Next Article in Journal
Room Temperature Electroluminescence from Tensile-Strained Si0.13Ge0.87/Ge Multiple Quantum Wells on a Ge Virtual Substrate
Previous Article in Journal
Inclined Fiber Pullout from a Cementitious Matrix: A Numerical Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Structural, Photocatalytic Property Characterization and Enhanced Photocatalytic Activities of Novel Photocatalysts Bi2GaSbO7 and Bi2InSbO7 during Visible Light Irradiation

1
State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210093, China
2
Department of Chemical Engineering, Technion-Israel Institute of Technology, Haifa 32000, Israel
*
Author to whom correspondence should be addressed.
Materials 2016, 9(10), 801; https://doi.org/10.3390/ma9100801
Submission received: 29 July 2016 / Revised: 18 September 2016 / Accepted: 21 September 2016 / Published: 27 September 2016
(This article belongs to the Section Advanced Materials Characterization)

Abstract

:
In order to develop original and efficient visible light response photocatalysts for degrading organic pollutants in wastewater, new photocatalysts Bi2GaSbO7 and Bi2InSbO7 were firstly synthesized by a solid-state reaction method and their chemical, physical and structural properties were characterized. Bi2GaSbO7 and Bi2InSbO7 were crystallized with a pyrochlore-type structure and the lattice parameter of Bi2GaSbO7 or Bi2InSbO7 was 10.356497 Å or 10.666031 Å. The band gap of Bi2GaSbO7 or Bi2InSbO7 was estimated to be 2.59 eV or 2.54 eV. Compared with nitrogen doped TiO2, Bi2GaSbO7 and Bi2InSbO7, both showed excellent photocatalytic activities for degrading methylene blue during visible light irradiation due to their narrower band gaps and higher crystallization perfection. Bi2GaSbO7 showed higher catalytic activity compared with Bi2InSbO7. The photocatalytic degradation of methylene blue followed by the first-order reaction kinetics and the first-order rate constant was 0.01470 min−1, 0.00967 min−1 or 0.00259 min−1 with Bi2GaSbO7, Bi2InSbO7 or nitrogen doped TiO2 as a catalyst. The evolution of CO2 and the removal of total organic carbon were successfully measured and these results indicated continuous mineralization of methylene blue during the photocatalytic process. The possible degradation scheme and pathway of methylene blue was also analyzed. Bi2GaSbO7 and Bi2InSbO7 photocatalysts both had great potential to purify textile industry wastewater.

Graphical Abstract

1. Introduction

Dye contaminants from textile wastewater were difficult to treat for their high chroma, high chemical oxygen demand content and complicated ingredients. Some conventional methods including biodegradation [1,2,3], electrochemistry [3,4,5,6], adsorption [7,8,9], and flocculation−precipitation [10,11] had been exploited to degrade those dye contaminates, but there still existed a serious of problems with them. Methylene blue (MB), usually adopted as dyestuff, was one of the most common dye contaminants.
Photocatalysis had gained great development since photocatalytic reaction was found in 1972 [12]. Photocatalytic degradation of the pollutants in wastewater entailed a chain of advantages including conserving energy and little secondary pollution; it had therefore gradually attracted more and more attention in textile wastewater treatment. Metal oxides [13,14,15,16,17,18,19,20,21,22,23,24] and metal sulfides [21,22,23,24,25,26,27,28,29,30,31,32,33] were the most common semiconductor photocatalysts. Among metal oxides, anatase TiO2 was investigated most repeatedly owing to its non-toxic property, excellent stability and low cost. However, with a wider band gap (3.2 eV), anatase TiO2 only efficiently absorbed ultraviolet light which occupied only 5% of the solar energy, and thus failed to make good use of optical energy. In order to make the best use of visible light which occupied 43% of sunlight, developing visible light responsive photocatalysts was an inevitable tendency in the field of photocatalysis research, which could be embodied from abundant endeavors of previous scholars in realizing the degradation of the pollutants during visible light irradiation by the method of iron doping [34,35,36], forming heterojunction [37,38,39,40,41,42] or photosensitization [43,44,45,46,47,48]. Several years ago, Zou and Arakawa [49,50] found that two types of metal oxides, ABO4 and A2B2O7, had great potential for photocatalytic H2 production during visible light irradiation. It was well known that minute changes in internal structure of the semiconductor photocatalysts would presumably promote the separation of photogenerated electrons and holes and thus improve photocatalytic activities. Zou et al. synthetized Bi2MNbO7 (M = Al, Ga, In, Y or Fe) [51,52,53] which was one remarkable representative of the family of A2B2O7 compounds with the A3+2B4+2O7 pyrochlore structure by substituting B4+ sites in A3+2B4+2O7 for M3+ (M3+ = Al3+, Ga3+, In3+) and Nb5+. Similarly, previous studies had reported Bi2GaVO7 [54] and Bi2SbVO7 [55] by element doping, which had realized visible-light photocatalytic degradation and H2 production. Previous works indicated that the Ga3+ and In3+ ions could influence the band gap and the electronic structure of the compound photocatalysts, which was expected to cause the different photocatalytic activity [56,57].
As an important element with higher electron drift velocity and mobility, antimony (Sb) has been extensively studied as a good dopant candidate for enhancing the electron transfer rate of semiconductors [58]. Omidi et al. [59] evaluated the photocatalytic activity of Sb-doped ZnO nanostructures (0 ≤ mol fraction of Sb3+ ions ≤ 0.15) for the photodegradation of MB. In addition, the acquired results showed that doping the ZnO nanostructures with 0.03 mol fraction of Sb3+ ions increased the reaction rate by about three times, indicating that the decreasing recombination of charge carriers could enhance the photocatalytic activity. Al-Hamdi et al. [60] reported that Sb-doped dioxide (SnO2) nanoparticles with different Sb concentrations (at % = 0, 2, 4 and 6), which was prepared by a sol–gel method, could degrade 12%, 45%, 71% and 97% of phenol in the mineralization process under UV irradiation for 120 min, which showed higher photocatalytic activity than the undoped SnO2 catalyst. These previous reports have shown that moderate Sb doped on the photocatalysts could greatly enhance the photocatalytic activity.
In this paper, new photocatalysts, Bi2GaSbO7 and Bi2InSbO7, were synthetized by doping element Ga or In with a solid-state reaction method. Meanwhile, the structural properties of Bi2GaSbO7 and Bi2InSbO7 were also characterized and their photocatalytic activities were also examined in degrading MB solution compared with N-doped TiO2, which had achieved the visible light response.

2. Materials and Methods

2.1. Synthesis of Bi2GaSbO7, Bi2InSbO7 and N-doped TiO2 Photocatalysts

New Bi2GaSbO7 and Bi2InSbO7 samples were firstly synthesized by a solid-state reaction method. Firstly, for the sake of the synthesis of Bi2GaSbO7, Bi2O3, Ga2O3 and Sb2O5 with a purity of 99.99% (Sinopharm Group Chemical Reagent Co., Ltd., Shanghai, China) were obtained by an atomic ratio of 2:1:1 to serve as raw materials. All powders were dried at 200 °C for 4 h before synthesis. In order to synthesize Bi2GaSbO7, the precursors were fully mingled with each other, then pressed into small columns and put into an alumina crucible (Shenyang Crucible Co., Ltd., Shenyang, China). Eventually, calcination was performed at 1100 °C for 40 h in an electric furnace (KSL 1700X, Hefei Kejing Materials Technology Co., Ltd., Hefei, China). Accordingly, Bi2O3, In2O3 and Sb2O5 with a purity of 99.99% (Sinopharm Group Chemical Reagent Co., Ltd., Shanghai, China) were obtained by an atomic ratio of 2:1:1 for the preparation of Bi2InSbO7. The synthesization procedure of Bi2InSbO7 was similar to that of Bi2GaSbO7, just the calcination was performed at 1070 °C for 30 h during mixed powder in the alumina crucible. The preparation of N-doped TiO2 was by the sol–gel method which was mentioned in our previous studies [61].

