AgNi双金属改性多面体钒酸铋的制备及光催化性能
English
Preparation and photocatalytic performance of AgNi bimetallic modified polyhedral bismuth vanadate
-
Key words:
- bismuth vanadate
- / bimetal
- / catalyst
- / photocatalysis
-
0. Introduction
In recent decades, the continuous growth of the global population, rapid industrialization, human activities, and improper management of various hazardous wastes have led to two major problems that require urgent attention: environmental pollution and the energy crisis[1]. When considering only water pollution, approximately 80% of industrial and municipal wastewater is discharged into the environment without proper treatment, leading to shortages of freshwater and drinking water. The development of sustainable energy sources is currently the most promising and effective method for addressing fossil fuel reserves and environmental pollution issues. Solar energy possesses advantages such as being an endless and renewable resource. In recent years, the use of solar energy in photocatalysis has been considered a simple, effective, and cost- efficient environmental remediation method, especially for the degradation of organic pollutants in water[2-4]. Since visible light constitutes the largest portion of the solar spectrum, photocatalysis has been widely utilized with semiconductor photocatalysts to decompose water or degrade organic pollutants under visible light irradiation. This approach has great promise and is sustainable in terms of energy[5-9]. However, conventional photocatalysts, such as titanium dioxide, have band gaps that are too large to respond to visible light, severely limiting their potential applications. Therefore, researchers are dedicated to developing new semiconductor materials, more than 150 semiconductor materials have been developed. Common bismuth-based compounds in photocatalytic materials include BiOX (X=F, Cl, Br, I), Bi4Ti3O12, Bi2O3, Bi2O2CO3, BiVO4, etc. BiVO4 exhibits three crystal structures: monoclinic scheelite, tetragonal zircon, and tetragonal scheelite. All three structures possess excellent photoelectrochemical properties and are non-toxic, making them widely applied in novel high-performance photocatalytic materials. However, BiVO4 faces challenges such as a narrow absorption range and high rates of photogenerated carrier recombination, significantly impacting its applications. To address this issue, researchers suggest that loading a single metal onto the surface of BiVO4 can greatly enhance photocatalytic performance by suppressing the recombination of electrons and holes. Kudo et al.[10] synthesized an Ag-BiVO4 catalyst using a dispersant, which exhibited high efficiency in the degradation of phenol-containing wastewater. Ge et al.[11-12] employed an impregnation method to synthesize Pt-BiVO4 and Pd-BiVO4, reporting significant photocatalytic efficiency in the degradation of methyl orange for both photocatalysts. Zhao et al.[13] synthesized Cu-BiVO4 metal-loaded catalysts using a hydrothermal and wet impregnation approach, and the metal-loaded BiVO4 catalyst outperformed BiVO4 in the photocatalytic degradation of ibuprofen. Wei et al.[14] synthesized Fe/BiVO4 through piezoelectric catalysis, and the addition of iron increased charge carrier density, thereby enhancing the piezoelectric catalytic activity of BiVO4 and effectively degrading dichlorophenol.
There is currently no research available regarding the improvement of photocatalytic performance for the heterojunction photocatalyst AgNi/BiVO4 (AgNi/BVO), which involves loading both Ag and Ni onto BVO. In this study, the simple hydrothermal and chemical reduction method was employed to load AgNi bimetallic nanoalloys onto the facets of BVO. The catalyst′s absorption of visible light was enhanced by the surface plasmon resonance effect of silver and the lattice interface effect of nickel, in turn, promoted the separation of photogenerated electrons. The morphology and physicochemical properties of AgNi/BVO composite catalysts were thoroughly examined. Degradation experiments were conducted with AgNi-loaded catalysts of varying mass ratios to investigate the synergistic effects between the two metals. The results showed that compared with BVO, Ag/BVO and Ni/BVO, and AgNi/BVO composites exhibited better and more stable photocatalytic performance.
1. Experimental
All chemicals were of reagen-grade quality obtained from TianJin Damao Chemical Reagent Factory and used as supplied unless otherwise stated. Deionized water was prepared in our laboratory.
1.1 Synthesis
BVO was synthesized by a typical hydrothermal method. 5 mmol Bi(NO3)3·5H2O was dissolved in 30 mL 0.3 mol·L-1 HNO3 with continuous stirring at room temperature. 5 mmol of NH4VO3 was dissolved in 30 mL of deionized water, and it was slowly added drop by drop into the Bi(NO3)3 solution. The resulting solution was stirred at room temperature for 30 min and then transferred to a 100 mL hydrothermal reactor, where it was hydrothermally treated at 180 ℃ for 12 h. After cooling to room temperature, the product was separated by centrifugation, washed, and dried at 60 ℃ for 10 h. The obtained pure BiVO4 sample was named BVO.
