Next Article in Journal
Tuning the Magnetic Moment of Small Late 3d-Transition-Metal Oxide Clusters by Selectively Mixing the Transition-Metal Constituents
Previous Article in Journal
Characterization of Commercial Metal Oxide Nanomaterials: Crystalline Phase, Particle Size and Specific Surface Area
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Spherical Bi2WO6/Bi2S3/MoS2 n-p Heterojunction with Excellent Visible-Light Photocatalytic Reduction Cr(VI) Activity

Key Laboratory of Fine Chemicals in Universities of Shandong, School of Chemistry and Chemical Engineering, Qilu University of Technology (Shandong Academy of Sciences), Jinan 250353, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(9), 1813; https://doi.org/10.3390/nano10091813
Submission received: 9 August 2020 / Revised: 3 September 2020 / Accepted: 8 September 2020 / Published: 11 September 2020

Abstract

:
Exploiting excellent photocatalytic activity and stable heterostructure composites are of critical importance for environmental sustainability. The spherical Bi2WO6/Bi2S3/MoS2 n-p heterojunction is first prepared via an in situ hydrothermal method using Bi2WO6, Na2MoO4·2H2O, and CH4N2S, in which the intermediate phase Bi2S3 is formed due to chemical coupling interaction of Bi2WO6 and CH4N2S. Scanning electron microscopy indicates that the compactness of the sample can be easily adjusted by changing the contents of S and Mo sources in the solution. The results of ultraviolet–visible (UV–vis) diffuse reflectance spectra, photoluminescence, transient photocurrent response, and electrochemical impedance spectra indicate that the formation of heterojunctions contributes to enhancing visible-light utilization and promoting photogenerated carrier separation and transfer. The composite material is used as a catalyst for the visible light photocatalytic reduction of Cr(VI). Remarkably, the optimal Bi2WO6/Bi2S3/MoS2 n-p heterojunction achieves the greatest Cr(VI) reduction rate of 100% within 75 min (λ > 420 nm, pH = 2); this rate is considerably better than the Cr(VI) reduction rate of pure Bi2WO6. The recycling experiment also reveals that the photocatalytic performance of the n-p heterojunction toward Cr(VI) is still maintained at 80% after three cycles, indicating that the n-p heterojunction has excellent structural stability. The capture experiment proves that the main active species in the system are electrons. The reasonable mechanism of Bi2WO6/Bi2S3/MoS2 photocatalytic reduction Cr(VI) is proposed. Our work provides new research ideas for the design of ternary heterojunction composites and new strategies for the development of photocatalysts for wastewater treatment.

Graphical Abstract

1. Introduction

With the rapid development of industrial production, the pollution of water and land resources by heavy metals is becoming increasingly serious [1,2,3]. Water-soluble, non-biodegradable hexavalent chromium Cr(VI), which easily penetrates through food chain enrichment, can induce cellular oxidative stress, leading to DNA damage, gene mutation, fetal malformation, and carcinogenesis, is one of the most dangerous heavy metals [4,5]. Cr(VI) is widely applied in smelting, electroplating, painting, chemical manufacturing, and tanned leather [6,7,8]. The World Health Organization stipulates that the maximum limit of pollutants in surface water is 0.1 mg L−1. However, the concentration of Cr(VI) in sewage is usually higher than 100 mg L−1 [9]. Therefore, the development of economical and efficient wastewater chromium removal technology is significant. At present, the methods for reducing the Cr(VI) concentration in aqueous solutions mainly include electrochemical precipitation, ion exchange, membrane filtration, photocatalytic degradation, and adsorption [10,11,12,13,14]. The reduction of Cr(VI) to less toxic Cr(III) by semiconductor photocatalytic technology is considered an effective, low-cost method that generates no harmful substances [15,16,17].
Bismuth-based photocatalysts have always been a research hotspot in the field of photocatalysis. Bismuth tungstate (Bi2WO6) has been widely studied for its safety, cheap, proper band gap (~2.8 eV), high stability, and excellent photocatalytic activity [18,19,20]. However, bare Bi2WO6 has the disadvantages of high photo-generated electron-holes recombination efficiency, narrow light absorption range, small specific surface area, and weak surface adsorption capacity, which makes it exhibit poor photocatalytic performance [21,22,23]. Therefore, a scientific strategy that enhances the performance of photocatalysts in practical applications must be developed.
At present, the methods used to improve the photocatalytic activity of semiconductors are as follows: nanostructure modification [24], surface engineering, and homojunction/heterostructure construction. Among these methods, the construction of heterojunctions is considered the most simple and efficient [25,26,27]. The reason is that the establishment of heterojunction cannot only effectively broaden the range of light response and enhance the light absorption of catalyst but also achieve the effective separation of photo-generated carriers under the action of internal electric field and improve the catalytic activity [28,29]. Many Bi2WO6-based heterojunction photocatalysts, such as Bi2WO6/MoS2 [30], Bi2WO6/Fe2O3 [25], Bi2WO6/Bi2S3 [31], CdS/Bi/Bi2WO6 [32], meso-tetra (4-carboxyphenyl) porphyrin/rGO/Bi2WO6 [33], and Co3O4/Ag/Bi2WO6 [34], have been used to improve the performance of pure Bi2WO6. Huang et al. successfully prepared a new flower-shaped AgBr/Bi2WO6 catalyst, which showed good catalytic performance in the degradation of tetracycline (TC) under visible light (vis-light) irradiation [35]. Wan et al. prepared Au/Bi2WO6–MoS2 heterojunction photocatalysts, which exhibited excellent vis-light photocatalytic activity in Cr(VI) and tetracycline hydrochloride degradation [23]. Xue et al. synthesized new g-C3N4/Bi2WO6/AgI catalyst by a hierarchical assembly method. Compared with bare Bi2WO6, the ternary heterojunction composite has stronger redox capacity and exhibits better catalytic activity during the photodegradation of organic pollutants such as TC [36]. Long et al. prepared 3D flower-like MoS2/Bi2S3 heterostructures with excellent photocatalytic activity toward the photodegradation of low concentrations of organic pollutants [37]. The above results show that the successful construction of multi-component heterostructure can effectively improve the photocatalytic activity compared with a single component. However, the Bi2WO6-based photocatalysts still have disadvantages, including a complex preparation method, narrow light response range, and rapid photogenerated charge carrier recombination. In recent years, the use of MoS2 and Bi2S3 coupled with other semiconductors for boosted photocatalytic performance has been widely investigated. To our knowledge, the preparation and application in photocatalysis of the Bi2WO6/Bi2S3/MoS2 ternary heterojunction has not been reported. Hence, seeking a facile and controllable preparation method to fabricate ternary heterojunctions containing Bi2WO6, Bi2S3, and MoS2 is of great importance for improving the photocatalytic performance of Bi2WO6 in environmental purification.
Based on the above considerations, in the present work, Bi2WO6/Bi2S3/MoS2 heterojunction ternary composite materials are prepared via hydrothermal method using the synthesized Bi2WO6 microspheres as substrate (Scheme 1) and which are used for photocatalytic reduction of Cr(VI) to Cr(III). Importantly, the formation of Bi2S3 does not require an additional Bi source, and S2− partially replaces WO66− in Bi2WO6. This process maintains a superior spherical structure and is beneficial to the uniform distribution of Bi2S3 in the composite. The compactness of composite nanoflakes can be easily adjusted by changing the content of sodium molybdate dihydrate (Na2MoO4·2H2O) and thiourea (CH4N2S) in the solution during the hydrothermal process. To the best of our knowledge, the Bi2WO6/Bi2S3/MoS2 heterojunction for Cr(VI) photocatalytic reduction under vis-light irradiation is investigated for the first time. The composites have higher adsorption capacity and photocatalytic activity than pure Bi2WO6 in terms of Cr(VI) reduction due to the successful construction of heterojunction structure. The synergistic effect among the three components enhances light absorption and realizes the effective separation and transmission of photogenerated carriers. The Cr(VI) reduction rate of Bi2WO6/Bi2S3/MoS2 reaches 100% within 75 min (λ > 420 nm, pH = 2) and is considerably better than that of the pure Bi2WO6. These results provide new research ideas for the design of ternary heterojunctions to develop highly efficient vis-light-driven photocatalysts for wastewater treatment.

2. Experimental Section

2.1. Materials and Chemicals

All chemicals and materials in this work were of analytical grade and purchased from commercial suppliers, which could be directly utilized without any further purification. Sodium tungstate dihydrate (Na2WO4·2H2O, AR, 99.5%), Na2MoO4·2H2O (AR, 99.0%), bismuth(III) nitrate pentahydrate (Bi(NO3)3·5H2O, AR, 99.0%), and CH4N2S (AR, 99.0%) were acquired from Sigma-Aldrich (St. Louis, MO, USA). Absolute ethanol (C2H5OH, AR, ≥99.7%), glacial acetic acid (CH3COOH, AR, ≥99.5%), polyvinyl pyrrolidone K30 (PVP K30, AR), and other chemicals used in the experiments were bought from Shanghai Chemical Reagent Co., Ltd. (Shanghai, China). Ultrapure water (18.2 MΩ cm−1) was served throughout the study and acquired from the Milli-Q water purifying system (Millipore Corp., Bedford, MA, USA).

2.2. Synthesis of Spherical Bi2WO6 Nanostructures

Solutions A and B were prepared in the synthesis of Bi2WO6 precursor. In solution A, 2 mmol Bi(NO3)3·5H2O and 4 g PVP K30 were added to a mixed solution of 50 mL ultrapure water, absolute ethanol, and glacial acetic acid with a 3:1:1 volume ratio and then stirred at room temperature for 60 min. In solution B, 1 mmol Na2WO4·2H2O was added to 20 mL H2O for 30 min of ultrasonication. After the solutions were clarified, solution B was dropped to solution A under agitation and stirred continuously for 60 min to obtain a white uniform suspension. The suspension was transferred to a Teflon-sealed autoclave (100 mL) for a solvothermal reaction at 180 °C for 18 h. After cooling, the light-yellow product was collected, washed thrice with absolute ethanol and ultrapure water in sequence, and finally dried overnight and ground for reserves.