2.2. Characterization

In our paper, we adopted the X-ray diffraction method (XRD, D/MAX-RB, Rigaku Corporation, Tokyo, Japan) with Cu Ka radiation (λ = 1.54056 angstrom) to confirm the crystal structures of Bi2GaSbO7 and Bi2InSbO7. The patterns of Bi2GaSbO7 and Bi2InSbO7 were recorded at 295 K with a step–scan procedure in the range of 2θ = 10°–100° (for Bi2GaSbO7) or 10°–95° (for Bi2InSbO7). The step interval was 0.02° and the time per step was 1 s. The transmission electron microscopy (TEM, Tecnal F20 S-Twin, FEI Corporation, Hillsboro, OR, USA) was used to observe the surface state and structure of the photocatalysts. The Malvern’s mastersize-2000 particle size analyzer (Malvern Instruments Ltd., Malvern, UK) was utilized to measure the particle size of the photocatalysts. We also utilized X-ray photoelectron spectroscopy (XPS, ESCALABMK-2, VG Scientific Ltd., London, UK) to determine the Bi3+ content, Ga3+ content, Sb5+ content, In3+ content and O2− content of Bi2GaSbO7 and Bi2InSbO7. The chemical composition of Bi2GaSbO7 and Bi2InSbO7 was determined by scanning the electron microscope-X-ray energy dispersion spectrum (SEM–EDS, LEO 1530VP, LEO Corporation, Dresden, Germany). The surface areas of Bi2GaSbO7 and Bi2InSbO7 were measured by the Brunauere–Emmette–Teller (BET) method (MS-21, Quantachrome Instruments Corporation, Boynton Beach, FL, USA) with N2 adsorption at liquid nitrogen temperature. Their diffuse reflectance spectrums were analyzed by a UV-visible spectrophotometer (Shimadzu UV-2550 UV-Visible spectrometer, Kyoto, Japan).

2.3. Photocatalytic Properties Test

MB (C16H18ClN3S) (Tianjin Bodi Chemical Co., Ltd., Tianjin, China) served as our objective pollutant. The whole photocatalytic activity process was as follows: firstly, we prepared 300 mL MB aqueous solution in quartz tubes whose initial concentration was 0.025 mmol·L−1 and initial PH value was 7.0. Then, 0.8 g photocatalyst powder of N-doped TiO2, Bi2GaSbO7 or Bi2InSbO7 was placed into every quartz tube, respectively. In order to ensure the establishment of an adsorption/desorption equilibrium among photocatalysts, the MB dye and atmospheric oxygen, above per solution was magnetically stirred in the dark for 45 min. In our paper, we employed a 500 W Xenon lamp (λ > 420 nm), which utilized a 420 nm cutoff filter as a visible-light source. The photoreaction was carried out in a photochemical reaction apparatus (Nanjing Xujiang Machine Plant, Nanjing, China). During visible light illumination, the MB dye pollution was stirred by a magnetic stirrer and the photocatalyst powder was kept suspended in the solution. The filtrate was subsequently measured by a Shimadzu UV-2450 UV-visible spectrometer (Kyoto, Japan) with the detecting wavelength at 665 nm. The identification of MB and the degradation intermediate products of MB were measured by a liquid chromatograph-mass spectrometer (LC–MS, Thermo Quest LCQ Duo, Silicon Valley, CA, USA, Beta Basic-C18 HPLC column: 150 × 2.1 mm2, ID of 5 μm, Finnigan, Thermo, Silicon Valley, CA, USA). Here, post-photocatalysis solution (20 μL) was injected automatically into the LC–MS system. The eluent contained 60% methanol and 40% water, and the flow rate was 0.2 mL·min−1. MS conditions included an electrospray ionization interface, a capillary temperature of 27 °C with a voltage of 19.00 V, a spray voltage of 5000 V and a constant sheath gas flow rate. The spectrum was acquired in the negative ion scan mode, sweeping the m/z range from 50 to 600. Evolution of CO2 was analyzed with an intersmat™ IGC120-MB gas chromatograph (6890 N, Agilent Technologies, Palo Alto, CA, USA) equipped with a porapack Q column (3 m in length and with an inner diameter of 0.25 in.), which was connected to a catharometer detector. The total organic carbon (TOC) concentration was determined with a TOC analyzer (TOC-5000, Shimadzu Corporation, Kyoto, Japan). The photonic efficiency was calculated according to the following equation [62,63]:
ξ = R / I 0
where ξ was the photonic efficiency (%), and R was the rate of MB degradation (mol·L−1·s−1), and I0 was the incident photon flux (Einstein·L−1·s−1). The incident photon flux I0 which was measured by a radiometer (Model FZ-A, Photoelectric Instrument Factory Beijing Normal University, Beijing, China) was determined to be 4.76 × 106 Einstein·L−1·s−1 under visible light irradiation (wavelength range of 400–700 nm).

3. Results and Discussion

3.1. Characterization

Figure 1a,b shows the TEM images of Bi2GaSbO7 and Bi2InSbO7 with high magnification. We could observe from the images of Bi2GaSbO7 and Bi2InSbO7 that their particles presented a similar oblate spheroid appearance and that their distribution was relatively uniform. The average particle size of Bi2GaSbO7 approached 190 nm, which was smaller than that of Bi2InSbO7, whose average particle size approached 390 nm. We could observe from the BET results that the specific surface area of Bi2GaSbO7 approached 2.36 m2/g, which was bigger than that of Bi2InSbO7, whose specific surface area approached 1.82 m2/g. It was clear that the BET results were consistent with the TEM results, indicating that the samples with small average particle size would have a higher specific surface area. Figure 2a,b shows the SEM–EDS spectra taken from Bi2GaSbO7 and Bi2InSbO7. It could be seen from Figure 2a,b that the superfluous peaks did not exist in the spectra of Bi2GaSbO7 and Bi2InSbO7, meaning that Bi2GaSbO7 and Bi2InSbO7 crystals were both pure phase without impure elements.
In this paper, X-ray photoelectron spectroscopy analysis techniques were utilized to reveal the surface chemical compositions and the valence states of various elements in Bi2GaSbO7 and Bi2InSbO7. The various elemental peaks which are corresponding to specific binding energies are given in Table 1. Analysis results of the full XPS spectra were as follows: the prepared Bi2GaSbO7 sample contained Bi, Ga, Sb and O elements. Similarly, the prepared Bi2InSbO7 sample contained Bi, In, Sb and O elements. These results also uncovered that Bi2GaSbO7 crystal or Bi2InSbO7 crystal were both at a high pure phase. Moreover, the analysis results of the XPS spectra also manifested that the valence of Bi, Ga, Sb, In or O from Bi2GaSbO7 and Bi2InSbO7 was +3, +3, +5, +3 or −2, respectively. Eventually, according to our comprehensive XPS and SEM–EDS analyses, as for Bi2GaSbO7, the mean atomic ratio of Bi, Ga, Sb and O was 2.00:0.98:1.02:6.98. As for Bi2InSbO7, the mean atomic ratio of Bi, In, Sb and O was 2.00:0.99:1.01:6.99.
Figure 3 presents the X-ray powder diffraction patterns of Bi2GaSbO7 and Bi2InSbO7, respectively. We could judge from Figure 3 that Bi2GaSbO7 crystal or Bi2InSbO7 crystal was single phase. Figure 4a,b shows the Pawley refinement results of XRD data for Bi2GaSbO7 and Bi2InSbO7. The refined outcomes from Figure 4a,b displayed that the actual intensities of Bi2GaSbO7 or Bi2InSbO7 were both highly in accordance with the intensities of the pyrochlore-type structure with a cubic crystal system and a space group Fd3m (O atoms were included in the model), indicating that Bi2GaSbO7 and Bi2InSbO7 indeed formed the same crystal structure. The atomic coordinates and structural parameters of Bi2GaSbO7 and Bi2InSbO7 are listed in Table 2 and Table 3, respectively. Above results showed that the lattice parameter a of Bi2GaSbO7 was 10.356497 Å, which was slightly lower than that of Bi2InSbO7 whose lattice parameter a was 10.666031 Å. From the SEM–EDS spectra and XPS spectra which were taken from Bi2GaSbO7 and Bi2InSbO7, we had known that Bi2GaSbO7 crystal or Bi2InSbO7 crystal was both pure phase. Therefore, excluding the effects of impurities, we could deduce that the difference between the lattice parameter a for Bi2GaSbO7 and Bi2InSbO7 was perhaps concerned with M ionic radii which belonged to Bi2MSbO7. The reason was that the ionic radii of Ga3+ (0.62 Å) was minutely lower than that of In3+ (0.92 Å). Lastly, all the diffraction peaks (222), (400), (440), (622), (444), (800), (662), (840), (844) for Bi2GaSbO7 and Bi2InSbO7 were successfully indexed according to the lattice constant and above space group.
Figure 5 presents the diffuse reflection spectra of Bi2GaSbO7 and Bi2InSbO7, respectively. Compared with N-doped TiO2 whose absorption edge was about 445 nm, the absorption spectrum of newly prepared photocatalyst Bi2GaSbO7 or Bi2InSbO7 was estimated to be 480 nm or 490 nm, respectively, implicating that they had sizable potential to realize visible light response. The maximum absorption wavelength of MB was detected by an ultraviolet spectrophotometer, while the diffuse reflection spectra of Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 was detected by ultraviolet spectrophotometer with integrating sphere. In addition, the above two testing methods were totally different. Furthermore, the absorbance was obtained from the reflectance data and scattering should also be taken into consideration in data conversion from reflectance into absorbance, which was the reason why the ordinate of the diffuse reflection spectra in Figure 5 was absorbance.
We realized that absorbance could not be proportional to 1-transmission, thus the absorbance was calculated using the Kubelka–Munk transformation method in our experiment. For a crystalline semiconductor compound, the optical absorption near the band edge followed the equation [64,65]:
αhν = A × (Eg)n
Here, A, α, Eg and ν denoted proportional constant, absorption coefficient, band gap and light frequency, respectively. In this equation, n determined the character of the transition in a semiconductor compound. Eg and n could be calculated by the following steps: (i) plotting ln(αhν) versus ln(Eg) assuming an approximate value of Eg; (ii) deducing the value of n according to the slope in this graph; (iii) refining the value of Eg by plotting (αhν)1/n versus and extrapolating the plot to (αhν)1/n = 0. According to this method, we first estimated that the value of n for Bi2GaSbO7 or Bi2InSbO7 was 2, indicating that the optical transition for Bi2GaSbO7 or Bi2InSbO7 is indirectly allowed. Figure 6 presents the plot of (αhν)1/n versus for Bi2GaSbO7 and Bi2InSbO7. It could be found that the value of Eg for Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 was calculated to be 2.59 eV, 2.54 eV or 2.78 eV.