0.5 g of BVO was dispersed in 30 mL of deionized water and subjected to ultrasonication. 0.05 g of AgNO3·6H2O and 0.016 6 g of Ni(NO3)2·6H2O were separately added to the above solution (in a mass ratio of 3∶1 for AgNO3·6H2O and Ni(NO3)2·6H2O) until uniformly dispersed via ultrasonication. After 30 min, a freshly prepared NaBH4 solution was added to the solution (with a molar ratio of 5∶1 for NaBH4 to Ag and 10∶1 for NaBH4 to Ni). After continuous stirring at room temperature for 2 h, the product was centrifuged, washed with deionized water and ethanol, and then dried overnight at 60 ℃ in a vacuum drying oven to obtain the AgNi/BVO catalyst. Using the same method, Ag/BVO and Ni/BVO were prepared under the conditions, where was added 0.05 g of AgNO3·6H2O or 0.016 6 g of Ni(NO3)2·6H2O.
1.2 Characterization
X-ray diffraction (XRD) analysis was performed using Bruker D8 advanced diffractometer at 40 kV and 40 mA with Cu Kα radiation (λ=0.154 18 nm) and a scanning range was 20° to 70°. The morphology and microstructure of the materials were characterized using a scanning electron microscope (SEM, Hitachi SU8010, operating voltage: 15 kV) and a transmission electron microscope (TEM, JEM-2100F, operating voltage: 200 kV) produced by Hitachi Corporation in Japan. UV-Vis diffuse reflectance spectrum (DRS) was measured using an Agilent Cary 5000 spectrometer in the United States. X-ray photoelectron spectroscopy (XPS, thermo escalab 250 XPS system) was used to analyze the surface state of the catalyst, and Al Kα radiation was used as a source of excitation. Brunauer-Emmett-Teller (BET) surface area and pore size distribution were analyzed by nitrogen adsorption. Photoluminescence (PL) was measured using the FluoroMax-4 photoluminescence spectrometer from the Japanese company HORIBA. The excitation wavelength was set at 280 nm.
1.3 Adsorption and photocatalytic performance test
A 500 W Xe lamp was used as the simulated light source. 50 mL solution of MB (methylene blue) (initial concentration ρ0=10 mg·L-1) was placed in the reaction vessel, and 50 mg of the prepared catalyst was added. Continuous stirring was maintained throughout the entire reaction process. A dark reaction was conducted for 30 min initially to reach adsorption equilibrium, followed by 90 min photoreaction. At 30-minute intervals, 3 mL solution was taken, and the absorbance was tested at the end through centrifugation.
2. Results and discussion
2.1 Structure and morphology
The crystal structure and composition of BVO, Ag/BVO, Ni/BVO, and AgNi/BVO were investigated through XRD. As shown in Fig. 1, the diffraction peaks of the BVO sample were in perfect agreement with the monoclinic structure of BiVO4, as indicated by the standard card (PDF No.97-003-3243). Compared to other crystal phases of BiVO4, monoclinic BiVO4 is known to exhibit the highest photocatalytic efficiency[15-16]. For the binary samples, Ag/BVO and Ni/BVO, as well as the ternary composite material AgNi/BVO, all the BVO diffraction peaks were identifiable in the XRD patterns. The intervention of AgNi nanoparticles did not alter the crystal structure of BVO. The diffraction peaks at 28.8°, 30.5°, 35.2°, and 50.3° correspond to the (112), (004), (020), and (220) crystal planes of AgNi/BVO. It was found that the relative intensity of (112) and (004) plane peaks of BVO and AgNi/BVO changed. The crystallinity of AgNi/BVO was superior to that of BVO.
Figure 1
Fig. 2a and 2b shows that BVO presented smooth polyhedral morphology. When Ag nanoparticles were introduced into BVO, Ag nanospheres were uniformly distributed on the BVO surface, as shown in Fig. 2c and 2d. Metallic Ni nanoparticles exhibited a random distribution on the BVO surface, as shown in Fig. 2e and 2f. In the AgNi/BVO sample, the particle size of metallic silver was smaller than that of metallic nickel, and its dispersion was notably better, as shown in Fig. 2g and 2h. The above explanation illustrates that the in-situ reduction method allows for the deposition of metal Ag and Ni nanoparticles on the surface of BVO.
