Next Article in Journal
Diacylglycerol Kinases in Signal Transduction
Next Article in Special Issue
Efficient Adsorption-Photocatalytic Removal of Tetracycline Hydrochloride over Octahedral MnS
Previous Article in Journal
Inhibition of Autophagy Negates Radiofrequency-Induced Adaptive Response in SH-SY5Y Neuroblastoma Cells
Previous Article in Special Issue
Cuprous Oxide Thin Films Implanted with Chromium Ions—Optical and Physical Properties Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Tin-Doped Three-Dimensional Flower-like Bismuth Tungstate with Enhanced Photocatalytic Activity

1
School of Mechanical Engineering, Chengdu University, Chengdu 610106, China
2
School of Science, Xichang University, Xichang 615013, China
3
College of Materials and Chemistry & Chemiacl Engineering, Chengdu University of Technology, Chengdu 610051, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(15), 8422; https://doi.org/10.3390/ijms23158422
Submission received: 22 June 2022 / Revised: 25 July 2022 / Accepted: 27 July 2022 / Published: 29 July 2022
(This article belongs to the Special Issue Feature Papers in Physical Chemistry and Chemical Physics 2022)

Abstract

:
Photocatalytic degradation of harmful organic matter is a feasible and environmentally friendly method. Bi2WO6 has become a hotspot of photocatalysts because of its unique layered structure and visible light response. In the present study, Sn doping was adopted to modified Bi2WO6 by hydrothermal method. The Sn-doped Bi2WO6 photocatalysts were characterized by XRD, SEM, TEM, BET, XPS, PL, and DRS, respectively. The results show that Sn-doped Bi2WO6 shows three-dimensional (3D) flower-like morphology, which is composed of two-dimensional (2D) nanosheets. Sn4+ ions enter into the Bi2WO6 lattice, producing a degree of Bi2WO6 lattice distortion, which is in favor of reducing the recombination of photogenerated electrons and holes. Moreover, the specific surface area of Bi2WO6 is significantly increased after doping, which is beneficial to providing more active sites. The photocatalytic results show that 2%Sn-Bi2WO6 exhibits the highest photocatalytic activity. After 60 min of irradiation, the photocatalytic degradation degree of methylene blue (MB) increases from 80.6% for pure Bi2WO6 to 92.0% for 2%Sn-Bi2WO6. The first-order reaction rate constant of 2%Sn-Bi2WO6 is 0.030 min−1, which is 1.7 times than that of pure Bi2WO6.

1. Introduction

With the rapid development of economy and society, the problem of water pollution has become increasingly prominent [1,2,3]. Semiconductor photocatalysis degradation technology is widely studied as it is a green and feasible way for environmental protection [4,5,6]. Traditional semiconductor photocatalysts (TiO2, ZnO, CeO2, etc.) usually only absorb ultraviolet light, but ultraviolet light accounts for less than 5% of sunlight, which greatly limits the utilization efficiency of solar energy [7,8,9,10]. Researchers have developed several high-efficiency visible light response photocatalysts to improve the utilization of sunlight [11,12,13]. Bi2WO6 has attracted much attention due to its unique layered structure and good visible light response [14,15,16]. Sharma [16] et al. prepared 2D Bi2WO6 nanosheets by hydrothermal method, which shows a band gap of 2.78 eV. The degradation degree of ciprofloxacin is 47% under 150 min natural light irradiation. However, in practical applications, the photocatalytic activity of pure Bi2WO6 is greatly limited due to its rapid recombination of photogenerated electrons and holes. Employing metal ion doping to modify Bi2WO6 is one of the methods used to enhance the photocatalytic activity [17,18,19]. On one hand, new impurity levels will be formed by cation doping, which reduces the excitation energy of transition [20,21]. On the other hand, ion doping will produce lattice distortion, trapping photogenerated charges and improving the separation of carriers [22,23].
Generally, photocatalysts that present three-dimensional (3D) morphology exhibit higher photocatalytic activity than other morphology as 3D morphology is beneficial to providing more reactive active sites due to its larger specific surface area and the utilization of visible light owing to the multiple reflection of light in 3D structures [24,25,26]. In our previous work, the 3D flower-like Bi2WO6 photocatalysts were prepared and it was found that the optimum hydrothermal temperature is 160 °C [27]. On this basis, in the present work, to improve the photogenerated charge separation and specific surface area of pure Bi2WO6, Sn doping modification was adopted and Sn-doped Bi2WO6 photocatalyst was prepared under the hydrothermal conditions of 160 °C and 12 h. The crystal structure, surface morphology, specific surface area, elemental composition and valence state, and optical properties of samples were analyzed using various characterization methods. The photocatalytic activity of samples was evaluated by the degradation of methylene blue (MB), and the mechanism of Sn doping improving the photocatalytic activity was comprehensively analyzed.

