DOI:
10.1039/C4RA13990A
(Paper)
RSC Adv., 2015,
5, 10688-10696
Efficient photocatalytic degradation of bisphenol A and dye pollutants over BiOI/Zn2SnO4 heterojunction photocatalyst†
Received
6th November 2014
, Accepted 5th January 2015
First published on 5th January 2015
Abstract
BiOI/Zn2SnO4 composites with a heterostructure have been synthesized via a chemical deposition method under mild conditions by tuning the BiOI mass ratios. The physicochemical characteristics were investigated by X-ray diffraction (XRD), diffuse reflectance ultraviolet-visible light spectroscopy (DRS), scanning electron microscopy (SEM), and high-resolution transmission electron microscopy (HRTEM). The XRD results show that two phases of BiOI and Zn2SnO4 co-existed in the composites. The HRTEM image showing clear lattice fringes proves the formation of a heterojunction at the interfaces of BiOI and Zn2SnO4. The photocatalytic degradation of the endocrine disruptor bisphenol A and dyes (MB and RhB) indicated that the BiOI/Zn2SnO4 composites were more photoactive than pure BiOI and Zn2SnO4. The activity enhancement was mainly ascribed to the formation of a heterojunction between BiOI and Zn2SnO4, which facilitated the transfer and separation of photogenerated electron–hole pairs. The photoelectrochemical measurement has also confirmed the enhancement of the separation efficiency of electron–hole pairs.
Introduction
Heterogeneous photocatalysis has become an exciting and rapidly growing research area in the last few years.1,2 Photocatalysis technology can be used to decompose organic compounds into inorganic substances efficiently and completely under mild conditions, so it has great application prospects for purifying industrial wastewater,3,4 or splitting water into hydrogen and oxygen gases for clean renewable energy.5 There are numerous papers on the photodegradation of organic compounds, and the reported photocatalysts mainly include oxides, nitrides, sulfides, etc., but most of them are focused on TiO2-based photocatalyst.6,7 As classic photocatalyst, TiO2 has a low quantum yield, poor solar energy utilization, and severe deactivation, which greatly limits the application. So, to develop novel non-TiO2 based photocatalysts, which were environmental friendly and highly efficient has become the focus of photocatalysis research. The n-type Zn2SnO4 semiconductor which is composed of the d10 p-block metal ion has potential applications in photoelectrical devices, batteries, functional coatings, and photocatalysis due to its high electron mobility and high electrical conductivity.8–12 Owing to its wide band-gap, Zn2SnO4 can only utilize a small fraction of solar light. In order to expand its optical response from the UV to visible light region and improve the separation efficiency of photo-generated carriers, narrow band gap semiconductors have been used to combine with it. Semiconductor coupling not only extend the light absorption range but also promote the photo-induced carriers separation efficiency, which ultimately result in excellent photocatalytic activity.13 In the past few years, the heterojunction fabricated by semiconductor coupling has also been used to prepare novel catalysts, such as Bi2O3/Bi2S3,14 g-C3N4/Bi2O2CO3,15 g-C3N4/BiOX (X = Br, I),16,17 Bi2O3/BiVO4,18 and Bi2O2CO3/BiOI19 have been developed for effective visible light driven photocatalyst.
Recently, BiOI has drawn extensive interests of researchers because of its unique electrical and optical properties.20–22 As a typical p-type semiconductor, BiOI has been naturally selected to couple with n-type TiO2,23,24 ZnSn(OH)6,25 ZnTiO326 and ZnO27 etc. Likewise, Fan and co-workers28 reported the synthesis of p-BiOI/n-Zn2SnO4 heterostructures with photocatalytic activity for the degradation of MO under visible light irradiation, but the activity was still not very well. Thus, it pushes us to thoroughly investigate the synthesis and activity enhancement of BiOI/Zn2SnO4 with heterostructure.
Herein, we report the characterization and photocatalytic properties of BiOI/Zn2SnO4 composite prepared by a facile and economical two-step method. Zn2SnO4 nanoparticles were first prepared by solvothermal method, and then, the BiOI/Zn2SnO4 composites were prepared by chemical deposition method under mild conditions. The photocatalytic activity and recycling ability were investigated by the degradation of dye pollutants (MB, RhB) and endocrine disruptor bisphenol A (BPA) under UV and visible light irradiation, respectively. The degradation efficiency of pollutants over BiOI/Zn2SnO4 composites was greatly enhanced compared with that over pure Zn2SnO4 or BiOI. This activity enhancement was mainly ascribed to the formation of heterojunction at the interface of BiOI and Zn2SnO4, which facilitated the transfer and separation of photo-generated carriers, as well as the strong visible light absorption originating from the sensitization role of BiOI to Zn2SnO4.