2.3. Synthesis of Bi2WO6/Bi2S3/MoS2 n-p Heterojunction Photocatalyst

Bi2WO6/Bi2S3/MoS2 n-p heterojunction nanocomposites were prepared by a simple hydrothermal reaction. First, 200 mg Bi2WO6 was dispersed in 40 mL water with ultrasonic treatment for 10 min. Further 1 h stirring treatment was needed after the addition of 200 mg Na2MoO4·2H2O and 400 mg CH4N2S as an ion source. Second, the dispersion was transferred to 100 mL Teflon-sealed autoclave for hydrothermal reaction at 200 °C for 24 h. Wait for cooling, the obtained sample (named as product BBM-3) was rinsed with ultrapure water and anhydrous ethanol for thrice and then dried overnight at 60 °C in a vacuum oven. As a control, the addition amount of Na2MoO4·2H2O:CH4N2S was adjusted (80 mg:160 mg, 120 mg:240 mg, 300 mg:600 mg), and the corresponding products with BBM-1, BBM-2, and BBM-4 were expressed. In addition, the preparation method of pure MoS2 nanosheets was similar to the above process except that Bi2WO6 was not added.

2.4. Characterization

Powder X-ray diffraction (XRD) patterns were obtained on a SmartLab SE X-ray diffractometer (Rigaku Corp., Tokyo, Japan) with Cu Kα (λ = 1.5046 Å) radiation. Raman spectra were measured using a Renishaw in Via9 Raman microscope system (Renishaw, London, UK) with a 50× objective and a 532 nm laser irradiation to focal point the laser beam into a spot with a diameter of approximately 1 µm. The morphology and energy-dispersive spectra (EDS) of the samples were tested and characterized by a field-emission scanning electron microscope (FESEM, Regulus 8220, Hitachi, Tokyo, Japan). The microstructure and lattice fringe of the samples were examined by a transmission electron microscope (TEM, JEM-2100, JEOL, Tokyo, Japan) and high-resolution TEM (HRTEM, JEM-2100). The elemental composition and chemical state of the samples were determined by X-ray photoelectron spectroscopy (XPS) (Perkin-Elmer PHI5300 spectrometer, Perkin Elmer, Waltham, MA, USA). The specific surface areas of the samples were obtained based on the N2 adsorption–desorption isotherm tested on a Micromeritics ASAP 2460 system (ASAP, Norcross, GA, USA). The ultraviolet–visible (UV–vis) diffuse reflectance spectra (DRS) were tested within a 200 nm to 800 nm wavelength range using a spectrometer (UV-2600, Shimadzu, Kyoto, Japan) with BaSO4 as a reference. A Hitachi F4500 fluorescence spectrophotometer (Hitachi, Tokyo, Japan) was used to test the photoluminescence (PL) measurements (λexcitation = 300 nm). The Mott–Schottky curves, electrochemical impedance spectra (EIS), and photocurrent response experiments were carried out using the electrochemical workstation (PARSTAT 4000, Ametec, Berwyn, PA, USA) with a conventional three-electrode configuration (working electrode: fluorine-doped tin oxide conducting glass; counter electrode: platinum plate; reference electrode: Ag/AgCl electrode) and Na2SO4 aqueous solution as the electrolyte (0.1 mol L−1).

2.5. Photocatalytic Activity Experiments

The specific operational steps of the prefabricated catalysts for photocatalytic reduction of Cr(VI) (Cr(VI) source: K2Cr2O7) were as follows: Cr(VI) solution with concentration of 40 mg L−1 was prepared using ultrapure water as solvent. Then, 50 mL of this initial solution was accurately measured and placed in the reaction vessel. Next, the initial solution pH to 2 was adjusted with HCl solution (1 mol L−1). Afterward, the 20 mg as-synthesized catalysts were evenly dispersed in the solution by ultrasonication. Before vis-light irradiation, the suspension was stored in the dark place and stirred for 60 min to reach equilibrium of adsorption and desorption. Under the irradiation by Xe lamp (300 W, 100 mW cm−2, λ > 420 nm), 3 mL suspension was taken out in the reaction container every 15 min and centrifuged (9000 r min−1, 10 min). Then, the supernatant was collected with microporous (0.22 μm) membrane filter syringe to eliminate residual particles. NaOH (1 mol L−1) or HCl (1 mol L−1) solution was used to adjust the pH of Cr(VI) solution to investigate the effect of solution pH on photocatalysis. Finally, the Cr(VI) concentration was obtained by measuring the absorbance of the supernatant at 540 nm (UV-2600, Shimadzu) with diphenylcarbazide approach (Electronic Supplementary Materials).

3. Results and Discussion

3.1. X-ray Diffraction (XRD) and Raman Analysis

Figure 1a shows the XRD patterns of the products. For Bi2WO6, the diffraction peaks at 2θ = 28.3°, 32.8°, 47.1°, 56.0°, 58.5°, 68.8°, 76.1°, and 78.5° correspond to the (131), (200), (202), (133), (262), (400), (2102), and (204) crystal faces of the Bi2WO6 orthorhombic phase (JCPDS Card No.39-0256), respectively [38,39]. The three diffraction peaks of bare MoS2 at 9.0°, 32.0°, and 58.0° correspond to the (002), (100), and (110) crystal faces of 2H-MoS2, respectively. The peaks at 9.0° and 17.0° indicate the formation of a layered structure with enlarged interlayer spacing [40]. For heterojunction photocatalysts, the XRD patterns display new diffraction peaks. The diffraction peaks located at 25.0° correspond to the (130) crystal plane of Bi2S3 [41]. Given the strong interaction force between Bi3+ and S2−, Bi2S3 will be formed at relatively high temperatures [42]. The Bi2WO6/Bi2S3/MoS2 samples show a discernible peak at approximately 32.0°, which is attributed to the (100) crystal plane of MoS2, indicating that the composite material contains MoS2 component. However, the absence of the highest MoS2 intensity peak (~9.0°) from the heterostructure samples indicates that MoS2 nanosheets may contain only a few layers that are too thin to be detected by XRD [43].
Raman measurements of the as-synthesized samples were performed in the range of 200–1200 cm−1, and the results are shown in Figure 1b. The black, green, and red dashed lines in the figure represent the Raman characteristic peaks of Bi2WO6, Bi2S3, and MoS2, respectively. The peaks of 308, 723, 798, and 829 cm−1 in the Raman spectra are characteristic Raman shifts of Bi2WO6 [44]. The Raman peaks of Bi2S3 are located at 234.8, 260, 590, and 970 cm−1, of which the peaks at 234.8 and 260 cm−1 matched the Ag1 and B1g vibration mode, respectively [45]. Meanwhile, the typical peaks at 383 and 408 cm−1 are ascribed to the E12g and A1g vibrations of MoS2, respectively [42,45]. Based on the above results, the Bi2WO6/Bi2S3/MoS2 ternary composites are successfully prepared.

3.2. Morphology

As shown in the scanning electron microscope (SEM) images (Figure 2a), Bi2WO6 microspheres with diameters of 2.6–3.0 μm are self-assembled from nanosheets. The SEM images in Figure 2b–e show that the degree of looseness of the Bi2WO6/Bi2S3/MoS2 microsphere increases accordingly with the increase in Mo and S sources concentration during the hydrothermal process. Nevertheless, with the further increase in concentration, MoS2 agglomerates are formed on the Bi2WO6/Bi2S3/MoS2 surface, and the corresponding results are shown in Figure 2e. This finding is consistent with the information expressed in the TEM diagram in Figure 2f–j. The causes of these phenomenon are as follows: I. A strong affinity exists between Bi3+ and S2−, which reacts under high temperature and pressure to form Bi2S3 (Bi2WO6 + 3S2− → Bi2S3 + WO66−) [42,45]. II. Bi2WO6 is consumed during this process, and Bi2S3 and MoS2 are generated. As the consumption of Bi2WO6 increases, the structure becomes looser. III. The formation of MoS2 agglomerates is mainly caused by the excessively high concentration of Mo and S sources, which promotes the nucleation speed to extreme degrees [46]. Soon afterward, BBM-3 is used as the model, and its composition is characterized by high-resolution transmission electron microscopy (HRTEM). Figure 2k shows the tight interface between Bi2WO6, Bi2S3, and MoS2 in the BBM-3. The measured interplanar distances of 0.315, 0.360, and 0.620 nm belong to the (131) crystal plane of orthorhombic Bi2WO6, the (130) plane of Bi2S3, and the (002) lattice plane of MoS2, respectively [47,48,49]. The high-magnification TEM of BBM-3 (Figure S1), which can intuitively illustrate the close interface contact between Bi2WO6, Bi2S3, and MoS2, confirms the successful construction of the Bi2WO6/Bi2S3/MoS2 heterojunction [42]. Meanwhile, element mapping is used to analyze BBM-3 in depth to further determine the distribution of Mo, S, Bi, W, and O in the material. The results (Figure 2l–p) coincide with the EDS characterization results (Figure S2), confirming that BBM-3 consists of Mo, S, Bi, W, and O. The above results confirm that Bi2WO6/Bi2S3/MoS2 n-p heterojunction with spherical structure can be synthesized by a simple in-situ hydrothermal method.