2. Results and Discussion

2.1. Catalyst Characterization

Figure 1a shows the XRD patterns of pure Bi2WO6 and Sn-Bi2WO6. The diffraction peaks appear at 28.3°, 32.8°, 47.1°, 56.0°, 58.5°, and 68.8°, corresponding to the (131), (200), (202), (133), (262), and (400) crystal planes of Bi2WO6 [15,17,18]. The Sn-Bi2WO6 patterns show that the shape and position of the diffraction peaks are similar to pure Bi2WO6. No diffraction peak related to Sn is detected in the patterns of 1%Sn-Bi2WO6, 2%Sn-Bi2WO6 and 4%Sn-Bi2WO6. A part of Sn4+ ions can enter the Bi2WO6 lattice to replace Bi3+ without reacting with Bi2WO6 to generate a new phase [18,23]. Lattice expansion or shrinkage caused by ion doping is controversial. Generally, when ions with a radius less than Bi3+ enter the lattice and replace them, the lattice will shrink. According to the Bragg equation, the XRD diffraction peak will shift to higher angle [18,21]. On the contrary, some research shows that the XRD diffraction peak shifts to a smaller angle after ions with an ionic radius less than Bi3+ doping, indicating that it causes lattice expansion [20,28]. In the study of Sm-doped Bi2WO6 reported by Liu et al. [20], it was found that when Sm3+ ions with an atomic radius (0.096 nm) less than Bi3+ (0.108 nm) were doped, the XRD diffraction peaks shifts to a lower angle, indicating that the lattice expansion is caused after doping. They believe that after Sm3+ is doped, part of Sm3+ will replace Bi in its lattice, and part of Sm3+ will enter the interstitial position of the Bi2WO6 lattice, causing lattice expansion. Alhadi et al. [28] believe that the same phenomenon appears in their Sn-doped Bi2WO6 research. In this work, it can be seen from the enlarged XRD pattern (Figure 1b) that the diffraction peak of the (131) crystal plane shifts slightly to a small angle after Sn doping, which indicates that the lattice of Bi2WO6 is expanding. The lattice parameters and cell volumes were calculated, as shown in Table 1. The results show that the lattice expands slightly after Sn doping, which is consistent with the literature [20,28]. Because the ion radius of Sn4+ (0.069 nm) [28] is smaller than Bi3+ (0.108 nm) [20], the lattice will shrink when it replaces Bi3+. However, lattice expansion occurred in this study, which may be due to the fact that in addition to the Sn4+ ions replacing Bi3+, the remaining Sn4+ ions will enter the interstitial position of the Bi2WO6 lattice, causing lattice expansion. Moreover, there is a view that the change in electronic environment will also cause lattice changes. Chen et al. [29] believe that when elements with large electronegativity are used to replace elements with small electronegativity, the mutual attraction between atoms will be enhanced, which will cause lattice contraction. Replacing elements with small electronegativity with elements with large electronegativity will weaken the mutual attraction between atoms and cause lattice expansion. In this study, since the electronegativity of Sn (Pauling, 1.96) [30] is lower than that of Bi (Pauling, 2.02) [31], when Sn4+ replaces Bi3+, the attractive force between Sn atoms and O, W atoms will be reduced, resulting in lattice expansion.
Compared with the diffraction peak of pure Bi2WO6, the wider FWHM of the diffraction peak and the lower intensity of the diffraction peak imply that the grain size decreases and the amorphous composition increases after Sn doping [28]. Bi-O-Sn bonds will be formed by Sn doping, which hinders the migration of Bi, O, and W atoms, delaying the nucleation and growth of Bi2WO6 [32]. The crystallite size of samples can be calculated using the Scherrer formula [22,33]. The grain sizes of pure Bi2WO6, 1%Sn-Bi2WO6, 2%Sn-Bi2WO6, 4%Sn-Bi2WO6, and 6%Sn-Bi2WO6 are 12.2, 7.4, 6.0, 6.1, and 4.7 nm, respectively, indicating that the grains are refined by Sn doping. The crystallinities of pure Bi2WO6, 1%Sn-Bi2WO6, 2%Sn-Bi2WO6, 4%Sn-Bi2WO6, and 6%Sn-Bi2WO6 are 90.8%, 88.9%, 87.7%, 86.3%, and 79.0%, respectively. The crystallinity of Sn-Bi2WO6 decreases gradually with the increase in Sn doping concentration. Remarkably, there is a diffraction peak near 26.8° in the pattern of the 6%Sn-Bi2WO6 sample, which corresponds to the SnO2 (110) crystal plane, which indicates that SnO2/Bi2WO6 composite photocatalyst forms at a high Sn doping concentration due to the excess of Sn4+ in the Bi2WO6 lattice solubility, forming a new phase SnO2 [33,34].
The SEM images of samples are shown in Figure 2a–j. In Figure 2a,b, it can be clearly observed that Bi2WO6 is composed of 2D nanoflakes interwoven with each other in the shape of 3D flower-like with an average diameter of 2–4 μm, and the length of the nanoflakes varies in size from a few tens to hundreds of nanometers. It is observed in Figure 2c–j that 3D morphology of Sn-Bi2WO6 is similar to Bi2WO6, and the length of 2D nanosheets is concentrated at 0.5–1 μm. Remarkably, the flake thickness of Sn-Bi2WO6 decreases, which is conducive to increasing the surface area. Figure 2k–p shows the element mappings of 2%Sn-Bi2WO6, from which it is known that 2%Sn-Bi2WO6 contains Bi, O, W, and Sn elements, and these elements are uniformly distributed. Figure S1 shows the EDS results of 1%Sn-Bi2WO6 (a), 4%Sn-Bi2WO6 (b), and 6%Sn-Bi2WO6 (c). The actual Sn/Bi molar ratio in the doped samples was measured. The molar ratios of Sn/Bi in 1%Sn-Bi2WO6, 2%Sn-Bi2WO6, 4%Sn-Bi2WO6, and 6%Sn-Bi2WO6 are 1.5%, 3.8%, 5.2%, and 6.6%, respectively. With the increasing Sn concentration, the Sn/Bi ratio shows an upward trend.
Figure 3 shows the TEM images of Bi2WO6 (a) and 2%Sn-Bi2WO6 (b). It can be seen that the three-dimensional morphology of Bi2WO6 is nearly a regular sphere, with a diameter of 3 μm (Figure 3a). After Sn doping, the diameter of the sample does not change obviously. According to HRTEM images of samples, the interplanar spacing of Bi2WO6 is about 0.316 nm (Figure 3c), which basically agrees with the theoretical value of the (131) plane of Bi2WO6. In Figure 3d, the interplanar spacing 0.318 nm is observed in 2%Sn-Bi2WO6, which corresponds to the (131) plane [35].
Figure 4 shows the pore size distribution curves and the N2 adsorption–desorption isotherms of Sn-Bi2WO6. The pore size of pure Bi2WO6 is mainly concentrated at 0–30 nm [27], while the pore sizes of Sn-Bi2WO6 are mainly concentrated at 0–15 nm. The specific surface areas of 1%Sn-Bi2WO6, 2%Sn-Bi2WO6, 4%Sn-Bi2WO6, and 6%Sn-Bi2WO6 are 37.1, 41.8, 37.6, and 43.5 m2/g, respectively. The specific areas of Sn-doped Bi2WO6 are significantly higher than pure Bi2WO6 (20.8 m2/g) [27]. Sn doping results in smaller grain sizes and thinner nanosheets than pure Bi2WO6, increasing the specific surface area.
Figure 5 presents the XPS spectra of pure Bi2WO6 and 2%Sn-Bi2WO6. The presence of signal peaks of Bi 4f, W 4f, O 1s, and C 1s in Bi2WO6 can be observed in Figure 5a. C element may be originated from the oil contamination [33]. The high-resolution spectra of Bi 4f are shown in Figure 5b. The spin-orbit of Bi 4f of pure Bi2WO6 splits into two characteristic peaks at 164.3 and 158.9 eV, corresponding to Bi 4f5/2 and Bi 4f7/2, indicating that the Bi element in the sample exists in the chemical state of Bi3+ [36,37,38]. The W 4f of pure Bi2WO6 has two characteristic peaks at 37.4 and 35.4 eV, corresponding to W 4f5/2 and W 4f7/2, verifying that the W element exists as the 6+ valence state [14,39,40]. The O 1s spectrum of pure Bi2WO6 (Figure 5d) is decomposed into three peaks at 529.5, 530.7, and 532.0 eV, corresponding to lattice oxygen (OL), surface hydroxyl (OH), and surface adsorbed oxygen (OA), respectively [41,42]. After Sn modification, the binding energies of Bi 4f, W 4f, and O 1s shift to a higher energy band, which can be ascribed to the interaction between Sn element and Bi, W, and O elements [38]. The Sn 3d shows two characteristic peaks at 486.8 and 495.5 eV, corresponding to Sn 3d5/2 and Sn 3d3/2, indicating that the Sn element exists in the chemical state of Sn4+ [28,43]. Figure S2 presents the XPS survey spectra of 1%Sn-Bi2WO6, 4%Sn-Bi2WO6, and 6%Sn-Bi2WO6. The binding energies of all the samples are shown in Table 2. Compared with pure Bi2WO6, the binding energies corresponding to the peaks of Bi 4f shift after Sn doping, which indicates that the electronic environment inside the Bi2WO6 lattice has changed, and proves that Sn4+ ions are incorporated into the Bi2WO6 lattice by doping [37,42].
Figure 6 shows the PL spectra of samples. Sn-Bi2WO6 demonstrates lower intensity than pure Bi2WO6, indicating that Sn doping is beneficial to inhibiting the recombination of photogenerated electrons and holes. Crystal defects will be formed by Sn doping, which captures photogenerated charges, improving the separation of carriers [28,35]. The PL peak intensity of 2%Sn-Bi2WO6 is the lowest, indicating that the inhibition effect is the best when Sn/Bi molar ratio is 2%. The 4%Sn-Bi2WO6 showing a higher PL peak intensity than 2%Sn-Bi2WO6 can be ascribed to the fact that a high level of doping will bring new recombination centers, reducing the carrier separation [44]. The PL peak intensity of 6%Sn-Bi2WO6 is lower than 4%Sn-Bi2WO6. The XRD results prove that a new phase SnO2 appears in 6%Sn-Bi2WO6 and SnO2/Bi2WO6 semiconductor composite structure forms. The coupling of semiconductors with different energy band positions is beneficial to accelerating the migration of photogenerated carriers at the phase interfaces, improving the separation of photogenerated charges [39,45].
The UV-visible absorption spectra of samples are shown in Figure 7a. The absorption edges of Sn-Bi2WO6 photocatalysts are similar, all around 440 nm, which is slightly smaller than that of pure Bi2WO6 (460 nm) [27]. It can be seen from Figure 7b that the band gap energies of pure Bi2WO6, 1%Sn-Bi2WO6, 2%Sn-Bi2WO6, 4%Sn-Bi2WO6, and 6%Sn-Bi2WO6 are 2.58, 2.62, 2.67, 2.59, and 2.62 eV, respectively.