Experimental section
Synthesis of Zn2SnO4 precursor
All chemicals with analytical purity were obtained from Sinopharm Chemical Reagent Co., Ltd and were used without further purification. Deionized water was employed in all experiments. In a typical procedure for the synthesis of Zn2SnO4, 2.10 g ZnCl2 and 2.70 g SnCl4·5H2O were added to 100 mL water/ethylene glycol = 1
:
1 (volume ratio) under magnetic stirring. Then 50 mL of 2 M n-butylamine aqueous solution was added dropwise to the above solution. The final concentration of n-butylamine in the solution was 0.4 M. After stirring for 30 min, the obtained white slurry was transferred to 100 mL autoclave and maintained at 180 °C for 20 h. The product formed at the bottom of the autoclave was centrifuged and rinsed thoroughly with deionized water and ethanol several times. Finally, the product denoted as ZSO was dried in air at 60 °C for 6 h.
Preparation of BiOI/Zn2SnO4 composites
The BiOI/Zn2SnO4 samples with different BiOI contents were prepared by a deposition method. In a typical experiment, different stoichiometric amounts of Bi(NO3)3·5H2O and KI were dissolved in 20 mL ethylene glycol to obtain a clear solution A. Zn2SnO4 was ultrasonically dispersed into 80 mL deionized water to form a homogeneous solution B. Then, the solution A was added drop-wise into the solution B under strong stirring. After being stirred for 1.0 h at room temperature, the resulting mixtures were heated at 80 °C for 2.0 h in an oil bath. Finally, the precipitates were collected, washed thoroughly with deionized water and ethanol, and dried at 60 °C in air. According to this method, BiOI/Zn2SnO4 composites with different mass ratios of 20, 25, 30, 35, and 40 wt% have been synthesized and named as BiOI/ZSO-20, BiOI/ZSO-25, BiOI/ZSO-30, BiOI/ZSO-35, and BiOI/ZSO-40, respectively. For comparison, pure BiOI was prepared by adopting the method in the absence of Zn2SnO4.
Characterization of photocatalysts
X-ray diffraction (XRD) patterns of the obtained products were recorded on a Bruker D8 Advance X-ray diffractometer under the conditions of generator voltage = 40 kV; generator current = 40 mA; divergence slit = 1.0 mm; Cu Kα (λ = 1.5406 Å); and polyethylene holder. The morphologies and microstructures of the samples were examined with a Hitachi S-4800 scanning electron microscopy (SEM). The transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) images were measured by a JEOL model JEM 2010 EX instrument at an accelerating voltage of 200 kV. Carbon-coated copper grid was used as the sample holder. Energy-dispersive X-ray spectra (EDS) were obtained on a JEOL-2010 at an accelerating voltage of 200 kV. Diffuse reflectance ultraviolet-visible light spectra (DRS) were measured at room temperature in the range of 200–700 nm on a UV-vis spectrophotometer (Cary 500 Scan Spectrophotometers, Varian, and U.S.A) equipped with an integrating sphere attachment.
Photocatalytic tests
Photocatalytic reactions were acted in a customized reactor with a cooling-water-cycle system. It can keep the reaction temperature of the aqueous solution to maintain at room temperature. The visible light activity of BiOI/Zn2SnO4 composites were evaluated by photo-degradation of dyes and BPA in aqueous solution using a 300 W xenon lamp (PLS-SXE300C, Beijing Perfect Light Co. Ltd., Beijing) with a cutoff filter (λ > 420 nm) as light source. The UV light activities of BiOI/Zn2SnO4 composites were tested using four 4 W UV lamps with a wavelength centered at 254 nm (Philips, TUV 4W/G4 T5) as the light source. In each experiment, 100 mg of photocatalyst was added into 100 mL dye (MB and RhB) solution (10 mg L−1) or BPA solution (20 mg L−1). Before irradiation, the suspensions were magnetically stirred in dark for 60 min to ensure the establishment of an adsorption/desorption equilibrium. During reaction under light irradiation, 3 mL of suspension was sampled at given time intervals and centrifuged to remove the photocatalyst particles. The resulting clear liquor was analyzed on a PerkinElmer UV-vis spectrophotometer (Model: Lambda 35) to record the concentration changes of pollutant solution.
