Next Article in Journal
Strain-Compensated InGaAsP Superlattices for Defect Reduction of InP Grown on Exact-Oriented (001) Patterned Si Substrates by Metal Organic Chemical Vapor Deposition
Next Article in Special Issue
New Insights into Sensitization Mechanism of the Doped Ce (IV) into Strontium Titanate
Previous Article in Journal
Efficiency Evaluation of Food Waste Materials for the Removal of Metals and Metalloids from Complex Multi-Element Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Insights into Mn1−xZnxFe2O4 via Fabricating Magnetic Photocatalyst Material BiVO4/Mn1−xZnxFe2O4

1
State Key Laboratory of Coal Mine Disaster Dynamics and Control, Chongqing University, Chongqing 400044, China
2
Chongqing Key Laboratory of Extraordinary Bond Engineering and Advanced Materials Technology (EBEAM), Yangtze Normal University, Chongqing 408100, China
3
College of Chemistry and Chemical Engineering, Chongqing University, Chongqing 400044, China
*
Author to whom correspondence should be addressed.
Materials 2018, 11(3), 335; https://doi.org/10.3390/ma11030335
Submission received: 3 January 2018 / Revised: 9 February 2018 / Accepted: 13 February 2018 / Published: 26 February 2018
(This article belongs to the Special Issue Photocatalysis for Wastewater Treatment)

Abstract

:
BiVO4/Mn1−xZnxFe2O4 was prepared by the impregnation roasting method. XRD (X-ray Diffractometer) tests showed that the prepared BiVO4 is monoclinic crystal, and the introduction of Mn1−xZnxFe2O4 does not change the crystal structure of BiVO4. The introduction of a soft-magnetic material, Mn1−xZnxFe2O4, was beneficial to the composite photocatalyst’s separation from the liquid solution using an extra magnet after use. UV-vis spectra analysis indicated that Mn1−xZnxFe2O4 enhanced the absorption intensity of visible light for BiVO4. EIS (electrochemical impedance spectroscopy) investigation revealed that the introduction of Mn1−xZnxFe2O4 enhanced the conductivity of BiVO4, further decreasing its electron transfer impedance. The photocatalytic efficiency of BiVO4/Mn1−xZnxFe2O4 was higher than that of pure BiVO4. In other words, Mn1−xZnxFe2O4 could enhance the photocatalytic reaction rate.

Graphical Abstract

1. Introduction

Photocatalysis and photocatalytic technology, which uses semiconductor materials to directly absorb and transmute renewable solar light energy into chemical energy, have been considered as promising methods to resolve environmental and energy problems facing the human population [1]. To date, cleaning up organic compounds via degradation and water splitting to produce H2 are the two most important applications of photocatalysis and its corresponding technology, which are aimed at environmental pollutant treatment and molecular hydrogen production, respectively [2].
In spite of the promising results regarding solar-light-driven photocatalysts, there are still some challenges inhibiting their practical application. The most pressing problem concerns the photocatalytic reaction kinetics using photocatalysts under solar light irradiation [3]. Most of the degradation reactions under solar light irradiation are very slow (they may take several hours). Therefore, it is very important to exploit new and highly efficient photocatalysts. In addition, the intrinsic relationship between photocatalytic activity and photocatalytic material structure can be elucidated by studying the photocatalytic mechanism, which will guide the synthesis and application of new and more efficient photocatalytic systems.
A monoclinic scheelite structure, BiVO4, with a better absorption ability of visible light, has attracted a great deal of attention due to its a relatively narrow band gap. The hybridization of Bi6s-O2p orbitals upshifted the valence band of monoclinic BiVO4 to a lower potential at about +2.4 eV [2,4,5,6,7]. Nonetheless, the photocatalytic efficiency of BiVO4 is generally low because of its poor electron transfer and slow reaction kinetics.
Given that the photocatalyst materials could not be thoroughly recycled from a liquid solution after reaction, secondary pollution would probably be induced by the residual photocatalyst. This prevents it application as a wastewater treatment. Fortunately, a magnetic photocatalyst could solve the above issue [8,9]. The exploitation and application of various magnetic photocatalyst materials have been boosting scientists’ morale.
Here, we use Mn1−xZnxFe2O4 as a soft-magnetic substrate to prepare magnetic photocatalyst BiVO4/Mn1−xZnxFe2O4. In the composite system, the magnetic substrate, Mn1−xZnxFe2O4, facilitated the recovery of the photocatalyst from the liquid reaction solution using an extra magnet after the reaction. Most importantly, Mn1−xZnxFe2O4 could enhance the photocatalytic reaction rate of BiVO4 by enhancing the absorption intensity of visible light for BiVO4 and heightening the conductivity of BiVO4. We sincerely hope to extend the application field of Mn1−xZnxFe2O4 according to this report.