3.3. X-ray Photoelectron Spectroscopy (XPS) Analysis

The survey XPS curves in Figure 3a indicate that BBM-3 is composed of Mo, Bi, S, W, and O. Figure 3b–e shows the high-resolution spectra of Mo 3d, Bi 4f, W 4f, and O1s, respectively. The Mo 3d (Figure 3b) shows two peaks centered at 227.6 and 230.8 eV, which correspond to Mo 3d5/2 and Mo 3d3/2 of Mo4+, respectively [50]. The satellite peak at approximately 234.7 eV represents Mo6+ [51]. In addition, the mid-strong peak at 225.4 eV can be well matched to S 2s [50]. The characteristic signal in Bi 4f diagram (Figure 3c) is formed by Bi 4f7/2 at 157.4 eV, Bi 4f5/2 at 162.7 eV, and S 2p at 160.6 eV [52]. The difference between the binding energy of Bi 4f7/2 and Bi 4f5/2 is 5.3 eV, indicating that Bi exists in BBM-3 as Bi3+. Figure 3d contains the peaks at 35.2 and 37.6 eV, which are characteristics of W 4f7/2 and W 4f5/2, respectively [53]. The three fitted peaks in the O 1s spectrum (Figure 3e) are located at 531.5, 530.6, and 529.7 eV, which indicates that three types of O are present in BBM-3. The peak at 531.5 eV represents the chemically adsorbed oxygen (O–H) on the surface of BBM-3, whereas the peaks at 530.6 and 529.7 eV correspond to the O–Bi and O–W lattice oxygen in BBM-3, respectively [54]. Compared with the peaks of pure Bi2WO6, Bi2S3, and MoS2, the Mo 3d and Bi 4f peaks of BBM-3 display a shift ~1.0 eV to the lower binding energy direction. Conversely, the W 4f peak of BBM-3 shows a shift ~1.0 eV to the higher binding energy direction. These results are primarily due to strong interactions and charge transfer among Bi2WO6, Bi2S3, and MoS2 in BBM-3 (Figure S3) [29]. The above XPS analyses confirm again that Bi2WO6, Bi2S3, and MoS2 coexist in the Bi2WO6/Bi2S3/MoS2 ternary heterojunction photocatalyst.

3.4. Brunauer–Emmett–Teller (BET) Specific Surface Area Analysis

As shown in Figure S4, the Bi2WO6 and Bi2WO6/Bi2S3/MoS2 ternary heterojunction samples exhibit type IV isotherms, which indicate the existence of mesoporous structures [11,22]. The Brunauer–Emmett–Teller (BET) surface area of Bi2WO6/Bi2S3/MoS2 composites is higher than that of bare Bi2WO6 (14.7 m2 g−1). The BET surface areas of BBM-1, BBM-2, BBM-3, and BBM-4 are 16.6, 19.7, 22.7, and 19.4 m2 g−1, respectively. Compared with pure Bi2WO6, Bi2WO6/Bi2S3/MoS2 composites have high BET surface areas and rich mesoporous structures, which facilitate the adsorption and reduction of Cr(VI).

3.5. Ultraviolet–Visible (UV–Vis) Absorption and Band Gap Positions

The DRS of the pristine Bi2WO6, pristine MoS2, and ternary composites are recorded to investigate the light absorption of the samples. As shown in Figure 4a, for the pristine Bi2WO6, its intrinsic light absorption edge is at 450 nm, which means that the material has light absorption only in the UV and partially visible regions. By contrast, pure MoS2 shows a strong absorption in the UV–vis region. As expected, Bi2WO6/Bi2S3/MoS2 ternary heterojunction photocatalyst extends the vis-light absorption range compared with the Bi2WO6. Thus, the good photocatalysis performance of the composite is predicted. Furthermore, the bandgap energy (Eg) of as-fabricated materials is obtained in accordance with Tauc’s equation (Equation (1)) [44,55]:
( α h ν )   =   A ( h ν     E g ) n / 2 ,
where α: absorption coefficient, h: Planck’s constant, ν: light frequency, and A: a constant.
The value of n depends on the type of electronic transition in the semiconductor (n values of direct/indirect transition: 1/4). According to the previous reports, the n of Bi2WO6 and MoS2 is 1, and their Eg are determined by the extrapolation of Tauc linear region [28,56]. The Eg of pure Bi2WO6 and MoS2 are ~2.74 and 1.30 eV, respectively (Figure 4b,c), which are close to previously reported values [30,36]. The Eg of BBM-1, BBM-2, BBM-3, and BBM-4 composites are ~1.40, 1.34, 1.15, and 1.32 eV, respectively (Figure 4d–g).
On the above basis, the valence band (VB) and conduction band (CB) edge potentials of the samples are calculated in accordance with the Mulliken atomic electronegativity theory (Equations (2) and (3), respectively) [30,51]:
E CB   +   0.5 E g   =   X     E e ,
E VB   =   E CB   +   E g ,
where Eg and X represent the band gap energy and absolute electronegativity, respectively. EVB and ECB are the VB and CB edge, respectively. Ee is energy of free electrons (~4.5 eV) on the hydrogen scale. Table 1 shows the calculation results of material-related parameters. The flat-band potentials of the related materials are studied by using Mott–Schottky curves (Figure S5) to verify the rationality of the calculation results. As presented in Figure S5a–c, Bi2WO6 is classified as an n-type semiconductor due to its positive slope, whereas Bi2S3 and MoS2 are confirmed as p-type semiconductors due to their negative slopes. When they are coupled to each other to form a n-p heterojunction (Bi2WO6/Bi2S3/MoS2), the Mott–Schottky curve shows an inverted ‘V-shape’ characteristic (Figure S5d). Generally, EVB for p-type semiconductors is very close to the flat-band potential, whereas ECB for n-type semiconductors is very close to the flat-band potential [57]. The flat-band potential in the n-type semiconductor is 0.1–0.3 eV higher than ECB, whereas that in the p-type semiconductor is 0.1–0.3 eV lower than EVB [58]. Figure S5a–c shows that the flat-band potentials of pure Bi2WO6, Bi2S3, and MoS2 can be confirmed to be 0.20 (0.40 eV vs. normal hydrogen electrode (NHE)), 1.06 (1.26 eV vs. NHE), and 1.14 (1.34 eV vs. NHE), respectively. Therefore, ECB of pure Bi2WO6 and EVB of Bi2S3 and MoS2 can be estimated to be 0.33, 1.37, and 1.47 eV. On the basis of Equation (3), the corresponding EVB of Bi2WO6 and ECB of Bi2S3 and MoS2 can occur at approximately 3.07, 0.18, and 0.17 eV. These results are in agreement with the result calculated in accordance with Mulliken atomic electronegativity theory.

3.6. Photoelectrochemical Performance

The catalyst is further characterized by PL spectroscopy and photocurrent response to explore its charge separation efficiency. High PL intensity indicates low charge separation efficiency and easy electron-hole recombination, whereas the photocurrent response shows the opposite [35,59,60]. The PL spectra (Figure 5a) of Bi2WO6 and Bi2WO6/Bi2S3/MoS2 composites are obtained under the condition of λexcitation = 300 nm. The emission intensities of all Bi2WO6/Bi2S3/MoS2 composites are significantly lower than that of bare Bi2WO6. Based on the intensity, the composites can be sorted as Bi2WO6 > BBM-1 > BBM-2 > BBM-4 > BBM-3. This result shows that the successful recombination of MoS2, Bi2S3, and Bi2WO6 improves the efficiency of charge separation. The photocurrent response (Figure S6) confirms this conclusion. In the experimental process of up to 400 s, BBM-3 consistently shows the highest photocurrent. The EIS test can be used to explore the interface charge transfer properties, with the small arc radius reflecting a fast charge transfer speed [61]. Figure 5b shows the Nyquist plots of Bi2WO6 and composites. The composites exhibit a smaller Nyquist plot semicircle radius compared with pure Bi2WO6. BBM-3 also shows a considerably smaller semicircle radius of EIS Nyquist plots than the other composites (BBM-1, BBM-2, and BBM-4), which is highly consistent with the PL and photocurrent test analysis results. Therefore, the following conclusions can be drawn. First, the construction of Bi2WO6, Bi2S3, and MoS2 heterostructures can significantly improve the charge separation efficiency. Second, only when Bi2WO6 is compounded with suitable amount of Bi2S3 and MoS2 can n-p heterojunction photocatalysts be formed effectively. Thus, extremely high and extremely low compounding ratios are not conducive to the formation of heterostructures and the separation and transfer of photogenerated carriers. Third, BBM-3 is expected to have the best vis-light catalytic activity because it enables the effective separation of photo-generated carriers.