Figure 2
Further detailed morphological information of BVO and AgNi/BVO samples was obtained through TEM (Fig. 3a and 3b). In Fig. 3a, the TEM image of the BVO sample still exhibited a distinct polyhedral morphology. In Fig. 3b, AgNi nanoparticles were observed to be distributed along the edges of the BVO polyhedra. Therefore, the SEM and TEM results both confirm the successful composite formation of the materials. In Fig. 3c-3g, it was the elemental distribution of the AgNi/BVO composite material, revealing the presence of the elements B, O, V, Ag, and Ni in the composite material. The dispersion of each element was good. Further substantiates the successful synthesis of the composite material.
Figure 3
To investigate the surface morphology and chemical composition of AgNi/BVO composite material, XPS testing was performed. As shown in Fig. 4a, the XPS survey spectrum of AgNi/BVO was primarily composed of five elements: Bi, V, O, Ag, and Ni. Individual elemental energy spectrum analyses were conducted for all elements. As shown in Fig. 4b, in the XPS spectrum of Bi4f, two distinct peaks appeared at binding energies of 159.06 and 164.35 eV, corresponding to the spin-orbit split peaks of Bi4f7/2 and Bi4f5/2. Fig. 4c displayed a high-resolution spectrum of V2p, showing the spin-orbit split peaks of V2p3/2 and V2p1/2, with binding energies of 516.65 and 524.09 eV. The above results indicate that the Bi and V in the AgNi/BVO sample exist in the +3 and +5 oxidation states. Fig. 4d represented a typical high-resolution spectrum of O1s in which a characteristic peak at a binding energy of 529.68 eV appeared. Fig. 4e displayed the XPS spectra of Ag3d. Two prominent peaks located at 367.83 and 374.01 eV, corresponding to Ag3d5/2 and Ag3d3/2. The presence of metallic Ag has been substantiated[17-18]. In Fig. 4f, the peak located at 855.64 eV corresponded to Ni2p3/2, while Ni2p1/2 appeared at 872.96 eV. The two faint peaks appearing at 861.38 and 879.26 eV may be attributed to some oxidation of Ni due to surface exposure or testing, which led to the formation of nickel oxides. Similar phenomena have been reported in other literature[19-20]. These XPS results align with the SEM findings, providing further evidence that both elemental Ag and Ni have been introduced into the BVO system simultaneously.
Figure 4
2.2 Specific surface area and optical properties
An N2 adsorption-desorption was employed for BET testing of BVO, Ag/BVO, Ni/BVO, and AgNi/BVO. As shown in Fig. 5a, all four samples exhibited type Ⅳ isotherms, with AgNi/BVO displaying a pronounced H3 hysteresis loop. This is because the metal AgNi nanoparticles generated by the in-situ reduction method are not only partially loaded on the surface of BVO but also adhere to the pore structure of BVO, further confirming the successful loading of AgNi on BVO. From the pore size distribution curve in Fig. 5b, it can be concluded that the pore sizes of the four samples were mainly distributed in the range of 1 to 10 nm. This indicates that all four materials belong to mesoporous materials. The specific surface area, pore volume, and average pore size can be seen in Table 1. From Table 1, it can be observed that the BET-specific surface areas of BVO, Ag/BVO, Ni/BVO, and AgNi/BVO were 0, 1, 2, and 1 m2·g-1. It can be concluded that the specific surface areas of all four samples were relatively small and had similar values. Considering the degradation data, it is evident that the specific surface area of the catalyst is not the sole influencing factor for catalyst activity. The loading of AgNi enhances the catalyst′s pore volume, thereby facilitating the diffusion of reactant molecules into the pores.
Figure 5
Table 1
Sample BET surface area/(m2·g-1) Average pore diameter/nm Pore volume/(cm3·g-1) BVO 0 1.347 0 0.000 7 Ag/BVO 1 1.168 2 0004 2 Ni/BVO 2 1.415 3 0.004 2 AgNi/BVO 1 1.098 7 0.004 7 The optical absorption performance of the prepared samples was tested using UV-Vis diffuse reflectance spectroscopy (UV-Vis DRS), as shown in Fig. 6a. Within the wavelength range of less than 600 nm, BVO exhibited light absorption characteristics in both the ultraviolet and visible light regions. In comparison to BVO, the absorption peaks of Ag/BVO, Ni/BVO, and AgNi/BVO composite material showed a redshift. The AgNi/BVO composite material exhibited a significantly broadened absorption band in the visible light region, with a noticeable enhancement in light absorption capacity. Indicating that the AgNi/BVO composite material possessed visible light-responsive properties[21-23].