2. Experimental Procedure

All reagents purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China) were of analytical grade purity and were used directly without further purification. The water used in all experimental processes was deionized water.

2.1. Preparation of BiVO4/Mn1−xZnxFe2O4

Preparation of Mn1xZnxFe2O4 [10]. ZnSO4, MnSO4, and FeCl3⋅6H2O with a given molar ratio of n (ZnO): n (MnO): n (Fe2O3) = 13.3:32.8:53.9 was separately weighed and dissolved in water to form three solutions. Subsequently, the prepared ZnSO4 solution and FeCl3 solution were added into the MnSO4 solution under continuous stirring condition, in order to form the mixture solution. A definite amount of (NH4)2C2O4⋅H2O was dissolved to form a (NH4)2C2O4 solution. The prepared mixture solution was slowly added into the (NH4)2C2O4 solution, and the mixture solution and (NH4)2C2O4 solution were heated to 80 °C, respectively. The pH value of the whole reaction system was adjusted to 7 by slowly adding NaOH solution, and a large quantity of precipitate was observed. The precipitate was washed using water and dried at 80 °C for 12 h after filtrating to gain the precursor. The precursor was sintered at 1200 °C for 3 h to form the resultant magnetic substrate, MnxZn1−xFe2O4.
Preparation of BiVO4 [11]. First, 0.01 mol Bi(NO3)3·5H2O and 0.02 mol tartaric acid were dissolved using a certain amount of 2 mol/L HNO3 solution to form a mixture solution A. Then, its pH value was adjusted to 7.5. Subsequently, 0.01 mol NH4VO3 was dissolved in 50 mL hot water (70 °C) to form solution B. Solution A and solution B were mixed to form the precursor of BiVO4. The precursor was dried at 80 °C for 24 h and then sintered at 500 °C for 4 h to obtain pure BiVO4.
Preparation of BiVO4/Mn1xZnxFe2O4. First, 0.01 mol Bi(NO3)3·5H2O and 0.02 mol tartaric acid were dissolved with a certain amount of 2 mol/L HNO3 solution to form a mixture solution A. Then, its pH value was adjusted to 7.5. Subsequently, 0.01 mol NH4VO3 was dissolved in 50 mL hot water (70 °C) to from solution B. Solution A and solution B were mixed to form C solution. Then, 15.0 wt % MnxZn1−xFe2O4 was added to solution C. BiVO4/MnxZn1−xFe2O4 (15.0 wt %) was obtained after the mixture was dried at 80 °C for 24 h and then sintered at 500 °C for 4 h. The composites BiVO4/MnxZn1−xFe2O4 (10.0%, 20.0%, 25.0 wt %) were prepared by adjusting the mass ratio of MnxZn1−xFe2O4.