3.7. Photocatalytic Activity

In this working system, the catalytic reduction of Cr(VI) under vis-light irradiation is used as the evaluation standard to evaluate the performance of the prefabricated materials. Previous reports have shown that the initial solution pH strongly influences the photocatalytic Cr(VI) reduction. Therefore, the pH of the initial solution is adjusted to show a linear gradient change, which is used to investigate the effect of pH on the catalytic activity of BBM-3. The photocatalytic reduction efficiency of BBM-3 is the highest under acidic conditions (Figure 6a). Under the condition of pH = 2.00, the Cr(VI) reduction rate of BBM-3 is as high as 100%. With the increase in pH, the reduction rate of Cr(VI) shows a strictly decreasing trend. When pH = 10.00, the reduction efficiency of Cr(VI) reaches 14%. This change is confirmed by the corresponding UV–vis absorption spectra (Figure 6b–d and Figure S7). The above situation is mainly caused by the following factors. First, Cr(VI) mainly exists in the form of HCrO4 and Cr2O72− in acidic environments and CrO42− in alkaline environments [62]. When the solution environment is strongly acidic, the hydroxyl groups on the surface of the catalyst will be protonated to become (–OH2+), which in turn enhances the electrostatic adsorption on Cr(VI) [61]. Second, the reactions in acidic conditions are as follows (Equations (4) and (5)) [17,61,63]:
HCrO 4   +   7 H +   +   3 e   Cr 3 +   +   4 H 2 O ,
Cr 2 O 7 2   +   14 H +   +   6 e     2 Cr 3 +   +   7 H 2 O ,
The reaction under alkaline conditions is as follows (Equation (6)) [17,61,63]:
CrO 4 2   +   4 H 2 O   + 3 e     Cr ( OH ) 3   +   5 OH ,
Equation (6) shows that Cr3+ will be converted into Cr(OH)3 and deposited on the catalyst surface under alkaline conditions, blocking the active sites that can be used for adsorption and photocatalytic reactions. In summary, acidic conditions are more conducive to the reduction of Cr(VI) than alkaline conditions.
The photocatalytic activities of Bi2WO6 and different Bi2WO6/Bi2S3/MoS2 composites are evaluated at the initial solution pH = 2.00, and the results are shown in Figure 7a. The composites exhibit high Cr(VI) adsorption performance compared with pure Bi2WO6, which benefits from the high surface area and abundant mesoporous structure of the composite. After 75 min of irradiation, the reduction rate of Bi2WO6 to Cr(VI) is 5%. By contrast, all the Bi2WO6/Bi2S3/MoS2 composites manifest remarkably high photocatalytic reduction activity under the same conditions. In particular, BBM-3 shows the best photocatalytic performance with a corresponding Cr(VI) reduction rate of up to 100%. It is not difficult to find that there is an optimal compounding ratio between Bi2S3, MoS2, and Bi2WO6 occurs in the removal of Cr(VI). Extremely low or extremely high compounding ratio is not conducive to enhancing the photoreduction activity of Bi2WO6/Bi2S3/MoS2. When the actual ratio is lower than the optimal ratio, the number of active sites used to capture carriers increases with the increase in recombination ratio, thus prolonging the carrier lifetime and then increasing the photocatalytic activity. However, when the compounding ratio is higher than the optimum, excess MoS2 will agglomerate to disrupt the effective construction of n-p heterojunction, as confirmed by the SEM result (Figure 2e).
Figure 7b shows the pseudo first-order kinetic curves of Bi2WO6, BBM-1, BBM-2, BBM-3, and BBM-4 for the photocatalytic reduction of Cr(VI) and the apparent reaction rate constant “k” (Figure 7c). The pseudo first-order model demonstrated here is shown by Equation (7) [41]:
l n ( C 0 C t )   =   k t ,
where t stands for the vis-light exposure time, C0 represents the original concentration of Cr(VI) solution, and Ct is the concentration of Cr(VI) solution at “t” irradiation time. The “k” values calculated by the linear fit of ln(C0/Ct) and irradiation time (min) plots are 0.037, 0.223, 0.628, 3.612, and 1.379 × 10−2 min−1 for Bi2WO6, BBM-1, BBM-2, BBM-3, and BBM-4, respectively. BBM-3 obtains the highest k value. Immediately afterwards, an active species capture experiment is carried out to further explore the mechanism involved in the reaction system (hole/electron trapping agent: citric acid/KBrO3). The obtained results are shown in Figure 7d. First, the Cr(VI) solution without photocatalyst shows good stability under vis-light exposure. However, KBrO3 significantly inhibited the photoreduction of Cr(VI) by BBM-3 with a final reduction rate of 75%. This finding indicates that the main active material in the catalytic reduction of Cr(VI) process is photogenerated electrons, which is consistent with previous reports [64]. By contrast, with the addition of citric acid, the adsorption and reduction rate of Cr(VI) by BBM-3 are significantly improved. The factors that cause this phenomenon are as follows: First, the surface of the catalyst becomes more positive by the addition of citric acid, which promotes the adsorption of HCrO4 or Cr2O72− ions [65]. Second, photogenerated holes can oxidize citric acid, which is equivalent to promoting the separation of photogenerated carriers and prolonging the lifetime of photogenerated electrons [61,66]. The above conclusion is confirmed by the corresponding UV–vis absorption spectra in Figure S8. Finally, compared with the other reported catalysts for vis-light reduction of Cr(VI) (Table 2), Bi2WO6/Bi2S3/MoS2 heterojunction composites show relatively satisfactory photocatalytic activity.
Issues such as photocatalysts separation, recovery, reuse, and stability, are important in practical applications. The stability and reusability of BBM-3 after the reaction is proven by recycling and reusing the same catalyst for three cycles. The photoreduction rate of Cr(VI) after three cycle tests is 80% (Figure 8a), which shows that the photocatalysts have enough stability and reusability. This is mainly attributed to the existence of electrostatic attraction that induces a strengthened coupling interaction among Bi2WO6, Bi2S3, and MoS2; this interaction is beneficial to improving structural stability [51,52]. Figure S9 shows the UV–vis absorption spectra after the second and third cycles. Furthermore, we collected composite samples after use in three photocatalytic cycles (BBM-3-a) and further characterized them by XRD and XPS. The positions of the characteristic peaks of the samples after circulation (Figure 8b) exhibit no change compared with the initial sample. The evident signals of Bi, W, S, Mo, O, and Cr contaminants can be observed in the survey XPS curves of BBM-3-a (Figure 8c). The XPS peak of 577.1 eV in the Cr 2p spectrum (Figure 8d) belongs to Cr 2p3/2, which highly matches Cr(III) in Cr(OH)3 [73]. To sum up, Bi2WO6/Bi2S3/MoS2 heterojunction composite has high structural stability, good reusability, and can effectively reduce the toxicity of Cr(VI) to reduce it to Cr(III).

3.8. Possible Photocatalytic Mechanism

We can now tentatively explain the mechanism underlying heterojunctions in photocatalytic reactions. When n-type Bi2WO6 is coupled with p-type Bi2S3 and MoS2, the n-p heterojunction is formed among semiconductors. The formation of the n-p heterojunction results in the equalization of their Fermi levels. This effect, in turn, induces band bending and a strong electric field at their interface. In this case, electrons and holes are prevented from coming into contact with each other due to the built-in electric field [74]. Meanwhile, according to the energy band structure, a type-I straddling heterojunction forms on the interface of MoS2 and Bi2S3, whereas a traditional type-II staggered heterojunction forms on the interface of Bi2S3 and Bi2WO6. MoS2 and Bi2S3 are excited when type-I MoS2/Bi2S3 are exposed to vis-light. The electrons on the CB of MoS2 will quickly transfer onto that of Bi2S3, and the holes on the VB of the MoS2 simultaneously hop onto that of Bi2S3. If no measures are taken, the electrons and holes accumulate in the Bi2S3 semiconductor and recombine rapidly. Interestingly, the existence of type-II Bi2S3/Bi2WO6 enables the electrons on the CB of the p-type Bi2S3 semiconductor to transfer directly onto the CB of n-type Bi2WO6, and the holes on the VB of Bi2WO6 can be spontaneously injected into the VB of Bi2S3. This phenomenon realizes the effective separation and transfer of photogenerated electron-hole pairs. The strong electric field generated by the n-p heterojunction further promotes this process. Compared with a single heterostructure, this system can better realize the separation and transfer of photogenerated electron–hole pairs under the joint action of multiple heterostructures. Given the above analysis, the proposed photocatalytic reduction mechanism of Cr(VI) by Bi2WO6/Bi2S3/MoS2 photocatalysis under vis-light is obtained (Scheme 2). When vis-light is irradiated on the Bi2WO6/Bi2S3/MoS2 heterojunction, electrons on the semiconductor VB are excited to CB and the corresponding numbers of holes are retained on the VB, thus forming photogenerated electron–hole pairs (Equations (8)–(10)). Electrons on the CB of type-p Bi2S3 and MoS2 are transferred to CB of type-n Bi2WO6, which is finally used for Cr(VI) reduction, whereas the holes remain in the VB of Bi2S3 (Equation (11)). Moreover, the CB edges of Bi2WO6, Bi2S3, and MoS2 are more negative than the reduction potential of E(Cr(VI)/Cr(III)) (0.51 eV) [61,67]. Theoretically, the reduction of Cr(VI) to Cr(III) can be feasibly achieved by this route. Eventually, the electrons in the system will reduce Cr2O72− to Cr(III), and the holes will oxidize H2O to produce O2 (Equations (12) and (13), respectively).
Bi 2 WO 6   +   h v     e CB   +   h + VB ,
Bi 2 S 3   +   h v     e CB   +   h + VB ,
MoS 2   +   h v     e CB   +   h + VB ,
Bi 2 WO 6 ( e CB )     Bi 2 S 3 ( e CB )     MoS 2 ( e CB )   +   Bi 2 WO 6 ( h + VB )     Bi 2 S 3 ( h + VB )     MoS 2 ( h + VB )     Bi 2 WO 6 ( total e CB )   +   Bi 2 S 3 ( total h + VB )   ,
Cr 2 O 7 2   +   14 H +   +   6 e     2 Cr 3 +   +   7 H 2 O ,
2 H 2 O   +   4 h +     O 2   +   4 H + ,

