Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Highly improved photocatalytic degradation of rhodamine B over Bi2Ga4−xFexO9 solid solutions under visible light irradiation

Jia Yang*a, Xiaorui Sun*a, Chunmei Zengb, Xiaoting Wanga, Yilan Hua, Ting Zenga and Jianwei Shia
aChongqing Key Laboratory of Inorganic Special Functional Materials, College of Chemistry and Chemical Engineering, Yangtze Normal University, Fuling, Chongqing 408100, P. R. China. E-mail: sunxiaoruiyznu@163.com; yangjiayznu@163.com; Tel: +86-18883876787 Tel: +86-18716372096
bChemical Synthesis and Pollution Control Key Laboratory of Sichuan Province, College of Chemistry and Chemical Engineering, China West Normal University, Nanchong 637002, P. R. China

Received 20th June 2019 , Accepted 13th August 2019

First published on 28th August 2019


Abstract

In this work, Bi2Ga4−xFexO9 (0 ≤ x ≤ 1.2) solid solutions were prepared via the traditional high-temperature solid-state reaction. The Le Bail fitting on the powder X-ray diffraction patterns shows that these solid solutions were successfully synthesized. Scanning electron microscopy showed that the Bi2Ga3.2Fe0.8O9 sample was composed of sub-micron particle crystallites. Energy dispersive spectroscopy analysis and X-ray photoelectron spectroscopy were used to identify that the Fe element is trivalent when doping into the crystal structure. Ultraviolet-visible diffused reflectance spectra suggested that the bandgap of Bi2Ga3.2Fe0.8O9 is narrower than that of the undoped Bi2Ga4O9 sample. Three strategies, including Fe3+ doping, addition of H2O2, and loading of the cocatalyst, were utilized to improve the photocatalytic degradation activity. The optimum photocatalytic performance was obtained over 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 sample in 20 ppm RhB aqueous solution (containing 1.5 mL H2O2) under visible light irradiation. Its photodegradation rate is 8.0 times that of Bi2Ga4O9 containing 0.5 mL H2O2. The 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 photocatalyst remained stable and active even after four cycles. Also, its photocatalytic conversion efficiency for RhB was nearly 100%, which was achieved in 3 hours. The photocatalytic mechanism indicated that ·OH and h+ played an important role in the photocatalytic degradation reaction.


1. Introduction

Dye effluents from textile industries and photographic industries are becoming a serious environmental problem because of their toxicity, unacceptable color, high chemical oxygen demand, and biological degradation.1–3 Many advanced oxidation technologies, especially photocatalytic technology, were developed to treat industrial wastewater.4 Photocatalysis is a green technique for the degradation of various dyes such as rhodamine B (RhB), brilliant blue, thionine, and methylene orange.5–8 The most important point is that the semiconducting material, such as single semiconductor, semiconducting heterostructure, or solid solution, was utilized as the photocatalyst in this type of reaction.9–14 However, a majority of semiconductors, such as TiO2, Ca2PbGa8O15, and ZnO, only have a response under ultraviolet light irradiation.15–17 The ultraviolet spectrum and visible spectrum account for 4% and 46% of the solar spectrum, respectively.18 Hence, discovering a new photocatalyst for the photocatalytic degradation reaction is an event of significance.

Usually, the Bi-based oxides have sensitive activity under visible light. The famous BiVO4 material was extensively studied for visible-light photocatalytic applications such as oxygen evolution, water splitting, carbon dioxide reduction, and dye degradation.19–22 The other Bi-based oxides, such as Bi2O3, Bi2MoO6, Bi2WO6, BiTaO4, PbBi2Nb2O9, and CdBiYO4, were studied as visible-light-driven photocatalysts for photocatalytic applications as well.23–28 The visible-light response of these materials arises due to the presence of the Bi3+ lone pair of electrons. It was found that the bandgap shrinking was due to the contribution of Bi 6s orbitals to the valence band composition. However, the conduction bands of these Bi-based oxide compounds are more positive than 0 V, which means that they cannot be used for hydrogen evolution by themselves. Recently, Bi2Ga4O9 was employed as the photocatalyst for water splitting under visible-light irradiation.29 Therefore, it is interesting to investigate the visible-light photocatalytic degradation performance of Bi2Ga4O9.

