
ratio. But for convenience, the N-doping concentration still used
2. Experimental the theoretical doping amount in text.
To investigate the doped nitrogen chemical states in the photo-
2.1. Photocatalyst preparation
catalyst, the N 1s high resolution XPS spectrum was analyzed, as
Bi(NO3)35H2O and NH4VO3 (Shenyang Chemical Company) were used as shown in Fig. 2(a). As shown in Fig. 2(a), the N 1s peak was found
received without further purification. All other chemicals used in experimentation at 399.8 eV only; no other peaks appeared. The binding energy at
were of analytical grade. 399.8 eV can be attributed to nitrogen replacing the common oxy-
In a typical preparation process, 0.01 mol Bi(NO3)35H2O was first dissolved in
gen in the crystal lattice of BiVO4 to form the O–Bi–N–V–O bond
50 mL of 10% (w/w) HNO3. The appropriate amount of citric acid was added until
the molar ratio of citric acid and Bi(NO3)3 was 2:1. Solution A was thereby obtained. [24]. This binding energy could not represent O–Bi–N or O–V–N
NH4VO3 was dissolved in 80 °C distilled water, and the solution was magnetically alone because both the binding energies of Bi 4f and V 2p, presented
stirred until the dissolution was complete. The appropriate amount of citric acid in Fig. 2(b) and (c), changed to some degree. Therefore, the common
was then added until the molar ratio of citric acid and NH4VO3 was 2:1. The mixture
O atom in the O–Bi–O–V–O bond was most likely replaced by a N
was completely stirred to dissolve the solid, generating solution B. Solution A was
then added drop-wise into solution B with continual magnetic stirring until the atom. Furthermore, because the high electronegativity of oxygen
two solutions were mixed at a 1:1 M ratio to obtain solution C. A given amount leads to the reduction of electron density on the nitrogen, a rela-
of C6H12N4 (C6H12N4 in strong acid solution is hydrolyzed when it is heated to more tively higher binding energy at 399.8 eV was observed compared
than 100 °C) was then added to solution C. Under vigorous stirring, the pH of the to V–N in VN (the N 1s binding energy is 397.4 eV [25]) or Bi–N
mixture was adjusted to approximately 6.5 using ammonia solution. The mixture
in BiN (the N 1s binding energy is 397.4 eV [25]). So, it can be con-
was stirred at 80 °C until the dark blue BiVO4 sol–gel was obtained and then dried
at 80 °C in a drying cabinet for 10 h to produce the BiVO4 precursor. The resulting cluded that the doping N only replaced the O in the crystal lattice of
powder was collected and calcined in air at different temperatures (between 350 BiVO4, but did not form the VN and (or) BiN species in the N-doped
and 500 °C) for 5 h, cooled to room temperature, and then crushed to obtain a fine BiVO4 sample; this interpretation is further supported by the XRD
powder. The sample was labeled as BVON-x, where x represents the initial molar
results.
ratio of N to Bi (between 0.13 and 0.30). Pure BiVO4 was prepared by the same
method as N-BiVO4, except that the ‘solution C’ corresponding to pure BiVO4 did Fig. 2(b) shows the XPS spectra for the Bi 4f region of BVO and
not contain C6H12N4. BVON-20. By fitting the curves, two strong symmetrical peaks were

2.2. Photocatalyst characterization


O1s
XPS spectra were measured on a MULTILAB2000 X-ray photoelectron spectro-
scope equipped with Mg Ka excitation. Detailed scans were recorded for N 1s, V
2p, Bi 4f and O 1s. The C 1s peak, set at 285.0 eV, was used as an internal reference
for absolute binding energy. XRD patterns were measured on a Scintag-XDS-2000 Bi4f V2p
diffractometer with Cu Ka radiation (k = 0.15418 nm). The morphologies and struc- Bi4d3/2
Intensity/a.u.