Photocurrent measurements (PC)
The photoelectrochemical experiment was conducted on CHI 760E electrochemical workstation (CHI 760E Chenhua Instrument Company, Shanghai, China) in a conventional three-electrode configuration with a Pt wire as the counter electrode and a saturated calomel electrode as reference electrode. Irradiation proceeded by a Xe arc lamp through a UV cut off filter (λ > 420 nm). Na2SO4 (0.2 M) aqueous solution was used as the electrolyte. The working electrodes were prepared as follows: an indium tin oxide (ITO) glass piece with a size of 2.0 × 1.0 cm. Then ITO glass piece was cleaned successively by acetone, ethanol, deionized water, and then dried in N2 stream. 5.0 mg of the ground sample was dispersed uniformly with 1.0 mL of distilled water under the condition of ultrasonic, and 10 μL of above solution was added to surface of the ITO, and then 5 μL of Nafion solution (0.5%) was dropping added on the modified working electrode surface as binder and dried at 120 °C for 1 h.
Results and discussion
Structural characterization
Fig. 1 shows the XRD patterns of pure BiOI, Zn2SnO4, and BiOI/Zn2SnO4 composites with different mass ratios. As we can see, all the diffraction peaks in Fig. 1a could be indexed to the tetragonal phase of BiOI (JCPDS no. 73-2062)29 while that for Zn2SnO4 was the cubic phase (JCPDS no. 74-2184).30 The sharp and intense diffraction peaks of BiOI and Zn2SnO4 indicated that the samples were well crystallized. For the BiOI/Zn2SnO4 composites (Fig. 1b–f), characteristic peaks for BiOI and Zn2SnO4 were both observed. Notably, the peaks assigned to BiOI became broader and weaker, suggesting that the presence of Zn2SnO4 could inhibit the crystal growth of BiOI.31 In addition, no other impurity peaks have been found from the diffraction patterns, indicating that the composites are only composed of BiOI and Zn2SnO4.
 |
| Fig. 1 XRD patterns of (a) pure Zn2SnO4, (b) BiOI/ZSO-20, (c) BiOI/ZSO-25, (d) BiOI/ZSO-30, (e) BiOI/ZSO-35, (f) BiOI/ZSO-40, and (g) pure BiOI. | |
The morphologies of Zn2SnO4, BiOI and BiOI/Zn2SnO4 composites have been characterized by SEM. As shown in Fig. 2, the pure Zn2SnO4 presents irregular aggregates of uniform nanoparticles with the diameter of approximately 5–10 nm, as reported by Kim et al.32 It also reveals that the pure BiOI was consisted of large number of irregular plates with smooth surfaces. As for the BiOI/ZSO-35 sample, it obviously shows that the Zn2SnO4 nanoparticles are uniformly embedded in the BiOI plates (Fig. 2e and f). Meanwhile, the morphology of BiOI nanoplate changed obviously after the addition of Zn2SnO4 nanoparticles, further revealing that the addition of Zn2SnO4 significantly affects the morphologies and crystal growth of BiOI, which is consistent with the XRD analysis. A similar phenomenon has also been reported by Zhang et al.31 Fig. 2g presents the typical EDS line scanning results. It indicates that the elements of Bi, O, I, Sn and Zn are uniformly distributed across the scanned distance. Fig. 2h displays the EDS spectrum for the as-prepared sample. It can be seen that only Bi, O, I, Sn and Zn elements existed in the composite, demonstrating that BiOI/ZSO-35 composite was composed of both BiOI and Zn2SnO4.
 |
| Fig. 2 SEM images of (a) and (b) pure Zn2SnO4; (c) and (d) pure BiOI; (e) and (f) BiOI/ZSO-35; (g) the EDS line-scans image of BiOI/ZSO-35; (h) the EDS spectrum of BiOI/ZSO-35. | |
In order to further ascertain the formation of heterojunction between BiOI and Zn2SnO4, the sample was further investigated by TEM and HRTEM. Fig. 3a shows the typical TEM image of BiOI/Zn2SnO4 composite, in which Zn2SnO4 nanoparticles were deposited on the surface of the plate-like BiOI substrates, it was consistent with the SEM observations. A magnified TEM image (Fig. 3b) reveals that some nanoparticles with sizes of about 7 nm are anchored on the plates. Fig. 3c presents the HRTEM image of the BiOI/Zn2SnO4 composite, the clear lattice fringes reveal that the materials are highly crystallized. Two sets of different lattice images are observed with d spaces of 0.28 and 0.30 nm corresponding to the (1 1 0) and (1 0 2) plane of tetragonal BiOI,33 and with d spaces of 0.26 and 0.31 nm belonging to the (3 1 1) and (2 2 0) planes of Zn2SnO4,28 respectively. The HRTEM analysis demonstrated that the existence of BiOI/Zn2SnO4 heterojunction. These results are in good accordance with the XRD results. And the corresponding EDS mapping images (Fig. 3d) of Zn, Sn, O, Bi and I elements further demonstrated that these elements distribute uniformly all through the composites.