Figure 6
Walsh et al.[24] employed density functional theory (DFT) calculation to demonstrate that BiVO4 is a direct bandgap semiconductor. The valence band formed by the hybridization of Bi6s and O2p orbitals results in a relatively small bandgap energy for BiVO4, which can be calculated from the plot of (αhν)2 versus photon energy (hν)[25-26]. The intercept of the X-axis on the plot can effectively approximate the bandgap value of the sample. As shown in Fig. 6b, the calculated bandgap widths (Eg) for BVO, Ni/BVO, Ag/BVO, and AgNi/BVO were 2.35, 2.33, 2.25, and 2.18 eV, respectively. Compared to BVO, the inclusion of AgNi narrows the bandgap of the AgNi/BVO material, which is related to the control of the metal′s morphology on BVO[27].
Fluorescence analysis is an analytical method employed to investigate and assess the performance of photocatalysts. PL testing was carried out at room temperature to assess the recombination rate of photo-generated charge carriers in BVO, Ni/BVO, Ag/BVO, and AgNi/BVO catalysts, to explore the role of metals in the electron-hole separation within composite materials. As shown in Fig. 7, all four samples exhibited their most intense photoluminescence at 383 nm. Compared to BVO, the loading of metal cocatalysts decreased the emission peak intensity in the following order: BVO > Ni/BVO > Ag/BVO > AgNi/BVO. This indicates that metal cocatalysts can effectively suppress the recombination of photo-generated electron-hole pairs at the interface. The AgNi/BVO alloy exhibited the lowest emission peak intensity, suggesting that in comparison to single-metal loading, dual-metal loading can further optimize the electron migration pathway through secondary electron transfer, thereby enhancing the separation efficiency of photo-generated charge carriers in BVO more effectively.
Figure 7
2.3 Photocatalytic properties
The photocatalytic degradation efficiency of the samples for MB was evaluated, and reaction kinetics constants were calculated based on the first-order kinetic equation. Fig. 8a illustrates the photocatalytic degradation efficiency of the samples for MB. The relatively smooth surface structure of BVO, which resulted in a higher degree of recombination of photogenerated charge carriers and lower utilization of visible light, its photocatalytic performance was the weakest, degrading only 29.63% of MB. Under the same conditions, the photocatalytic performance was slightly enhanced after loading single metals, Ag and Ni, onto BVO catalysts. After 120 min reaction, Ag/BVO and Ni/BVO could degrade 57.49% and 50.76% of MB. After loading AgNi alloy, the degradation performance of the catalyst could be elevated to over 93%. This is the synergistic effect between the AgNi bimetallic catalyst, which effectively promotes the separation of photogenerated electrons and holes, while enhancing the catalyst′s absorption of visible light, thereby improving the photocatalytic performance of the BVO catalyst. The influence of loading various mass ratios of Ag to Ni (3∶1, 2∶1, 1∶1, 1∶2, 1∶3) on the degradation performance of MB by BVO was investigated, as shown in Fig. 8b. When the ratios were 3∶1, 2∶1, 1∶1, 1∶2, and 1∶3, the degradation rates for MB under the same conditions were 93.00%, 69.54%, 66.94%, 22.55%, and 8.38%, respectively. It was observed that the ratio of 3∶1 resulted in the highest degradation rate. Under this ratio, the electron transfer between Ag and Ni bimetals was maximized, enhancing the interaction between reactants and the catalyst, thus facilitating the photocatalytic reaction. As shown in Fig. 8c, according to the Langmuir-Hinshelwood model, the photocatalytic reaction follows pseudo-first-order kinetics. The initial concentration after adsorption equilibrium was ρ0′, and ρ was the concentration at the time of reaction (t) after illumination. The apparent first-order reaction rate constant can be obtained. The rate constant for AgNi/BVO was the highest, being 5.4 times that of BVO. The UV-Vis DRS of the MB solution on the AgNi/BVO photocatalyst at different illumination times are shown in Fig. 8d. The absorption peak intensity at 664 nm rapidly decreased, indicating the effective decomposition of MB. Throughout the degradation process, there was no shift in the absorption peak position of the MB solution, and no additional peaks were formed in the ultraviolet-visible absorption spectra of the MB solution. Fig. 8e and 8f present the degradation profiles and first-order kinetic linear relationship graphs of rhodamine B (RhB, 10 mg·L-1) under the influence of BVO, Ni/BVO, Ag/BVO, and AgNi/BVO catalysts. It was observed that the degradation trend of RhB was consistent with that of MB. Furthermore, upon the loading of the AgNi bimetallic catalyst, the degradation rate of RhB was significantly improved. This method demonstrates that the modification of BVO catalysts is suitable for degrading a wide range of organic pollutants. The synthesized AgNi/BVO catalyst was compared with other bismuth vanadate catalysts reported in existing literature for photocatalytic degradation efficiency, as shown in Table 2. Through this comparison, it is found that the photocatalytic performance of AgNi/BVO obtained in this study has a certain competitive advantage.