tures of the prepared samples were observed using scanning electron microscopy
Bi4d5/2 a
(SEM), Hitachi S-3000 N. UV–Vis DRS spectra were obtained on a Hitachi U-3010
spectrophotometer equipped with an integrating sphere. Absorption spectra were
referenced to BaSO4. The BET specific surface area was measured using ASAPZOLOM C1s
Micrometrics, with nitrogen as the adsorbent, at 196 °C. Prior to measurement, all
samples were degassed at 130 °C for 5 h. b

2.3. Photocatalytic activity test


Bi5d
The photocatalytic activities of the samples were evaluated through the degra- N1s
dation of MO under visible light. A 250 W halogen lamp was used as the visible light
source and was placed 16 cm from the reactor. A cut-off filter was placed under the
lamp to exclude all wavelengths lower than 400 nm to ensure irradiation with only
visible light (For visible light, k > 400 nm.). The experiments were performed as fol-
0 200 400 600 800 1000
lows. Equal amounts (10 mg) of photocatalyst and 0.1 mL H2O2 (10%) were added to
50 mL of 10 mg/L MO solution. The solution was stirred for 20 min in the dark prior Binding Energy/eV
to illumination to attain the adsorption–desorption equilibrium between the dye
and the catalyst surface under air-equilibrated room conditions. At given time Fig. 1. XPS spectra of (a) BVO and (b) BVON-20 calcined at 500 °C.
72 M. Wang et al. / Journal of Alloys and Compounds 548 (2013) 70–76

a:N1s 399.8eV Bi4f7/2 159.7eV


b:Bi4f
Bi4f5/2 165.0eV

Intensity/a.u.
Intensity/a.u.

Bi4f7/2 159.1eV
Bi4f5/2 164.4eV

BVON-20

BVO

385 390 395 400 405 410 154 156 158 160 162 164 166 168 170
Binding Energy/eV Binding Energy/eV

c1:V2p c2:V2p

516.8eV
516.6eV
Intensity / a.u.

Intensity/a.u.
516.1eV 515.8eV

512 514 516 518 520 513 514 515 516 517 518 519
Binding Energy / eV Binding Energy/eV

d2:O1s
d1:O1s 529.6eV

529.1eV
Intensity/a.u.

529.9eV
Intensity / a.u.

531.4eV

526 528 530 532 534 536 526 528 530 532 534 536
Binding Energy / eV Binding Energy/eV

Fig. 2. N 1s (a), Bi 4f (b), V 2p (c) and O 1s (d) XPS spectra of BVO and BVONF-6 samples. (1) BVO and (2) BVONF-6.

obtained at 159.1 and 164.4 eV, corresponding to the Bi 4f7/2 and Bi From Fig. 2(c), the asymmetric V 2p3/2 signals were decomposed
4f5/2 signals of BVO; peaks at 159.7 and 165.0 eV corresponded to into two peaks at Eb = 515.1 and 516.8 eV for BVO and Eb = 515.8
the Bi 4f7/2 and Bi 4f5/2 signals of BVON-20, respectively, which and 516.6 eV for BVON-20, attributable to the surface V4+ and V5+
were characteristic of the Bi3+ species [26]. We can conclude that species, respectively [28,29]. This result indicated that the banding
N-doping had no effect on the chemical state of the bismuth ion energy of V changed after N-doping. In addition, the molar ratio of
in BiVO4. However, the banding energy of Bi 4f increased by surface V4+/V5+ of the N-doped BiVO4 sample was higher than that
0.6 eV when compared with the pure sample. This result reveals of the undoped BiVO4 sample. This result is shown in Table 1.
that doping N can lead to a decrease of electron density on Bi According to the electroneutrality principle, the BiVO4 samples
due to the lower electronegativity of nitrogen compared to that were oxygen-deficient, and the amount of nonstoichiometric
of oxygen [27]. oxygen at the surface was dependent on the molar ratio of surface
M. Wang et al. / Journal of Alloys and Compounds 548 (2013) 70–76 73

Table 1
Analysis of O 1s and V 2p of BVO and BVON-20 samples.