3.2. Photocatalytic Properties of Bi2GaSbO7 and Bi2InSbO7 Photocatalysts

From the UV-vis spectra of Bi2GaSbO7 and Bi2InSbO7, we had analyzed that both of the novel photocatalysts sent a strong absorption signal in the visible light region. Therefore, we expected that they could have the potential to degrade organic pollutants under visible light irradiation. In order to evaluate their visible light photocatalytic degradation capabilities, we listed N-doped TiO2 as a referential photocatalyst. Figure 7a presents the kinetics of MB degradation with Bi2GaSbO7, Bi2InSbO7, N-doped TiO2 as well as in the absence of a photocatalyst under visible light irradiation (>420 nm). Consistent with our expectations, as time went by, the color of the MB solution gradually shallowed and the concentration of MB gradually declined in our measurements in the absence of a photocatalyst. After visible light irradiation for 400 min, the removal rate of MB was estimated to be 99.75%, 98.95%, 59.92% or 40.6% with Bi2GaSbO7, Bi2InSbO7, N-doped TiO2 as catalyst, as well as in the absence of a photocatalyst, respectively. The sharp decrease in the concentration of MB under visible light irradiation from 0 to 120 min was mainly due to the adsorption of MB on the surface of Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a photocatalyst [66]. In the meantime, the photocatalytic degradation of MB with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst also played a significant role compared with the absence of a photocatalyst under visible light irradiation in this sharp decrease. In addition, the slower speed of MB degradation by using Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a photocatalyst during the later reaction process could be the result of as-prepared samples surface blocking by adsorbed MB degradation byproducts [67]. Moreover, the photocatalytic degradation rate of MB was 1.039 × 10−9 mol·L−1·s−1, 1.031 × 10−9 mol·L−1·s−1 or 0.624 × 10−9 mol·L−1·s−1 with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst during 400 min of visible light irradiation, respectively. The self-degradation rate of MB was 0.422 × 10−9 mol·L−1·s−1 without a catalyst. Furthermore, the photonic efficiency was estimated to be 0.0218% (λ = 420 nm), 0.0217% (λ = 420 nm) or 0.0131% (λ = 420 nm) with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst, indicating that the sufficient use of a large number of photons could lead to the production of a large number of electron/hole pairs which were responsible for the photocatalytic degradation reaction directly and/or indirectly [68]. According to above results, it was apparent that Bi2GaSbO7 and Bi2InSbO7 harvested the highest photocatalytic degradation rate and photonic efficiency compared with N-doped TiO2 for degrading MB. The decolored MB solution and the decrease of MB concentration reflected from Figure 7a might ascribe to the destruction of chromophore and the thorough degradation of the whole MB molecular [69]. We have verified our conjecture by detecting the mount variation of TOC and CO2 during MB degradation.
Figure 7b presents the UV-vis spectral changes during the photodegradation of MB with Bi2GaSbO7 as a photocatalyst. Noticeably, we could observe a subtle blue shift in the maximum absorbance of MB in the spectral changes by using Bi2GaSbO7 as a photocatalyst under visible light irradiation, indicating the rather facile cleavage of the whole conjugated chromophore structure [70]. This blue shift in the maximum absorbance of MB also proved the existence of some photodegradation intermediate products of MB during the photocatalytic degradation of MB under visible light irradiation in the presence of Bi2GaSbO7.
Figure 8 shows the change of TOC for the photocatalytic degradation of MB during visible light irradiation with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a photocatalyst, which is consistent with the tendency shown in Figure 7. The gradual decrease of TOC represented the gradual disappearance of organic carbon when the MB solution which contained Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 was exposed under visible light irradiation and the removal rate of TOC was 98.23%, 96.42% or 58.08% with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst after visible light irradiation for 400 min. In addition, the reactions stopped when the light was turned off in this experiment, which showed the obvious light response, suggesting that MB had been converted to other kinds of byproducts and the organic carbon in the MB had not been decomposed to CO2 [71].
Figure 9 shows the amount of variation of CO2 produced during the photocatalytic degradation of MB by using Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a photocatalyst under visible light irradiation. It could be distinctly seen from Figure 9 that the amount of CO2 gradually augmented along the light irradiation time and increased less during the last 100 min when much TOC was eliminated according to the results of Figure 8. In addition, after visible light irradiation of 400 min, the CO2 production of 0.11711 mmol or 0.11512 mmol with Bi2GaSbO7 or Bi2InSbO7 as a catalyst was higher than that of 0.06875 mmol with N-doped TiO2 as a catalyst. In addition, the amount of CO2 production was nearly equivalent to that of the removed TOC; at the same time, the amount of CO2 production or the removed TOC was slightly lower than the amount of reduced MB by using different catalysts with respect to the C element equilibrium, which indicated that MB was mainly degraded into some inorganic products including CO2 and eventually H2O.
Figure 10 presents the first order nature of the photocatalytic degradation kinetics with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst, which exhibits a linear correlation between ln(C/C0) or ln(TOC/TOC0) and the irradiation time for the photocatalytic degradation of MB under visible light irradiation by using the aforementioned catalysts. The pseudo-first-order kinetic curves of MB photodegradation were plotted to quantitatively compare the degradation rate of MB [72]. In the above expression, C and TOC represented the MB concentration and the total organic carbon concentration at time t, respectively. Likewise, C0 and TOC0 represented the initial concentration of MB and the initial total organic carbon concentration, respectively. By a linear fit for the relationship between ln(C/C0) and the irradiation time, the first-order rate constant kC was estimated to be 0.01470 min−1 with Bi2GaSbO7 as a catalyst, 0.00967 min−1 with Bi2InSbO7 as a catalyst or 0.00259 min−1 with N-doped TiO2 as a catalyst, which distinctly showed that Bi2GaSbO7 and Bi2InSbO7, with the highest and the second highest value of kC, respectively, exhibited more excellent visible light photocatalytic activities for degrading MB compared with N-doped TiO2. Similarly, by a linear fit for the relationship between ln(TOC/TOC0) and the irradiation time, the first-order rate constant kTOC was estimated to be 0.00881 min−1 with Bi2GaSbO7 as a catalyst, 0.00745 min−1 with Bi2InSbO7 as a catalyst or 0.00239 min−1 with N-doped TiO2 as a catalyst. The difference between kC and kTOC reflected that there might be some photodegradation intermediate products of MB which were produced during the photocatalytic degradation of MB under visible light irradiation.
Figure 11 presents the photocatalytic degradation rate of phenol under visible light irradiation in the presence of Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a photocatalyst with respect to time. It could be seen from Figure 11 that improved activity was obtained when colorless phenol was selected as a contaminant model with Bi2GaSbO7 or Bi2InSbO7 as a photocatalyst in comparison with the N-doped TiO2. The photocatalytic degradation efficiency of phenol by using Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a photocatalyst under visible light irradiation after 400 min was estimated to be 75.00%, 69.76% or 47.08%, respectively, indicating that Bi2GaSbO7 or Bi2InSbO7 itself had photocatalytic activity and that the photodegradation process of MB by using Bi2GaSbO7 or Bi2InSbO7 as a photocatalyst was not mainly due to the photosensitive effect [73]. Moreover, we could observe that the photodegradation efficiency or apparent rate constant of phenol or MB in the presence of Bi2GaSbO7 or Bi2InSbO7 was much higher than that in the presence of N-doped TiO2, meaning that the visible-light photocatalytic activity of Bi2GaSbO7 or Bi2InSbO7 was higher than that of N-doped TiO2.
The specific surface area of Bi2GaSbO7 or Bi2InSbO7 was measured to be 2.36 m2·g−1 or 1.82 m2·g−1, which was much smaller than that of N-doped TiO2, whose specific surface area was 45.53 m2·g−1. Generally speaking, a larger specific surface area would facilitate higher photocatalytic activities at the same experimental condition [74,75]. However, according to preceding results and discussions, Bi2GaSbO7 and Bi2InSbO7 showed higher activities than N-doped TiO2 for degrading MB during visible light irradiation, which sufficiently highlighted the excellent photocatalytic properties of Bi2GaSbO7 and Bi2InSbO7, and the above results might ascribe to two explanations. Firstly, as already mentioned, the calculated band gap for Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 was 2.59 eV, 2.54 eV or 2.78 eV. Apparently, Bi2GaSbO7 or Bi2InSbO7 possessed a narrower band gap than N-doped TiO2, meaning that Bi2GaSbO7 or Bi2InSbO7 could utilize more visible light energy than N-doped TiO2 [76,77]. Secondly, according to the XRD results of Bi2GaSbO7 and Bi2InSbO7, we could find that Bi2GaSbO7 and Bi2InSbO7 were both obtained with high crystallization perfection, which might more efficiently inhibit the recombination of photoinduced electrons and holes than N-doped TiO2.
Meanwhile, the photocatalytic degradation rate and photonic efficiency of Bi2GaSbO7 were slightly higher than that of Bi2InSbO7. There were perhaps two probable reasons to explain it. As we all know, the greater mobility of the photoinduced electrons and holes indicated the greater chance that the photoinduced electrons and holes would reach the reactive sites of the catalyst surface, which would bring higher photocatalytic activities. As we previously mentioned, the lattice parameter α = 10.356497 Å for Bi2GaSbO7 was lower than the lattice parameter α = 10.666031 Å for Bi2InSbO7. Generally speaking, the smaller the ionic radius was, the smaller the size of the particles could be; and the lower the lattice parameter was, the larger the specific surface area could be, which could increase more reactive sites on the photocatalyst surface and absorb more reactive species to improve the photocatalytic activities [78]. In addition, according to previous luminescent studies, the closer the M–O–M bond angle was to 180°, the more delocalized the excited state was [79]. As a result, the charge carriers could move easily in the matrix. In this experiment, for Bi2GaSbO7, the Ga–O–Ga bond angle was 131.302°; accordingly, for Bi2InSbO7, the In–O–In bond angle was 128.640°. Obviously, the bond angle of the Ga–O–Ga bond angle of Bi2GaSbO7 was larger than the bond angle of Bi2InSbO7, which induced that Bi2GaSbO7 exhibited higher photocatalytic activity than Bi2InSbO7.