Figure 8
Table 2
Catalyst Light source Amount of catalyst/mg Pollutant Time/min Degradation rate/% Ref. AgNi/BVO Xe lamp (500 W) 50 MB 90 93.0 This work PANI-BiVO4-GO Xe lamp (500 W) 100 MB 180 73 [28] BiVO4-Cement composites Two Havells brand bulbs (each 15 W) 100 MB 240 58 [29] rGO-BiVO4 Two Havells brand bulbs (each 15 W) 50 MB 180 52 [30] BiVO4-Al2O3 25 W fluorescent lamp 200 MB 60 86 [31] BiVO4-SiO2 Three 18 W halogen lamps 200 MB 120 88 [32] Stability tests were conducted on AgNi/BVO, and the post-experimental catalyst underwent centrifugation, filtration, washing, and drying, as shown in Fig. 9. After four cycles of experiments, a slight decrease in the degradation performance of MB by AgNi/BVO catalyst was observed. This could be attributed to the minor loss of the catalyst during the washing and centrifugation processes, or it might be a small amount of residual MB occupying some active sites on the catalyst′s surface. This still demonstrates that AgNi/BVO can serve as a photocatalyst with excellent cycling stability.
Figure 9
To investigate the active species in the photocatalytic reaction, isopropanol (IPA), p-benzoquinone (BQ), and acid disodium salt (EDTA-2Na) were used as scavengers for hydroxyl radicals (·OH), superoxide radicals (·O2-), and holes (h+). The mass concentrations of the three kinds of traps were 1 g·L-1. As shown in Fig. 10, the photocatalytic activity of AgNi/BVO significantly decreased when EDTA-2Na was introduced into the reaction system, indicating that h+ was the primary active species. The addition of BQ and IPA resulted in a relatively minor decrease in the degradation rate, suggesting that ·O2- and ·OH had a relatively smaller impact on the photocatalytic system.
Figure 10
2.4 Mechanism of removal of MB by photocatalytic degradation
Based on the above-mentioned research, the photocatalytic degradation mechanism of MB by AgNi/BVO material was proposed, as illustrated in Fig. 11. When the AgNi bimetal was loaded onto the BVO material, under visible light irradiation, electrons in BVO were excited, causing the transition of electrons (e-) from the valence band (VB) to the conduction band (CB). Under the influence of the AgNi bimetal, e- reacted with the oxygen (O2) adsorbed on its surface, thereby generating active free radicals ·O2-. Simultaneously, a portion of h+ in BVO oxidized H2O into ·OH. The substantial presence of h+ remaining in the VB facilitated the effective separation of electrons and holes. In the end, h+, ·OH, and ·O2- degrade MB into small molecules such as CO2 and H2O, which is in basic agreement with the results of free radical capture experiments.
Figure 11
3. Conclusions
In this study, a novel silver-nickel bimetal-loaded polyhedral bismuth vanadate heterojunction was successfully prepared for the first time through a two-step process involving a simple hydrothermal method and chemical reduction. The enhanced photocatalytic degradation performance of AgNi/BVO compared to BVO can be attributed to two factors: Firstly, the AgNi alloy significantly enhances the utilization of visible light by BVO, and secondly, the synergistic interaction between the bimetal components facilitates more efficient separation of photo-generated electron-hole pairs. Compared to BVO, the introduction of AgNi bimetal resulted in a photodegradation rate of approximately 93% for MB after 1.5 h of illumination, which was 5.4 times more efficient than BVO. Additionally, it exhibited excellent degradation performance for RhB as well. Even after four cycles of photocatalytic experiments, AgNi/BVO maintained its stability and photocatalytic activity. The preparation of AgNi/BVO was simple, environmentally friendly, and exhibited a degree of universality in its synthesis process. This study provides valuable guidance for the development of other catalysts with similar structures that possess high-efficiency photocatalytic performance.