To the best of our knowledge, there has been no report regarding the application of Bi2Ga4O9 in the visible-light driven degradation of dyes. Herein, the photocatalytic degradation performance of mullite type Bi2Ga4O9 was improved by doping Fe3+ ions, adding H2O2, and loading cocatalysts (such as Cu, Ag, Au, Pt, Ni, and Ru). RhB was used as the model organic dye to study the photocatalytic property of the as-prepared samples. In particular, the composition, morphology, physical property, and photocatalytic mechanism of Bi2Ga3.2Fe0.8O9 were systematically investigated.

2. Materials and methods

2.1 Preparation of the catalysts

The solid-state method was applied to synthesize Bi2Ga4−xFexO9 (0 ≤ x ≤ 1.2) samples. The starting materials, Bi2O3 (99.9%, Sinopharm Chemical Reagent Co., Ltd.), Ga2O3 (99.99%, Alfa Aesar), and Fe2O3 (99.5%, Sinopharm Chemical Reagent Co., Ltd.), were used after pre-calcination at 600 °C to remove the possible absorbed moisture or CO2. Typically, for the synthesis of Bi2Ga3.2Fe0.8O9, stoichiometric amounts of the reagents were homogenized using an agate mortar, followed by preheating at 600 °C for 10 h. The resultant powder was manually re-ground thoroughly. Finally, it was heated at 850 °C for another 15 h in air.

2.2 Preparation of the cocatalysts

The metal cocatalyst was loaded on Bi2Ga3.2Fe0.8O9 by a photo-deposition method.29 For instance, 100 mg Bi2Ga3.2Fe0.8O9 sample together with 7.8 mL of CuCl2 (0.32 mg mL−1) was mixed in 200 mL 2 vol% methanol aqueous solution. This solution, in a 300 mL Pyrex glass reactor, was stirred for 10 minutes and then the mixture was irradiated by a 300 W high-pressure Hg-lamp (CEL-LAM300, Beijing AuLight Ltd. Co., intensity of light is 49.5 mW cm−2) for 1 hour. After this, the sample was washed with distilled water and dried to be used as a photocatalyst. This sample was denoted as 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9.

2.3 Characterization

The crystal structure of the as-prepared sample was characterized by X-ray diffraction (XRD, XRD-6100AS, Cu Kα irradiation, Shimadzu, Japan). Le Bail refinements were performed to obtain the cell parameters using TOPAS software package. The morphology and size of the samples were characterized by scanning electron microscopy (SEM FEI verios 460) equipped with an energy dispersive spectrometer (EDS). The compositions of the Bi2Ga3.2Fe0.8O9 sample were analyzed with inductively coupled plasma-atomic emission spectroscopy (ICP-OES, PerkinElmer 8300). X-ray photoelectron spectroscopy (XPS) was carried out using a PHI5000 spectrometer with an Al Kα (15 kV) X-ray source. Ultraviolet-visible diffused reflectance spectra (DRS) were obtained on a Shimadzu UV-3600 spectrometer equipped with an integrating sphere attachment. The analysis range was from 200 to 1100 nm and BaSO4 was used as the reflectance standard. The number of incident photons was measured by using a calibrated Si photodiode (CEL-NP2000).

2.4 Photocatalytic performance evaluation

The photocatalytic performance was tested on the reaction equipment, which contains a 300 mL Pyrex glass reactor and a quartz cold trap. In a typical run, 100 mg of the catalyst was dispersed by a magnetic stirrer in 250 mL of 20 ppm RhB aqueous solution. The solution was continuously stirred and a self-made circulating cooling water system maintained the temperature of the reaction system at 25 ± 5 °C. The light irradiation source was generated by an external 300 W long-arc xenon lamp (CEL-LAX300, Beijing AuLight Ltd. Co., intensity of light is 38.5 mW cm−2).

These photocatalysts with 250 mL Rh (20 ppm) solution were stirred in dark for 30 min to achieve the adsorption–desorption equilibrium. After light illumination at regular time intervals, the concentration of the RhB solution was monitored with a Shimadzu UV2600 UV-Vis spectrophotometer at 10 minute intervals.

2.5 Determination of the reactive species

The photocatalytic mechanism of 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 was studied in accordance with part 2.4, but before the start of the reaction, some quenchers (1 mmol L−1) were added into the RhB solution, such as p-benzoquinone (BQ; ·O2 radical scavenger), disodium ethylenediaminetetraacetate (EDTA-2Na; h+ scavenger), and isopropanol (IPA; ·OH radical scavenger).