2.2. Photocatalytic Performance

Figure 8a shows the degradation curves of MB by Sn-Bi2WO6 photocatalysts. The results show that the degradation degrees of 1%Sn-Bi2WO6, 2%Sn-Bi2WO6, and 4%Sn-Bi2WO6 are 83.7%, 92.0%, and 90.5%, respectively, which are higher than pure Bi2WO6 (80.6%) [27]. Sn doping improving the photocatalytic activity of Bi2WO6 can be attributed to the following two points: (1) The recombination of photogenerated charges is suppressed and the separation is improved by Sn doping. (2) The specific surface area of Bi2WO6 is enhanced after Sn doping, which is conducive to providing more active reaction sites, improving the photocatalytic efficiency [42]. It is worth noting that the degradation degrees of Sn-Bi2WO6 present a trend of first increasing and then decreasing, indicating that Sn doping level has an optimal doping concentration. The decreased photodegradation efficiency of 4%Sn-Bi2WO6 and 6%Sn-Bi2WO6 (75.5%) can be attributed to the formation of new recombination centers and excessive amorphous components due to excessive doping, respectively [28,45].
Figure 8b shows the first order kinetics curves of −ln(C/C0) versus time. The first-order reaction rate constants of pure Bi2WO6, 1%Sn-Bi2WO6, 2%Sn-Bi2WO6, 4%Sn-Bi2WO6, and 6%Sn-Bi2WO6 are 0.018, 0.021, 0.030, 0.031, and 0.013 min−1, respectively [27]. The first-order reaction rate constant of 2%Sn-Bi2WO6 is 1.7 times that of pure Bi2WO6.
Figure 9 shows the active species experiment of 2%Sn-Bi2WO6. During the experiments, holes (h+), ·O2 and ·OH radicals could be quenched by benzoquinone (BQ), ammonium oxalate (AO), and isopropanol (IPA), respectively [25,41]. When BQ, AO, and IPA were added to the radical scavengers, the degradation degrees of 2%Sn-Bi2WO6 decreased from 92.0% to 77.8%, 59.0% and 78.6%, respectively. The results verify that holes (h+) are the main active species in the degradation process, and ·O2 and ·OH radicals also contribute to the degradation.
According to Formulas (1) and (2), where EVB is the valence band potential, ECB is the conduction band potential, E0 is the free electron energy on the hydrogen scale (E0 = 4.5 eV); Eg is the bandgap energy of the photocatalytic material, and X is the absolute electronegativity of the semiconductor [46].
ECB = X − E0 − Eg/2
EVB = Eg + ECB
The valence band potential (EVB) and conduction band potential (ECB) of 2%Sn-Bi2WO6 are 3.20 and 0.53 eV, respectively. Based on the results, a schematic diagram of the photodegradation MB by 2%Sn-Bi2WO6 is proposed, as shown in Figure 10. Sn doping introduces lattice distortion, forming more crystal defects, capturing photogenerated electrons, and preventing their recombination with holes. Consequently, the holes react with H2O/OH and generates strong oxidizing hydroxyl radicals (·OH). The holes and ·OH hydroxyl radicals directly decompose MB molecules into inorganic small molecules [47,48].