-
-
[1]
Queiroz B D, Fernandes J A, Martins C A, Wender H. Photocatalytic fuel cells: From batch to microfluidics[J]. J. Environ. Chem. Eng., 2022, 10(3): 107611. doi: 10.1016/j.jece.2022.107611
-
[2]
Sahoo D P, Das K K, Patnaik S, Parida K. Double charge carrier mechanism through 2D/2D interface-assisted ultrafast water reduction and antibiotic degradation over architectural S, P co-doped g-C3N4/ ZnCr LDH photocatalyst[J]. Inorg. Chem. Front., 2020, 7(19): 3695-3717. doi: 10.1039/D0QI00617C
-
[3]
赵东启, 顾贵洲, 李政. 水滑石类光催化剂在污水处理中的应用[J]. 石油化工高等学校学报, 2022,35,(3): 36-42. ZHAO D Q, GU G Z, LI Z. Application of hydrotalcite-like photocatalyst in wastewater treatment[J]. Journal of Petrochemical Universities, 2022, 35(3): 36-42.
-
[4]
张若男, 何仕婕, 王晓蓉, 王芳芳, 陈常东. BaTiO3-TiO2复合体的软化学制备及其性能研究[J]. 石油化工高等学校学报, 2022,35,(4): 46-51. ZHANG R N, HE S J, WANG X R, WANG F F, CHEN C D. Study on soft chemical preparation and properties of BatiO3-TiO2 complex[J]. Journal of Petrochemical Universities, 2022, 35(4): 46-51.
-
[5]
Kublik N, Gomes L E, Plaça L F, Lima T H N, Abelha T F, Ferencz J A P, Caires A R L, Wender H. Metal-free g-C3N4/nanodiamond heterostructures for enhanced photocatalytic pollutant removal and bacteria photoinactivation[J]. Photochem, 2021, 1(2): 302-318. doi: 10.3390/photochem1020019
-
[6]
Stelo F, Kublik N, Ullah S, Wender H. Recent advances in Bi2MoO6 based Z-scheme heterojunctions for photocatalytic degradation of pollutants[J]. J. Alloy. Compd., 2020, 829: 154591. doi: 10.1016/j.jallcom.2020.154591
-
[7]
Nogueira A C, Gomes L E, Ferencz J A P, Rodrigues J E F S, Gonçalves R V, Wender H. Improved visible light photoactivity of CuBi2O4/CuO heterojunctions for photodegradation of methylene blue and metronidazole[J]. J. Phys. Chem. C, 2019, 123: 25680-25690. doi: 10.1021/acs.jpcc.9b06907
-
[8]
Gomes L E, Silva M F, Gonçalves R V, Machado G, Alcantara G B, Caires A R L, Wender H. Synthesis and visible-light-driven photocatalytic activity of Ta4+ self-doped gray Ta2O5 nanoparticles[J]. J. Phys. Chem. C, 2018, 122: 6014-6025.
-
[9]
Long M, Cai W M, Cai J, Zhou B X, Chai X Y, Wu Y H. Efficient photocatalytic degradation of phenol over Co3O4/BiVO4 composite under visible light irradiation[J]. J. Phys. Chem. B, 2006, 110(41): 20211-20216. doi: 10.1021/jp063441z
-
[10]
Kohtani S, Hiro J, Yamamoto N, Kudo A, Tokumura K, Nakagaki R. Adsorptive and photocatalytic properties of Ag-loaded BiVO4 on the degradation of 4-n-alkylphenols under visible light irradiation[J]. Catal. Commun., 2008, 6(3): 185-189.
-
[11]
Ge L. Novel Pd/BiVO4 composite photocatalysts for efficient degradation of methyl orange under visible light irradiation[J]. Mater. Chem. Phys., 2008, 107(2): 465-470.
-
[12]
Ge L. Novel visible-light-driven Pt/BiVO4 photocatalyst for efficient degradation of methyl orange[J]. J. Mol. Catal. A-Chem., 2008, 282(1/2): 62-66.