2.2. Materials Characterization

The structure characterization of the samples was determined by X-ray Diffractometer (Shimadzu, XRD-6000, Kyoto, Japan), Fourier transform infrared spectroscopy (FTIR, Perkin-Elmersystem 2000, Akron, OH, USA), and INVIA Raman microprobe (Renishaw Instruments, Wotton-under-Edge, UK). The light absorption properties, magnetization, and surface performances of products were examined by an ultraviolet-visible diffuse reflectance spectrophotometer (UV-vis, DRS, TU1901, Company, Beijing, China), vibrating sample magnetometer (VSM 7410, Lake Shore, Carson, CA, USA), Brunauer−Emmett−Teller (BET, ASAP-2020, Micromeritics, Norcross, GA, USA), and scanning electron microscopy (SEM, EVO-LS15X, ZEISS, Upper Cohen, Germany). The electrochemical workstation (PGSTAT30) was employed to measure the electrochemical impedance spectroscopy (EIS) of samples. The following are the test EIS parameters: K3[Fe(CN)6]:K4[Fe(CN)6] (1:1)-KCl electrolyte solution was employed. The work electrode content contained the prepared photocatalytic materials, acetylene black, and polytetrafluoroethylene (mass ratio, 85:10:5); the counter electrode was platinum foil; and the reference electrode was a saturated calomel electrode (SCE). Finally, the AC (Alternating Current) voltage amplitude was set at 5 mV and the frequency range was 1 × 105~1 × 10−2 Hz.

2.3. Photocatalytic Tests

The photocatalytic activity of the prepared photocatalysts was investigated by the degradation of simulated dye wastewater (Rhodamine B, RhB) under visible light irradiation. One hundred milligrams photocatalyst was put into 100 mL RhB solution with a 5 mg/L concentration, then the suspension liquid was placed in the dark for 0.5 h with stirring to reach the adsorption-desorption equilibrium. A 500 W Xe lamp, equipping with a UV cut-off filter, was used as the visible light source (λ ≥ 420 nm). At the given irradiation time intervals, a series of the reaction solution was withdrawn and the absorbance was measured using the UV-vis spectrophotometer (TU-1901).

3. Results and Discussion

Primary analysis of photodegradation revealed that BiVO4/Mn1−xZnxFe2O4 (15 wt %) was the most efficient in the RhB degradation process under visible light irradiation.