4. Conclusions

In summary, spherical Bi2WO6/Bi2S3/MoS2 n-p heterojunction ternary composites with vis-light response are prepared by a hydrothermal method. The HRTEM and Mott–Schottky curves confirm the formation of the n-p heterojunction. The XPS spectra show the existence of strong interaction and charge transfer among Bi2WO6, Bi2S3, and MoS2 in the n-p heterojunction. The effects of various factors on the catalytic activity of Bi2WO6/Bi2S3/MoS2 photocatalysts are investigated. For vis-light photocatalytic reduction of Cr(VI), the composites show higher photocatalytic reduction capacity than pure Bi2WO6, where BBM-3 exhibits the highest photocatalytic activity with a corresponding Cr(VI) reduction rate of up to 100% within 75 min. After three cycles of experiments, XRD and XPS analyses verify that the heterojunction possesses structural stability and can effectively reduce Cr(VI) into Cr(III). The improvement of photocatalytic activity of composite materials mainly benefits from the following points: First, the successful construction of a heterojunction structure forms a good interface contact, which promotes the effective separation of photogenerated electrons and holes. Second, the effective assembly and interfacial synergy between the three components enhance the vis-light absorption capacity of the samples and expand their light absorption range. Third, the increased surface area and abundant mesoporous structure endow the composites with more reactive sites and strong adsorption capacity of pollutants. The successful construction and application of Bi2WO6/Bi2S3/MoS2 n-p heterojunction in this work provide new ideas and strategies for the development of photocatalysts for wastewater treatment.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/9/1813/s1: 1. Determination of Cr(VI) concentration using the DPC method; 2. Characterization data. Figure S1: High-magnification TEM image of BBM-3, Figure S2: EDS analysis of the BBM-3, Figure S3: XPS spectra of Bi2WO6, Bi2S3, MoS2, and BBM-3: (a) survey, (b) Mo 3d, (c) Bi 4f, and (d) W 4f spectra, Figure S4: N2 adsorption–desorption isotherms of the Bi2WO6, BBM-1, BBM-2, BBM-3, and BBM-4, Figure S5: The Mott–Schottky curves of (a) Bi2WO6, (b) Bi2S3, (c) MoS2, and (d) BBM-3, Figure S6: Photocurrent responses of Bi2WO6, BBM-1, BBM-2, BBM-3, and BBM-4, Figure S7: The UV–vis absorption spectra of Cr(VI) solution over the BBM-3 at (a) pH 4.00, and (b) pH 8.00, Figure S8: The UV–vis absorption spectra of Cr(VI) solution over the BBM-3 in the presence of (a) hole scavenger (citric acid) and (b) electron scavenger (KBrO3), Figure S9: The UV–vis absorption spectra of Cr(VI) solution over the BBM-3 in the cycle experiments: (a) 2nd run and (b) 3rd run.