-
[13]
Bian Z Y, Zhu Y Q, Zhang G X, Ding A Z, Wang H. Visible-light driven degradation of ibuprofen using abundant metal-loaded BiVO4 photocatalysts[J]. Chemosphere, 2014, 117: 527-531. doi: 10.1016/j.chemosphere.2014.09.017
-
[14]
Wei Y, Zhang Y W, Miao J, Geng W, Long M C. In-situ utilization of piezo-generated hydrogen peroxide for efficient p-chlorophenol degradation by Fe loading bismuth vanadate[J]. Appl. Surf. Sci., 2021, 543: 148791. doi: 10.1016/j.apsusc.2020.148791
-
[15]
Kudo A, Omori K, Kato H. A novel aqueous process for the preparation of crystal form-controlled and highly crystalline BiVO4 powder from layered vanadates at room temperature and its photocatalytic and photophysical properties[J]. J. Am. Chem. Soc., 1999, 121(49): 11459-11467. doi: 10.1021/ja992541y
-
[16]
Tokunaga S, Kato H, Kudo A. Selective preparation of monoclinic and tetragonal BiVO4 with scheelite structure and their photocatalytic properties[J]. Chem. Mater., 2001, 13(12): 4624-4628. doi: 10.1021/cm0103390
-
[17]
Ong W J, Putri L K, Tan L L, Chai S P, Yong S T. Heterostructured AgX/g-C3N4 (X=Cl and Br) nanocomposites via a sonication-assisted deposition precipitation approach: The emerging role of halide ions in the synergistic photocatalytic reduction of carbon dioxide[J]. Appl. Catal. B-Environ., 2016, 180: 530-543. doi: 10.1016/j.apcatb.2015.06.053
-
[18]
徐鹏, 李佑稷, 刘晨, 李铭, 邓瑞成, 匡美容. 银掺杂介孔二氧化钛的制备及可见光催化性能[J]. 硅酸盐学报, 2014,42,(9): 1196-1202. XU P, LI Y J, LIU C, LI M, DENG R C, KUANG M R. Preparation and visible light catalytic performance of silver doped mesoporous titanium dioxide[J]. J. Chin. Ceram. Soc., 2014, 42(9): 1196-1202.
-
[19]
Zhang B, Zhang X M, Wei Y, Xia L, Pi C R, Song H, Zheng Y, Gao B, Fu J J, Chu P K. General synthesis of NiCo alloy nanochain arrays with thin oxide coating: A highly efficient bifunctional electrocatalyst for overall water splitting[J]. J. Alloy. Compd., 2019, 797: 1216-1223. doi: 10.1016/j.jallcom.2019.05.036
-
[20]
Guo Z W, Liu T, Wang Q T, Gao G H. Construction of cost-effective bimetallic nanoparticles on titanium carbides as a superb catalyst for promoting hydrolysis of ammonia borane[J]. RSC Adv., 2018, 8(2): 843-847. doi: 10.1039/C7RA10568A
-
[21]
张晓燕, 陈飘, 赵莹鑫, 李雄建, 杨水金, 杨赟. MOF-808/Bi2MoO6复合材料的构筑及其光催化性能[J]. 无机化学学报, 2023,39,(5): 805-814. ZHANG X Y, CHEN P, ZHAO Y X, LI X J, YANG S J, YANG Y. Fabrication and photocatalytic properties of MOF-808/Bi2MoO6 composites[J]. Chinese J. Inorg. Chem., 2023, 39(5): 805-814.
-
[22]
赵莹鑫, 胡豪, 周鑫, 杨水金, 杨赟. MOF-808/BiOCl复合材料的制备及其光催化性能[J]. 无机化学学报, 2023,39,(8): 1553-1563. ZHAO Y X, HU H, ZHOU X, YANG S J, YANG Y. Preparation and photocatalytic degradation performance of MOF-808/BiOCl composites[J]. Chinese J. Inorg. Chem., 2023, 39(8): 1553-1563.
-
[23]
梁梦君, 邓楠, 向心怡, 梅英, 杨志远, 杨赟, 杨水金. Bi/BiVO4 & Bi4V2O11复合催化剂的制备及其光催化性能[J]. 无机化学学报, 2019,35,(2): 263-270. LIANG M J, DENG N, XIANG X Y, MEI Y, YANG Z Y, YANG Y, YANG S J. Bi/BiVO4 & Bi4V2O11 composite catalysts: Preparation and photocatalytic performance[J]. Chinese J. Inorg. Chem., 2019, 35(2): 263-270.