3.1. Structure and Specific Surface Property

Figure 1 shows the XRD patterns of the as-prepared samples. The characteristic spectra of monoclinic crystal BiVO4 was well indexed with the standard card (JCPDS file 14-0688), corresponding to the characteristic diffraction phases of (110), (011), (121), (040), (200), (002), (211), (150), (132), and (042) [12,13]. The diffraction pattern of Mn1-xZnxFe2O4 was fully matched with the standard card (JCPDS file 74-2400), with the characteristic reflection phases (220), (311), (222), (400), (422), (511), (440), (620), and (622) [10].
The diffraction peaks of the Mn1−xZnxFe2O4 pattern were difficult to observe in the pattern of BiVO4/Mn1−xZnxFe2O4 (15 wt %). One the one hand, the intensity of the diffraction peaks of Mn1−xZnxFe2O4 was relatively weak compared with that of the diffraction peaks of BiVO4. On the other hand, the diffraction patterns location of Mn1−xZnxFe2O4 overlapped with the domain diffraction patterns of BiVO4.
The peaks at (121) for BiVO4 in both patterns of pure BiVO4 and BiVO4/Mn1−xZnxFe2O4 (15 wt %) were clear, even in terms of their intensities. This phenomenon revealed that the introduction of Mn1−xZnxFe2O4 did not alter the growth orientation of BiVO4.
In addition, no impurity phases were found in the BiVO4/Mn1−xZnxFe2O4 (15 wt %) sample, confirming that there was no appreciable decomposition reaction for BiVO4 and Mn1−xZnxFe2O4 and no perceptible chemical reaction between the two components even though they were sintered at 500 °C.
To further elucidate the structure of BiVO4/Mn1−xZnxFe2O4 (15 wt %), we carried out the measurement of Fourier transform infrared spectroscopy. Figure 2 illustrates the FTIR spectra of the as-prepared samples. The vibration absorption peaks of Mn-O, Zn-O, and Fe-O bands of Mn1−xZnxFe2O4 were at 560.1 cm−1, 473.7 cm−1, and 412.4 cm−1, respectively [14], while the V-O vibration absorption peaks of BiVO4 were at 734.3 cm−1 and 823.4 cm−1 [15]. This spectrum of BiVO4/Mn1−xZnxFe2O4 (15 wt %) could confirm the coexistence of Mn1−xZnxFe2O4 and BiVO4 in the prepared composite, which indicated that BiVO4/Mn1−xZnxFe2O4 (15 wt %) was prepared successfully. In addition, the absorption patterns at 2341.7 cm−1 and 3433.6 cm−1 were attributed to CO2 and the surface adsorption H2O [9].
Raman spectroscopy can provide structural information for materials, and is also a sensitive method to study the crystallization, local structure, and electronic properties of materials. The Raman spectra of the synthesized samples are shown in Figure 3. It can be seen from Figure 3 that the Raman band at 120 cm−1, 210 cm−1, 324 cm−1, 366 cm−1, and 826 cm−1 is the typical vibrational band of BiVO4 [16]. These bands could be also observed in the spectroscopy of BiVO4/ Mn1−xZnxFe2O4 (15 wt %), which further revealed that the BiVO4/Mn1−xZnxFe2O4 (15 wt %) was synthesized successfully, in agreement with the of XRD and FTIR analysis results. In addition, the intensity of BiVO4 was much larger than that of Mn1−xZnxFe2O4. So, Mn1−xZnxFe2O4 signal was not obvious in the spectra of BiVO4/Mn1−xZnxFe2O4 (15 wt %).
In order to further observe the morphology of the sample, we characterized the material using SEM. Figure 4 is the SEM diagram of the prepared sample: (a) BiVO4, (b) Mn1−xZnxFe2O4, and (c) BiVO4/Mn1−xZnxFe2O4 (15 wt %). It can be seen from Figure 4a that pure BiVO4 is a three-dimensional spherical structure, and Figure 4b shows that the prepared Mn1−xZnxFe2O4 is six-square-like structure. The larger sphere seen in Figure 4c is the core-shell structure of BiVO4 coated with Mn1−xZnxFe2O4, which indicated that the introduction of Mn1−xZnxFe2O4 caused the agglomeration of the resultant composite to some extent.
Figure 5 shows the N2 adsorption-desorption isotherm and the pore size distribution curve of BiVO4/Mn1−xZnxFe2O4 sample. The absorption-desorption isotherm could be categorized as a typical Type III absorption-desorption isotherm [17], indicating that the sample has a porous structure, which was convex to the P/P0 axis over its entire range, revealing that the as-prepared composite BiVO4/Mn1−xZnxFe2O4 (15 wt %) belonged to the nonporous structure. This sharp increase in the adsorption isotherm was attributed to the presence of macropores. The most probable pore size of BiVO4/Mn1−xZnxFe2O4 (15 wt %) is 6.0 nm. In addition, the specific surface area of BiVO4/Mn1−xZnxFe2O4 sample calculated by BET measurement is 2.22 m2/g.

3.2. Magnetic Properties

Figure 6 depicts the hysteresis loops of the magnetic substrate and magnetic photocatalyst BiVO4/Mn1−xZnxFe2O4 (15 wt %). The saturation magnetization (Ms) and remanent magnetization (Ms) of BiVO4/Mn1−xZnxFe2O4 (15 wt %) were 84.03 emu/g and 7.03 emu/g, respectively. Compared with pure Mn1−xZnxFe2O4, the Ms of the magnetic photocatalyst decreased by 75.4% and 91.6%, respectively, due to the reduction to the magnetic material content per unit mass of the magnetic photocatalyst [18]. Overall, the magnetic properties of BiVO4/Mn1−xZnxFe2O4 (15 wt %) were beneficial to its separation from the liquid solution and its recycling from the reaction solution using an extra magnet after use. In addition, the magnetic photocatalyst showed magnetic properties similar to pure Mn1−xZnxFe2O4, which indicated that the magnetic photocatalyst included Mn1−xZnxFe2O4.