Samples Binding energy (eV) Relative content Binding energy (eV) Relative content
5+ 4+
Lattice oxygen Hydroxyl oxygen Lattice oxygen Hydroxyl oxygen V V V5+ V4+
BVO 529.6 531.4 0.88 0.12 516.8 516.1 0.54 0.46
BVON-20 529.1 529.9 0.62 0.38 516.6 515.8 0.50 0.50

V4+/V5+. N3 substituting for O2 must cause partial electron trans- of the pure BiVO4 and N-doped BiVO4 series photocatalysts. The dif-
formation from the N3 to another cation (Bi3+ or V5+) due to the fraction peaks of all samples conform to the reference data for
structure of the O–Bi–N–V–O bond. From the former XPS analysis monoclinic BiVO4 (JCPDS cards No. 75-1866). Compared with that
of Bi 4f, there was no Bi2+ appearing after N-doping, confirming of the pure BiVO4, there were no impurity peaks arising in the
that there was no electron transfer to Bi3+; therefore, it is likely diffraction pattern of the N-doped BiVO4, which indicated that
that the electron transfer only occurred from V5+ to form more N-doping did not change the crystal type of BiVO4. Additionally,
V4+ sites in the crystal lattice. As for the O 1s XPS spectra no significant peaks characteristic of nitrogen oxides were detected.
(Fig. 2(d)), the asymmetric peak centered at approximately 530 This may be attributed to low concentrations of N species in these
eV decomposed into two components at Eb = 529.6 and 531.4 eV samples existing beyond the detection limit of the XRD technique.
for undoped BiVO4 and at Eb = 529.1 and 529.9 eV for BVON-20; The average crystallite sizes of these samples were estimated
these peaks were due to the surface lattice oxygen (Olatt) and the according to the line width analysis of the (1 2 1) diffraction peak
adsorbed oxygen (Oads) species, respectively [30]. This result indi- based on the Scherrer formula and were summarized in Table 2.
cates that the concentration of surface oxygen vacancies increases As shown in Table 2, the crystallite sizes of N-doped BiVO4 are
after N-doping. slightly bigger than those of pure BiVO4; it can therefore be pro-
The above results show that nitrogen is incorporated into the posed that N-doping may lead to a slight lattice distortion in the
lattice and is substituted for oxygen. As this process occurs, the structure of BiVO4. To further investigate the effect of N-doping
additional oxygen vacancies and V4+ sites would become the active on the crystal structure of BiVO4, the lattice parameters of all
sites for BiVO4, both of which are advantageous for the enhance- N-doped samples calculated using Bragg’s law (2dsinh = k) and the
 
ment of photocatalytic performance of BiVO4. 2
formula dl2 ¼ h aþk
2 2
þ cl 2 , which is applicable to tetragonal systems,
2

were listed in Table 2. These data reveal that the crystallite sizes and
3.2. XRD analysis
the lattice parameter increased with the amount of N-doping when
the initial molar ratio of N/Bi was less than 0.2, whereas the crystal-
XRD was carried out to investigate the effect of N-doping on the
lite sizes and crystal lattice volumes decreased with increasing
phase structure of BiVO4. Fig. 3 shows the XRD diffraction patterns
N-doping when the molar ratio of N/Bi was greater than 0.2.

e
Intensity/a.u.

e
Intensity / a.u.

d
d
c
c
b
b
a
a

0 20 40 60 80 0 10 20 30 40 50 60 70 80 90
2-Theta/degrees 2-Theta / degrees

Fig. 3. XRD patterns of BiVO4 products: (a) BVO; (b) BVON-13; (c) BVON-20; Fig. 4. XRD patterns of BVON-20 sample under different calcination temperatures:
(d) BVON-25 and (e) BVON-30. (a) 350, (b) 400, (c) 450, (d) 500 and (e) 550 °C.

Table 2
Some selected properties of BiVO4.