3. Materials and Methods

Bismuth nitrate (Bi(NO3)3·5H2O, Analytical Reagent, AR, ≥99.0%), sodium tungstate (Na2WO4·2H2O, AR, ≥99.5%), anhydrous ethanol (C2H5OH, AR, ≥99.7%), glacial acetic acid (C2H4O2, AR, ≥99.5%), and tin tetrachloride (SnCl4·5H2O, AR, ≥99.0%) were purchased from Chengdu Kelong Chemical Co., Ltd., Chengdu, China.

3.1. Sample Preparation

Bi2WO6: Bi(NO3)3·5H2O and CH3COOH were dispersed in 20 mL deionized water to form solution A, and disperse Na2WO4·2H2O in 12 mL deionized water to form solution B. The mass ratio of Bi(NO3)3·5H2O to Na2WO4·2H2O was 2.94:1. Solution B was added dropwise to solution A, and stirring was continued for 30 min to form white flocs. The resulting mixture was transformed to a polytetrafluoroethylene liner, put in a reaction kettle, tightened, and heated at 160 °C 24 h. Alternately it was washed with deionized water and absolute ethanol until neutral, placed in a drying oven at 100 °C for drying for 10 h, and finally fully ground to obtain pure Bi2WO6 powder.
X%Sn-Bi2WO6: SnCl4·5H2O was added to solution A, and other experimental conditions were consistent with the preparation of Bi2WO6, the Sn-doped Bi2WO6 photocatalytic material could be synthesized, and the molar ratio of Sn/Bi was controlled to be 1%, 2%, 4%, and 6%. The Sn-doped Bi2WO6 with different concentration gradients were labeled as X%Sn-Bi2WO6 (X = 1, 2, 4, 6).

3.2. Sample Characterization

The crystal structure and phase information were studied using X-ray diffraction (XRD) using a DX-2700 X-ray diffractometer with Cu Kα radiation as the X-ray source. The scan range 2θ was 20°–70° and scan speed was 0.06°/s (Dandong Haoyuan Instrument Co., Ltd., Dandong, China). The XRD data were analyzed using jade 6.0 software. FEI-nspect F50 scanning electron microscope (SEM) and FEI-Tecnai G2 F20 transmission electron microscope (TEM and HRTEM) were used to observe the morphology (FEI Company, Hillsboro, OR, USA); the specific surface area were measured using a V-sorb 2800S analyzer (BET) (Mike Instrument Company, Atlanta, GA, USA); the composition and valence of elements were analyzed using an XSAM800 multifunctional surface analysis system (XPS) (Thermo Scientific K-Alpha, Kratos Ltd., Manchester, UK); the photoluminescence (PL) spectra were measured using an F-4600 fluorescence spectrum analyzer with a Xe lamp at an excitation wavelength of 320 nm (Shimadzu Group Company, Kyoto, Japan); the optical absorption was tested using a UV-3600 ultraviolet–visible photometer (DRS) (Shimadzu Group Company, Kyoto, Japan).

3.3. Photocatalysis Experiment

In total, 100 mL (10 mg/L) MB aqueous solution and 0.025 g samples were mixed into a beaker, placed in a dark state, and stirred for 30 min to reach the equilibrium of adsorption and desorption, and then a 250 W xenon lamp was used as the light source. Sampling occurred every 10 min, the samples were centrifuged to collect the upper clear solution, its absorbance A was measured at λ = 664 nm, and the degradation degree was calculated using the formula (A0 − At)/A0 × 100%, where A0 and At are the initial and absorbance at time.
On the basis of the MB degradation system, 2 mL (0.1 mol/L) of benzoquinone (BQ, ·O2 trapping agent), ammonium oxalate (AO, h+ trapping agent), and isopropanol (IPA, ·OH trapping agent) were added to investigate the active species.

4. Conclusions

The pure and Sn-doped Bi2WO6 nanomaterials with different concentrations were prepared by hydrothermal method, and the effects of Sn doping on the structure and photocatalytic performance of Bi2WO6 were studied. The results show that Sn-Bi2WO6 exhibits 3D flower-like morphology, and the nanosheets are thinner than pure Bi2WO6, which greatly increases the specific surface area. Sn doping does not cause a significant red shift; however, it promotes the separation of photogenerated electron–hole pairs. The photocatalytic degradation results show that the photocatalytic activity of 2%Sn-Bi2WO6 is the highest, and the degradation degree of MB is 92.0% after illumination for 60 min, which is higher than that of pure Bi2WO6 (80.6%). The first-order reaction rate constant of 2%Sn-Bi2WO6 is 0.030 min1, which is 1.7 times that of pure Bi2WO6.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23158422/s1.