3.3. UV-Vis DRS Analysis

The optical absorption properties of semiconductors are considered to be a key factor affecting their photocatalytic activity. Figure 7 shows the UV-Vis diffuse reflectance spectra of BiVO4, BiVO4/Mn1−xZnxFe2O4, and the curve of (Ahv)2 vs. hv. It can be seen from the diagram that BiVO4 and BiVO4/Mn1−xZnxFe2O4 mainly absorb light in the wavelength range below 500 nm. Compared with pure BiVO4, the maximum absorption wavelength of BiVO4/Mn1−xZnxFe2O4 (lambda max) increases, and the absorption of visible light is enhanced to some extent. In addition, the band gap plays a very important role in the determination of photocatalytic activity. The relationship between absorbance and incident light intensity hv can be expressed by the following formula [8,9]:
Ahv = C (hv − Eg)n/2
In the above equation, A, Eg, h, and v represent the absorption coefficient, band gap width, Planck constant, and incident light frequency, respectively, and C is defined as a constant. The band gap energy of the sample can be obtained from the (Ahv)2~hv curve. The band gaps (Eg) of BiVO4 and BiVO4/Mn1−xZnxFe2O4 were estimated. They have nearly the same Eg values (2.36 eV), which are in agreement with the literature [5,19]. Though the incorporation of Mn1−xZnxFe2O4 could not extend the range of absorption light, it is worth noting that the anther function of Mn1−xZnxFe2O4 enhanced the absorption intensity of visible light for BiVO4 by charge transfer transitions (2Fe3+ → Fe2+ + Fe4+) [18].

3.4. Conductivity and Electrochemical Performance

Electrochemical impedance spectroscopy (EIS) is an effective method to evaluate the electron transfer between solid and electrolyte interfaces. Figure 8 shows the Nyquist diagram of the prepared samples. It can be seen from the diagram that the diameter of the semicircle decreases obviously with the introduction of Mn1−xZnxFe2O4, which indicates that the resistance of the solid interface layer and the surface electron transfer impedance decrease. The BiVO4/Mn1−xZnxFe2O4 charge transfer impedance (206 Ω·cm2) is less than pure BiVO4 (351 Ω·cm2) [20]. This is due to the doping effect of Fe ions in the Mn1−xZnxFe2O4 crystal lattice system. The charge transfer was easier, probably due to the charge transfer transitions of Fe ions (2Fe3+ → Fe2+ + Fe4+) and Mn ions (2Mn4+ → Mn2+ + Mn6+). The electron transfer impedance of the composite solid /electrolyte interface decreases, which may be because the incorporation of magnetic Mn1−xZnxFe2O4 could enhance the conductivity of BiVO4 and heighten the quantum efficiency of BiVO4. Thus, we can preliminarily predict that the photocatalytic activity of BiVO4/Mn1−xZnxFe2O4 must be higher than that of pure BiVO4 under the same light irradiation conditions.