 |
| Fig. 3 (a) and (b) TEM and (c) HRTEM images of the BiOI/ZSO-35 composite, and (d) the corresponding EDS mapping images of Zn, Sn, O, Bi and I elements. | |
Optical characterization
The optical absorption properties of Zn2SnO4, BiOI, and BiOI/Zn2SnO4 composites are shown in Fig. 4. The absorption edges of Zn2SnO4 and BiOI were roughly located at 345 and 670 nm, respectively. When Zn2SnO4 combines with different amounts of BiOI, the optical absorption edges shifted from 345 to 650 nm. It has been reported that the inhibited recombination between photoelectrons and holes could result in the strong response in the visible region.34,35 Moreover, the steep shape of the spectrum indicated that the visible light absorption was due to the band gap transition. The band gap energy of a semiconductor could be calculated by the following equation:36
where α, A, ν and Eg are the absorption coefficient, a constant, light frequency and band-gap energy, respectively. Among them, n depends on the characteristics of the transition in a semiconductor (direct transition: n = 1; indirect transition: n = 4). For pure BiOI and Zn2SnO4, the value of n is 437 and 1,38 respectively. The band gap energies of the BiOI and Zn2SnO4 can be estimated from the plots of (αhν)2 or (αhν)1/2 versus photon energy (hν), respectively. Fig. S1† shows that the calculated band gaps (Eg) of Zn2SnO4 and BiOI is 3.59 and 1.85 eV, respectively.
 |
| Fig. 4 UV-vis diffuse reflectance spectra of Zn2SnO4, BiOI and BiOI/Zn2SnO4 composites. | |
Photocatalytic activity
The photocatalytic activities of BiOI/Zn2SnO4 composites were first investigated by the degradation of MB (10 mg L−1). Fig. 5a shows the degradation of MB over Zn2SnO4, BiOI, and BiOI/Zn2SnO4 composites under visible (λ > 420 nm) light irradiation. As we can see, MB degradation was not observed in the dark condition. However, in the absence of any photocatalyst, the concentrations of MB showed a little decrease under visible light irradiation. It is noteworthy that the MB decomposition over pure Zn2SnO4 has also been observed under visible light irradiation (Fig. 5a). This phenomenon may be ascribed to the weak dye photosensitization role of MB.39 When BiOI/Zn2SnO4 composites were added, the degradation rates of MB were significant improved compared with that over pure BiOI and Zn2SnO4. Furthermore, the BiOI content contributes great influences on the photocatalytic activities of BiOI/Zn2SnO4 composites. Concretely, the photocatalytic activities were first remarkably enhanced along with the increasing of BiOI content, when the BiOI content was larger than 35%, the activities would decrease gradually, suggesting that the optimal BiOI mass ratio in BiOI/Zn2SnO4 composite was 35 wt%. Under the optimum conditions, nearly 100% of MB could be decomposed within 100 min under visible light irradiation. The photocatalytic degradations of MB under UV light (254 nm) irradiation have also investigated, and a very similar phenomenon has been observed. As shown in Fig. 5b, the BiOI content also greatly influenced the UV activities of BiOI/Zn2SnO4 composites, and the BiOI/ZSO-35 sample exhibited the best activity. Under UV light irradiation, nearly 100% of MB could be decomposed within 90 min of reaction. As a comparison, the degradation of MB over PM-BiOI/ZSO-35 (the physical mixture of BiOI and Zn2SnO4, Labeled as PM-BiOI/ZSO-35) and BiOI/ZSO-R (prepared according the ref. 28) was also investigated under the same conditions, and the degradation results were shown in Fig. S1.† As we can see, whatever under UV or visible light irradiation, the activity of PM-BiOI/ZSO-35 is getting worse compared with that of BiOI/ZSO-35. Besides, the degradation rates of MB over BiOI/ZSO-35 was remarkably enhanced compared with BiOI/ZSO-R. The enhancement of photocatalytic activity could be attributed the synergetic effects of light absorption and heterojunction structure of BiOI/Zn2SnO4. It has also been found that the photocatalytic degradation of MB dye followed the first-order kinetics, and the kinetic constant (k) values were shown in Fig. 6. As we can see, whatever under UV or visible light irradiation, the BiOI content greatly affected the kinetic constant k. It reaches the maximum value when the mass ratio of BiOI was 35%. The kinetic constant k of BiOI/ZSO-35 under UV and visible light irradiation is 0.3124 and 0.3675 mim−1, respectively.