Author Contributions

Conceptualization, X.Z. (Xiaodong Zhu), Y.J. and W.F.; methodology, J.W.; data analysis, F.Q., X.Z. (Xiuping Zhang) and Y.Z.; investigation, X.Z. (Xiuping Zhang) and W.F.; writing—original draft preparation, F.Q. and X.Z. (Xiuping Zhang); writing—review and editing, J.W. and Y.L.; performed experiments, Y.Z.; supervision, Y.J. and W.F.; project administration, X.Z. (Xiaodong Zhu). All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Higher Education Talent Quality and Teaching Reform Project of Sichuan Province (JG2021-1104), the Talent Training Quality and Teaching Reform Project of Chengdu University (cdjgb2022033), and the Key Research and Development Projects of Liangshan Prefecture Science and Technology Bureau of Sichuan Province (21ZDYF0202).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saravanan, A.; Kumar, P.S.; Jeevanantham, S.; Anubha, M.; Jayashree, S. Degradation of toxic agrochemicals and pharmaceutical pollutants: Effective and alternative approaches toward photocatalysis. Environ. Pollut. 2022, 298, 118844. [Google Scholar] [CrossRef] [PubMed]
  2. Ahmed, S.; Khan, F.S.A.; Mubarak, N.M.; Khalid, M.; Tan, Y.H.; Mazari, S.A.; Karri, R.R.; Abdullah, E.C. Emerging pollutants and their removal using visible-light responsive photocatalysis—A comprehensive review. J. Environ. Chem. Eng. 2021, 9, 106643. [Google Scholar] [CrossRef]
  3. Wang, Y.Z.; Duan, X.J.; Wang, L.Q.; Zou, H. Spatial temporal patterns and driving factors of industrial pollution and structures in the Yangtze River Economic Belt. Chemosphere 2022, 303, 134996. [Google Scholar] [CrossRef] [PubMed]
  4. Zhu, X.D.; Zhou, Q.; Xia, Y.W.; Wang, J.; Chen, H.J.; Xu, Q.; Liu, J.W.; Feng, W.; Chen, S.H. Preparation and characterization of Cu-doped TiO2 nanomaterials with anatase/rutile/brookite triphasic structure and their photocatalytic activity. J. Mater. Sci. Mater. Electron. 2021, 32, 21511–21524. [Google Scholar] [CrossRef]
  5. Dou, L.; Li, J.J.; Long, N.; Lai, C.X.; Zhong, J.B.; Li, J.Z.; Huang, S.T. Fabrication of 3D flower-like OVs-Bi2SiO5 hierarchical microstructures for visible light-driven removal of tetracycline. Surf. Interfaces 2022, 29, 101787. [Google Scholar] [CrossRef]
  6. Dou, L.; Jin, X.Y.; Chen, J.F.; Zhong, J.B.; Li, J.Z.; Zeng, Y.; Duan, R. One-pot solvothermal fabrication of S-scheme OVs-Bi2O3/Bi2SiO5 microsphere heterojunctions with enhanced photocatalytic performance toward decontamination of organic pollutants. Appl. Surf. Sci. 2020, 527, 146775. [Google Scholar] [CrossRef]
  7. Sun, Y.; Gao, Y.; Zhao, B.S.; Xu, S.; Luo, C.H.; Zhao, Q. One-step hydrothermal preparation and characterization of ZnO–TiO2 nanocomposites for photocatalytic activity. Mater. Res. Express 2020, 7, 085010. [Google Scholar] [CrossRef]
  8. Zhu, X.D.; Wang, J.; Yang, D.X.; Liu, J.W.; He, L.L.; Tang, M.; Feng, W.; Wu, X.Q. Fabrication, characterization and high photocatalytic activity of Ag-ZnO heterojunctions under UV-visible light. RSC Adv. 2021, 11, 27257. [Google Scholar] [CrossRef]
  9. Sun, Y.; Gao, Y.; Zeng, J.Y.; Guo, J.; Wang, H. Enhancing visible-light photocatalytic activity of Ag-TiO2 nanowire composites by one-step hydrothermal process. Mater. Lett. 2020, 279, 128506. [Google Scholar] [CrossRef]
  10. Zhu, X.D.; Xu, H.Y.; Yao, Y.; Liu, H.; Wang, J.; Pu, Y.; Feng, W.; Chen, S.H. Effects of Ag0-modification and Fe3+-doping on the structural, optical and photocatalytic properties of TiO2. RSC Adv. 2019, 9, 40003. [Google Scholar] [CrossRef] [Green Version]
  11. Jia, X.C.; Jiang, D.H.; Gouma, P.I. Facile synthesis of self-supported WO3/PANI hybrid photocatalyst for methylene blue degradation under visible light. Mater. Lett. 2022, 314, 131869. [Google Scholar] [CrossRef]
  12. Sun, X.R.; Gu, M.S.; Yang, J.; Ye, G.H.; Xiao, X.G.; Chen, M.; Liu, M.M.; Chen, Z.L.; Huang, H.S. The photocatalytic performances of Bi2MTaO7 (M = Ga, In) photocatalysts for environmental cleaning under visible-light. Inorg. Chem. Commun. 2022, 139, 109390. [Google Scholar] [CrossRef]
  13. Karthikeyan, N.; Sivaranjani, T.; Dhanavel, S.; Gupta, V.K.; Narayanan, V.; Stephen, A. Visible light degradation of textile effluent by electrodeposited multiphase CuInSe2 semiconductor photocatalysts. J. Mol. Liq. 2017, 227, 194–201. [Google Scholar] [CrossRef]