3.3. Photocatalytic Degradation Pathway of MB with Bi2GaSbO7 and Bi2InSbO7 as Photocatalysts

The photodegradation intermediate products of MB in our experiment were identified as azure A, azure C, thionine, phenothiazine, leucomethylene blue, N,N-dimethyl-p-phenylenediamine, benzenesulfonic acid, phenol and aniline. There generated holes h+, ·O2 and ·OH radicals, as oxidative agents in the photocatalytic reactions. According to previous studies [80,81], the photodegradation of MB might occur by demethylation. Besides, there were also reports [82] which pointed out that ·OH radicals would first attack C − S+ = C functional group bonds to open the central aromatic ring which contained both heteroatoms S and N. Therefore, according to previous studies and our test results, a possible photocatalytic degradation pathway for MB was proposed. Figure 12 shows the suggested photocatalytic degradation pathway scheme for MB under visible light irradiation with Bi2GaSbO7 or Bi2InSbO7 as a catalyst. The MB molecule was converted to small organic species, which were subsequently mineralized into inorganic products such as SO42− ions, NO3 ions, CO2 and ultimately water.

3.4. Photocatalytic Degradation Mechanism

Figure 13 presents the action spectra of MB degradation with Bi2GaSbO7 or Bi2InSbO7 as a catalyst under visible light irradiation. A clear photonic efficiency (0.00964% for Bi2GaSbO7 and 0.00942% for Bi2InSbO7 at their respective maximal point) at wavelengths which corresponded to sub-Eg energies of the photocatalysts (λ from 480 to 700 nm for Bi2GaSbO7 and λ from 490 to 700 nm for Bi2InSbO7) was observed. The existence of photonic efficiency at this region revealed that the photons were not absorbed by the photocatalysts. Enlightened by the correlation between the low-energy action spectrum and the absorption spectrum of MB, we speculated that any photodegradation which results at wavelengths above 480 nm, should be attributed to photosensitization effect by the dye MB itself (Scheme 1). According to the photosensitization scheme, MB which was adsorbed on Bi2GaSbO7 or Bi2InSbO7 was excited by visible light irradiation. Subsequently, an electron was injected from the excited MB to the conduction band of Bi2GaSbO7 or Bi2InSbO7 where the electron was scavenged by molecular oxygen. This explained the results which were gained with Bi2GaSbO7 or Bi2InSbO7 as a catalyst under visible light irradiation, where the catalyst could serve to reduce recombination of photoinduced electrons and photoinduced holes by scavenging of electrons.
The situation was different below 480 nm, where the photonic efficiency correlated well with the absorption spectra of Bi2GaSbO7 or Bi2InSbO7. This result evidently indicated that the mechanism was the photodegradation of MB by the band gap excitation of Bi2GaSbO7 or Bi2InSbO7. As already mentioned, holes of h+, ·O2 and OH· radicals served as oxidative agents in the photocatalytic reactions. Although the detailed experiments about the effect of oxygen and water on the degradation mechanism of MB were not performed, it was sensible to assume that the mechanism in the first step was similar to the observed mechanism for Bi2GaSbO7 or Bi2InSbO7 under supra-bandgap irradiation, and the production scheme of oxidative radicals commonly was shown below (Scheme 2).
Figure 14 shows the suggested band structures of Bi2GaSbO7 and Bi2InSbO7. The positions and width of the conduction band (CB) and the valence band (VB) were studied by calculating the electronic band structure of Bi2GaSbO7 or Bi2InSbO7 with the plane-wave-based density functional method. The band structure calculations of Bi2GaSbO7 and Bi2InSbO7 were carried out with the program of Cambridge serial total energy package (CASTEP) and first-principles simulation. It could be seen from Figure 14 that the conduction band of Bi2GaSbO7 was composed of Ga 4p and Sb 5p orbital component, meanwhile, the valence band of Bi2GaSbO7 was composed of a small dominant O 2p and Bi 6s orbital component. Similarly, the conduction band of Bi2InSbO7 was composed of In 5p and Sb 5p orbital component. In addition, the valence band of Bi2InSbO7 was composed of a small dominant O 2p and Bi 6s orbital component. Direct absorption of photons by Bi2GaSbO7 or Bi2InSbO7 could produce electron–hole pairs within the catalyst, indicating that the larger energy than the band gap of Bi2GaSbO7 or Bi2InSbO7 was necessary for decomposing MB by the photocatalysis method.

4. Conclusions

New photocatalysts Bi2GaSbO7 and Bi2InSbO7 were firstly prepared by the solid-state reaction method. The structural properties and optical absorption properties of Bi2GaSbO7 and Bi2InSbO7 were characterized by some material characterization methods, the photocatalytic properties of Bi2GaSbO7 and Bi2InSbO7 were also verified in comparison with N-doped TiO2. XRD results indicated that Bi2GaSbO7 and Bi2InSbO7 crystallized with the pyrochlore-type structure, cubic crystal system and space group Fd3m. The lattice parameter a for Bi2GaSbO7 or Bi2InSbO7 was a = 10.356497 Å or a = 10.666031 Å. According to the results from the UV-vis absorption spectra of Bi2GaSbO7 and Bi2InSbO7, the band gap of Bi2GaSbO7 or Bi2InSbO7 was estimated to be about 2.59 eV or 2.54 eV, indicating that Bi2GaSbO7 and Bi2InSbO7 showed a strong optical absorption in the visible light region (λ > 420 nm). Photocatalytic degradation of aqueous MB was realized under visible light irradiation in the presence of Bi2GaSbO7 or Bi2InSbO7 accompanied with the formation of final products such as CO2 and water. The complete removal of organic carbon from MB was obtained as indicated from TOC and CO2 yield measurements with Bi2GaSbO7 or Bi2InSbO7 as a catalyst under visible light irradiation. Compared with N-doped TiO2, Bi2GaSbO7 and Bi2InSbO7 exhibited higher photocatalytic activities for MB degradation under visible light irradiation. Consequently, according to the above analyses, Bi2GaSbO7 and Bi2InSbO7 both had great potential to degrade MB in textile industry wastewater. In addition, Bi2GaSbO7 exhibited slightly higher photocatalytic activities for the degradation of MB than Bi2InSbO7.