Sample Crystal size/nm Crystal lattice parameter Band gap/eV BET/m2 g1 Absorption rate%
a/nm b/nm c/nm V/nm3
BVO 49.52 5.19202 5.09148 11.69728 309.21 2.29 2.09 7.83
BVON-13 50.14 5.19284 5.09331 11.70429 309.56 2.27 2.91 10.06
BVON-20 49.84 5.19237 5.09447 11.69839 309.62 2.23 3.03 12.43
BVON-25 49.43 5.19298 5.09427 11.69986 309.51 2.25 3.15 12.41
BVON-30 45.90 5.18104 5.09178 11.69552 308.89 2.26 3.63 13.20
74 M. Wang et al. / Journal of Alloys and Compounds 548 (2013) 70–76

Based on the XPS analysis, we know the N ion is likely to be on Powder Diffraction Standards (JCPDS 75-1866) card for mono-
incorporated into the interstitial BiVO4 structure and to lead to an clinic scheelite BiVO4 with space group I2/b. No diffraction peak
increasing V4+/V5+ molar ratio, creating more V4+ sites in an of the VN and (or) BiN3 phases are observed, which may be attrib-
N-doped BiVO4 lattice than in that of undoped BiVO4. This means uted to the nitrogen concentration existing below the level of
that when the initial molar ratio of N/Bi was less than 0.2, the detection. According to the line width analysis of the (1 2 1) diffrac-
N-doped BiVO4 samples had larger crystallite sizes and crystal lat- tion peak based on the Scherrer formula, the average crystallite
tice volumes because the radius of V4+ (0.074 nm) is larger than that sizes of these samples are as follows: 36.62, 35.93, 42.30, 49.84,
of V5+ (0.059 nm), and the radius of N3 (0.171 nm) is larger than and 50.76 nm. Thus, it appears that the crystal sizes of the samples
that of O2 (0.140 nm). When N-doping content increased further, increase with the calcination temperature, implying that high tem-
however, the crystallite sizes and crystal lattice volumes of samples peratures favored grain growth in accordance with the theory of
decreased. This may be attributed to the probable separation of N thermodynamic nucleation and growth.
ions from the BiVO4 lattice as other nitrogen phases appeared.
Moreover, as these phases are generated within minutes, their 3.3. Scanning electron microscopy (SEM) and BET specific surface area
timescale may be below the detection limit of the XRD technique.
The XRD patterns of BVON-20 catalyst samples calcined at dif- Fig. 5 shows the morphologies of the BVO, BVON-13, BVON-20
ferent temperatures are shown in Fig. 4. The diffraction peaks of and BVO-25 photocatalysts. A spherical structure can be observed
all as-prepared samples fit well with the standard Joint Committee for all samples and the particles were uniform in size. There is little

Fig. 5. SEM micrographs of (a) and (b) for BVO; (c) and (d) for BVON-13; (e) and (f) for BVON-20; (e) and (f) for BVON-20; (g) and (h) for BVON-25.
M. Wang et al. / Journal of Alloys and Compounds 548 (2013) 70–76 75