-
[24]
Walsh A, Yan Y F, Huda M N, Al-Jassim M M, Wei S H. Band edge electronic structure of BiVO4: Elucidating the role of the Bi s and V d orbitals[J]. Chem. Mater., 2009, 21(3): 547-551.
-
[25]
Sun S M, Wang W Z, Zhang L, Yin W Z, Shang M. Visible light- induced efficient contaminant removal by Bi5O7I[J]. Environ. Sci. Technol., 2009, 43(6): 2005-2010.
-
[26]
Chang X F, Huang J, Cheng C, Sui Q, Sha W, Ji G B, Deng S B, Yu G. BiOX (X=Cl, Br, I) photocatalysts prepared using NaBiO3 as the Bi source: Characterization and catalytic performance[J]. Catal. Commun., 2010, 11(5): 460-464.
-
[27]
Zhang L, Wang W Z, Zhou L, Xu H L. Bi2WO6 nano and microstructures: Shape control and associated visible-light-driven photocatalytic activities[J]. Small, 2007, 3(9): 1618-1625.
-
[28]
Oh W C, Liu Y, Cho K Y, Biswas M R U D. Sonochemical synthesis of PANI-BiVO4-GO semiconductor nanocomposite highly efficient visible-light photocatalytic performance[J]. Fuller. Nanotub. Carbon Nanostruct., 2020, 28(11): 945-958.
-
[29]
Kumar M, Elqahtani Z M, Alrowaili Z A, Al-Buriahi M S, Kebaili I, Boukhris I, Vaish R. Photocatalytic BiVO4-cement composites for dye degradation[J]. J. Electron. Mater., 2023, 52: 4672-4685.
-
[30]
Kumar M, Ansari M N M, Boukhris I, Al-Buriahi M S, Alrowaili Z A, Alfryyan N, Thomas P, Vaish R. Sonophotocatalytic dye degradation using rGO-BiVO4 composites[J]. Global Challenges, 2022, 6(6): 2100132.
-
[31]
Sánchez-Albores R M, Pérez-Sariñana B Y, Meza-Avendaño C A, Sebastian P G, Reyes-Vallejo O, Robles-Ocampo J B. Hydrothermal synthesis of bismuth vanadate-alumina assisted by microwaves to evaluate the photocatalytic activity in the degradation of methylene blue[J]. Catal. Today, 2020, 353(15): 126-133.
-
[32]
Channei D, Nakaruk A, Khanitchaidecha W, Jannoey P, Phanichphant S. Adsorption and photocatalytic processes of mesoporous SiO2-coated monoclinic BiVO4[J]. Front. Chem., 2018, 6: 415.
-
[1]
-
Figure 8 (a) Degradation activity of MB by different catalysts; (b) Degradation activity of MB by different masses of AgNi/BVO catalysts; (c) Kinetic fitting of MB degradation of different catalysts; (d) UV-Vis DRS of MB by AgNi/BVO catalyst with irradiation time; (e) Degradation activity of RhB by different catalysts; (f) Kinetic fitting of RhB degradation of different catalysts
Table 1. BET specific surface areas, average pore diameters, and pore volumes of the samples
Sample BET surface area/(m2·g-1) Average pore diameter/nm Pore volume/(cm3·g-1) BVO 0 1.347 0 0.000 7 Ag/BVO 1 1.168 2 0004 2 Ni/BVO 2 1.415 3 0.004 2 AgNi/BVO 1 1.098 7 0.004 7 Table 2. Comparison of photocatalytic degradation of MB by different composites under visible light irradiation
Catalyst Light source Amount of catalyst/mg Pollutant Time/min Degradation rate/% Ref. AgNi/BVO Xe lamp (500 W) 50 MB 90 93.0 This work PANI-BiVO4-GO Xe lamp (500 W) 100 MB 180 73 [28] BiVO4-Cement composites Two Havells brand bulbs (each 15 W) 100 MB 240 58 [29] rGO-BiVO4 Two Havells brand bulbs (each 15 W) 50 MB 180 52 [30] BiVO4-Al2O3 25 W fluorescent lamp 200 MB 60 86 [31] BiVO4-SiO2 Three 18 W halogen lamps 200 MB 120 88 [32]
计量
- PDF下载量: 0
- 文章访问数: 191
- HTML全文浏览量: 20