3.5. Photocatalytic Activity and Stability

The results of the RhB photodegradation by the prepared BiVO4 and BiVO4/Mn1−xZnxFe2O4 samples are shown in Figure 9. Obviously, the photocatalytic efficiency of BiVO4/Mn1−xZnxFe2O4 was higher than that of pure BiVO4. Under the same visible light irradiation, the degradation rate of RhB for pure BiVO4 reached 97% after 3 h, while BiVO4/Mn1−xZnxFe2O4 only took 2 h to achieve the same degradation rate. This result was in good agreement with the EIS analysis. On the one hand, the introduction of Mn1−xZnxFe2O4 enhanced the conductivity of BiVO4, further heightening the electron transfer ability. On the other hand, Mn1−xZnxFe2O4 enhanced the absorption intensity of visible light for BiVO4. Thus, BiVO4/Mn1−xZnxFe2O4 could produce more photoinduced electron-hole pairs under same visible light irradiation. The two factors generate a synergistic effect to boost the quantum efficiency of BiVO4.
The repeatability of the magnetic photocatalyst was detected by cycling tests. After each cycle, the catalyst was separated by an external magnet, and then was washed and dried for the next cycle. The degradation rate of RhB using the fifth recycled photocatalyst was still more than 93% (see Figure 10). This indicated that the as-prepared magnetic photocatalyst had an excellent stability. In general, the introduction of a magnetic substrate can enhance the stability of a single component photocatalyst.

4. Conclusions

BiVO4/Mn1−xZnxFe2O4 was prepared by the impregnation roasting method. The synthesis method was suitable for the mass production of various composites. XRD tests showed that the prepared BiVO4 is monoclinic crystal, and the introduction of Mn1−xZnxFe2O4 does not change the crystal structure of BiVO4. The introduction of a soft-magnetic material, Mn1−xZnxFe2O4, was beneficial to the composite photocatalyst’s separation from the liquid solution using an extra magnet after the reaction. UV-vis spectra analysis indicated that Mn1−xZnxFe2O4 enhanced the absorption intensity of visible light for BiVO4. EIS investigation revealed that the introduction of Mn1−xZnxFe2O4 could decrease the electron transfer impedance of BiVO4, further enhancing its conductivity and heightening its quantum efficiency. The photocatalytic efficiency of BiVO4/Mn1−xZnxFe2O4 was higher than that of pure BiVO4.

Acknowledgments

This research was financially supported by the Fundamental and advanced research projects of Chongqing Science and Technology Commission (No. CSTC2015jcyjBX0015), the Scientific & Technologic Program of Chongqing Land resources and Housing Authority (No. CQGT-KJ-2014012) and the Youth research talents′ growth support program of Yangtze Normal University. In addition, we gratefully acknowledge many important contributions from the researchers of all reports cited in our paper.