Acknowledgments

This work was supported by a grant from the Natural Science Foundation of Jiangsu Province (No. BK20141312), by a Project of Science and Technology Development Plan of Suzhou City of China from 2014 (No. ZXG201440), by a grant from China-Israel Joint Research Program in Water Technology and Renewable Energy (No. 5).

Author Contributions

Jingfei Luan were involved with all aspects of the study including conceiving, designing, data interpretation and writing the manuscript. Yue Shen, Yanyan Li and Yaron Paz performed the experiments and analyzed data. Jingfei Luan, Yue Shen and Yanyan Li wrote the paper. All authors read and approved the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yao, J.T.; Jia, R.; Zheng, L.L.; Wang, B.X. Rapid decolorization of azo dyes by crude manganese peroxidase from Schizophyllum sp. F17 in solid-state fermentation. Biotechnol. Bioprocess Eng. 2013, 18, 868–877. [Google Scholar] [CrossRef]
  2. Apostol, L.C.; Pereira, L.; Pereira, R.; Gavrilescu, M.; Alves, M.M. Biological decolorization of xanthene dyes by anaerobic granular biomass. Biodegradation 2012, 23, 725–737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Aravind, P.; Subramanyan, V.; Ferro, S.; Gopalakrishnan, R. Eco-friendly and facile integrated biological-cum-photo assisted electrooxidation process for degradation of textile wastewater. Water Res. 2016, 93, 230–241. [Google Scholar] [CrossRef] [PubMed]
  4. Korbahti, B.K.; Artut, K.; Gecgel, C.; Ozer, A. Electrochemical decolorization of textile dyes and removal of metal ions from textile dye and metal ion binary mixtures. Chem. Eng. J. 2011, 173, 677–688. [Google Scholar] [CrossRef]
  5. Khedr, A.M.; Abu, G.N.; Salem, M.F.; Gaber, M. Determination of the efficiency of different modified electrodes in electrochemical degradation of reactive Red 24 dyes in wastewater dyestuff solutions. Int. J. Electrochem. Sci. 2012, 7, 8779–8793. [Google Scholar]
  6. Haque, M.M.; Smith, W.T.; Wong, D.K.Y. Conducting polypyrrole films as a potential tool for electrochemical treatment of azo dyes intextile wastewaters. J. Hazard. Mater. 2015, 283, 164–170. [Google Scholar] [CrossRef] [PubMed]
  7. Shirmardi, M.; Mahvi, A.H.; Mesdaghinia, A.; Nasseri, S.; Nabizadeh, R. Adsorption of acid red18 dye from aqueous solution using single-wall carbon nanotubes: Kinetic and equilibrium. Desalin. Water Treat. 2013, 51, 6507–6516. [Google Scholar] [CrossRef]
  8. Hayati, B.; Mahmoodi, N.M. Modification of activated carbon by the alkaline treatment to remove the dyes from wastewater: Mechanism, isotherm and kinetic. Desalin. Water Treat. 2012, 47, 322–333. [Google Scholar] [CrossRef]
  9. Nascimento, G.E.; Duarte, M.M.M.B.; Campos, N.F.; da Rocha, O.R.S.; da Silva, V.L. Adsorption of azo dyes using peanut hull and orange peel: A comparative study. Environ. Technol. 2014, 35, 1436–1453. [Google Scholar] [CrossRef] [PubMed]
  10. Tang, H.; Zhang, D.; Tang, G.G.; Ji, X.R.; Li, C.S.; Yan, X.H.; Wu, Q. Low temperature synthesis and photocatalytic properties of mesoporous TiO2 nanospheres. J. Alloy. Compd. 2014, 591, 52–57. [Google Scholar] [CrossRef]
  11. Dragan, E.S.; Dinu, I.A. Removal of azo dyes from aqueous solution by coagulation/flocculation with strong polycations. Res. J. Chem. Environ. 2008, 12, 5–11. [Google Scholar]
  12. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, J.; Wan, Y.P.; Huang, Y.L.; Wang, Y.R.; Qin, L.; Seo, H.J. Layered oxide semiconductor In2Fe2CuO7: Optical properties and visible-light responsive photocatalytic abilities. Mater. Lett. 2016, 179, 175–178. [Google Scholar] [CrossRef]
  14. Bu, Y.Y.; Chen, Z.Y.; Sun, C.J. Highly efficient Z-Scheme Ag3PO4/Ag/WO3-x photocatalyst for its enhanced photocatalytic performance. Appl. Catal. B 2015, 179, 363–371. [Google Scholar] [CrossRef]
  15. Zhao, B.; Wang, M.; Lin, L.; Zeng, Q.Q.; He, D.N. Synthesis of parallel squared nanosheet-assembled Bi2WO6 microstructures under alkalescent hydrothermal treatment. Ceram. Int. 2014, 40, 5831–5835. [Google Scholar] [CrossRef]
  16. Alemi, A.A.; Kashfi, R.; Shabani, B. Preparation and characterization of novel Ln (Gd3+, Ho3+ and Yb3+)-doped Bi2MoO6 with Aurivillius layered structures and photocatalytic activities under visible light irradiation. J. Mol. Catal. A Chem. 2014, 392, 290–298. [Google Scholar] [CrossRef]
  17. Nazim, S.; Kousar, T.; Shahid, M.; Khan, M.A.; Nasar, G.; Sher, M.; Warsi, M.F. New graphene-CoxZn1-xFe2O4 nano-heterostructures: Magnetically separable visible lightphotocatalytic materials. Ceram. Int. 2016, 42, 7647–7654. [Google Scholar] [CrossRef]
  18. Ghaffar, I.; Warsi, M.F.; Shahid, M.; Shakir, I. Unprecedented photocatalytic activity of carbon coated/MoO3 core-shell nanoheterostructurs under visible light irradiation. Phys. E Low-Dimens. Syst. Nanostruct. 2016, 79, 1–7. [Google Scholar] [CrossRef]
  19. Kiransan, M.; Khataee, A.; Karaca, S.; Sheydaei, M. Artificial neural network modeling of photocatalytic removal of a disperse dye using synthesized of ZnO nanoparticles on montmorillonite. Spectrochim. Acta A 2015, 140, 465–473. [Google Scholar] [CrossRef] [PubMed]
  20. Khataee, A.; Karimi, A.; Arefi-Oskoui, S.; Soltani, R.D.C.; Hanifehpour, Y.; Soltani, B.; Joo, S.W. Sonochemical synthesis of Pr-doped ZnO nanoparticles for sonocatalytic degradation of Acid Red 17. Ultrason. Sonochem. 2015, 22, 371–381. [Google Scholar] [CrossRef] [PubMed]
  21. Yi, X.; Li, J.L. Synthesis and optical property of NaTaO3 nanofibers prepared by electrospinning. J. Sol-Gel Sci. Technol. 2010, 53, 480–484. [Google Scholar] [CrossRef]
  22. Yang, J.X.; Akbarzadeh, J.; Maurer, C.; Peterlik, H.; Schubert, U. Sol-gel synthesis of ZnTiO3 using a single-source precursor based on p-carboxybenzaldehyde oxime as a linker. J. Mater. Chem. 2012, 22, 24034–24041. [Google Scholar] [CrossRef]
  23. Suresh, R.; Giribabu, K.; Manigandan, R.; Munusamy, S.; Kumar, S.P.; Muthamizh, S.; Stephen, A.; Narayanan, V. Doping of Co into V2O5 nanoparticles enhances of methylene blue. J. Alloy. Compd. 2014, 598, 151–160. [Google Scholar] [CrossRef]
  24. Ge, L.; Zhang, X.H. Synthesis of novel visible light driven BiVO4 photocatalysts via microemulsion process and its photocatalytic performance. J. Inorg. Mater. 2009, 24, 453–456. [Google Scholar] [CrossRef]
  25. Yi, X.F.; Zheng, J.S.; Zhao, Y.B. Hydrothermal synthesis of CdS nanorods in NaOH solution. Chem. J. Chin. Univ. 2012, 33, 2597–2603. [Google Scholar]
  26. Yang, S.X.; Yue, Q.; Wu, F.M.; Huo, N.J.; Chen, Z.H.; Yang, J.H.; Li, J.B. Synthesis of the nanostructured Cd4GeS6 photocatalysts and their visible-light-driven photocatalytic degradation property. J. Alloy. Compd. 2014, 597, 91–94. [Google Scholar] [CrossRef]