 |
| Fig. 5 Photocatalytic activities of Zn2SnO4, BiOI and BiOI/Zn2SnO4 composites for the degradation of MB (10 mg L−1) under (a) visible (λ > 420 nm) light and (b) UV (λ = 254 nm) light irradiation. | |
 |
| Fig. 6 Photocatalytic degradation rate constant k of MB as a function of BiOI content (wt%) under visible (λ > 420 nm) light and UV (λ = 254 nm) light irradiation. | |
BPA and RhB were also used to evaluate the photocatalytic performances of BiOI/ZSO-35 composite. As a comparison, pure BiOI and Zn2SnO4 were also tested under identical conditions. Fig. 7 shows the BPA and RhB degradation curve under UV and visible light irradiation, respectively. As it can be seen, whatever under UV or visible light irradiation the degradation rates of BPA and RhB over BiOI/ZSO-35 was remarkably enhanced compared with pure BiOI and Zn2SnO4. For the BPA degradation, the physical mixture of BiOI and Zn2SnO4 (labeled as PM-35 wt% BiOI/ZSO) was also used as a comparison. It can be clearly seen that the photocatalytic performance of BiOI/ZSO-35 was obviously higher than that of PM-35 wt% BiOI/ZSO. Under visible light irradiation, the degradation ratio of BPA was up to 99% after 180 min of reaction, while that over PM-35 wt% BiOI/ZSO was only 21%. Under UV light irradiation, the degradation rate of BPA was greatly accelerated, and about 97% of BPA was photocatalytical decomposed within 100 min of reaction. As for the degradation of RhB, a very similar phenomenon was observed.
 |
| Fig. 7 Photo-degradation activities of Zn2SnO4, BiOI, BiOI/ZSO-35, and PM-35 wt% BiOI/ZSO for the degradation of (a and b) BPA and (c and d) RhB under UV and visible light irradiation. | |
The photocatalytic kinetic rate constants for the degradation of different pollutants were shown in Fig. 8. It can be seen that the k for RhB photo-degradation over the BiOI/ZSO-35 catalyst under UV and visible light irradiation was about 2.1 and 8.4 times, 35 and 10 times higher than that over pure Zn2SnO4 and BiOI, respectively. The k for BPA photodegradation over the BiOI/ZSO-35 catalyst under UV and visible light irradiation was about 6 and 10 times, 53 and 37 times higher than that over pure Zn2SnO4 and BiOI, respectively.
 |
| Fig. 8 Photo-degradation rate constant k of MB, RhB and BPA upon BiOI/ZSO-35 under (a) visible (λ > 420 nm) light and (b) UV (λ = 254 nm) light irradiation. | |
These results indicated clearly that the activity enhancement of BiOI/Zn2SnO4 under UV and visible light irradiation may be due to the synergistic effect of heterojunction formed between Zn2SnO4 and BiOI accelerates separating electron–hole pairs at the interface.
Regeneration and reusability
To test the stability of BiOI/Zn2SnO4 composites for the photocatalytic degradation of MB, the catalyst recycled after reaction was reused for photocatalytic reaction five times under the same conditions and the results are shown in Fig. 9. It can be seen that the photocatalytic efficiency does not exhibit any significant loss after five recycles, which indicate that the catalyst is stable for the photodegradation of MB. Additionally, the XRD patterns in Fig. 10 reveal that there is no observable structural difference between the samples before and after reaction, indicating that the phase and structure of BiOI/Zn2SnO4 heterostructure remain intact and stable. These results indicate that the BiOI/Zn2SnO4 composite prepared by this facile method is stable and completely recyclable for the photocatalysis degradation of pollutant, which is important for its practical application.
 |
| Fig. 9 Recycling test on the BiOI/ZSO-35 composite for the degradation of MB under (a) visible (λ > 420 nm) light and (b) UV (λ = 254 nm) light irradiation. | |
 |
| Fig. 10 XRD patterns of the BiOI/ZSO-35 composite before and after photocatalytic reaction. | |
Possible photocatalytic mechanism
Separation efficiency of photo-generated electron–hole pairs. To investigate the interface charge separation efficiencies of BiOI/Zn2SnO4 composites, photocurrent and EIS characterizations have been carried out. As shown in Fig. 11a, a fast photocurrent response can be observed for each switch-on and switch-off event in all the electrodes. The BiOI/Zn2SnO4 composites exhibit higher transient photocurrent intensity than that on single-component semiconductor, and the photocurrent changes tendency of the samples correspond to the photocatalytic performance variations. The transient photocurrent enhancements of the BiOI/Zn2SnO4 composites indicate the high separation rates of photo-induced electron–hole pairs. Among these, the BiOI/ZSO-35 composite exhibits the highest current intensity, demonstrating that a more effective separation of photon-generated carriers and faster interfacial charge transfer occurred in the BiOI/ZSO-35 composite. As shown in Fig. 11b, the diameter of the Nyquist circle of BiOI/ZSO-35 composite was smaller than that of pure BiOI and Zn2SnO4, indicating that the BiOI/ZSO-35 composite had a relatively lower resistance compared with one-component catalyst. The results demonstrate that the effective transfer and separation of photo-induced carriers is critical factors to improve the activity of heterojunction.