Author Contributions

Taiping Xie and Chenglun Liu conceived and designed the experiments; Taiping Xie and Hui Li performed the experiments; Taiping Xie and Longjun Xu analyzed the data; Chenglun Liu and Longjun Xu contributed reagents/materials/analysis tools; Taiping Xie and Chenglun Liu wrote the paper.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Nakata, K.; Ochiai, T.; Murakami, T.; Fujishima, A. Photoenergy conversion with TiO2 photocatalysis: New materials and recent applications. Electrochim. Acta 2012, 84, 103–111. [Google Scholar] [CrossRef]
  2. Tan, H.L.; Amal, R.; Ng, Y.H. Alternative strategies in improving the photocatalytic and photoelectrochemical activities of visible light-driven BiVO4: A review. J. Mater. Chem. A 2017, 5, 16498–16521. [Google Scholar] [CrossRef]
  3. Mamba, G.; Mishra, A. Advances in magnetically separable photocatalysts: Smart, recyclable materials for water pollution mitigation. Catalysts 2016, 6, 79. [Google Scholar] [CrossRef]
  4. Saison, T.; Chemin, N.; Chanéac, C. New insights into BiVO4 properties as visible light photocatalyst. J. Phys. Chem. C 2015, 119, 12967–12977. [Google Scholar] [CrossRef]
  5. Li, X.; Yu, J.G.; Wageh, S.; Al-Ghamdi, A.A.; Xie, J. Graphene in photocatalysis: A review. Small 2016, 12, 6640–6696. [Google Scholar] [CrossRef] [PubMed]
  6. Hong, S.J.; Lee, S.; Jang, J.S.; Lee, J.S. Heterojunction BiVO4/WO3 electrodes for enhanced photoactivity of water oxidation. Energy Environ. Sci. 2011, 4, 1781–1787. [Google Scholar] [CrossRef]
  7. Sun, J.J.; Li, X.Y.; Zhao, Q.D.; Ke, J.; Zhang, D.K. Novel V2O5/BiVO4/TiO2 nanocomposites with high Visible-light-induced photocatalytic activity for the degradation of toluene. J. Phys. Chem. C 2014, 118, 10113–10121. [Google Scholar] [CrossRef]
  8. Xie, T.P.; Xu, L.J.; Liu, C.L. Magnetic composite BiOCl-SrFe12O19: A novel p-n type heterojunction with its enhanced photocatalytic activity. Dalton Trans. 2014, 43, 2211–2220. [Google Scholar] [CrossRef] [PubMed]
  9. Xie, T.P.; Liu, C.L.; Xu, L.J. Novel heterojunction Bi2O3/SrFe12O19 magnetic photocatalyst with highly enhanced photocatalytic activity. J. Phys. Chem. C 2013, 117, 24601–24610. [Google Scholar] [CrossRef]
  10. Zhang, Z.D.; Xu, L.J.; Liu, C.L. Preparation and characterization of composite magnetic photocatalyst MnxZn1−xFe2O4/β-Bi2O3. RSC Adv. 2015, 5, 79997–80004. [Google Scholar] [CrossRef]
  11. Liu, C.L.; Li, H.; Ye, H.P.; Xu, L.J. Preparation and visible-light-driven photocatalytic performance of magnetic SrFe12O19/BiVO4. J. Mater. Eng. Perform. 2015, 24, 771–777. [Google Scholar]
  12. Liu, H.B.; Hou, H.L.; Gao, F.M.; Yao, X.H.; Yang, W.Y. Tailored fabrication of thoroughly mesoporous BiVO4 nanofibers and their visible-light photocatalytic activities. Appl. Mater. Interfaces 2016, 8, 1929–1936. [Google Scholar] [CrossRef] [PubMed]
  13. Grigioni, I.; Stamplecoskie, K.G.; Selli, E.; Kamat, P.V. Dynamics of photogenerated charge carriers in WO3/BiVO4 heterojunction photoanodes. J. Phys. Chem. C 2015, 119, 20792–20800. [Google Scholar] [CrossRef]
  14. Wang, S.H.; Zhou, Q.S. Titania deposited on soft magnetic activated carbon as a magnetically separable photocatalyst with enhanced activity. Appl. Surf. Sci. 2016, 256, 6191–6198. [Google Scholar] [CrossRef]
  15. Wang, A.L.; Shen, S.; Zhao, Y.B.; Wu, W. Preparation and characterizations of BiVO4/reduced graphene oxide nanocomposites with higher visible light reduction activities. J. Colloid Interface Sci. 2015, 445, 330–336. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Z.D.; Lu, Q.F.; Wang, C.Q.; Liu, J.H.; Liu, G.S. Preparation of bamboo-shaped BiVO4 nanofibers by electrospinning method and the enhanced visible-light photocatalytic activity. J. Alloy. Compd. 2015, 651, 29–33. [Google Scholar] [CrossRef]
  17. Zhou, F.Q.; Fan, J.C.; Xu, Q.J.; Min, Y.L. BiVO4 nanowires decorated with CdS nanoparticles as Z-scheme photocatalyst with enhanced H2 generation. Appl. Catal. B 2017, 210, 77–83. [Google Scholar] [CrossRef]