  14. Gu, Y.F.; Guo, B.B. Promoting photocatalytic performance of Bi2WO6 nanosheet incorporated with a 3D-Succulent plant-like SrMoO4 modified by Ag under simulated sunlight. Opt. Mater. 2021, 121, 111473. [Google Scholar] [CrossRef]
  15. Lai, M.T.L.; Lai, C.W.; Lee, K.M.; Chook, S.W.; Yang, T.C.K.; Chong, S.H.; Juan, J.C. Facile one-pot solvothermal method to synthesize solar active Bi2WO6 for photocatalytic degradation of organic dye. J. Alloys Compd. 2019, 801, 502–510. [Google Scholar] [CrossRef]
  16. Sharma, V.; Kumar, A.; Kumar, A.; Krishnan, V. Enhanced photocatalytic activity of two dimensional ternary nanocomposites of ZnO-Bi2WO6-Ti3C2 MXene under natural sunlight irradiation. Chemosphere 2022, 287, 132119. [Google Scholar] [CrossRef]
  17. Wang, H.; Wang, M.; Chi, H.T.; Zhang, S.T.; Wang, Y.G.; Wu, D.; Wei, Q. Sandwich-type photoelectrochemical immunosensor for procalcitonin detection based on Mn2+ doped CdS sensitized Bi2WO6 and signal amplification of NaYF4:Yb, Tm upconversion nanomaterial. Anal. Chim. Acta 2021, 1188, 339190. [Google Scholar] [CrossRef]
  18. Zhang, Y.Y.; Ma, Y.C.; Liu, Q.Z.; Jiang, H.Y.; Wang, Q.; Qu, D.; Shi, J.S. Synthesis of Er3+/Zn2+ co-doped Bi2WO6 with highly efficient photocatalytic performance under natural indoor weak light illumination. Ceram. Int. 2017, 43, 2598–2605. [Google Scholar] [CrossRef]
  19. Li, J.; Ni, G.; Han, Y.; Ma, Y.M. Synthesis of La doped Bi2WO6 nanosheets with high visible light photocatalytic activity. J. Mater. Sci. Mater. Electron. 2017, 28, 10148–10157. [Google Scholar] [CrossRef]
  20. Liu, Z.; Liu, X.Q.; Wei, L.F.; Yu, C.L.; Yi, J.H.; Ji, H.B. Regulate the crystal and optoelectronic properties of Bi2WO6 nanosheet crystals by Sm3+ doping for superior visible-light-driven photocatalytic performance. Appl. Surf. Sci. 2020, 508, 145309. [Google Scholar] [CrossRef]
  21. Huang, J.; Tan, G.Q.; Ren, H.J.; Xia, A.; Luo, Y.Y. Multi-factors on photocatalytic properties of Y-doped Bi2WO6 crystallites prepared by microwave-hydrothermal method. Cryst. Res. Technol. 2014, 49, 467–473. [Google Scholar] [CrossRef]
  22. Guo, S.; Li, X.F.; Wang, H.Q.; Dong, F.; Wu, Z.B. Fe-ions modified mesoporous Bi2WO6 nanosheets with high visible light photocatalytic activity. J. Colloid Interf. Sci. 2012, 369, 373–380. [Google Scholar] [CrossRef] [PubMed]
  23. Gu, H.D.; Yu, L.; Wang, J.; Ni, M.; Liu, T.T.; Chen, F. Tunable luminescence and enhanced photocatalytic activity for Eu (III) doped Bi2WO6 nanoparticles. Spectrochim. Acta A 2017, 177, 58–62. [Google Scholar] [CrossRef] [PubMed]
  24. Chen, T.; Liu, L.Z.; Hu, C.; Huang, H.W. Recent advances on Bi2WO6-based photocatalysts for environmental and energy applications. Chin. J. Catal. 2021, 42, 1413–1438. [Google Scholar] [CrossRef]
  25. Ma, D.M.; Zhong, J.B.; Li, J.Z.; Wang, L.; Peng, R.F. Enhanced photocatalytic activity of BiOCl by C70 modification and mechanism insight. Appl. Surf. Sci. 2018, 443, 497–505. [Google Scholar] [CrossRef]
  26. Chen, B.; Zhang, J.; Yu, J.; Wang, R.; He, B.B.; Jin, J.; Wang, H.W.; Gong, Y.S. Rational design of all-solid-state TiO2−x/Cu/ZnO Z-scheme heterojunction via ALD-assistance for enhanced photocatalytic activity. J. Colloid Interface Sci. 2022, 607, 760–768. [Google Scholar] [CrossRef]
  27. Wang, J.; Zhu, X.D.; Qin, F.Q.; Wang, Y.X.; Sun, Y.; Feng, W. Synthesis and photocatalytic performance of three-dimensional flower-like bismuth tungstate: Influence of hydrothermal temperature. Mater. Lett. 2022, 314, 131892. [Google Scholar] [CrossRef]
  28. Alhadi, A.; Ma, S. Synthesis of Sn doped-Bi2WO6 nanoslices for enhanced isopropanol sensing properties. Phys. B 2022, 635, 413819. [Google Scholar] [CrossRef]
  29. Chen, Y.; Pen, Z.H.; Wang, Q.Y.; Zhu, J.G. Crystalline structure, ferroelectric properties, and electrical conduction characteristics of W/Cr co-doped Bi4Ti3O12 ceramics. J. Alloys Compd. 2014, 612, 120–125. [Google Scholar] [CrossRef]
  30. Lafi, O.A. Correlation of some opto-electrical properties of Se-Te-Sn glassy semiconductors with the average single bond energy and the average electronegativity. J. Alloys Compd. 2016, 660, 503–508. [Google Scholar] [CrossRef]
  31. Sacher, E.; Currie, J.F. A comparison of electronegativity series. J. Electron Spectrosc. 1988, 46, 173–177. [Google Scholar] [CrossRef]