 |
| Fig. 11 (a) Photocurrent responses and (b) EIS profiles of pure Zn2SnO4, BiOI, and BiOI/Zn2SnO4 composites under visible light irradiation. | |
Detection of active species
To investigate the photocatalytic degradation mechanism of pollutants over these heterostructured catalysts, diagnostic experiments have also been performed. Hence, various scavengers, such as t-BuOH (˙OH scavenger),40 (NH4)2C2O4 (AO) (h+ scavenger),41 p-benzoquinone (BQ) (˙O2− scavenger), and NaN3 (1O2 scavenger),42 were introduced to evaluate the contributions of different active species on the photo-degradation efficiency. The photocatalytic degradation results in the presence of different scavengers under UV or visible light irradiation are shown in Fig. 12. When IPA (1.0 mL) was added, the degradation ratios of MB under visible and UV light illumination decreased to 70 and 60%, respectively. When AO (0.2 g) added, the degradation ratios decreased to 30 and 80% accordingly. Furthermore, the addition of BQ (0.05 g) or NaN3 (0.05 g) could also decrease the photo-degradation rates of MB under the same reaction conditions. Summarizing the results above, the reactive species ˙OH, ˙O2−, h+, and 1O2 participated in the photo-degradation reaction and played equivalent role in the photocatalytic process under visible light irradiation. However, under UV light irradiation, the effect of h+ and 1O2 played vital important contributions, but the participation of ˙OH and ˙O2− played a minor role for the MB degradation. In order to study the effect of molecular oxygen on the degradation of MB, N2 was bubbled into the catalytic system to ensure that the reaction was operated without O2 as an electron scavenger to produce a variety of active oxygen species. It shows that the degradation efficiency of MB decreased obviously, which meant that the molecular oxygen may play a non-ignorable role in the degradation process.
 |
| Fig. 12 Effects of different scavengers on the degradation of MB in the presence of 35 wt% BiOI/ZSO under (a) visible (λ > 420 nm) and (b) UV (λ = 254 nm) light irradiation. | |
The proposed photocatalytic mechanism
In order to understand the band structures of BiOI/Zn2SnO4, the CB and VB positions of the materials were calculated by a simple theoretical method. The CB and VB positions were calculated through the equation:
where X is the Mulliken's electronegativity, Ee is the energy of free electrons on the hydrogen scale (4.50 eV), and Eg is the band gap.43 The X values for Zn2SnO4 and BiOI were 7.0 and 5.94 eV, respectively. Accordingly, the top of VB and bottom of CB of Zn2SnO4 was calculated to be +4.3 and +0.7 eV (vs. NHE), respectively. The VB and CB of BiOI was estimated to be +0.57 and +2.42 eV (vs. NHE), respectively. When BiOI and Zn2SnO4 are in contact, since the CB potential of Zn2SnO4 is more negative than that of BiOI, the photo-generated electrons will transfer from Zn2SnO4 to BiOI, resulting in the accumulation of negative charges in BiOI close to the junction. Meanwhile, the photo-generated holes transfer from BiOI to Zn2SnO4, leaving a positive section in Zn2SnO4 near the junction. With equilibration of BiOI and Zn2SnO4 Fermi levels (Ef), the diffusion of electrons from Zn2SnO4 to BiOI stops. At the same time, the energy bands of BiOI shift upward along the Fermi level (Ef–p) and those of the Zn2SnO4 shift downward along its Fermi level (Ef–n), as shown in Scheme 1.
 |
| Scheme 1 Photocatalytic degradation mechanism over BiOI/Zn2SnO4 heterostructure under (a) visible and (b) UV light irradiation. | |
Under visible light irradiation, only BiOI could be excited to generate electron–hole pairs. Due to the effect of the internal field with the direction from Zn2SnO4 to BiOI, the photoelectrons on the CB of BiOI efficiently transferred to that of Zn2SnO4, and holes remained on the VB of BiOI. Under UV light irradiation, both BiOI and Zn2SnO4 absorb photons of energy greater than the corresponding band gap energy, which excite the electrons in the VB to the CB and leave holes in the VB. The photo-generated electrons in the CB of BiOI are then transferred to the CB of Zn2SnO4, and the photo-generated holes in the VB of Zn2SnO4 transferred to the VB of BiOI. Therefore, the photo-generated electron–hole pairs will be effectively separated due to the formation of a junction between the BiOI and Zn2SnO4 interface, resulting in a suppressed electron–hole recombination.