Author Contributions

Investigation, J.R., T.H., Q.G. and T.G.; Methodology, J.R., B.S. and G.Z.; Software, J.R., Q.G. and P.C.; Data curation, J.R., T.H. and Q.W.; Formal analysis, J.R., Q.G. and P.C.; Writing—original draft, J.R. and Q.G.; Visualization, J.R., Q.W., B.S. and T.G.; Writing—review and editing, B.S. and G.Z.; Conceptualization, G.Z.; Resources, G.Z.; Supervision, B.S. and G.Z.; Project administration, G.Z.; Funding acquisition, G.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant Nos. 51972180, 51572134), Key Technology Research and Development Program of Shandong (Grant No. 2019GGX102070), and the Program for Scientific Research Innovation Team in Colleges and Universities of Jinan (Grant No. 2018GXRC006).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cheng, C.; Chen, D.Y.; Li, N.J.; Xu, Q.F.; Li, H.; He, J.H.; Lu, J.M. ZnIn2S4 grown on nitrogen-doped hollow carbon spheres: An advanced catalyst for Cr(VI) reduction. J. Hazard. Mater. 2020, 391, 122205. [Google Scholar] [CrossRef] [PubMed]
  2. Lei, C.S.; Zhu, X.F.; Zhu, B.C.; Jiang, C.J.; Le, Y.; Yu, J.G. Superb adsorption capacity of hierarchical calcined Ni/Mg/Al layered double hydroxides for Congo red and Cr(VI) ions. J. Hazard. Mater. 2017, 321, 801–811. [Google Scholar] [CrossRef]
  3. Anceschi, A.; Caldera, F.; Bertasa, M.; Cecone, C.; Trotta, F.; Bracco, P.; Zanetti, M.; Malandrino, M.; Mallon, P.E.; Scalarone, D. New poly(β-Cyclodextrin)/poly(vinyl alcohol) electrospun sub-micrometric fibers and their potential application for wastewater treatments. Nanomaterials 2020, 10, 482. [Google Scholar] [CrossRef] [Green Version]
  4. Gu, T.Y.; Dai, M.; Young, D.J.; Ren, Z.G.; Lang, J.P. Luminescent Zn(II) coordination polymers for highly selective sensing of Cr(III) and Cr(VI) in water. Inorg. Chem. 2017, 56, 4668–4678. [Google Scholar] [CrossRef]
  5. Li, M.Q.; Mu, Y.; Shang, H.; Mao, C.L.; Cao, S.Y.; Ai, Z.H.; Zhang, L.Z. Phosphate modification enables high efficiency and electron selectivity of nZVI toward Cr(VI) removal. Appl. Catal. B Environ. 2020, 263, 118364. [Google Scholar] [CrossRef]
  6. Pradhan, D.; Sukla, L.B.; Sawyer, M.; Rahman, P.K.S.M. Recent bioreduction of hexavalent chromium in wastewater treatment: A review. J. Ind. Eng. Chem. 2017, 55, 1–20. [Google Scholar] [CrossRef] [Green Version]
  7. Lu, H.J.; Wang, J.K.; Hao, H.X.; Wang, T. Magnetically separable MoS2/Fe3O4/nZVI nanocomposites for the treatment of wastewater containing Cr(VI) and 4-Chlorophenol. Nanomaterials 2017, 7, 303. [Google Scholar] [CrossRef] [Green Version]
  8. Dong, R.F.; Zhong, Y.L.; Chen, D.Y.; Li, N.J.; Xu, Q.F.; Li, H.; He, J.H.; Lu, J.M. Morphology-controlled fabrication of CNT@MoS2/SnS2 nanotubes for promoting photocatalytic reduction of aqueous Cr(VI) under visible light. J. Alloys Compd. 2019, 784, 282–292. [Google Scholar] [CrossRef]
  9. Xu, Y.L.; Chen, J.Y.; Chen, R.; Yu, P.L.; Guo, S.; Wang, X.F. Adsorption and reduction of chromium(VI) from aqueous solution using polypyrrole/calcium rectorite composite adsorbent. Water Res. 2019, 160, 148–157. [Google Scholar] [CrossRef]
  10. Zheng, Y.Q.; Cheng, B.; You, W.; Yu, J.G.; Ho, W.K. 3D hierarchical graphene oxide-NiFe LDH composite with enhanced adsorption affinity to Congo red, methyl orange and Cr(VI) ions. J. Hazard. Mater. 2019, 369, 214–225. [Google Scholar] [CrossRef]
  11. Zheng, X.G.; Chen, Q.; Lv, S.H.; Fu, X.J.; Wen, J.; Liu, X.H. Enhanced visible-light photocatalytic activity of Ag QDs anchored on CeO2 nanosheets with a carbon coating. Nanomaterials 2019, 9, 1643. [Google Scholar] [CrossRef] [Green Version]
  12. Yang, X.; Liu, L.H.; Zhang, M.Z.; Tan, W.F.; Qiu, G.H.; Zheng, L.R. Improved removal capacity of magnetite for Cr(VI) by electrochemical reduction. J. Hazard. Mater. 2019, 374, 26–34. [Google Scholar] [CrossRef]
  13. Gherasim, C.V.; Bourceanu, G.; Olariu, R.I.; Arsene, C. A novel polymer inclusion membrane applied in chromium(VI) separation from aqueous solutions. J. Hazard. Mater. 2011, 197, 244–253. [Google Scholar] [CrossRef]
  14. Shao, Z.C.; Huang, C.; Wu, Q.; Zhao, Y.J.; Xu, W.J.; Liu, Y.Y.; Dang, J.; Hou, H.W. Ion exchange collaborating coordination substitution: More efficient Cr(VI) removal performance of a water-stable CuII-MOF material. J. Hazard. Mater. 2019, 378, 120719. [Google Scholar] [CrossRef]
  15. Yuan, X.Y.; Zhou, C.; Jing, Q.Y.; Tang, Q.; Mu, Y.H.; Du, A.K. Facile synthesis of g-C3N4 nanosheets/ZnO nanocomposites with enhanced photocatalytic activity in reduction of aqueous chromium(VI) under visible light. Nanomaterials 2016, 6, 173. [Google Scholar] [CrossRef] [Green Version]
  16. Li, X.K.; Li, C.X.; Xiang, D.; Zhang, C.M.; Xia, L.; Liu, X.Y.; Zheng, F.Q.; Xie, X.Y.; Zhang, Y.L.; Chen, W. Self-limiting synthesis of Au–Pd core–shell nanocrystals with a near surface alloy and monolayer Pd shell structure and their superior catalytic activity on the conversion of hexavalent chromium. Appl. Catal. B Environ. 2019, 253, 263–270. [Google Scholar] [CrossRef]
  17. Wang, Y.H.; Kang, C.L.; Xiao, K.K.; Wang, X.Y. Fabrication of Bi2S3/MOFs composites without noble metals for enhanced photoreduction of Cr(VI). Sep. Purif. Technol. 2020, 241, 116703. [Google Scholar] [CrossRef]
  18. Huang, H.W.; Cao, R.R.; Yu, S.X.; Xu, K.; Hao, W.C.; Wang, Y.G.; Dong, F.; Zhang, T.R.; Zhang, Y.H. Single-unit-cell layer established Bi2WO6 3D hierarchical architectures: Efficient adsorption, photocatalysis and dye-sensitized photoelectrochemical performance. Appl. Catal. B Environ. 2017, 219, 526–537. [Google Scholar] [CrossRef]
  19. Zhang, K.; Wang, J.; Jiang, W.J.; Yao, W.Q.; Yang, H.P.; Zhu, Y.F. Self-assembled perylene diimide based supramolecular heterojunction with Bi2WO6 for efficient visible-light-driven photocatalysis. Appl. Catal. B Environ. 2018, 232, 175–181. [Google Scholar] [CrossRef]
  20. Ma, Y.C.; Lv, C.; Hou, J.H.; Yuan, S.T.; Wang, Y.R.; Xu, P.; Gao, G.; Shi, J.S. 3D hollow hierarchical structures based on 1D BiOCl nanorods intersected with 2D Bi2WO6 nanosheets for efficient photocatalysis under visible light. Nanomaterials 2019, 9, 322. [Google Scholar] [CrossRef] [Green Version]
  21. Cao, R.R.; Huang, H.W.; Tian, N.; Zhang, Y.H.; Guo, Y.X.; Zhang, T.R. Novel Y doped Bi2WO6 photocatalyst: Hydrothermal fabrication, characterization and enhanced visible-light-driven photocatalytic activity for Rhodamine B degradation and photocurrent generatio. Mater. Charact. 2015, 101, 166–172. [Google Scholar] [CrossRef]
  22. Wan, J.; Du, X.; Wang, R.M.; Liu, E.Z.; Jia, J.; Bai, X.; Hu, X.Y.; Fan, J. Mesoporous nanoplate multi-directional assembled Bi2WO6 for high efficient photocatalytic oxidation of NO. Chemosphere 2018, 193, 737–744. [Google Scholar] [CrossRef] [PubMed]
  23. Wan, J.; Zhang, Y.; Wang, R.M.; Liu, L.; Liu, E.Z.; Fan, J.; Fu, F. Effective charge kinetics steering in surface plasmons coupled two-dimensional chemical Au/Bi2WO6-MoS2 heterojunction for superior photocatalytic detoxification performance. J. Hazard. Mater. 2020, 384, 121484. [Google Scholar] [CrossRef] [PubMed]
  24. Gong, Q.H.; Gao, T.T.; Hu, T.T.; Zhou, G.W. Synthesis and electrochemical energy storage applications of Micro/Nanostructured spherical materials. Nanomaterials 2019, 9, 1207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Adhikari, S.; Selvaraj, S.; Kim, D.H. Construction of heterojunction photoelectrode via atomic layer deposition of Fe2O3 on Bi2WO6 for highly efficient photoelectrochemical sensing and degradation of tetracycline. Appl. Catal. B Environ. 2019, 244, 11–24. [Google Scholar] [CrossRef]
  26. Liu, X.T.; Gu, S.N.; Zhao, Y.J.; Zhou, G.W.; Li, W.J. BiVO4, Bi2WO6, Bi2MoO6 photocatalysis: A brief review. J. Mater. Sci. Technol. 2020, 56, 45–68. [Google Scholar] [CrossRef]
  27. Xu, Q.L.; Zhang, L.Y.; Cheng, B.; Fan, J.J.; Yu, J.G. S-Scheme heterojunction photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  28. Xie, T.P.; Liu, Y.; Wang, H.Q.; Wu, Z.B. Layered MoSe2/Bi2WO6 composite with P-N heterojunctions as a promising visible-light induced photocatalyst. Appl. Surf. Sci. 2018, 444, 320–329. [Google Scholar] [CrossRef]
  29. Jo, W.K.; Kumar, S.; Eslava, S.; Tonda, S. Construction of Bi2WO6/RGO/g-C3N4 2D/2D/2D hybrid Z-scheme heterojunctions with large interfacial contact area for efficient charge separation and high-performance photoreduction of CO2 and H2O into solar fuels. Appl. Catal. B Environ. 2018, 239, 586–598. [Google Scholar] [CrossRef]
  30. Meng, X.C.; Li, Z.Z.; Zeng, H.M.; Chen, J.; Zhang, Z.S. MoS2 quantum dots-interspersed Bi2WO6 heterostructures for visible light-induced detoxification and disinfection. Appl. Catal. B Environ. 2017, 210, 160–172. [Google Scholar] [CrossRef]
  31. Adhikari, S.; Kim, D.H. Synthesis of Bi2S3/Bi2WO6 hierarchical microstructures for enhanced visible light driven photocatalytic degradation and photoelectrochemical sensing of ofloxacin. Chem. Eng. J. 2018, 354, 692–705. [Google Scholar] [CrossRef]
  32. Pan, J.B.; Liu, J.J.; Ma, H.C.; Zuo, S.L.; Khan, U.A.; Yu, Y.C.; Li, B.S. Structure of flower-like hierarchical CdS QDs/Bi/Bi2WO6 heterojunction with enhanced photocatalytic activity. New J. Chem. 2018, 42, 7293–7300. [Google Scholar] [CrossRef]
  33. Hu, K.; Chen, C.Y.; Zhu, Y.; Zeng, G.M.; Huang, B.B.; Chen, W.Q.; Liu, S.H.; Lei, C.; Li, B.S.; Yang, Y. Ternary Z-scheme heterojunction of Bi2WO6 with reduced graphene oxide (rGO) and meso-tetra (4-carboxyphenyl) porphyrin (TCPP) for enhanced visible-light photocatalysis. J. Colloid Interface Sci. 2019, 540, 115–125. [Google Scholar] [CrossRef]
  34. Wan, J.; Xue, P.; Wang, R.M.; Liu, L.; Liu, E.Z.; Bai, X.; Fan, J.; Hu, X.Y. Synergistic effects in simultaneous photocatalytic removal of Cr(VI) and tetracycline hydrochloride by Z-scheme Co3O4/Ag/Bi2WO6 heterojunction. Appl. Surf. Sci. 2019, 483, 677–687. [Google Scholar] [CrossRef]
  35. Huang, D.L.; Li, J.; Zeng, G.M.; Xue, W.J.; Chen, S.; Li, Z.H.; Deng, R.; Yang, Y.; Cheng, M. Facile construction of hierarchical flower-like Z-scheme AgBr/Bi2WO6 photocatalysts for effective removal of tetracycline: Degradation pathways and mechanism. Chem. Eng. J. 2019, 375, 121991. [Google Scholar] [CrossRef]
  36. Xue, W.J.; Huang, D.L.; Li, J.; Zeng, G.M.; Deng, R.; Yang, Y.; Chen, S.; Li, Z.H.; Gong, X.M.; Li, B. Assembly of AgI nanoparticles and ultrathin g-C3N4 nanosheets codecorated Bi2WO6 direct dual Z-scheme photocatalyst: An efficient, sustainable and heterogeneous catalyst with enhanced photocatalytic performance. Chem. Eng. J. 2019, 373, 1144–1157. [Google Scholar] [CrossRef]
  37. Long, L.L.; Chen, J.J.; Zhang, X.; Zhang, A.Y.; Huang, Y.X.; Rong, Q.; Yu, H.Q. Layer-controlled growth of MoS2 on self-assembled flower-like Bi2S3 for enhanced photocatalysis under visible light irradiation. NPG Asia Mater. 2016, 8, e263. [Google Scholar] [CrossRef]
  38. Jonjana, S.; Phuruangrat, A.; Thongtem, S.; Thongtem, T. Synthesis, characterization and photocatalysis of heterostructure AgBr/Bi2WO6 nanocomposites. Mater. Lett. 2018, 216, 92–96. [Google Scholar] [CrossRef]
  39. Babu, B.; Koutavarapu, R.; Shim, J.; Yoo, K. Enhanced solar light-driven photocatalytic degradation of tetracycline and organic pollutants by novel one-dimensional ZnWO4 nanorod-decorated two-dimensional Bi2WO6 nanoflakes. J. Taiwan Inst. Chem. E 2020, 110, 58–70. [Google Scholar] [CrossRef]
  40. Xie, J.F.; Zhang, J.J.; Li, S.; Grote, F.; Zhang, X.D.; Zhang, H.; Wang, R.X.; Lei, Y.; Pan, B.C.; Xie, Y. Controllable disorder engineering in oxygen-incorporated MoS2 ultrathin nanosheets for efficient hydrogen evolution. J. Am. Chem. Soc. 2013, 135, 17881–17888. [Google Scholar] [CrossRef]
  41. Chen, D.D.; Fang, J.Z.; Lu, S.Y.; Zhou, G.Y.; Feng, W.H.; Yang, F.; Chen, Y.; Fang, Z.Q. Fabrication of Bi modified Bi2S3 pillared g-C3N4 photocatalyst and its efficient photocatalytic reduction and oxidation performances. Appl. Surf. Sci. 2017, 426, 427–436. [Google Scholar] [CrossRef]
  42. Ke, J.; Liu, J.; Sun, H.Q.; Zhang, H.Y.; Duan, X.G.; Liang, P.; Li, X.Y.; Tade, M.O.; Liu, S.M.; Wang, S.B. Facile assembly of Bi2O3/Bi2S3/MoS2 n-p heterojunction with layered n-Bi2O3 and p-MoS2 for enhanced photocatalytic water oxidation and pollutant degradation. Appl. Catal. B Environ. 2017, 200, 47–55. [Google Scholar] [CrossRef]
  43. Zhou, W.J.; Yin, Z.Y.; Du, Y.P.; Huang, X.; Zeng, Z.Y.; Fan, Z.X.; Liu, H.; Wang, J.Y.; Zhang, H. Synthesis of Few-Layer MoS2 Nanosheet-Coated TiO2 Nanobelt Heterostructures for Enhanced Photocatalytic Activities. Small 2013, 9, 140–147. [Google Scholar] [CrossRef] [PubMed]
  44. Zhu, Y.Y.; Wang, Y.J.; Ling, Q.; Zhu, Y.F. Enhancement of full-spectrum photocatalytic activity over BiPO4/Bi2WO6 composites. Appl. Catal. B Environ. 2017, 200, 222–229. [Google Scholar] [CrossRef] [Green Version]
  45. Chen, Y.J.; Wang, G.F.; Li, H.L.; Zhang, F.F.; Jiang, H.Y.; Tian, G.H. Controlled synthesis and exceptional photoelectrocatalytic properties of Bi2S3/MoS2/Bi2MoO6 ternary hetero-structured porous film. J. Colloid Interface Sci. 2019, 555, 214–223. [Google Scholar] [CrossRef] [PubMed]
  46. Zhou, H.M.; Xia, X.; Lv, P.F.; Zhang, J.; Pang, Z.Y.; Li, D.W.; Cai, Y.B.; Wei, Q.F. Wintersweet Branch-Like C/C@SnO2/MoS2 Nanofibers as High-Performance Li and Na-Ion Battery Anodes. Part. Part. Syst. Charact. 2017, 34, 1700295. [Google Scholar] [CrossRef]
  47. Zhou, Y.X.; Lv, P.F.; Zhang, W.; Meng, X.D.; He, H.; Zeng, X.H.; Shen, X.S. Pristine Bi2WO6 and hybrid Au-Bi2WO6 hollow microspheres with excellent photocatalytic activities. Appl. Surf. Sci. 2018, 457, 925–932. [Google Scholar] [CrossRef]
  48. Shao, B.B.; Liu, X.J.; Liu, Z.F.; Zeng, G.M.; Liang, Q.H.; Liang, C.; Cheng, Y.; Zhang, W.; Liu, Y.; Gong, S.X. A novel double Z-scheme photocatalyst Ag3PO4/Bi2S3/Bi2O3 with enhanced visible-light photocatalytic performance for antibiotic degradation. Chem. Eng. J. 2019, 368, 730–745. [Google Scholar] [CrossRef]
  49. Song, S.S.; Wang, J.M.; Peng, T.Y.; Fu, W.L.; Zan, L. MoS2-MoO3−x hybrid cocatalyst for effectively enhanced H2 production photoactivity of AgIn5S8 nano-octahedrons. Appl. Catal. B Environ. 2018, 228, 39–46. [Google Scholar] [CrossRef]
  50. Pan, Q.C.; Zhang, Q.B.; Zheng, F.H.; Liu, Y.Z.; Li, Y.P.; Ou, X.; Xiong, X.H.; Yang, C.H.; Liu, M.L. Construction of MoS2/C hierarchical tubular heterostructures for high-performance sodium ion batteries. ACS Nano 2018, 12, 12578–12586. [Google Scholar] [CrossRef]
  51. Wang, J.Z.; Jin, J.; Wang, X.G.; Yang, S.N.; Zhao, Y.L.; Wu, Y.W.; Dong, S.Y.; Sun, J.Y.; Sun, J.H. Facile fabrication of novel BiVO4/Bi2S3/MoS2 n-p heterojunction with enhanced photocatalytic activities towards pollutant degradation under natural sunlight. J. Colloid Interface Sci. 2017, 505, 805–815. [Google Scholar] [CrossRef] [PubMed]
  52. Cao, L.; Liang, X.H.; Ou, X.; Yang, X.F.; Li, Y.Z.; Yang, C.H.; Lin, Z.; Liu, M.L. Heterointerface engineering of hierarchical Bi2S3/MoS2 with self-generated rich phase boundaries for superior sodium storage performance. Adv. Funct. Mater. 2020, 30, 1910732. [Google Scholar] [CrossRef]
  53. Huang, H.W.; Zhou, C.; Jiao, X.C.; Yuan, H.F.; Zhao, J.W.; He, C.Q.; Hofkens, J.; Roeffaers, M.B.J.; Long, J.L.; Steele, J.A. Subsurface defect engineering in single-unit-cell Bi2WO6 monolayers boosts solar-driven photocatalytic performance. ACS Catal. 2020, 10, 1439–1443. [Google Scholar] [CrossRef]
  54. Qian, X.F.; Yue, D.T.; Tian, Z.Y.; Reng, M.; Zhu, Y.; Kan, M.; Zhang, T.Y.; Zhao, Y.X. Carbon quantum dots decorated Bi2WO6 nanocomposite with enhanced photocatalytic oxidation activity for VOCs. Appl. Catal. B Environ. 2016, 193, 16–21. [Google Scholar] [CrossRef]
  55. Song, N.N.; Zhang, M.H.; Zhou, H.; Li, C.Y.; Liu, G.; Zhong, S.; Zhang, S.Y. Synthesis and properties of Bi2WO6 coupled with SnO2 nano-microspheres for improved photocatalytic reduction of Cr6+ under visible light irradiation. Appl. Surf. Sci. 2019, 495, 143551. [Google Scholar] [CrossRef]
  56. Bai, X.; Du, Y.Y.; Hu, X.Y.; He, Y.D.; He, C.L.; Liu, E.Z.; Fan, J. Synergy removal of Cr(VI) and organic pollutants over RP-MoS2/rGO photocatalyst. Appl. Catal. B Environ. 2018, 239, 204–213. [Google Scholar] [CrossRef]
  57. Zheng, J.H.; Zhang, L. Incorporation of CoO nanoparticles in 3D marigold flower-like hierarchical architecture MnCo2O4 for highly boosting solar light photo-oxidation and reduction ability. Appl. Catal. B Environ. 2018, 237, 1–8. [Google Scholar] [CrossRef]
  58. Yang, L.J.; Hu, Y.D.; Zhang, L. Architecting Z-scheme Bi2S3@CoO with 3D chrysanthemums-like architecture for both photoeletro-oxidization and -reduction performance under visible light. Chem. Eng. J. 2019, 378, 122092. [Google Scholar] [CrossRef]
  59. Babu, B.; Koutavarapu, R.; Shim, J.; Yoo, K. Enhanced visible-light-driven photoelectrochemical and photocatalytic performance of Au-SnO2 quantum dot-anchored g-C3N4 nanosheets. Sep. Purif. Technol. 2020, 240, 116652. [Google Scholar] [CrossRef]
  60. Babu, B.; Koutavarapu, R.; Shim, J.; Yoo, K. Facile one-pot synthesis of gold/tin oxide quantum dots for visible light catalytic degradation of methylene blue: Optimization of plasmonic effect. J. Alloys Compd. 2020, 812, 152081. [Google Scholar]
  61. Zhang, G.P.; Chen, D.Y.; Li, N.J.; Xu, Q.F.; Li, H.; He, J.H.; Lu, J.M. Fabrication of Bi2MoO6/ZnO hierarchical heterostructures with enhanced visible-light photocatalytic activity. Appl. Catal. B Environ. 2019, 250, 313–324. [Google Scholar] [CrossRef]
  62. Zhang, F.; Zhang, Y.C.; Zhou, C.Q.; Yang, Z.J.; Xue, H.G.; Dionysiou, D.D. A new high efficiency visible-light photocatalyst made of SnS2 and conjugated derivative of polyvinyl alcohol and its application to Cr(VI) reduction. Chem. Eng. J. 2017, 324, 140–153. [Google Scholar] [CrossRef]
  63. Wang, Y.J.; Bao, S.Y.; Liu, Y.Q.; Yang, W.W.; Yu, Y.S.; Feng, M.; Li, K.F. Efficient photocatalytic reduction of Cr(VI) in aqueous solution over CoS2/g-C3N4-rGO nanocomposites under visible light. Appl. Surf. Sci. 2020, 510, 145495. [Google Scholar] [CrossRef]
  64. Alam, U.; Khan, A.; Bahnemann, D.; Muneer, M. Synthesis of Co doped ZnWO4 for simultaneous oxidation of RhB and reduction of Cr(VI) under UV-light irradiation. J. Environ. Chem. Eng. 2018, 6, 4885–4898. [Google Scholar] [CrossRef]
  65. Patnaik, S.; Swain, G.; Parida, K.M. Highly efficient charge transfer through a double Z-scheme mechanism by a Cu-promoted MoO3/g-C3N4 hybrid nanocomposite with superior electrochemical and photocatalytic performance. Nanoscale 2018, 10, 5950–5964. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, X.; Li, Y.X.; Yi, X.H.; Zhao, C.; Wang, P.; Deng, J.G.; Wang, C.C. Photocatalytic Cr(VI) elimination over BUC-21/N-K2Ti4O9 composites: Big differences in performance resulting from small differences in composition. Chin. J. Catal. 2021, 42, 259–270. [Google Scholar] [CrossRef]
  67. Ren, Z.X.; Liu, X.J.; Zhuge, Z.H.; Gong, Y.Y.; Sun, C.Q. MoSe2/ZnO/ZnSe hybrids for efficient Cr(VI) reduction under visible light irradiation. Chin. J. Catal. 2020, 41, 180–187. [Google Scholar] [CrossRef]
  68. Xia, Y.; Gang, R.Q.; Xu, L.; Huang, S.J.; Zhou, L.X.; Wang, J. Nanorod-pillared mesoporous rGO/ZnO/Au hybrids for photocatalytic Cr(VI) reduction: Enhanced Cr(VI) adsorption and solar energy harvest. Ceram. Int. 2020, 46, 1487–1493. [Google Scholar] [CrossRef]
  69. Yang, L.; Xu, C.; Wan, F.C.; He, H.H.; Gu, H.S.; Xiong, J. Synthesis of RGO/BiOI/ZnO composites with efficient photocatalytic reduction of aqueous Cr(VI) under visible-light irradiation. Mater. Res. Bull. 2019, 112, 154–158. [Google Scholar] [CrossRef]
  70. Xu, F.; Chen, H.M.; Xu, C.Y.; Wu, D.P.; Gao, Z.Y.; Zhang, Q.; Jiang, K. Ultra-thin Bi2WO6 porous nanosheets with high lattice coherence for enhanced performance for photocatalytic reduction of Cr(VI). J. Colloid Interface Sci. 2018, 525, 97–106. [Google Scholar] [CrossRef]
  71. Luo, S.; Ke, J.; Yuan, M.Q.; Zhang, Q.; Xie, P.; Deng, L.D.; Wang, S.B. CuInS quantum dots embedded in Bi2WO6 nanoflowers for enhanced visible light photocatalytic removal of contaminants. Appl. Catal. B Environ. 2018, 221, 215–222. [Google Scholar] [CrossRef]
  72. Zhang, C.M.; Chen, G.; Li, C.M.; Sun, J.X.; Lv, C.D.; Fan, S.; Xing, W.N. In situ fabrication of Bi2WO6/MoS2/RGO heterojunction with nanosized interfacial contact via confined space effect toward enhanced photocatalytic properties. ACS Sustain. Chem. Eng. 2016, 4, 5936–5942. [Google Scholar] [CrossRef]
  73. Jiang, Z.K.; Chen, K.X.; Zhang, Y.C.; Wang, Y.Y.; Wang, F.; Zhang, G.S.; Dionysioue, D.D. Magnetically recoverable MgFe2O4/conjugated polyvinyl chloride derivative nanocomposite with higher visible-light photocatalytic activity for treating Cr(VI)-polluted water. Sep. Purif. Technol. 2020, 236, 116272. [Google Scholar] [CrossRef]
  74. Hua, E.B.; Jin, S.; Wang, X.R.; Ni, S.; Liu, G.; Xu, X.X. Ultrathin 2D type-II p-n heterojunctions La2Ti2O7/In2S3 with efficient charge separations and photocatalytic hydrogen evolution under visible light illumination. Appl. Catal. B Environ. 2019, 245, 733–742. [Google Scholar] [CrossRef]
Scheme 1. Schematic diagram of preparing Bi2WO6/Bi2S3/MoS2 heterojunction ternary composites.
Scheme 1. Schematic diagram of preparing Bi2WO6/Bi2S3/MoS2 heterojunction ternary composites.
Nanomaterials 10 01813 sch001
Figure 1. X-ray diffraction (XRD) patterns (a), and Raman spectra (b) of bare Bi2WO6, MoS2, BBM-1, BBM-2, BBM-3, and BBM-4 heterojunction.
Figure 1. X-ray diffraction (XRD) patterns (a), and Raman spectra (b) of bare Bi2WO6, MoS2, BBM-1, BBM-2, BBM-3, and BBM-4 heterojunction.
Nanomaterials 10 01813 g001
Figure 2. Scanning electron microscope (SEM) and transmission electron microscope (TEM) images of (a,f) bare Bi2WO6, (b,g) BBM-1, (c,h) BBM-2, (d,i) BBM-3, and (e,j) BBM-4. High-resolution TEM (HRTEM) image (k), and energy-dispersive spectra (EDS) mapping images (lp) of BBM-3.
Figure 2. Scanning electron microscope (SEM) and transmission electron microscope (TEM) images of (a,f) bare Bi2WO6, (b,g) BBM-1, (c,h) BBM-2, (d,i) BBM-3, and (e,j) BBM-4. High-resolution TEM (HRTEM) image (k), and energy-dispersive spectra (EDS) mapping images (lp) of BBM-3.
Nanomaterials 10 01813 g002
Figure 3. X-ray photoelectron spectroscopy (XPS) analysis of BBM-3 showing (a) survey, (b) Mo 3d, (c) Bi 4f, (d) W 4f, and (e) O 1s spectra.
Figure 3. X-ray photoelectron spectroscopy (XPS) analysis of BBM-3 showing (a) survey, (b) Mo 3d, (c) Bi 4f, (d) W 4f, and (e) O 1s spectra.
Nanomaterials 10 01813 g003
Figure 4. (a) Ultraviolet–visible (UV–vis) diffuse reflectance spectra (DRS) spectra of Bi2WO6, MoS2, BBM-1, BBM-2, BBM-3, and BBM-4. (bg) Plot of (αhν)2 versus photon energy of Bi2WO6, MoS2, BBM-1, BBM-2, BBM-3, and BBM-4.
Figure 4. (a) Ultraviolet–visible (UV–vis) diffuse reflectance spectra (DRS) spectra of Bi2WO6, MoS2, BBM-1, BBM-2, BBM-3, and BBM-4. (bg) Plot of (αhν)2 versus photon energy of Bi2WO6, MoS2, BBM-1, BBM-2, BBM-3, and BBM-4.
Nanomaterials 10 01813 g004
Figure 5. (a) Photoluminescence spectra (λexcitation = 300 nm) and (b) electrochemical impedance spectra (EIS) spectra of the Bi2WO6, BBM-1, BBM-2, BBM-3, and BBM-4.
Figure 5. (a) Photoluminescence spectra (λexcitation = 300 nm) and (b) electrochemical impedance spectra (EIS) spectra of the Bi2WO6, BBM-1, BBM-2, BBM-3, and BBM-4.
Nanomaterials 10 01813 g005
Figure 6. (a) Effect of pH on the catalytic activity of BBM-3. At the pH of 2.00, 6.00, and 10.00 (bd, respectively), the UV–vis absorption spectra of Cr(VI) solution on BBM-3 sample changed.
Figure 6. (a) Effect of pH on the catalytic activity of BBM-3. At the pH of 2.00, 6.00, and 10.00 (bd, respectively), the UV–vis absorption spectra of Cr(VI) solution on BBM-3 sample changed.
Nanomaterials 10 01813 g006
Figure 7. (a) Visible light (vis-light) catalytic reduction of Cr(VI) by different catalysts. (b) Corresponding pseudo first-order kinetic curves, and (c) rate constant “k” by different catalysts. (d) BBM-3 vis-light catalytic reduction Cr(VI) in the presence of electron scavengers (KBrO3) and hole scavengers (citric acid), respectively. Dosages of KBrO3 and citric acid aqueous solutions: 100 μL of 50 mg mL−1.
Figure 7. (a) Visible light (vis-light) catalytic reduction of Cr(VI) by different catalysts. (b) Corresponding pseudo first-order kinetic curves, and (c) rate constant “k” by different catalysts. (d) BBM-3 vis-light catalytic reduction Cr(VI) in the presence of electron scavengers (KBrO3) and hole scavengers (citric acid), respectively. Dosages of KBrO3 and citric acid aqueous solutions: 100 μL of 50 mg mL−1.
Nanomaterials 10 01813 g007
Figure 8. (a) Vis-light catalytic reduction of Cr(VI) by BBM-3 at different recycling runs. (b) XRD pattern of the BBM-3 and BBM-3-a. XPS spectra of BBM-3-a: (c) survey and (d) Cr 2p.
Figure 8. (a) Vis-light catalytic reduction of Cr(VI) by BBM-3 at different recycling runs. (b) XRD pattern of the BBM-3 and BBM-3-a. XPS spectra of BBM-3-a: (c) survey and (d) Cr 2p.
Nanomaterials 10 01813 g008
Scheme 2. Proposed photocatalytic reduction Cr(VI) mechanism of Bi2WO6/Bi2S3/MoS2 under vis-light conditions.
Scheme 2. Proposed photocatalytic reduction Cr(VI) mechanism of Bi2WO6/Bi2S3/MoS2 under vis-light conditions.
Nanomaterials 10 01813 sch002
Table 1. Summary of the band gap energy (Eg), conduction band edge (ECB), and valence band edge (EVB) of Bi2WO6, Bi2S3, and MoS2.
Table 1. Summary of the band gap energy (Eg), conduction band edge (ECB), and valence band edge (EVB) of Bi2WO6, Bi2S3, and MoS2.
MaterialsEg (eV)X (eV)Ee (eV)ECB vs. NHE 1 (eV)EVB vs. NHE 1 (eV)
Bi2WO62.746.204.500.333.07
Bi2S31.19 [51]5.274.500.181.37
MoS21.305.324.500.171.47
1 NHE: normal hydrogen electrode.
Table 2. Performance comparison with other materials used for vis-light catalytic reduction of Cr(VI).
Table 2. Performance comparison with other materials used for vis-light catalytic reduction of Cr(VI).
Materials/Amount (mg)Cr(VI) Solution Volume (mL)/Concentration (mg L−1)Time (min)Photocatalytic Removal RatePublication DateRef.
CoS2/g-C3N4-rGO/1020/2012099.82020[63]
MoSe2/ZnO/ZnSe/8080/201801002020[67]
rGO/ZnO/Au/5050/1040972020[68]
Bi2MoO6/ZnO/10050/501501002019[61]
RGO/BiOI/ZnO/100150/10180922019[69]
Bi2WO6+Oxalic/6050/101201002018[70]
CuInS2 QDs/Bi2WO6/2040/10300902018[71]
Bi2WO6/MoS2/RGO+Lactic acid/30100/10801002016[72]
Bi2WO6/Bi2S3/MoS2/2050/4075100 This work