  32. Fu, H.B.; Zhang, S.C.; Xu, T.G.; Zhu, Y.F.; Chen, J.M. Photocatalytic degradation of RhB by fluorinated Bi2WO6 and distributions of the intermediate products. Environ. Sci. Technol. 2008, 42, 2085–2091. [Google Scholar] [CrossRef] [PubMed]
  33. 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 nanomicrospheres for improved photocatalytic reduction of Cr6+ under visible light irradiation. Appl. Surf. Sci. 2019, 495, 143551. [Google Scholar] [CrossRef]
  34. Wu, C.H.; Kuo, C.Y.; Dong, C.D.; Chen, C.W.; Lin, Y.L.; Wang, W.J. Single-step solvothermal process for synthesizing SnO2/Bi2WO6 composites with high photocatalytic activity in the photodegradation of C. I. Reactive Red 2 under solar light. React. Kinet. Mech. Cat. 2019, 126, 1097–1113. [Google Scholar] [CrossRef]
  35. Li, H.; Hao, H.S.; Jin, S.S.; Guo, W.H.; Hu, X.F.; Hou, H.W.; Zhang, G.L.; Yan, S.; Gao, W.Y.; Liu, G.S. Hydrothermal synthesis and infrared to visible up-conversion luminescence of Ho3+/Yb3+ co-doped Bi2WO6 nanoparticles. Adv. Powder Technol. 2018, 29, 1216–1221. [Google Scholar] [CrossRef]
  36. Kumar, B.V.; Prasad, M.D.; Vithal, M. Enhanced visible light photocatalytic activity of Sn doped Bi2WO6 nanocrystals. Mater. Lett. 2015, 152, 200–202. [Google Scholar] [CrossRef]
  37. Yan, F.Y.; Wang, Y.; Yi, C.H.; Xu, J.X.; Wang, B.W.; Ma, R.; Xu, M. Construction of carbon dots modified Cl-doped Bi2WO6 hollow microspheres for boosting photocatalytic degradation of tetracycline under visible light irradiation. Ceram. Int. 2022, in press. [Google Scholar] [CrossRef]
  38. Jiang, R.R.; Lu, G.H.; Yan, Z.H.; Wu, D.H.; Liu, J.C.; Zhang, X.D. Enhanced photocatalytic activity of a hydrogen bond-assisted 2D/2D Z-scheme SnNb2O6/Bi2WO6 system: Highly efficient separation of photoinduced carriers. J. Colloid Interf. Sci. 2019, 552, 678–688. [Google Scholar] [CrossRef]
  39. Issarapanacheewin, S.; Wetchakun, K.; Phanichphant, S.; Kangwansupamonkon, W.; Wetchakun, N. Efficient photocatalytic degradation of Rhodamine B by a novel CeO2/Bi2WO6 composite film. Catal. Today 2016, 278, 280–290. [Google Scholar] [CrossRef]
  40. Koutavarapu, R.; Babu, B.; Reddy, C.V.; Reddy, I.N.; Reddy, K.R.; Rao, M.C.; Aminabhavi, T.M.; Cho, M.; Kim, D.; Shim, J. ZnO nanosheets-decorated Bi2WO6 nanolayers as efficient photocatalysts for the removal of toxic environmental pollutants and photoelectrochemical solar water oxidation. J. Environ. Manag. 2020, 265, 110504. [Google Scholar] [CrossRef]
  41. Zhang, Y.B.; Xu, C.; Wan, F.C.; Zhou, D.; Yang, L.; Gu, H.S.; Xiong, J. Synthesis of flower-like Bi2Sn2O7/Bi2WO6 hierarchical composites with enhanced visible light photocatalytic performance. J. Alloys Compd. 2019, 788, 1154–1161. [Google Scholar] [CrossRef]
  42. Li, H.; Zhang, J.C.; Yu, J.G.; Cao, S.W. Ultra-Thin carbon-doped Bi2WO6 nanosheets for enhanced photocatalytic CO2 reduction. Trans. Tianjin Univ. 2021, 27, 338–347. [Google Scholar] [CrossRef]
  43. Zhu, X.D.; Zhu, R.R.; Pei, L.X.; Liu, H.; Xu, L.; Wang, J.; Feng, W.; Jiao, Y.; Zhang, W.M. Fabrication, characterization, and photocatalytic activity of anatase/rutile/SnO2 nanocomposites. J. Mater. Sci. Mater. Electron. 2019, 30, 21210–21218. [Google Scholar] [CrossRef]
  44. Hoang, L.H.; Phu, N.D.; Peng, H.; Chen, X.B. High photocatalytic activity N-doped Bi2WO6 nanoparticles using a two-step microwave-assisted and hydrothermal synthesis. J. Alloys Compd. 2018, 744, 228–233. [Google Scholar] [CrossRef]
  45. Hou, X.; Shi, T.L.; Wei, C.H.; Zeng, H.; Hu, X.G.; Yan, B. A 2D-2D heterojunction Bi2WO6/WS2-x as a broad-spectrum bactericide: Sulfur vacancies mediate the interface interactions between biology and nanomaterials. Biomaterials 2020, 243, 119937. [Google Scholar] [CrossRef]
  46. Ren, F.Z.; Zhang, J.H.; Wang, Y.X. Enhanced photocatalytic activities of Bi2WO6 by introducing Zn to replace Bi lattice sites: A first-principles study. RSC Adv. 2015, 5, 29058–29065. [Google Scholar] [CrossRef]
  47. Munisamy, M.; Yang, H.W.; Perumal, N.; Kang, N.; Kang, W.S.; Kim, S.J. A flower-like In2O3 catalyst derived via metal-organic frameworks for photocatalytic applications. Int. J. Mol. Sci. 2022, 23, 4398. [Google Scholar] [CrossRef]
  48. Almofty, S.A.; Nawaz, M.; Qureshi, F.; Al-Mutairi, R. Hydrothermal synthesis of β-Nb2ZnO6 nanoparticles for photocatalytic degradation of methyl orange and cytotoxicity study. Int. J. Mol. Sci. 2022, 23, 4777. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns of samples (a) and the peak of (131) crystal plane (b).
Figure 1. The XRD patterns of samples (a) and the peak of (131) crystal plane (b).
Ijms 23 08422 g001
Figure 2. SEM images of pure Bi2WO6 (a,b), 1%Sn-Bi2WO6 (c,d), 2%Sn-Bi2WO6 (e,f), 4%Sn-Bi2WO6 (g,h), and 6%Sn-Bi2WO6 (i,j), element mappings of Bi, O, W, Sn (ko), and EDS analysis of 2%Sn-Bi2WO6 (p).
Figure 2. SEM images of pure Bi2WO6 (a,b), 1%Sn-Bi2WO6 (c,d), 2%Sn-Bi2WO6 (e,f), 4%Sn-Bi2WO6 (g,h), and 6%Sn-Bi2WO6 (i,j), element mappings of Bi, O, W, Sn (ko), and EDS analysis of 2%Sn-Bi2WO6 (p).
Ijms 23 08422 g002aIjms 23 08422 g002b
Figure 3. TEM images of pure Bi2WO6 (a) and 2%Sn-Bi2WO6 (b), HRTEM images of pure Bi2WO6 (c) and 2%Sn-Bi2WO6 (d).
Figure 3. TEM images of pure Bi2WO6 (a) and 2%Sn-Bi2WO6 (b), HRTEM images of pure Bi2WO6 (c) and 2%Sn-Bi2WO6 (d).
Ijms 23 08422 g003
Figure 4. Pore size distribution curves and N2 adsorption–desorption isotherms of 1%Sn-Bi2WO6 (a,b), 2%Sn-Bi2WO6 (c,d), 4%Sn-Bi2WO6 (e,f), and 6%Sn-Bi2WO6 (g,h).
Figure 4. Pore size distribution curves and N2 adsorption–desorption isotherms of 1%Sn-Bi2WO6 (a,b), 2%Sn-Bi2WO6 (c,d), 4%Sn-Bi2WO6 (e,f), and 6%Sn-Bi2WO6 (g,h).
Ijms 23 08422 g004
Figure 5. XPS survey of pure Bi2WO6 and 2%Sn-Bi2WO6 (a), high resolution spectra of Bi 4f (b), W 4f (c), O 1s (d), and Sn 3d (e).
Figure 5. XPS survey of pure Bi2WO6 and 2%Sn-Bi2WO6 (a), high resolution spectra of Bi 4f (b), W 4f (c), O 1s (d), and Sn 3d (e).
Ijms 23 08422 g005aIjms 23 08422 g005b
Figure 6. PL spectra of samples.
Figure 6. PL spectra of samples.
Ijms 23 08422 g006
Figure 7. Diffuse reflectance spectra (a) and band gap energy (b) of samples.
Figure 7. Diffuse reflectance spectra (a) and band gap energy (b) of samples.
Ijms 23 08422 g007
Figure 8. Photocatalytic degradation degree curves of MB (a) and kinetic fitting curves for pure Bi2WO6 and Sn-Bi2WO6 (b).
Figure 8. Photocatalytic degradation degree curves of MB (a) and kinetic fitting curves for pure Bi2WO6 and Sn-Bi2WO6 (b).
Ijms 23 08422 g008
Figure 9. The degradation degrees of 2%Sn-Bi2WO6 in the presence of different scavengers.
Figure 9. The degradation degrees of 2%Sn-Bi2WO6 in the presence of different scavengers.
Ijms 23 08422 g009
Figure 10. Schematic diagram of charge transfer and photodegradation of MB for 2%Sn-Bi2WO6.
Figure 10. Schematic diagram of charge transfer and photodegradation of MB for 2%Sn-Bi2WO6.
Ijms 23 08422 g010
Table 1. The lattice parameters of samples.
Table 1. The lattice parameters of samples.
SampleLattice ConstantCrystal Vol
/Å3
a/Åb/Åc/Å
Pure-Bi2WO65.487216.31205.4407486.98
1%Sn-Bi2WO65.465216.53655.4354491.23
2%Sn-Bi2WO65.443016.50685.4339488.22
4%Sn-Bi2WO65.510616.72875.3707495.10
6%Sn-Bi2WO65.451816.40355.6174502.36
Table 2. Binding energies of samples.
Table 2. Binding energies of samples.
BiWOSn
Sample4f7/24f5/24f7/24f5/2OLOHOA3d5/23d3/2
Pure Bi2WO6158.9164.335.437.4529.5530.7532.0
1%Sn-Bi2WO6158.9164.335.437.4529.8531.0532.9486.6495.3
2%Sn-Bi2WO6159.0164.435.537.5530.0531.5533.2486.8495.5
4%Sn-Bi2WO6159.2164.635.737.6530.2531.7533.4487.0495.7
6%Sn-Bi2WO6159.3164.735.837.7530.3531.8533.4487.1495.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhu, X.; Qin, F.; Zhang, X.; Zhong, Y.; Wang, J.; Jiao, Y.; Luo, Y.; Feng, W. Synthesis of Tin-Doped Three-Dimensional Flower-like Bismuth Tungstate with Enhanced Photocatalytic Activity. Int. J. Mol. Sci. 2022, 23, 8422. https://doi.org/10.3390/ijms23158422

AMA Style

Zhu X, Qin F, Zhang X, Zhong Y, Wang J, Jiao Y, Luo Y, Feng W. Synthesis of Tin-Doped Three-Dimensional Flower-like Bismuth Tungstate with Enhanced Photocatalytic Activity. International Journal of Molecular Sciences. 2022; 23(15):8422. https://doi.org/10.3390/ijms23158422

Chicago/Turabian Style

Zhu, Xiaodong, Fengqiu Qin, Xiuping Zhang, Yuanyuan Zhong, Juan Wang, Yu Jiao, Yuhao Luo, and Wei Feng. 2022. "Synthesis of Tin-Doped Three-Dimensional Flower-like Bismuth Tungstate with Enhanced Photocatalytic Activity" International Journal of Molecular Sciences 23, no. 15: 8422. https://doi.org/10.3390/ijms23158422

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