The migration of photo-generated carriers can be promoted by the inner electric field established at the heterojunction interfaces. The photoelectrons were further trapped by molecular oxygen44 to produce ˙O2− or ˙OH reactive species.45 Meanwhile, the photo-generated holes would decompose the pollutants adsorbed on the surface of the composite or react with ˙O2− to form 1O2 reactive species, which continue participate in the degradation reaction.
Conclusions
BiOI/Zn2SnO4 composites with different mass ratios have been synthesized via a facile chemical deposition method. The absorption edge of the BiOI/Zn2SnO4 heterojunction shifted significantly to longer wavelengths compared to Zn2SnO4. The BiOI/ZSO-35 composite shows excellent photocatalytic activities in the degradation of MB, RhB, and BPA under visible or UV light irradiation, respectively. The formation of heterostructure with intimate contacts between Zn2SnO4 and BiOI resulted in the easier charge transfer and more efficient separation of electron–hole pairs, which finally improved the photocatalytic activities of the composites. This study provides a simple and economical route to couple two different semiconductors with different physicochemical properties for developing novel high-performance photocatalysts.
Acknowledgements
This work was financially supported by the National Natural Science Foundation of China (no. 21271027, 21103069, and 40672158), the 863 Project of China (no. 20140AA065103-1), Scientific Research Reward Fund for Excellent Young and Middle-Aged Scientists of Shandong Province (BS2012HZ001), and Scientific Research Foundation for Doctors of University of Jinan (XBS1037, XKY1321, and XKY1043).
References
- A. Fujishima and K. Honda, Nature, 1972, 238, 37 CrossRef CAS.
- Z. Zou, J. Ye, K. Sayama and H. Arakawa, Nature, 2001, 414, 625 CrossRef CAS PubMed.
- M. Xie, L. Jing, J. Zhou, J. Lin and H. Fu, J. Hazard. Mater., 2010, 176, 139 CrossRef CAS PubMed.
- M. Shannon, P. Bohn, M. Elimelech, J. Georgiadis, B. Mari as and A. Mayes, Nature, 2008, 452, 301 CrossRef CAS PubMed.
- D. Yamasita, T. Takata, M. Hara, J. N. Kondo and K. Domen, Solid State Ionics, 2004, 172, 591 CrossRef CAS PubMed.
- R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and T. Taga, Science, 2001, 293, 269 CrossRef CAS PubMed.
- H. G. Yang, C. H. Sun, S. Z. Qiao, J. Zou, G. Liu, S. C. Smith, H. M. Cheng and G. Q. Lu, Nature, 2008, 453, 638 CrossRef CAS PubMed.
- X. S. Liu, J. Q. Wang, E. J. Liang and W. F. Zhang, Appl. Surf. Sci., 2013, 280, 556 CrossRef CAS PubMed.
- K. Wang, Y. Huang, H. J. Huang, Y. Zhao, X. L. Qin, X. Sun and Y. L. Wang, Ceram. Int., 2014, 40, 8021 CrossRef CAS PubMed.
- X. Y. Liu, H. W. Zheng, Z. L. Zhang, X. S. Liu, R. Q. Wan and W. F. Zhang, J. Mater. Chem., 2011, 21, 4108 RSC.
- M. Peiteado, Y. Iglesias, J. F. Fernándeza, J. De Frutos and A. C. Caballero, Mater. Chem. Phys., 2007, 101, 1 CrossRef CAS PubMed.
- E. L. Folettoa, J. M. Sim, M. A. Mazutti, S. L. Jahn and E. I. Muller, Ceram. Int., 2013, 39, 4569 CrossRef PubMed.
- X. B. Chen, S. H. Shen, L. J. Guo and S. S. Mao, Chem. Rev., 2011, 110, 6503 CrossRef PubMed.
- L. Chen, J. He, Q. Yuan, Y. Liu, C. T. Au and S. F. Yin, J. Mater. Chem. A, 2015, 3, 1096 CAS.
- M. Xiong, L. Chen, Q. Yuan, J. He, S. L. Luo, C. T. Au and S. F. Yin, Dalton Trans., 2014, 43, 8331 RSC.
- J. Di, J. X. Xia, S. Yin, H. Xu, M. Q. He, H. M. Li, L. Xu and Y. P. Jiang, RSC Adv., 2013, 3, 19624 RSC.
- J. Di, J. X. Xia, S. Yin, H. Xu, Y. G. Xu, M. Q. He and H. M. Li, J. Mater. Chem. A, 2014, 2, 5340 CAS.
- L. Chen, Q. Zhang, R. Huang, S. F. Yin, S. L. Luo and C. T. Au, Dalton Trans., 2012, 41, 9513 RSC.
- L. Chen, S. F. Yin, S. L. Luo, R. Huang, T. Hong and C. T. Au, Ind. Eng. Chem. Res., 2012, 51, 6760 CrossRef CAS.
- J. X. Xia, S. Yin, H. M. Li, H. Xu, L. Xua and Q. Zhang, Colloids Surf., A, 2011, 387, 23 CrossRef CAS PubMed.
- Y. Y. Li, J. S. Wang, H. C. Yao, L. Y. Dang and Z. J. Li, J. Mol. Catal. A: Chem., 2011, 334, 116 CrossRef CAS PubMed.
- R. Hao, X. Xiao, X. X Zuo, J. M. Nan and W. D. Zhang, J. Hazard. Mater., 2012, 209–210, 137 CrossRef CAS PubMed.
- G. P. Dai, J. G. Yu and G. Liu, J. Phys. Chem. C, 2011, 115, 7339 CAS.
- X. Zhang, L. Z. Zhang, T. F. Xie and D. J. Wang, J. Phys. Chem. C, 2009, 113, 7371 CAS.
- H. Q. Li, Y. M. Cui, W. S. Hong and B. L. Xu, Chem. Eng. J., 2013, 228, 1110 CrossRef CAS PubMed.
- K. H. Reddy, S. Martha and K. M. Parida, Inorg. Chem., 2013, 52, 6390 CrossRef CAS PubMed.
- J. Jiang, X. Zhang, P. B. Sun and L. Z. Zhang, J. Phys. Chem. C, 2011, 115, 20555 CAS.
- H. R. Li, Z. Jin, H. G. Sun, L. M. Sun, Q. B. Li, X. Zhao, C. J. Jia and W. L. Fan, Mater. Res. Bull., 2014, 55, 196 CrossRef CAS PubMed.
- J. Cao, B. Y. Xu, B. D. Luo, H. L. Lin and S. F. Chen, Catal. Commun., 2011, 13, 63 CrossRef CAS PubMed.
- X. D. Lou, X. H. Jia, J. Q. Xu, S. Z Liu and Q. H. Gao, Mater. Sci. Eng., A, 2006, 432, 221 CrossRef PubMed.
- J. Jiang, X. Zhang, P. B. Sun and L. Z. Zhang, J. Phys. Chem. C, 2011, 115, 20555 CAS.
- K. Kim, A. Annamalai, S. H. Park, T. H. Kwon, M. W. Pyeon and M. J. Lee, Electrochim. Acta, 2012, 76, 192 CrossRef CAS PubMed.
- Y. Y. Li, J. S. Wang, H. C. Yao, L. Y. Dang and Z. J. Li, J. Mol. Catal. A: Chem., 2011, 334, 116 CrossRef CAS PubMed.
- X. Zhang and L. Z. Zhang, J. Phys. Chem. C, 2010, 114, 18198 CAS.
- X. Fu, X. Wang and J. Long, J. Solid State Chem., 2009, 182(3), 517 CrossRef CAS PubMed.
- X. Zhang, Z. H. Ai, F. L. Jia and L. Z. Zhang, J. Phys. Chem. C, 2008, 112, 747 CAS.
- J. Cao, B. Y. Xu, H. L. Lin, B. D. Luo and S. F. Chen, Chem. Eng. J., 2012, 185(186), 91 CrossRef PubMed.
- H. Lin, S. Liao and S. Hung, Mater. Chem. Phys., 2009, 117(1), 9 CrossRef CAS PubMed.
- J. Cao, B. Y. Xu, H. L. Lin and S. F. Chen, Chem. Eng. J., 2013, 228, 482 CrossRef CAS PubMed.
- Y. Hu, D. Z. Li, Y. Zheng, W. Chen, Y. H. He, Y. Shao, X. Z. Fu and G. C. Xiao, Appl. Catal., B, 2011, 104, 30 CrossRef CAS PubMed.
- J. Cao, B. Y. Xu, H. L. Lin, B. D Luo and S. F Chen, Catal. Commun., 2012, 20, 204 CrossRef PubMed.
- C. Chang, L. Y. Zhu, Y. Fu and X. L. Chu, Chem. Eng. J., 2013, 233, 305 CrossRef CAS PubMed.
- Y. Xu and M. A. A. Schoonen, Am. Mineral., 2000, 85, 543 CAS.
- H. F. Cheng, B. B. Huang, X. Y. Qin, X. Y. Zhang and Y. Dai, Chem. Commun., 2012, 48, 97 RSC.
- L. Chen, R. Huang, S. F. Yin, S. L. Luo and C. T. Au, Chem. Eng. J., 2012, 193–194, 123 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra13990a |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.