  18. Xie, T.P.; Xu, L.J.; Liu, C.L. Magnetic composite ZnFe2O4/SrFe12O19: Preparation, characterization, and photocatalytic activity under visible light. Appl. Surf. Sci. 2013, 273, 684–691. [Google Scholar] [CrossRef]
  19. Singh, S.; Sharma, R.; Mehta, B.R. Enhanced surface area, high Zn interstitial defects and band gap reduction in N-doped ZnO nanosheets coupled with BiVO4 leads to improved photocatalytic performance. Appl. Surf. Sci. 2017, 411, 321–330. [Google Scholar] [CrossRef]
  20. Chang, X.X.; Wang, T.; Zhang, P. Enhanced surface reaction kinetics and charge separation of p−n heterojunction Co3O4/BiVO4 photoanodes. J. Am. Chem. Soc. 2015, 137, 8356–8359. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD (X-ray Diffractometer) patterns of the prepared samples.
Figure 1. XRD (X-ray Diffractometer) patterns of the prepared samples.
Materials 11 00335 g001
Figure 2. FTIR (Fourier transform infrared) spectrum of the prepared samples.
Figure 2. FTIR (Fourier transform infrared) spectrum of the prepared samples.
Materials 11 00335 g002
Figure 3. Raman spectra of the prepared samples.
Figure 3. Raman spectra of the prepared samples.
Materials 11 00335 g003
Figure 4. SEM (scanning electron microscopy) photographs of materials: (a) BiVO4, (b) Mn1−xZnxFe2O4, (c) BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Figure 4. SEM (scanning electron microscopy) photographs of materials: (a) BiVO4, (b) Mn1−xZnxFe2O4, (c) BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Materials 11 00335 g004
Figure 5. N2 adsorption-desorption isotherms and pore size distribution curves (insert) of BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Figure 5. N2 adsorption-desorption isotherms and pore size distribution curves (insert) of BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Materials 11 00335 g005
Figure 6. Hysteresis loops of (a) Mn1−xZnxFe2O4 and (b) BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Figure 6. Hysteresis loops of (a) Mn1−xZnxFe2O4 and (b) BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Materials 11 00335 g006
Figure 7. UV-Vis diffuse reflectance spectra of pure BiVO4 and BiVO4/Mn1−xZnxFe2O4 (15 wt %) samples. Inset: the plot of (Ahv)2 vs. hv to estimate the Eg value.
Figure 7. UV-Vis diffuse reflectance spectra of pure BiVO4 and BiVO4/Mn1−xZnxFe2O4 (15 wt %) samples. Inset: the plot of (Ahv)2 vs. hv to estimate the Eg value.
Materials 11 00335 g007
Figure 8. The Nyquist plots of BiVO4 and BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Figure 8. The Nyquist plots of BiVO4 and BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Materials 11 00335 g008
Figure 9. Photocatalytic degradation ratios of RhB with BiVO4 and BiVO4/Mn1−xZnxFe2O4 (15 wt %) samples.
Figure 9. Photocatalytic degradation ratios of RhB with BiVO4 and BiVO4/Mn1−xZnxFe2O4 (15 wt %) samples.
Materials 11 00335 g009
Figure 10. Recycling experiments of degrading RhB with BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Figure 10. Recycling experiments of degrading RhB with BiVO4/Mn1−xZnxFe2O4 (15 wt %).
Materials 11 00335 g010

Share and Cite

MDPI and ACS Style

Xie, T.; Liu, C.; Xu, L.; Li, H. New Insights into Mn1−xZnxFe2O4 via Fabricating Magnetic Photocatalyst Material BiVO4/Mn1−xZnxFe2O4. Materials 2018, 11, 335. https://doi.org/10.3390/ma11030335

AMA Style

Xie T, Liu C, Xu L, Li H. New Insights into Mn1−xZnxFe2O4 via Fabricating Magnetic Photocatalyst Material BiVO4/Mn1−xZnxFe2O4. Materials. 2018; 11(3):335. https://doi.org/10.3390/ma11030335

Chicago/Turabian Style

Xie, Taiping, Chenglun Liu, Longjun Xu, and Hui Li. 2018. "New Insights into Mn1−xZnxFe2O4 via Fabricating Magnetic Photocatalyst Material BiVO4/Mn1−xZnxFe2O4" Materials 11, no. 3: 335. https://doi.org/10.3390/ma11030335

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