3. Results and discussion

Fig. 1 presents the XRD patterns of Bi2Ga4−xFexO9 (0 < x < 1.2) solid solutions and the sharp peaks indicate the high crystallinity of these samples. Compared to the simulated XRD pattern (ICSD-250413) of Bi2Ga4O9, it is evident that the pure phases of the Bi2Ga4−xFexO9 (0 ≤ x ≤ 1.2) solid solutions were synthesized without any impurity peak. As shown in the inserted red rectangle of Fig. 1, we selected the section of 2θ from 28° to 29.5° to observe the peak shift by Fe3+-doping. The peaks of the solid solutions shifted linearly to a smaller angle. This means that the Fe3+ ions were incorporated successfully into the crystal structure of Bi2Ga4O9.30 We can determine the change in the cell lattice parameters (a, b, c, and V) by Le Bail fitting of the whole XRD patterns. The plot of the crystallographic data for a, b, c, and V along with an increase in x suggests a linear swell (see Fig. S1).31 The crystal structure of Bi2Ga3.2Fe0.8O9 can be confirmed by TOPAS (see Fig. S2). These results strongly suggest that Fe3+ has been successfully doped into Bi2Ga4O9 without any structural change.
image file: c9ra04632a-f1.tif
Fig. 1 The powder XRD patterns for the solid solutions, where the simulated XRD pattern for the un-doped Bi2Ga4O9 is also given at the bottom. Also, the inserted red rectangle was magnified to the right of the XRD patterns.

The as-prepared Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9 were observed by electron microscopy. The SEM images of the Bi2Ga3.2Fe0.8O9 powders were composed of uneven sized particles in the range of 0.4–1 μm (see Fig. 2). The particle size of the Bi2Ga3.2Fe0.8O9 powder was remarkably smaller than that of the undoped Bi2Ga4O9 powder, which is in the range of 0.8–1.5 μm (see Fig. S3). Both the particles were well-crystallized, which is consistent with the XRD analysis. Elemental analysis was performed on the Bi2Ga3.2Fe0.8O9 sample, which gave an average atomic ratio of Bi[thin space (1/6-em)]:[thin space (1/6-em)]Ga[thin space (1/6-em)]:[thin space (1/6-em)]Fe[thin space (1/6-em)]:[thin space (1/6-em)]O = 2.00[thin space (1/6-em)]:[thin space (1/6-em)]3.35[thin space (1/6-em)]:[thin space (1/6-em)]0.83[thin space (1/6-em)]:[thin space (1/6-em)]9.18 (see Fig. S4). Also, this sample was analyzed by ICP, which gave a close metallic atomic ratio of Bi[thin space (1/6-em)]:[thin space (1/6-em)]Ga[thin space (1/6-em)]:[thin space (1/6-em)]Fe = 2[thin space (1/6-em)]:[thin space (1/6-em)]3.27[thin space (1/6-em)]:[thin space (1/6-em)]0.73. Besides, the XPS spectra for Bi2Ga3.2Fe0.8O9 indicated that not only the elemental compositions are in agreement with the EDS result but also support the fact that the Fe element is trivalent (see Fig. S5).


image file: c9ra04632a-f2.tif
Fig. 2 SEM images for Bi2Ga3.2Fe0.8O9.

Fig. 3 shows the DRS of the Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9 samples. The spectrum of Bi2Ga4O9 showed a typical bandgap transition similar to the other semiconductors, such as C3N4 and CuFeS2.32,33 The spectrum of Bi2Ga3.2Fe0.8O9 showed an obvious red shift and enhancement in comparison to the spectrum of Bi2Ga4O9 (see Fig. 3a). There is an absorption peak in the range of 625–750 nm, which is due to the d–d transition of the Fe3+ ion.29 For most semiconductors, the bandgap energy Eg can be calculated by the following equation: αhν = A(Eg)n/2, where h, v, and A are the Planck's constant, light frequency, and proportionality, respectively; n is determined on the basis of the transition type (i.e., n = 1 for direct transition, n = 4 for indirect transition).13 The best fit of (αhν)2 vs. Eg was obtained only when n is 1, suggesting that direct transitions across the energy bandgaps of Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9 are allowed (see Fig. 3b). The extrapolated value of at α = 0 gives an absorption edge energy corresponding to Eg, which are 2.99 eV and 2.59 eV for Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9, respectively. Besides, on comparing the absorption spectra of Bi2Ga4−xFexO9 with different Fe contents (0 < x < 1.2), the absorption edge appeared to show a “red shift” with increasing Fe3+ concentration (see Fig. S6).


image file: c9ra04632a-f3.tif
Fig. 3 (a) UV-Vis diffused reflectance spectrum for Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9. (b) The estimated bandgap energy Eg with a plot of (αhν)2 against the photon energy ().