difference in the morphologies and particle shapes of the samples, absorption to occur, significant photocatalytic activity can be
which means that N-doping has little effect on the morphology of expected from N-doped BiVO4 under visible light irradiation.
BiVO4. The BET specific surface areas of as-prepared samples were
measured by a nitrogen adsorption BET method (shown in Table 2). 3.5. Photocatalytic activity
All of the powders showed relatively small BET surface areas,
although the specific areas of N-doped samples were slightly high- The photocatalytic activity of N-doped BiVO4 was evaluated by
er than that of the pure BiVO4. The BET results of this work indicate measuring the decomposition of MO solution under visible light
that N-doping has little impact on increasing the specific surface irradiation. No degradation of MO was observed in the absence
area, which is consistent with the increased particle grain sizes of photocatalyst or illumination. To determine the effect of calcina-
confirmed by XRD results. tion temperature of the BiVO4 precursor on the photocatalytic
activity of N-doped BiVO4, the calcination temperature of BiVO4
was varied from 350 to 550 °C while holding the molar ratio of
3.4. UV–visible diffuse reflectance spectroscopy (DRS) analysis N/Bi at 0.2. Fig. 7 shows that the MO decomposition rates de-
pended on the calcination temperature of BiVO4. The photocata-
The UV–Vis diffuse reflectance spectra for the pure BiVO4 and lytic activity of N-doped BiVO4 increased when the calcination
N-doped BiVO4 samples are shown in Fig. 6. All the samples temperature was as low as 500 °C; however, the catalytic activity
showed a strong absorption in the UV- and visible-light regions; remarkably decreased when the calcination temperature reached
this visible light absorption was due to the bandgap transition. 550 °C. In the latter case, the high temperature heat treatment
N-doping caused a red shift in the BiVO4 adsorption edge. The may have resulted in grain growth, which can lower photocatalytic
sharp peaks of the spectra indicated that the visible light adsorp- activity. The catalyst calcined at 500 °C exhibits the highest photo-
tion is due to the band gap transition [31]; the prolonged absorp- catalytic activity because of its highly crystallized phase. In con-
tion tail extending to approximately 600 nm may be the result of trast, the calcination temperature above 500 °C promotes further
the oxygen vacancies formed and the increased presence of V4+ grain growth, potentially leading to lower photocatalytic activity.
sites that occurred as the nitrogen concentration increased [32]. Therefore, the optimum calcination temperature was found to be
The band gaps of the samples were calculated from the DRS; the 500 °C.
absorption onsets of the samples were measured by linear extrap- Fig. 8 shows the effect of N dopant concentration on the photo-
olation from the inflection point of the curve to the baseline, reach- catalytic activity of BiVO4 catalysts calcined at 500 °C for MO
ing a quantitative estimate of the band gap energies. The band gap decomposition. The degradation rate of MO on the pure BiVO4
energy (Eg, eV) values for the different samples were calculated under visible light irradiation was very low, approximately 38%
from the UV–Vis spectra using the equation Eg (eV) = 1240/k in 50 min. Obviously, the photocatalytic activity of N-doped BiVO4
(nm), where k represents the wavelength of the absorption onset is superior to that of pure BiVO4 for degradation of MO. The photo-
[33]. As shown in the spectra, the absorption onsets located at dif- catalytic activities of the photocatalysts increased with heightened
ferent wavelengths and their corresponding band gap energies are N-doping in cases where the initial molar ratio of N/Bi was less
summarized in Table 2. A reduction in the band gap energy for than 0.2; whereas the activities decreased at greater initial molar
N-doped samples occurred, compared to the undoped BiVO4. The ratios of N/Bi. The highest photocatalytic activity resulted in a
band gap energy values of N-doped BiVO4 photocatalyst decreased MO degradation rate reaching approximately 85% at the optimum
with increasing N-doping content when the initial N/Bi molar ratio N/Bi molar ratio of 0.2. The higher photocatalytic activity of nitro-
reached values below 0.2. On the other hand, the band gap energy gen-modified BiVO4 herein observed may be attributed to the fol-
values decreased when the molar ratio was above 0.2. The absorp- lowing reasons. On the one hand, the enhanced photocatalytic
tion of BVON-20 is the highest and corresponds to the lowest band activity of N-doped BiVO4 may be attributed to the increasing mo-
gap energy of 2.23 eV. In accordance with the XPS and XRD data, lar ratio of V4+/V5+ and the increasing amount of oxygen vacancies,
this result might be due to the N atom and the increased presence as confirmed by XPS results; this may create a shallow trap for the
of oxygen vacancies in the crystal lattice. Because its band gap lies photogenerated electron and hole, inhibiting recombination and
within the visible-light region and caused considerable light extending the lifetime of charge carriers [34]. Meanwhile, the

1.0 90
0.4
Absorbance

b 80
0.8 e 0.2
70
Absorbance

Decoloration rate%

60
0.6 0.0
500 520 540 560 580 600 50
Wavelength/nm
40
0.4
30
350
c 20 400
0.2 d 450
a 10 500
0 550
0.0
400 500 600 700 800 -10
Wavelength/nm 0 10 20 30 40 50
Irradiation time/ min
Fig. 6. UV–Vis spectra of different samples calcined at 500 °C: (a) BVO; (b) BVON-
13; (c)BVON-20; (d) BVON-25; (e) BVON-30. Fig. 7. Photocatalytic activity of BVON-20 calcined at different temperatures.
76 M. Wang et al. / Journal of Alloys and Compounds 548 (2013) 70–76