  27. Wu, W.M.; Lin, R.; Shen, L.J.; Liang, R.W.; Yuan, R.S.; Wu, L. Visible-light-induced photocatalytic hydrogenation of 4-nitroaniline over In2S3 photocatalyst in water. Catal. Commun. 2013, 40, 1–4. [Google Scholar] [CrossRef]
  28. Wei, R.J.; Hu, J.C.; Zhou, T.F.; Zhou, X.L.; Liu, J.X.; Li, J.L. Ultrathin SnS2 nanosheets with exposed {001} facets and enhanced photocatalytic properties. Acta Mater. 2014, 66, 163–171. [Google Scholar] [CrossRef]
  29. Han, Q.F.; Chen, L.; Wang, M.J.; Yang, X.J.; Lu, L.D.; Wang, X. Low-temperature synthesis of uniform Sb2S3 nanorods and its visible-light-driven photocatalytic activities. Mater. Sci. Eng. B 2010, 166, 118–121. [Google Scholar] [CrossRef]
  30. Ding, J.J.; Sun, S.; Yan, W.H.; Bao, J.; Gao, C. Photocatalytic H-2 evolution on a novel CaIn2S4 photocatalyst under visible light irradiation. Int. J. Hydrogen Energy 2013, 38, 13153–13158. [Google Scholar] [CrossRef]
  31. Hanifehpour, Y.; Soltani, B.; Amani-Ghadim, A.R.; Hedayati, B.; Khomami, B.; Joo, S.W. Praseodymium-doped ZnS nanomaterials: Hydrothermal synthesis and characterization with enhanced visible light photocatalytic activity. J. Ind. Eng. Chem. 2016, 34, 41–50. [Google Scholar] [CrossRef]
  32. Peng, W.C.; Chen, Y.; Li, X.Y. MoS2/reduced graphene oxide hybrid with CdS nanoparticles as a visible light-driven photocatalyst for the reduction of 4-nitrophenol. J. Hazard. Mater. 2016, 309, 173–179. [Google Scholar] [CrossRef] [PubMed]
  33. Kato, T.; Hakari, Y.; Ikeda, S.; Jia, Q.X.; Iwase, A.; Kudo, A. Utilization of metal sulfide material of (CuGa)(1-x)Zn2xS2 solid solution with visible light response in photocatalytic and photoelectrochemical solar water splitting systems. J. Phys. Chem. Lett. 2015, 6, 1042–1047. [Google Scholar] [CrossRef] [PubMed]
  34. Eder, D.; Motta, M.; Windle, A.H. Iron-doped Pt-TiO2 nanotubes for photo-catalytic water splitting. Nanotechnology 2009, 20, 055602. [Google Scholar] [CrossRef] [PubMed]
  35. Biswas, S.K.; Baeg, J.O. Enhanced photoactivity of visible light responsive W incorporated FeVO4 photoanode for solar water splitting. Int. J. Hydrogen Energy 2013, 38, 14451–14457. [Google Scholar] [CrossRef]
  36. Anandan, S.; Rao, T.N.; Gopalan, R.; Ikuma, Y. Fabrication of visible-light-driven N-doped ordered mesoporous TiO2 photocatalysts and their photocatalytic applications. J. Nanosci. Nanotechnol. 2014, 14, 3181–3186. [Google Scholar] [CrossRef] [PubMed]
  37. Zong, X.; Yan, H.J.; Wu, G.P.; Ma, G.J.; Wen, F.Y.; Wang, L.; Li, C. Enhancement of photocatalytic H-2 evolution on CdS by loading MOS2 as cocatalyst under visible light irradiation. J. Am. Chem. Soc. 2008, 130, 7176–7177. [Google Scholar] [CrossRef] [PubMed]
  38. Zong, X.; Han, J.F.; Ma, G.J.; Yan, H.J.; Wu, G.P.; Li, C. Photocatalytic H-2 evolution on CdS loaded with WS2 as cocatalyst under visible light irradiation. J. Phys. Chem. C 2011, 115, 12202–12208. [Google Scholar] [CrossRef]
  39. Dai, K.; Lv, J.L.; Lu, L.H.; Liang, C.H.; Geng, L.; Zhu, G.P. A facile fabrication of plasmonic g-C3N4/Ag2WO4/Ag ternary heterojunction visible-light photocatalyst. Mater. Chem. Phys. 2016, 177, 529–537. [Google Scholar] [CrossRef]
  40. Wang, X.F.; Hu, H.M.; Chen, S.H.; Zhang, K.H.; Zhang, J.; Zou, W.S.; Wang, R.X. One-step fabrication of BiOCl/CuS heterojunction photocatalysts with enhanced visible-light responsive activity. Mater. Chem. Phys. 2015, 158, 67–73. [Google Scholar] [CrossRef]
  41. Xie, T.P.; Liu, C.L.; Xu, L.J.; Yang, J.; Zhou, W. Novel heterojunction Bi2O3/SrFe12O19 magnetic photocatalyst with highly enhanced photocatalytic activity. J. Phys. Chem. C 2013, 117, 24601–24610. [Google Scholar] [CrossRef]
  42. Han, C.C.; Ge, L.; Chen, C.F.; Li, Y.J.; Xiao, X.L.; Zhang, Y.N.; Guo, L.L. Novel visible light induced Co3O4-g-C3N4 heterojunction photocatalysts for efficient degradation of methyl orange. Appl. Catal. B Environ. 2014, 147, 546–553. [Google Scholar] [CrossRef]
  43. Liu, X.; Li, Y.X.; Peng, S.Q.; Lu, G.X.; Li, S.B. Photosensitization of SiW11O398—Modified TiO2 by Eosin Y for stable visible-light H-2 generation. Int. J. Hydrogen Energy 2013, 38, 11709–11719. [Google Scholar] [CrossRef]
  44. Biswas, S.; Sundstrom, V.; De, S. Facile synthesis of luminescent TiO2 nanorods using an anionic surfactant: Their photosensitization and photocatalytic efficiency. Mater. Chem. Phys. 2014, 147, 761–771. [Google Scholar] [CrossRef]
  45. Pirhashemi, M.; Habibi-Yangjeh, A. Photosensitization of ZnO by AgBr and Ag2CO3: Nanocomposites with tandem n-n heterojunctions and highly enhanced visible-light photocatalytic activity. J. Colloid Interface Sci. 2016, 474, 103–113. [Google Scholar] [CrossRef] [PubMed]
  46. Lu, H.J.; Xu, L.L.; Wei, B.; Zhang, M.Y.; Gao, H.; Sun, W.J. Enhanced photosensitization process induced by the p-n junction of Bi2O2CO3/BiOCl heterojunctions on the degradation of rhodamine B. Appl. Surf. Sci. 2014, 303, 360–366. [Google Scholar] [CrossRef]
  47. Chen, X.F.; Zhang, J.; Huo, Y.N.; Li, H.X. Preparation and visible light catalytic activity of three-dimensional ordered macroporous CdS/TiO2 films. Chin. J. Catal. 2013, 34, 949–955. [Google Scholar] [CrossRef]
  48. Cao, J.; Xu, B.Y.; Lin, H.L.; Luo, B.D.; Chen, S.F. Novel Bi2S3-sensitized BiOCl with highly visible light photocatalytic activity for the removal of rhodamine B. Catal. Commun. 2012, 26, 204–208. [Google Scholar] [CrossRef]
  49. Zou, Z.G.; Ye, J.H.; Arakawa, H. Photophysical and photocatalytic properties of InMO4 (M = Nb5+, Ta5+) under visible light irradiation. Mater. Res. Bull. 2001, 36, 1185–1193. [Google Scholar] [CrossRef]
  50. Zou, Z.G.; Arakawa, H. Direct water splitting into H2 and O2 under visible light irradiation with a new series of mixed oxide semiconductor photocatalysts. J. Photochem. Photobiol. A Chem. 2003, 158, 145–162. [Google Scholar] [CrossRef]
  51. Zou, Z.G.; Ye, J.H.; Arakawa, H. Role of R in Bi2RNbO7 (R = Y, rare earth): Effect on band structure and photocatalytic properties. J. Phys. Chem. B 2002, 106, 517–520. [Google Scholar] [CrossRef]
  52. Zou, Z.G.; Ye, J.H.; Arakawa, H. Photocatalytic and photophysical properties of a novel series of solid photocatalysts, Bi2MNbO7 (M = Al3+, Ga3+ and In3+). Chem. Phys. Lett. 2001, 333, 57–62. [Google Scholar] [CrossRef]
  53. Zou, Z.G.; Ye, J.H.; Arakawa, H. Substitution effects of In3+ by Fe3+ on photocatalytic and structural properties of Bi2InNbO7 photocatalysts. J. Mol. Catal. A Chem. 2001, 168, 289–297. [Google Scholar] [CrossRef]