The photocatalytic activities were evaluated for the Bi2Ga4−xFexO9 (0 < x < 1.2) samples using the catalytic degradation of RhB as a probe reaction under visible light irradiation. The absorption–desorption equilibrium was established by stirring the reaction solution mixed with the photocatalyst powder in a darkroom for 30 min before illumination. Fig. 4 displays the results for RhB degradation using the Bi2Ga4−xFexO9 (0 < x < 1.2) samples under the same reaction conditions. As shown in Fig. 4a, the absorbance of the RhB reaction solution decreased rapidly after the visible light irradiation. To quantitatively assess the photocatalytic performance of the solid solutions, the reaction kinetics of RhB degradation were calculated by the first-order model with the ln(C0/C) versus time plot (see Fig. 4b).34 The calculated k values of Bi2Ga4−xFexO9, where x = 0, 0.2, 0.4, 0.6, 0.8, 1.0, and 1.2, are 0.0022 min−1, 0.0026 min−1, 0.0030 min−1, 0.0053 min−1, 0.0091 min−1, 0.0082 min−1, and 0.0060 min−1, respectively. The k value of Bi2Ga3.2Fe0.8O9 is 4.1 times larger than that of the undoped Bi2Ga4O9. Also, the kinetic constant for RhB degradation catalyzed by the undoped sample is the smallest. The conversion efficiency of the solid solutions displayed that the photocatalytic activity of Bi2Ga3.2Fe0.8O9 is better than that of the undoped Bi2Ga4O9 as well (see Fig. 4c). The above results illustrate that doping a slight amount of Fe3+ into the Bi2Ga4O9 crystal structure not only makes the photocatalyst economical but also significantly promotes the photocatalytic performance.35 However, an excessive amount of Fe3+ plays the role of a recombination center for photo-generated electrons and holes.18,36 Owing to the two effects, the optimum Fe3+ amount was 20 wt%; hence, the Bi2Ga3.2Fe0.8O9 sample exhibits the highest photocatalytic performance among the solid solutions. Fig. 4d displays that RhB was photocatalytically oxidized in one hour using the Bi2Ga3.2Fe0.8O9 sample.


image file: c9ra04632a-f4.tif
Fig. 4 (a) Photocatalytic degradation of RhB using Bi2Ga4−xFexO9 solid solutions under visible light irradiation. (b) First-order kinetic constants for the degradation of RhB under visible light irradiation. (c) The conversion efficiency of RhB over the solid solutions under visible light irradiation. (d) The UV-Vis absorption spectra of RhB at different concentrations using Bi2Ga3.2Fe0.8O9 as the photocatalyst under visible light irradiation. Photocatalytic conditions: 100 mg photocatalyst, 250 mL solution, RhB 20 ppm, pH = 6, V(H2O2) = 0.5 mL, 300 W long-arc xenon lamp.

Fig. 5 presents the conversion efficiency of RhB over the Bi2Ga3.2Fe0.8O9 sample under visible light irradiation. Usually, H2O2 plays a key role in the degradation of various dyes.37,38 In this work, using the above optimum Bi2Ga3.2Fe0.8O9 sample without H2O2 displayed a very low conversion efficiency of RhB in the photocatalytic reaction. The conversion efficiency of RhB was significantly improved on increasing the usage amount of H2O2 by comparing the photocatalytic activity of Bi2Ga3.2Fe0.8O9 with 0 mL H2O2 (see Fig. 5a). Since excess H2O2 acts as a scavenger of ·OH and exhausts ·OH in the solution, an optimum amount of H2O2 is used.39 The optimum k value of Bi2Ga3.2Fe0.8O9 with 1.5 mL H2O2 is 0.0108 min−1 (see Fig. S7). Besides, the photodecomposition activity for RhB using only 1.5 mL H2O2 is better (see Fig. S8). Six kinds of cocatalysts were loaded on the Bi2Ga3.2Fe0.8O9 sample to improve its photocatalytic performance. Generally, the metal cocatalyst played the role of an active site, which is beneficial for the separation of photo-generated electrons and holes.40,41 However, only the Cu and Ag cocatalysts have a positive influence on the photocatalytic activity in our experiment (see Fig. 5b). Also, the optimum k value of Bi2Ga3.2Fe0.8O9 with 1 wt% Cu cocatalyst is 0.0131 min−1 (see Fig. S9). The possible reason will be discussed in the photocatalytic mechanism section.


image file: c9ra04632a-f5.tif
Fig. 5 The conversion efficiency of RhB over the Bi2Ga3.2Fe0.8O9 sample under visible light irradiation: (a) addition of different volumes of H2O2 in 250 mL solution; (b) loading 1 wt% of different cocatalysts on the photocatalyst. Photocatalytic conditions: 100 mg photocatalyst, 250 mL solution, RhB 20 ppm, pH = 6, 300 W long-arc xenon lamp.

The usage amount of Cu cocatalyst was tested for the photocatalytic degradation of RhB under visible light. The 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 displays the optimum photocatalytic activity (see Fig. 6a). Also, the corresponding optimum k value is 0.0175 min−1 (see Fig. S10). The negative effect of excessive Cu may be caused by the interfacial charge recombination.42 Other than photocatalytic efficiency, the stability and reusability of the photocatalyst is also an important factor to evaluate the property of the catalyst.43 Therefore, the stability of the best performing 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 sample by photocatalytic degradation of RhB for four cycles under the same conditions was studied. Specifically, each cycle consisted of 60 min of photocatalytic reaction, followed by the separation of the photocatalyst by centrifugation, then washing with distilled water, and finally drying of the photocatalyst in an oven at 60 °C. The change in the normalized concentration of RhB (C/C0) over four cycles is displayed in Fig. 6b. The photocatalytic conversion efficiency of 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 sample decreased slightly after each cycle, such that 65% and 50% of RhB was degraded after the first and fourth cycles, respectively. The reduction in the photocatalytic activity is attributable to the inevitable loss of the photocatalyst during the separation and cleaning steps.6 This demonstrates that the sample is relatively photo-stable in the photocatalytic reaction. Also, the XRD patterns of the photocatalysts after the photocatalytic reaction display the photo-stability of these samples (see Fig. S11). Fig. 6c and d display that nearly 100% RhB was completely degraded in 3 hours.


image file: c9ra04632a-f6.tif
Fig. 6 (a) The conversion efficiency of RhB using Bi2Ga3.2Fe0.8O9 with different amounts of Cu-cocatalyst under visible light irradiation. (b) Cyclic photocatalytic degradation of RhB using 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 under visible light irradiation. (c) The UV-Vis absorption spectra of RhB with different concentrations using 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 as the photocatalyst under visible light irradiation. (d) Photocatalytic degradation of RhB using 0.5 wt% Cu/Bi2Ga3.2Fe0.8O9 under visible light irradiation. Photocatalytic conditions: 100 mg photocatalyst, 250 mL solution, RhB 20 ppm, pH = 6, V(H2O2) = 1.5 mL, 300 W long-arc xenon lamp.

The photocatalytic mechanism for the 1.5 wt% Cu/Bi2Ga3.2Fe0.8O9 sample was tested. We employ electronegativity to estimate the energy potential of the conduction band (CB) and valence band (VB) semi-quantitatively. The applied equations are EVB = χEe + 1/2Eg and ECB = EVBEg, where, χ is the Mulliken electronegativity; Ee is the energy of free electrons on the hydrogen scale, which is 4.50 eV.29 Eg is the observed bandgap energy for Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9, which are 2.99 eV and 2.59 eV from DRS, respectively. χ for Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9 can be calculated by their molecular formulas, which were found to be 5.63 eV and 5.70 eV, respectively. So, the ECB and EVB of Bi2Ga4O9 are −0.37 V and 2.62 V, respectively. The ECB and EVB of Bi2Ga3.2Fe0.8O9 are −0.09 V and 2.50 V, respectively. It can be clearly seen that the bandgap of Bi2Ga3.2Fe0.8O9 is narrower than that of Bi2Ga4O9, and the absolute values of the CB and VB potentials for Bi2Ga3.2Fe0.8O9 are smaller than those of Bi2Ga4O9. This shows that the photocatalytic activity of Bi2Ga3.2Fe0.8O9 is better than that of Bi2Ga4O9. As we know, the photocatalytic active species, including h+, ·OH, and ·O2, play a very important role in the photocatalytic degradation reaction.44,45 Therefore, the relative redox potentials of O2/·O2, OH/·OH, and H2O/·OH are displayed in Fig. 7a.43 We can conclude that ·O2 has a slight influence over the Bi2Ga3.2Fe0.8O9 sample in the photocatalytic reaction. EDTA-2Na, IPA, and BQ were used as the quenchers for h+, ·OH, and ·O2, respectively (see Fig. 7b). It can be seen that the photocatalytic performance was slightly affected by adding BQ and the photo-degradation rate was 50.6%, indicating that ·O2 played a small role in the photocatalytic degradation reaction. In contrast, compared with that without a quencher (64.6%), the photo-degradation rate reduced to 18.7% and 7.3%, respectively, which demonstrated that ·OH and h+ played an important role in the photocatalytic degradation reaction. All the relevant reactions are described as follows:

Bi2Ga3.2Fe0.8O9 + → Bi2Ga3.2Fe0.8O9(e + h+)

e(Bi2Ga3.2Fe0.8O9) + H+ → 1/2H2

h+(Bi2Ga3.2Fe0.8O9) + OH → ·OH

h+(Bi2Ga3.2Fe0.8O9) + H2O → ·OH + H+

Fe3+ + e(Bi2Ga3.2Fe0.8O9) → Fe2+

Fe2+ + H2O2 → Fe3+ + ·OH + OH

H2O → H+ + OH

·OH + RhB → product

h+(Bi2Ga3.2Fe0.8O9) + RhB → product


image file: c9ra04632a-f7.tif
Fig. 7 (a) Schematic illustration of the band structure diagram of Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9. (b) The photocatalytic conversion efficiency of 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 for the degradation of RhB in the presence of different scavengers (1 mmol L−1) under visible light irradiation.

These processes can be used to explain the improvement in the photocatalytic activity of Bi2Ga3.2Fe0.8O9 using Cu and Ag. The active species other than the ·O2 play an important role in the photocatalytic degradation reaction. This means that the photo-generated electron plays a small role in the photocatalytic reaction. Owing to the presence of methanol in the photo-deposition process, these cocatalysts mainly tend to be metals on the surface of the photocatalyst. As we know, the introduced metal mostly functions as the electron acceptor to manifest electron accumulation at the metal and hole localization at the semiconductor.46,47 Besides, the cocatalyst has an influence on the reduction of the chemical barrier. Hence, with the assistance of the cocatalyst, the photo-generated electron was used to form H2 and even ·O2. However, RhB is a cationic dye, which is not easily absorbed on the surface of the photocatalyst. On the contrary, without the cocatalyst, the photo-generated electron can hardly produce H2. Hence, the electron is mainly transferred on the RhB, which leads to a relatively large blank (see Fig. 5b). According this opinion, the cocatalyst has more positive influence on the reducing capacity, which decreases the decomposition of RhB. Usually, Au, Ni, Pd, and Pt cocatalysts are good for reduction reactions, such as H2 production.18

4. Conclusions

We prepared Bi2Ga4−xFexO9 (0.0 ≤ x ≤ 1.2) solid solutions by high-temperature solid-state reaction. Le Bail fitting on powder XRD verified the high purity and crystallinity of the as-prepared samples, and the SEM images show that the Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9 samples are composed of micron and submicron crystallites, respectively. EDS, ICP, and XPS analyses indicated that Fe element is trivalent when incorporated into the Bi2Ga4O9 crystal structure. DRS of Bi2Ga4O9 and Bi2Ga3.2Fe0.8O9 samples suggest wide bandgap characteristics and the observed bandgaps are 2.99 eV and 2.59 eV, respectively, assuming the direct semiconductor model. The photocatalytic degradation performance of Bi2Ga4O9 was improved by doping Fe3+, adding H2O2, and loading a cocatalyst. The optimum photocatalytic performance was obtained using 2.5 wt% Cu/Bi2Ga3.2Fe0.8O9 sample in the presence of 1.5 mL H2O2 under visible light irradiation, whose photo-degradation rate is 8.0 times that of Bi2Ga4O9 in the presence of 0.5 mL H2O2. This photocatalyst remained stable and active even after four cycles (4 hours in total). The photocatalytic mechanism was suggested as well; the ·OH and h+ played important roles in the photocatalytic degradation reaction.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was financially supported by the Science and Technology Project of Chongqing Municipal Education Commission (KJQN201801407), Talent Introduction Project of Yangtze Normal University (2017KYQD22).

References

  1. W. L. Wang, Y. Z. Cai, H. Y. Hu, J. Chen, J. Wang, G. Xue and Q. Y. Wu, Chem. Eng. J., 2019, 359, 168–175 CrossRef CAS.
  2. J. Y. Liang, X. A. Ning, J. Sun, J. Song, J. Lu, H. L. Cai and Y. X. Hong, Ecotoxicol. Environ. Saf., 2018, 166, 56–62 CrossRef CAS PubMed.
  3. M. Ahmad, M. Yousaf, A. Nasir, I. A. Bhatti, A. Mahmood, X. C. Fang, X. Jian, K. Kalantar-Zadeh and N. Mahmood, Environ. Sci. Technol., 2019, 53, 2161–2170 CrossRef CAS PubMed.
  4. L. Clarizia, D. Russoa, I. D. Somma, R. Marottaa and R. Andreozzia, Appl. Catal., B, 2017, 209, 358–371 CrossRef CAS.
  5. Y. C. Chen, T. C. Liu and Y. J. Hsu, ACS Appl. Mater. Interfaces, 2015, 7(3), 1616–1623 CrossRef CAS PubMed.
  6. Y. Wu, H. Wang, W. G. Tu, Y. Liu, Y. Z. Tan, X. Z. Yuan and J. W. Chew, J. Hazard. Mater., 2018, 347, 412–422 CrossRef CAS PubMed.
  7. W. T. Chen and Y. J. Hsu, Langmuir, 2010, 26(8), 5918–5925 CrossRef CAS PubMed.
  8. Y. C. Pu, H. Y. Chou, W. S. Kuo, K. H. Wei and Y. J. Hsu, Appl. Catal., B, 2017, 204, 21–32 CrossRef CAS.
  9. A. Fujishima and K. Honda, Nature, 1972, 238, 37–38 CrossRef CAS PubMed.
  10. Y. F. Lin and Y. J. Hsu, Appl. Catal., B, 2013, 130–131, 93–98 CrossRef CAS.
  11. Y. C. Pu, W. H. Lin and Y. J. Hsu, Appl. Catal., B, 2015, 163, 343–351 CrossRef CAS.
  12. Y. H. Chiu, T. F. M. Chang, C. Y. Chen, M. Sone and Y. J. Hsu, Catalysts, 2019, 9, 430 CrossRef CAS.
  13. J. Yang, X. R. Sun, T. Zeng, Y. L. Hu and J. W. Shi, Materials, 2019, 12(9), 1487 CrossRef PubMed.
  14. J. Yang, H. Fu, D. F. Yang, W. L. Gao, R. H. Cong and T. Yang, Inorg. Chem., 2015, 54, 2467–2473 CrossRef CAS PubMed.
  15. Y. H. Chiu and Y. J. Hsu, Nano Energy, 2017, 31, 286–295 CrossRef CAS.
  16. M. Y. Chen and Y. J. Hsu, Nanoscale, 2013, 5, 363–368 RSC.
  17. P. F. Jiang, F. W. Jiang, M. F. Yue, J. Ju, C. L. Xu, R. H. Cong and T. Yang, Angew. Chem., Int. Ed., 2019, 58(18), 5978–5982 CrossRef CAS PubMed.
  18. X. B. Chen, S. H. Shen, L. J. Guo and S. S. Mao, Chem. Rev., 2010, 110, 6503–6570 CrossRef CAS PubMed.
  19. R. G. Li, F. X. Zhang, D. G. Wang, J. X. Yang, M. R. Li, J. Zhu, X. Zhou, H. X. Han and C. Li, Nat. Commun., 2012, 4, 1432–1438 CrossRef PubMed.
  20. Q. X. Jia, A. Iwase and A. Kudo, Chem. Sci., 2014, 5, 1513–1519 RSC.
  21. Q. J. Shi, Z. J. Li, L. Chen, X. L. Zhang, W. H. Han, M. Z. Xie, J. L. Yang and L. Q. Jing, Appl. Catal., B, 2019, 244, 541–649 CrossRef.
  22. M. B. Tahir, T. Lqbal, H. Kiran and A. Hasan, Int. J. Energy Res., 2019, 43(6), 2410–2417 CrossRef CAS.
  23. H. Sudrajat and S. Hartuti, Adv. Powder Technol., 2019, 30(5), 983–991 CrossRef CAS.
  24. F. Fu, H. D. Shen, X. Sun, W. W. Xue, A. Shineye, J. N. Ma, L. Luo, D. J. Wang, J. G. Wang and J. W. Tang, Appl. Catal., B, 2019, 247, 150–162 CrossRef CAS.
  25. X. J. Yuan, D. Y. Shen, Q. Zhang, H. B. Zou, Z. L. Liu and F. Peng, Chem. Eng. J., 2019, 369, 292–301 CrossRef CAS.
  26. Y. D. Hu, G. Chen, C. M. Li, Y. S. Zhou, J. X. Sun, S. Hao and Z. H. Han, J. Mater. Chem. A, 2016, 4(14), 5274–5281 RSC.
  27. C. H. Lee, H. G. Kim, Y. Gu and D. H. Lim, Nanosci. Nanotechnol. Lett., 2018, 10, 1179–1186 CrossRef.
  28. H. Y. Du and J. F. Luan, Solid State Sci., 2012, 14(9), 1295–1305 CrossRef CAS.
  29. J. Yang, P. F. Jiang, M. F. Yue, D. F. Yang, R. H. Cong, W. L. Gao and T. Yang, J. Catal., 2017, 345, 236–244 CrossRef CAS.
  30. S. X. Ouyang and J. H. Ye, J. Am. Chem. Soc., 2011, 133, 7757–7763 CrossRef CAS PubMed.
  31. Z. G. Zou, J. H. Ye, K. Sayama and H. Arakawa, Nature, 2001, 414, 625–627 CrossRef CAS PubMed.
  32. X. C. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen and M. Antonietti, Nat. Mater., 2009, 8, 76–80 CrossRef CAS PubMed.
  33. J. Yang, M. F. Yue, J. Ju, R. H. Cong, W. L. Gao and T. Yang, Dalton Trans., 2014, 43, 15385–15390 RSC.
  34. H. X. Yang, B. Q. Shan and L. Zhang, RSC Adv., 2014, 4, 61226–61231 RSC.
  35. J. R. Ran, J. Zhang, J. G. Yu, M. Jaroniec and S. Z. Qiao, Chem. Soc. Rev., 2014, 43, 7787–7812 RSC.
  36. Z. G. Zou, J. H. Ye and H. Arakawa, Catal. Lett., 2001, 75, 3–4 CrossRef.
  37. J. Singh, S. Sharma, Aanchal and S. Basu, J. Photochem. Photobiol., A, 2019, 376, 32–42 CrossRef CAS.
  38. M. Mahanthappa, N. Kottam and S. Yellappa, Appl. Surf. Sci., 2019, 475, 828–838 CrossRef CAS.
  39. L. Zhang, B. Q. Shan, H. X. Yang, D. S. Wu, R. Zhu, J. H. Nie and R. Cao, RSC Adv., 2015, 5, 23556–23562 RSC.
  40. R. B. Wei, Z. L. Huang, G. H. Gu, Z. Wang, L. X. Zeng, Y. B. Chen and Z. Q. Liu, Appl. Catal., B, 2018, 231, 101–107 CrossRef CAS.
  41. K. Z. Qi, B. Cheng, J. G. Yu and W. K. Ho, Chin. J. Catal., 2017, 38(12), 1936–1955 CrossRef CAS.
  42. Y. C. Chen, Y. C. Pu and Y. J. Hsu, J. Phys. Chem. C, 2012, 116(4), 2967–2975 CrossRef CAS.
  43. Y. Y. Shang, X. Chen, W. W. Liu, P. F. Tan, H. Y. Chen, L. D. Wu, C. Ma, X. Xiong and J. Pan, Appl. Catal., B, 2017, 204, 78–88 CrossRef CAS.
  44. Z. Wan, G. K. Zhang, X. Y. Wu and S. Yin, Appl. Catal., B, 2017, 207, 17–26 CrossRef CAS.
  45. F. Chen, Q. Yang, X. M. Li, G. M. Zeng, D. B. Wang, C. G. Niu, J. W. Zhao, H. X. An, T. Xie and Y. C. Deng, Appl. Catal., B, 2017, 200, 330–342 CrossRef CAS.
  46. W. H. Lin, Y. H. Chiu, P. W. Shao and Y. J. Hsu, ACS Appl. Mater. Interfaces, 2016, 8(48), 32754–32763 CrossRef CAS PubMed.
  47. T. T. Yang, W. T. Chen, Y. J. Hsu, K. H. Wei, T. Y. Lin and T. W. Lin, J. Phys. Chem. C, 2010, 114, 11414–11420 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c9ra04632a

This journal is © The Royal Society of Chemistry 2019
Click here to see how this site uses Cookies. View our privacy policy here.