100 of N-doped BiVO4 photocatalysts extended a red shift when com-


BVO
pared with that of pure BiVO4. These were the main reasons that
BVON-13
the appropriate concentration of N dopant can significantly increase
BVON-20
80 the photocatalytic activity. However, the photocatalytic activity was
BVON-25
BVON-30 found to decrease with excessive N-doping. In addition, the some-
Decoloration rate%

what improved adsorption ability of MO molecules to the catalytic


60
surfaces after N-doping affected the photocatalytic activity slightly.

40
Acknowledgements

The authors gratefully thank the financially supports of the Na-


20 tional Natural Science Foundation of China (21207093), Education
Department of Liaoning Province (L2010471) and the Doctoral In-
novation Fund of Shenyang Ligong University, China
0 (2010BS0605).

0 10 20 30 40 50 References
Irradiation time/min
[1] B.H. Hameed, A.A. Ahmad, N. Aziz, Chem. Eng. J. 133 (2007) 195–203.
Fig. 8. Photocatalytic activity of N-doped BiVO4 with different amount of N doping. [2] A.L. Ahmad, S.W. Puasa, Chem. Eng. J. 132 (2007) 257–265.
[3] J.H. Mo, Y.H. Lee, J. Kim, J.Y. Jeong, J. Jegal, Dyes Pigm. 76 (2008) 429–434.
[4] A.N.M. Bagyo, H. Arai, T. Miyata, Appl. Radiat. Isot. 48 (1997) 175–181.
[5] H.F. Yu, S.T. Yang, J. Alloys Comp. 492 (2010) 695–700.
absorption spectra of N-doped samples showed stronger absorp- [6] R.M. Mohamed, I.A. Mkhalid, J. Alloys Comp. 501 (2010) 301–306.
tion in the visible range and a red shift in the band gap transition, [7] A.P. Zhang, J.Z. Zhang, J. Alloys Comp. 491 (2010) 631–635.
[8] T. Yang, D.G. Xia, G. Chen, Y. Chen, Mater. Chem. Phys. 114 (2009) 69–72.
as discussed previously. Therefore, N-doping increased the number [9] H.B. Li, G.C. Liu, X.C. Duan, Mater. Chem. Phys. 115 (2009) 9–13.
of photogenerated electrons and holes participating in the photo- [10] C.Y. Chung, C.H. Lu, J. Alloys Comp. 502 (2010) L1–L5.
catalytic reaction, which could partly explain the enhanced photo- [11] S. Eda, M. Fujishima, H. Tada, Appl. Catal. B Environ. 125 (2012) 288–293.
[12] X. Meng, L. Zhang, H.X. Dai, Z.X. Zhao, R.Z. Zhang, Y.X. Liu, Mater. Chem. Phys.
catalytic activity. Third, the increased number of V4+ surface states 125 (2011) 59–65.
in BiVO4 form a donor level between the band gaps of BiVO4, which [13] Xiao, Q.T. Zhou, J. Zhang, L.L. Ouyang, J. Alloys Comp. 468 (2009) L9–L12.
would improve its visible absorption and photocatalytic activity [14] B. Zhou, X. Zhao, H.J. Liu, J.H. Qu, C.P. Huang, Appl. Catal. B Environ. 99 (2010)
214–221.
[35]. Additionally, the lattice expansion due to N-doping (as con-
[15] X.F. Zhang, X. Quan, S. Chen, Y.B. Zhang, J. Hazard. Mater. 177 (2010) 914–917.
firmed by XRD analysis) may lead to some lattice distortion. This [16] W. Liu, Y.Q. Yu, L.X. Cao, G. Su, X.Y. Liu, L. Zhang, Y.G. Wang, J. Hazard. Mater.
distortion may be the other reason for enhanced photocatalytic 181 (2010) 1102–1108.
activity. Based on SEM analysis and the BET specific surface areas [17] Y.C. Chiou, U. Kumar, J.C.S. Wu, Appl. Catal. A Gen. 357 (2009) 73–78.
[18] H.Q. Jiang, M. Nagai, K. Kobayashi, J. Alloys Comp. 479 (2009) 821–827.
of our experiments, the heightened ability of serial N-doped BiVO4 [19] M.C. Long, W.M. Cai, J. Cai, B.X. Zhou, X.Y. Chai, Y.H. Wu, J. Phys. Chem. B 110
samples to adsorb compared to pure BiVO4 would not have con- (2006) 20211–20216.
tributed significantly to the photocatalytic activity. [20] J. Suo, L.F. Liu, F.L. Yang, Chin. J. Catal. 30 (2009) 323–327.
[21] P. Chatchai, Y. Murakami, S.Y. Kishioka, A.Y. Nosaka, Y. Nosaka, Electrochim.
However, when the molar ratio reached 0.25, the photocatalytic Acta 54 (2009) 1147–1152.
activity decreased remarkably, which may be due to the decreased [22] X.F. Zhang, Y.B. Zhang, X. Quan, S. Chen, J. Hazard. Mater. 167 (2009) 911–914.
light absorption when excess nitrogen doping. Thus, there is an [23] C.L. Yu, K. Yang, J.C. Yu, F.F. Cao, X. Li, X.C. Zhou, J. Alloys Comp. 509 (2011)
4547–4552.
optimum concentration of nitrogen doping. [24] A.P. Zhang, J.Z. Zhang, Acta Phys. Chim. Sin. 26 (2010) 1337–1342.
[25] R. Trehan, Y. Lifshitz, J.W. Rabalais, J. Vac. Sci. Technol. A 8 (1990) 4026–4031.
[26] Y. Chen, K.C. Chou, S.P. Huang, Z.Y. Li, G.C. Liu, J. Inorg. Mater. 27 (2012) 19–25.
4. Conclusion
[27] H.Q. Jiang, C.Y. Wang, P. Wang, J.S. Li, Z.Y. Lu, J. Mater. Sci. Eng. 29 (2011) 161–
166.
The N-doped BiVO4 photocatalysts were synthesized by com- [28] W. Liu, S.Y. Lai, H.X. Dai, S.J. Wang, H.Z. Sun, C.T. Au, Catal. Lett. 113 (2007)
plexing sol–gel method using citric acid as chelate and hexamethyl- 147–151.
[29] H.Y. Jiang, H.X. Dai, X. Meng, L. Zhang, J.G. Deng, K.M. Ji, Chin. J. Catal. 32 (2011)
ene tetramine (C6H12N4) as a nitrogen source; these photocatalysts 939–949.
were then characterized using many state-of-the-art techniques. [30] L. Ge, L.S. Cui, J. Chin. Ceram. Soc. 36 (2008) 320–324.
XRD and XPS analysis showed that the N-doping improves the abil- [31] L. Ge, J. Inorg. Mater. 23 (2008) 449–453.
[32] A.P. Zhang, J.Z. Zhang, Spectrochim. Acta, Part A Mol. Biomol. Spectrosc. 73
ity of the compounds to form a monoclinic BiVO4 structure and leads (2009) 336–341.
to slight lattice expansion. The doping N atom may weave into the [33] Y. He, Y.F. Zhu, N.Z. Wu, J. Solid State Chem. 177 (2004) 3868–3872.
interstitial BiVO4 structure and replace an O atom to form an [34] A.P. Zhang, J.Z. Zhang, J. Hazard. Mater. 173 (2010) 265–272.
[35] H.Y. Jiang, H.X. Dai, X. Meng, K.M. Ji, L. Zhang, J.G. Deng, Appl. Catal. B Environ.
O–Bi–N–V–O bond, which contributed to the appearance of more 105 (2011) 326–334.
V4+ and oxygen vacancies. Additionally, the light absorption edges

You might also like