Share and Cite

MDPI and ACS Style

Ren, J.; Hu, T.; Gong, Q.; Wang, Q.; Sun, B.; Gao, T.; Cao, P.; Zhou, G. Spherical Bi2WO6/Bi2S3/MoS2 n-p Heterojunction with Excellent Visible-Light Photocatalytic Reduction Cr(VI) Activity. Nanomaterials 2020, 10, 1813. https://doi.org/10.3390/nano10091813

AMA Style

Ren J, Hu T, Gong Q, Wang Q, Sun B, Gao T, Cao P, Zhou G. Spherical Bi2WO6/Bi2S3/MoS2 n-p Heterojunction with Excellent Visible-Light Photocatalytic Reduction Cr(VI) Activity. Nanomaterials. 2020; 10(9):1813. https://doi.org/10.3390/nano10091813

Chicago/Turabian Style

Ren, Jing, Tingting Hu, Qinghua Gong, Qian Wang, Bin Sun, Tingting Gao, Pei Cao, and Guowei Zhou. 2020. "Spherical Bi2WO6/Bi2S3/MoS2 n-p Heterojunction with Excellent Visible-Light Photocatalytic Reduction Cr(VI) Activity" Nanomaterials 10, no. 9: 1813. https://doi.org/10.3390/nano10091813

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