  54. Luan, J.F.; Cai, H.L.; Zheng, S.R.; Hao, X.P.; Luan, G.Y.; Wu, X.S.; Zou, Z.G. Structural and photocatalytic properties of novel Bi2GaVO7. Mater. Chem. Phys. 2007, 104, 119–124. [Google Scholar] [CrossRef]
  55. Luan, J.F.; Pan, B.C.; Paz, Y.; Li, Y.M.; Wu, X.S.; Zou, Z.G. Structural, photophysical and photocatalytic properties of new Bi2SbVO7 under visible light irradiation. Phys. Chem. Chem. Phys. 2009, 11, 6289–6298. [Google Scholar] [CrossRef] [PubMed]
  56. Robertson, J.; Peacock, P.W.; Towler, M.D.; Needs, R. Electronic structure of p-type conducting transparent oxides. Thin Solid Films 2002, 411, 96–100. [Google Scholar] [CrossRef]
  57. Yu, X.L.; An, X.Q.; Shavel, A.; Ibanez, M.; Cabot, A. The effect of the Ga content on the photocatalytic hydrogen evolution of CuIn1−xGaxS2 nanocrystals. J. Mater. Chem. 2014, 2, 12317–12322. [Google Scholar] [CrossRef]
  58. Zhao, H.P.; Tang, J.L.; Lai, Q.S.; Cheng, G.; Liu, Y.L.; Chen, R. Enhanced visible light photocatalytic performance of Sb-doped (BiO)2CO3 nanoplates. Catal. Commun. 2015, 58, 190–194. [Google Scholar] [CrossRef]
  59. Omidi, A.; Habibi-Yangjeh, A.; Pirhashemi, M. Application of ultrasonic irradiation method for preparation of ZnO nanostructures doped with Sb+3 ions as a highly efficient photocatalyst. Appl. Surf. Sci. 2013, 276, 468–475. [Google Scholar] [CrossRef]
  60. Al-Hamdi, A.M.; Sillanpaa, M.; Bora, T.; Dutta, J. Efficient photocatalytic degradation of phenol in aqueous solution by SnO2:Sb nanoparticles. Appl. Surf. Sci. 2016, 370, 229–236. [Google Scholar] [CrossRef]
  61. Du, H.Y.; Luan, J.F. Synthesis, characterization and photocatalytic activity of new photocatalyst CdBiYO4. Solid State Sci. 2012, 14, 1295–1305. [Google Scholar] [CrossRef]
  62. Sakthivel, S.; Shankar, M.V.; Palanichamy, M.; Arabindoo, B.; Bahnemann, D.W.; Murugesan, V. Enhancement of photocatalytic activity by metal deposition: Characterisation and photonic efficiency of Pt, Au and Pd deposited on TiO2 catalyst. Water Res. 2004, 38, 3001–3008. [Google Scholar] [CrossRef] [PubMed]
  63. Marugan, J.; Hufschmidt, D.; Sagawe, G.; Selzer, V.; Bahnemann, D. Optical density and photonic efficiency of silica-supported TiO2 photocatalysts. Water Res. 2006, 40, 833–839. [Google Scholar] [CrossRef] [PubMed]
  64. Zou, Z.G.; Ye, J.H.; Arakawa, H. Preparation, structural and photophysical properties of Bi2InNbO7 compound. J. Mater. Sci. Lett. 2000, 19, 1909–1911. [Google Scholar] [CrossRef]
  65. Tauc, J.; Grigorov, R.; Vancu, A. Optical properties and electronic structure of amorphous germanium. Phys. Status Solidi (B) 1966, 15, 627–637. [Google Scholar] [CrossRef]
  66. Vaiano, V.; Sacco, O.; Sannino, D.; Ciambelli, P. Nanostructured N-doped TiO2 coated on glass spheres for the photocatalytic removal of organic dyes under UV or visible light irradiation. Appl. Catal. B Environ. 2015, 170, 153–161. [Google Scholar] [CrossRef]
  67. Klein, M.; Nadolna, J.; Golabiewska, A.; Mazierski, P.; Klimczuk, T.; Remita, H.; Zaleska-Medynska, A. The effect of metal cluster deposition route on structure and photocatalytic activity of mono- and bimetallic nanoparticles supported on TiO2 by radiolytic method. Appl. Surf. Sci. 2016, 378, 37–48. [Google Scholar] [CrossRef]
  68. Guesh, K.; Marquez-Alvarez, C.; Chebude, Y.; Diaz, I. Enhanced photocatalytic activity of supported TiO2 by selective surface modification of zeolite Y. Appl. Surf. Sci. 2016, 378, 473–478. [Google Scholar] [CrossRef]
  69. Wang, Q.; Tian, S.L.; Cun, J.; Ning, P. Degradation of methylene blue using a heterogeneous Fenton process catalyzed by ferrocene. Desalin. Water Treat. 2013, 51, 5821–5830. [Google Scholar] [CrossRef]
  70. Wang, Q.; Chen, C.C.; Zhao, D.; Ma, W.H.; Zhao, J.C. Change of adsorption modes of dyes on fluorinated TiO2 and its effect on photocatalytic degradation of dyes under visible irradiation. Langmiur 2008, 24, 7338–7345. [Google Scholar] [CrossRef] [PubMed]
  71. Sun, H.; Qiu, G.H.; Wang, Y.; Feng, X.H.; Yin, H.; Liu, F. Effects of Co and Ni co-doping on the physicochemical properties of cryptomelane and its enhanced performance on photocatalytic degradation of phenol. Mater. Chem. Phys. 2014, 148, 783–789. [Google Scholar] [CrossRef]
  72. Tian, N.; Zhang, Y.H.; Huang, H.W.; He, Y.; Guo, Y.X. Influences of Gd substitution on the crystal structure and visible-light-driven photocatalytic performance of Bi2WO6. J. Phys. Chem. C 2014, 118, 15640–15648. [Google Scholar] [CrossRef]
  73. Calza, P.; Rigo, L.; Sangermano, M. Investigations of photocatalytic activities of photosensitive semiconductors dispersed into epoxy matrix. Appl. Catal. B Environ. 2011, 106, 657–663. [Google Scholar] [CrossRef]
  74. Zhang, H.; Lv, X.J.; Li, Y.M.; Wang, Y.; Li, J.H. P25-graphene composite as a high performance photocatalyst. ACS Nano 2010, 4, 380–386. [Google Scholar] [CrossRef] [PubMed]
  75. Ahmad, M.; Ahmed, E.; Hong, Z.L.; Xu, J.F.; Khalid, N.R.; Elhissi, A.; Ahmed, W. A facile one-step approach to synthesizing ZnO/graphene composites for enhanced degradation of methylene blue under visible light. Appl. Surf. Sci. 2013, 274, 273–281. [Google Scholar] [CrossRef]
  76. Zhai, H.F.; Li, A.D.; Kong, J.Z.; Li, X.F.; Zhao, J.; Guo, B.L.; Yin, J.; Li, Z.S.; Wu, D. Preparation and visible-light photocatalytic properties of BiNbO4 and BiTaO4 by a citrate method. J. Solid State Chem. 2013, 202, 6–14. [Google Scholar] [CrossRef]
  77. Feng, H.Y.; Hou, D.F.; Huang, Y.H.; Hu, X.L. Facile synthesis of porous InNbO4 nanofibers by electrospinning and their enhanced visible-light-driven photocatalytic properties. J. Alloy. Compd. 2014, 592, 301–305. [Google Scholar] [CrossRef]
  78. Buvaneswari, K.; Karthiga, R.; Kavitha, B.; Rajarajan, M.; Suganthi, A. Effect of FeWO4 doping on the photocatalytic activity of ZnO under visible light irradiation. Appl. Surf. Sci. 2015, 356, 333–340. [Google Scholar] [CrossRef]
  79. Wiegel, M.; Middel, W.; Blasse, G. Influence of ns2 ions on the luminescence of niobates and tantalates. J. Mater. Chem. 1995, 5, 981–983. [Google Scholar] [CrossRef]
  80. Yogi, C.; Kojima, K.; Wada, N.; Tokumoto, H.; Takai, T.; Mizoguchi, T.; Tamiaki, H. Photocatalytic degradation of methylene blue by TiO2 film and Au particles-TiO2 composite film. Thin Solid Films 2008, 516, 5881–5884. [Google Scholar] [CrossRef]
  81. Rauf, M.A.; Meetani, M.A.; Khaleel, A.; Ahmed, A. Photocatalytic degradation of Methylene Blue using a mixed catalyst and product analysis by LC/MS. Chem. Eng. J. 2010, 157, 373–378. [Google Scholar] [CrossRef]
  82. Sharma, S.D.; Saini, K.K.; Kant, C.; Sharma, C.P.; Jain, S.C. Photodegradation of dye pollutant under UV light by nano-catalyst doped titania thin films. Appl. Catal. B Environ. 2008, 84, 233–240. [Google Scholar] [CrossRef]
Figure 1. TEM images of (a) Bi2GaSbO7 and (b) Bi2InSbO7 with high magnification.
Figure 1. TEM images of (a) Bi2GaSbO7 and (b) Bi2InSbO7 with high magnification.
Materials 09 00801 g001
Figure 2. SEM–EDS spectra taken from (a) Bi2GaSbO7 and (b) Bi2InSbO7.
Figure 2. SEM–EDS spectra taken from (a) Bi2GaSbO7 and (b) Bi2InSbO7.
Materials 09 00801 g002
Figure 3. X-ray powder diffraction patterns of Bi2GaSbO7 and Bi2InSbO7.
Figure 3. X-ray powder diffraction patterns of Bi2GaSbO7 and Bi2InSbO7.
Materials 09 00801 g003
Figure 4. The Pawley refinement results of XRD data for (a) Bi2GaSbO7 and (b) Bi2InSbO7.
Figure 4. The Pawley refinement results of XRD data for (a) Bi2GaSbO7 and (b) Bi2InSbO7.
Materials 09 00801 g004
Figure 5. Diffuse reflection spectra of Bi2GaSbO7, Bi2InSbO7 and N-doped TiO2.
Figure 5. Diffuse reflection spectra of Bi2GaSbO7, Bi2InSbO7 and N-doped TiO2.
Materials 09 00801 g005
Figure 6. Plot of (αhν)1/2 versus for (a) Bi2GaSbO7; (b) Bi2InSbO7 and (c) N-doped TiO2.
Figure 6. Plot of (αhν)1/2 versus for (a) Bi2GaSbO7; (b) Bi2InSbO7 and (c) N-doped TiO2.
Materials 09 00801 g006
Figure 7. (a) Photocatalytic degradation of methylene blue under visible light irradiation in the presence of Bi2GaSbO7, Bi2InSbO7, N-doped TiO2 as well as in the absence of a photocatalyst; (b) Temporal UV-vis absorption spectral changes during the photocatalytic degradation of MB (0.025 mmol/L, pH = 7) in aqueous Bi2GaSbO7 suspensions.
Figure 7. (a) Photocatalytic degradation of methylene blue under visible light irradiation in the presence of Bi2GaSbO7, Bi2InSbO7, N-doped TiO2 as well as in the absence of a photocatalyst; (b) Temporal UV-vis absorption spectral changes during the photocatalytic degradation of MB (0.025 mmol/L, pH = 7) in aqueous Bi2GaSbO7 suspensions.
Materials 09 00801 g007
Figure 8. Disappearance of the total organic carbon (TOC) during the photocatalytic degradation of methylene blue with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst under visible light irradiation.
Figure 8. Disappearance of the total organic carbon (TOC) during the photocatalytic degradation of methylene blue with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst under visible light irradiation.
Materials 09 00801 g008
Figure 9. CO2 production kinetics during the photocatalytic degradation of methylene blue with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst under visible light irradiation.
Figure 9. CO2 production kinetics during the photocatalytic degradation of methylene blue with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst under visible light irradiation.
Materials 09 00801 g009
Figure 10. Observed first-order kinetic plots for the photocatalytic degradation of methylene blue with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst under visible light irradiation.
Figure 10. Observed first-order kinetic plots for the photocatalytic degradation of methylene blue with Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a catalyst under visible light irradiation.
Materials 09 00801 g010
Figure 11. Photocatalytic degradation of phenol under visible light irradiation in the presence of Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a photocatalyst.
Figure 11. Photocatalytic degradation of phenol under visible light irradiation in the presence of Bi2GaSbO7, Bi2InSbO7 or N-doped TiO2 as a photocatalyst.
Materials 09 00801 g011
Figure 12. Suggested photocatalytic degradation pathway scheme for methylene blue under visible light irradiation in the presence of Bi2GaSbO7 or Bi2InSbO7.
Figure 12. Suggested photocatalytic degradation pathway scheme for methylene blue under visible light irradiation in the presence of Bi2GaSbO7 or Bi2InSbO7.
Materials 09 00801 g012
Scheme 1. The photosensitization effect by the dye MB.
Scheme 1. The photosensitization effect by the dye MB.
Materials 09 00801 sch001
Scheme 2. The production scheme of oxidative radicals with Bi2GaSbO7 or Bi2InSbO7 as catalyst.
Scheme 2. The production scheme of oxidative radicals with Bi2GaSbO7 or Bi2InSbO7 as catalyst.
Materials 09 00801 sch002
Figure 13. Action spectra of methylene blue degradation with Bi2GaSbO7 or Bi2InSbO7 as a catalyst under visible light irradiation.
Figure 13. Action spectra of methylene blue degradation with Bi2GaSbO7 or Bi2InSbO7 as a catalyst under visible light irradiation.
Materials 09 00801 g013
Figure 14. Suggested band structures of Bi2GaSbO7 and Bi2InSbO7.
Figure 14. Suggested band structures of Bi2GaSbO7 and Bi2InSbO7.
Materials 09 00801 g014
Table 1. Binding energies (BE) for key elements of Bi2InSbO7 and Bi2GaSbO7.
Table 1. Binding energies (BE) for key elements of Bi2InSbO7 and Bi2GaSbO7.
CompoundBi4f7/2 BE (eV)Sb3d5/2 BE (eV)Ga3d5/2 BE (eV)In3d5/2 BE (eV)O1s BE (eV)
Bi2InSbO7159.70531.20444.60530.85
Bi2GaSbO7159.60531.4020.60531.10
Table 2. Structural parameters of Bi2GaSbO7 prepared by the solid state reaction method.
Table 2. Structural parameters of Bi2GaSbO7 prepared by the solid state reaction method.
AtomxyzOccupation Factor
Bi0.000000.000000.000001.0
Ga0.500000.500000.500000.5
Sb0.500000.500000.500000.5
O(1)−0.185000.125000.125001.0
O(2)0.125000.125000.125001.0
Table 3. Structural parameters of Bi2InSbO7 prepared by the solid state reaction method.
Table 3. Structural parameters of Bi2InSbO7 prepared by the solid state reaction method.
AtomxyzOccupation Factor
Bi0.000000.000000.000001.0
In0.500000.500000.500000.5
Sb0.500000.500000.500000.5
O(1)−0.165000.125000.125001.0
O(2)0.125000.125000.125001.0

Share and Cite

MDPI and ACS Style

Luan, J.; Shen, Y.; Li, Y.; Paz, Y. The Structural, Photocatalytic Property Characterization and Enhanced Photocatalytic Activities of Novel Photocatalysts Bi2GaSbO7 and Bi2InSbO7 during Visible Light Irradiation. Materials 2016, 9, 801. https://doi.org/10.3390/ma9100801

AMA Style

Luan J, Shen Y, Li Y, Paz Y. The Structural, Photocatalytic Property Characterization and Enhanced Photocatalytic Activities of Novel Photocatalysts Bi2GaSbO7 and Bi2InSbO7 during Visible Light Irradiation. Materials. 2016; 9(10):801. https://doi.org/10.3390/ma9100801

Chicago/Turabian Style

Luan, Jingfei, Yue Shen, Yanyan Li, and Yaron Paz. 2016. "The Structural, Photocatalytic Property Characterization and Enhanced Photocatalytic Activities of Novel Photocatalysts Bi2GaSbO7 and Bi2InSbO7 during Visible Light Irradiation" Materials 9, no. 10: 801. https://doi.org/10.3390/ma9100801

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop