Next Article in Journal
Improved UV Photoresponse Performance of ZnO Nanowire Array Photodetector via Effective Pt Nanoparticle Coupling
Previous Article in Journal
Potential of Enzymatically Synthesized Hemozoin Analog as Th1 Cell Adjuvant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Formulating InVO4/α-Fe2O3 Heterojunction Composites for Photocatalytic Tetracycline Hydrochloride Degradation

1
Manchester Metropolitan Joint Institute, Hubei University, Wuhan 430062, China
2
School of Materials Science and Engineering, Hubei University, Wuhan 430062, China
3
Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai 202104, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(17), 1441; https://doi.org/10.3390/nano14171441
Submission received: 31 July 2024 / Revised: 29 August 2024 / Accepted: 2 September 2024 / Published: 4 September 2024

Abstract

:
This study reports the synthesis of InVO4/α-Fe2O3 heterojunction photocatalysts with different stoichiometric ratios via a two-step hydrothermal synthesis reaction. The prepared photocatalysts were characterized by XRD, SEM, TEM, XPS, and other methods. The prepared composites exhibited good photocatalysis of tetracycline hydrochloride. Among the InVO4/α-Fe2O3 heterojunction photocatalysts with different ratios, the InVO4/0.25α-Fe2O3 photocatalyst showed the highest degradation rate for 20 mg L−1 tetracycline hydrochloride. After three photocatalytic runs, it still exhibited excellent stability and reusability. Meanwhile, this study also found that superoxide radical anion (-O2−), electron (e), hydroxyl radical (·OH), and photogenerated hole (h+) are the basic active substances in the photocatalytic process.

1. Introduction

Over the past few years, the detection of pharmaceuticals, notably antibiotics, in water sources has risen because of their extensive use in human and agricultural sectors [1]. Despite their importance for health and safety, their overuse threatens ecosystems and human health [2,3,4,5,6]. Tetracycline hydrochloride (TC) is among the most frequently used antibiotics worldwide. It is widely used in human and animal infection treatment and other medical aspects, offering the benefits of broad-spectrum antibacterial activity at a moderate cost [7,8]. However, its lengthy half-life and poor metabolic efficiency pose a significant and urgent environmental threat. Therefore, from the perspective of a healthy human life and socially sustainable development, it is crucial to develop technology that will help achieve environmentally friendly, green, and clean production to degrade TC [9,10].
Compared with traditional technologies of removing contaminants, such as flocculation, membrane separation, coagulation, and biodegradation, photocatalysis technology has attracted attention because of its high efficiency, recyclability, and economic advantages [11,12,13]. Photocatalysis is a type of chemical reaction that uses a unique material called a semiconductor photocatalyst and visible light to break down harmful substances. Using this technology to create potent oxidation free radicals, oxidize, and eliminate contaminants is the basic idea behind photocatalytic oxidation technology. The use of photocatalysis techniques, which combine solar technology and heterogeneous catalysis, has made it possible to remove antibiotics from wastewater economically and sustainably [14]. The use of semiconductor-based photocatalysis technology as an environmentally friendly means of eliminating contaminants from the environment has grown during the last several decades [15,16,17].
Even though semiconductor photocatalysis technology has advanced quickly, the low separation efficiency of photogenerated electron–hole pairs, restricted mass transfer, weak visible light absorption, and catalyst inactivation caused by pollutants and their byproducts make it difficult for current photocatalysis technology to meet industrial-scale needs [16]. Therefore, achieving the successful utilization of solar energy and large-scale deployment of TC removal from wastewater using ecological means requires choosing a suitable catalyst that is essential for photocatalytic degradation [17,18]. Currently, scientists are concentrating on using active materials in visible light. Narrow-band gap oxides are better suited for photocatalysis because they have a broader absorption spectrum in visible light, such as BiVO4 [19,20], TiO2 [21], CdS [22], and g-C3N4 [23], which have been widely studied for wastewater treatment under visible light. Indium vanadate (InVO4), a narrow bandgap substance, is a successful photocatalyst due to its unique properties. These include beneficial optical and electronic characteristics, biocompatibility, and broad visible light absorption. Notably, it is non-toxic and demonstrates outstanding resistance to chemicals and photo corrosion [18,19,20,21,22,23].
However, single-component photocatalysts typically exhibit lower light absorption and higher electron–hole pair complexation rates [24]. This is also true for the one-component InVO4 catalyst material, which does not have a high enough photocatalytic efficiency in visible light because of the rapid rate of electron–hole pair complexation generated in the narrow band gap [25]. In fact, many scientists have recently adopted heterogeneous composite materials to treat wastewater. Ivanov et al. prepared Fe3O4 and Zn-Al-LDH composite adsorbents with high exchange capacity and selectivity by the precipitation method, and the maximum adsorption and exchange capacity for U(VI) reached 268.65 ± 22.22 mg g−1, which is very suitable as adsorbents for U(VI) removal [26]. Balybina et al. investigated a series of adsorbent materials based on layered double hydroxide (LDH). They found that Co-Fe LDH showed the highest adsorption capacity in seawater, which is important for the nuclear energy industry and environmental protection [27]. Furthermore, it has been demonstrated that doping and heterojunction construction can enhance the photoelectrochemical performance of a photocatalyst [18,20,28,29]. Zhang et al. investigated Bi-Au/SiO2 tandem bimetallic catalysts, which were monitored using near-environmental pressure X-ray photoelectron spectroscopy and were of great significance for an in-depth understanding of the catalytic mechanism and optimization of catalyst design [30]. Performance under visible light illumination degradation can be enhanced by doping certain metals (i.e., Fe, Ag, and Ce). In order to enhance photocatalytic activity, several researchers and scholars have compounded InVO4 with other materials. For example, InVO4/CeVO4 empty nanoribbons with non-homogeneous interfaces and internal cavities and a TOC removal rate of 90.4% TC after 135 min of radiation were synthesized by Ding et al., which provided a new strategy for the design of high-efficiency photocatalysts [28]. Tamtam et al. successfully synthesized ZnWO4 nanorod-modified InVO4 nanosheets, which were used for the highly efficient decomposition of tetracycline (TC) with an enhanced performance of 7.6 times, demonstrating their potential in the photocatalytic field [25].
Previous studies have shown that a heterojunction coupled with α-Fe2O3 has a low photogenerated carrier complexation rate and a strong oxidation capability and can directly burst reducing electrons and oxidizing holes simultaneously by charge transfer at low potentials [31,32,33]. As an n-type narrow semiconductor, α-Fe2O3 has an excellent optical bandgap (1.9–2.2 eV) and can absorb light in the visible region [34,35,36]. Wang et al. prepared and obtained α-Fe2O3 using a straightforward hydrothermal method. Additionally, α-Fe2O3 is an excellent option for photocatalysts in wastewater treatment because of its low cost, good chemical stability, high abundance, high visible light utilization efficiency, non-toxicity, and environmental friendliness [37,38,39,40]. The rGO/InVO4/Fe2O3 Z-scheme heterostructured photocatalysts developed by Kumar et al. showed excellent results in CO2 photoreduction to methanol, demonstrating the potential of InVO4/Fe2O3 heterostructures in photocatalysts [41]. Consequently, the combination of InVO4 and α-Fe2O3 can overcome the limitations of the two materials and further enhance the utilization of sunlight by improving the photoelectrochemical properties of the photocatalyst. The aim of this paper is to gain insight into the structural and optical properties of heterojunction materials to improve the photocatalytic activity of InVO4 materials and to show the effect of InVO4/α-Fe2O3 materials in degrading TC.
In this study, InVO4 composite materials doped with varying ratios of α-Fe2O3 were prepared via the hydrothermal method, as shown in Figure 1 and the Experimental Methods Section. The impact of α-Fe2O3 doping on the properties of InVO4, including structure, morphology, and photoelectrochemical properties, was investigated. The visible photocatalytic activity of the prepared α-Fe2O3-doped InVO4 composites was evaluated by intensively investigating the photocatalytic degradation of a tetracycline hydrochloride solution, for example, by further studying the degradation properties of the materials by varying the conditions of catalyst addition and initial TC concentration, respectively.

2. Experimental Methods

2.1. Material Synthesis

2.1.1. Synthesis of α-Fe2O3

A hydrothermal method was used to synthesize α-Fe2O3. First, 0.540 g (0.2 mmol) of FeCl3·6H2O was dissolved in 22 mL of a solvent mixture of deionized water and ethanol in a volume ratio of 1:10. Then, 1.6 g of CH3COONa was added and stirred for 20 min. Then, the mixture was transferred to a 50 mL Teflon-lined stainless autoclave and heated at 180 °C for 8 h in a blast drying oven. After cooling, the solids were washed thrice with ultrapure water and anhydrous ethanol by high-speed centrifugation. The precipitate was dried in a vacuum drying oven at 60 °C for 18 h to obtain α-Fe2O3.

2.1.2. Synthesis of InVO4/α-Fe2O3 Composites and InVO4

InVO4/α-Fe2O3 composites with different ratios were synthesized by a hydrothermal method. Firstly, 0.293 g (0.1 mmol) of InCl3·4H2O and 0.117 g (0.2 mmol) of NH4VO3 were dissolved in two beakers with deionized water and heated to 60 °C. Then, different amounts of α-Fe2O3 particles were added proportionally to the NH4VO3 solution, stirred, and ultrasonicated for 30 min. Next, the mixed solution was added to the InCl3·4H2O solution with 0.2 g of PVP and was stirred continuously for 40 min, and then, its pH was adjusted to about 4 with 25 wt% NH3-H2O. Then, the mixture was transferred to a 100 mL Teflon-lined stainless autoclave and heated at 160 °C for 12 h in a blast drying oven. After cooling, the solids were washed thrice with ultrapure water and anhydrous ethanol by high-speed centrifugation. The precipitate was dried in a vacuum drying oven at 60 °C for 18 h to obtain α-Fe2O3/InVO4 composites. Using this process, InVO4/α-Fe2O3 composites were produced as InVO4/0.25α-Fe2O3 InVO4/0.5α-Fe2O3, InVO4/0.75α-Fe2O3, and so on, with varying molar ratios of α-Fe2O3 to InVO4. By the same synthesis procedure, pure InVO4 photocatalysts were synthesized for control studies without the addition of α-Fe2O3 particles. The synthetic route is shown in Figure 1.

2.2. Characterization

The size and morphology of the catalysts were studied by scanning electron microscopy (SEM, ZEISS Gemini 300, Jena, Germany) and EDS elemental analysis. The sample’s crystal structure was analyzed by an X-ray diffractometer (XRD) at beamline BL14B1, Shanghai Synchrotron Radiation Facility (SSRF), Shanghai, China. The wavelength used in the XRD analysis was 0.6888 Å (18 keV X-ray energy). The beam size was about 400 μm × 400 μm. The detector used to collect data was Mythen 1K (Dectris Inc., Baden, Switzerland). Al-Kα monochromatic X-ray photoelectron spectroscopy (XPS) was used to calibrate C 1s peak energy (equal to 284.8 eV). Ultraviolet–visible diffuse reflectance spectroscopy (UV-vis) was conducted with an ultraviolet–visible spectrophotometer (Shanghai Jinghua, Shanghai, China, UV-1800).

2.3. Photocatalytic Experiments

The visible light irradiation-degraded TC served as a proxy for the materials’ photocatalytic activity. The light source was a 300 W xenon lamp (HSX-F300, NBeT, Patna, India) with a 400 nm cut-off filter. For every photocatalysis experiment, 50 mg of photocatalyst particles and 100 mL of TC solution (20 mg L−1) were introduced to a quartz reactor. Next, the reactor was stirred in the dark for 40 min to equilibrate the adsorption and desorption reactions to construct the catalyst. Subsequently, the xenon lamp light source was switched on, and magnetic stirring was timed for ninety minutes. During the xenon lamp irradiation, 3 mL of water samples were taken at 15 min intervals, and the solid catalyst particles were removed via centrifugation. Using a UV–visible spectrophotometer (UV-1800, Shanghai Jinghua), the TC in the filtrate was examined. The photocatalytic degradation efficiency calculation formula is as follows:
% Photocatalytic degradation = [(C0 − Ct)/C0] × 100
The initial concentration of TC, designated as C0, was contrasted with the instantaneous concentration, designated as C, following the sampling analysis.

3. Results and Discussion

3.1. The Wavelength Used in XRD Analysis at the Synchrotron Beamline

Figure 2a displays the XRD spectra of InVO4, α-Fe2O3, InVO4/0.25α-Fe2O3, and *InVO4/0.25α-Fe2O3 (InVO4/α-Fe2O3 after degradation once). The InVO4/α-Fe2O3 composites exhibit diffraction peaks at 8.2°, 9.2°, 10.2°, 11.1°, 12.0°, and 15.54°, which are comparable to the pristine InVO4 and α-Fe2O3 and correlate to an orthorhombic crystalline structure with an orthorhombic InVO4 plane (110), (020), (111), (021), (002), and (130) (JCPDS 48-0898). The displacement of (002) planar InVO4 at roughly 12° is displayed in Figure 2b. The hematite (α-Fe2O3) planes (012), (104), (110), (113), (024), (116), (214), and (300) correspond to values of approximately 10.7°, 14.7°, 15.7°, 17.9°, 21.5°, 23.5°, 26.8°, and 27.4° at 2θ (JCPDS 33-0664). The above results indicate the successful synthesis of the InVO4/α-Fe2O3 complex photocatalysts.

3.2. Surface Area and Pore Structure

The BET-specific surface area and the corresponding pore size distribution of the synthesized samples were analyzed by N2 adsorption–desorption isotherms, as shown in Figure 3. According to the IUPAC classification, the N2 adsorption–desorption isotherms of all the synthesized samples can be categorized as type IV adsorption isotherms, indicating that the samples are mesoporous [41]. The pore size distribution curves calculated by the BJH method further reveal that the samples have mesoporous structures, and the average pore sizes of InVO4/0.25α-Fe2O3, InVO4, and α-Fe2O3 are 27.1070 nm, 21.7435 nm, and 22.7819 nm, respectively, with a more centralized distribution of particle sizes, as shown in Figure 3b. A comparison of the specific surface area of these samples shows that the specific surface area of InVO4/0.25α-Fe2O3 139.1711 m2 g−1 is greater than that of InVO4 (103.1375 m2 g−1) in the presence of α-Fe2O3 doping. According to the data, it can be seen that the surface area of the catalyst was prompted to be enlarged after the complexation of InVO4 with α-Fe2O3. In the photocatalytic process, the good specific surface area and internal pore structure facilitated the multiple reflections and refractions of light, which improved the utilization efficiency of light, thus effectively increasing the photocatalytic activity [42,43].

3.3. Morphologic Structure Analysis

The SEM images of α-Fe2O3 in Figure 4a reveal that α-Fe2O3 has an overall regular shape, micro agglomerates, and a hexagonal flake-like shape. In contrast, InVO4 has a one-dimensional nanorod-like morphology or irregular microsphere-like agglomerates with nanometer sizes in Figure 4b. Figure 4c demonstrates that the InVO4 microspheres are attached to the hexagonal structure of α-Fe2O3, and InVO4 is well combined with α-Fe2O3. The transmission electron micrograph (TEM) of the InVO4/0.25α-Fe2O3 material is displayed in Figure 5. InVO4 point-like crystals are present on the α-Fe2O3 nanosheets, as shown in Figure 5a. There are two distinct crystal plane spacings in α-Fe2O3 (Figure 5b), which can be identified from the high-magnification transmission electron microscope image. These can be compared to the standard card and shown to correspond to the (012) and (002) crystal planes of α-Fe2O3, respectively, and the (002) crystal plane of InVO4 with the 0.321 nm lattice spacing.

3.4. Composition and Valence State Analysis

The EDS elemental mapping in Figure 6 further verified the structure of the InVO4/0.25α-Fe2O3 heterojunction. The elemental maps of In, V, O, and Fe were identified from photographs of certain hues (Figure 6b), and these elemental maps demonstrated the presence of the InVO4/α-Fe2O3 heterojunction. The results showed that In accounted for 67.49% and Fe accounted for 32.96%; α-Fe2O3 was distributed in InVO4, forming an InVO4/0.25α-Fe2O3 catalyst.
X-ray photoelectron spectroscopy (XPS) was applied to examine the chemical composition and coordination environment of the InVO4/0.25α-Fe2O3 catalyst. As can be seen, the measured spectrum of InVO4/0.25α-Fe2O3 shows signals for the In, V, O, and Fe elements (Figure 7). The integration of 0.25α-Fe2O3 into InVO4 was confirmed by two peaks centered at 711 eV and 724.3 eV and two satellite peaks at 710.9 eV and 724.6 eV in the high-resolution spectra of Fe 2p (Figure 7e). Among them, Fe2+ and Fe3+ accounted for 56.24% and 43.74%, respectively. To uncover the impact of 0.25α-Fe2O3 doping on InVO4, additional analysis was conducted on the high-resolution spectra of V 2p, O 1s, and In 3d. The peaks at 444.07 eV and 451.62 eV for In 3d (Figure 7d) are ascribed to In 3d5/2 and In 3d2/3, respectively. The proportions were 38.68% and 61.32%, respectively. Lattice oxygen (OL) and chemisorbed oxygen (OV) (Figure 7c) are responsible for the two O 1s deconvolution peaks at 529.61 eV and 531.33 eV, respectively. The OL peak of the InVO4/0.25α-Fe2O3 catalyst exhibits a comparable positive shift behavior to Fe-InVO4 synthesized by predecessors. The increase in oxygen vacancy is caused by the injection of α-Fe2O3 [20]. The typical V5+ peaks for V 2p (Figure 7b) are located at 516.57 eV and 523.74 eV, suggesting that V5+ ions have the potential to pick up electrons from adjacent oxygen vacancies.

3.5. Optical Property Analysis

To obtain a more profound comprehension of the impact of the coupling between InVO4 and α-Fe2O3 on the photocatalytic activity, UV-Vis DRS was utilized to analyze the optical characteristics of the photocatalysts. Figure 8a displays the UV-vis absorption spectra of InVO4, α-Fe2O3, and InVO4/0.25α-Fe2O3. The material with InVO4/0.25α-Fe2O3 has a strong absorption capacity in both the visible and infrared spectrums, with a UV spectral region of 300~400 nm and a visible spectral region of 400~600 nm.
Furthermore, as seen in Figure 8b–d, the band gaps of InVO4, α-Fe2O3, and InVO4/0.25α-Fe2O3 were computed using the Kubelka–Munk technique. These band gaps were determined to be 1.95 eV, 2.02 eV, and 1.6 eV, respectively [29]. It is simple to determine that the bandgap of the successful InVO4 and α-Fe2O3 composite narrows, improving the composites’ photoresponse to visible light irradiation.

3.6. Photocatalytic Degradation Performance

As illustrated in Figure 9, the photocatalytic activity of the material was evaluated by the degradation of TC under visible light at room temperature under dark adsorption for 40 min and visible light irradiation for 90 min. Control experiments were conducted at a catalyst dosage of 0 mg, and negligible fluctuations in TC concentration were observed. The dark adsorption efficiency of all samples was less than 5%. Furthermore, the photodegradation efficiency of the InVO4/0.25α-Fe2O3 composite was significantly enhanced compared with that of pure InVO4 or pure α-Fe2O3. The InVO4/0.25α-Fe2O3 catalyst exhibited the highest photocatalytic activity among the samples. The photocatalytic activity of the InVO4/0.5α-Fe2O3 and InVO4/0.75α-Fe2O3 composites decreased sequentially as the Fe content increased. Opportune lattice defects can be produced by doping α-Fe2O3 into InVO4, providing enough holes for the adsorption of TC and trapping photogenerated electrons. In addition, the heterojunction structure that forms at the junction interface facilitates the efficient migration and separation of the photogenerated electron–hole pair.
The degradation rate of TC showed a negative connection with the initial TC content (20 mg L−1 to 40 mg L−1) (Figure 9b). TC with an initial concentration of 40 mg L−1 degraded more quickly, presumably because of the catalyst’s increased propensity to bind TC in solution. On the other hand, the TC concentration dropped over time, and its ultimate degradation rate was the lowest. This is partially because the catalytic effectiveness decreased when the by-products of TC breakdown at 40 mg L−1 were more adhered to the catalyst surface than those at 20 mg L−1. Since tetracycline hydrochloride is an amphoteric antibiotic, another primary parameter influencing photocatalytic efficiency is the pH of the tetracycline solution. The photocatalytic breakdown of the catalyst at varying pH values is depicted in (Figure 9c). The figure shows that the catalyst’s rate of tetracycline hydrochloride degradation is greatest at a pH of 6.3, and in the solution at a pH of 4.5, the rate shows little variation. At pH 7.5, the rate of deterioration is reduced. Because TC is transformed into an anion or cation and electrostatic repulsion is created in front of the catalyst, the degrading efficiency of the catalyst may be diminished at high or low pH values [44].

3.7. Stability and Recyclability Properties

Three cycles of TC degradation experiments were conducted (Figure 10a), which aimed to examine the stability and recoverability of the InVO4/0.25α-Fe2O3 photocatalyst. The outcomes show that the product’s absorption into the catalyst’s active sites may cause the modest drop in photocatalytic activity seen in the three cycles (Figure 10a). Nonetheless, the photocatalyst’s ability to catalyze remains potent. The XRD diffractograms of InVO4/0.25α-Fe2O3 before and after recycling are shown in Figure 10b. The findings of the XRD examination show that during the photocatalytic process, the phase composition of the InVO4/0.25α-Fe2O3 photocatalyst stays virtually unchanged, which suggests its outstanding stability after three cycles.

3.8. Photocatalytic Pollution Removal Mechanism

Reactive oxygen capture studies were conducted using various trapping agents to investigate the impact of distinct reactive groups on the photocatalytic mechanism (Figure 11a). Every trap prevented TC from degrading. Benzoquinone was found to have a notable impact on TC degradation, suggesting that superoxide radical oxidation is the primary mechanism involved in TC degradation, with the oxidation of hole and hydroxyl radicals also playing a role. As previously noted, a suggested mechanism for the degradation of TC was formulated based on the findings of the studies. It is suggested that TC is adsorbed both inside the internal spaces and on the catalyst’s surface at first. The catalyst’s variation in band gap makes it easier for photogenerated hole–electron pairs to form when exposed to visible light. Part of the O2 in solution is reduced to O2 by the electrons in the conduction band, and part of the OH- in solution is oxidized to ·OH by the holes in the valence band. TC degrades gradually because of its reactions with reactive groups such as h*, O2, OH-, and so on [25,45,46]. The photo-induced electrons of the InVO4/0.25α-Fe2O3 catalyst are shown to flow between VB and CB in Figure 11b. During this electron transfer, a superoxide radical anion (·O2) is formed because of the reduction of oxygen molecules. By contrast, when water molecules or hydroxyl ions are oxidized in the presence of Valance band gaps, hydroxyl radicals (·OH) are created. Tetracycline is broken down into oxidation products by these superoxide anions (·O2) and hydroxyl radicals (·OH).

4. Conclusions

Using a two-step hydrothermal synthesis process, InVO4/α-Fe2O3 heterojunctions were successfully prepared with varying ratios of α-Fe2O3 to InVO4, and their ability to enhance the photocatalytic degradation of TC under visible light irradiation was examined. The InVO4/α-Fe2O3 composites have a higher degradation rate when compared with that of the single material, α-Fe2O3 or InVO4, alone. InVO4/0.25α-Fe2O3 is the optimized photocatalyst with the highest photocatalytic activity. In conclusion, there is exciting future research potential for the green photocatalytic material, the InVO4/α-Fe2O3 heterojunction photocatalysts.

Author Contributions

Conceptualization, Z.W. and H.C.; methodology, H.C.; investigation, Z.W.; resources, Z.W., B.S. and F.S.; data curation, Y.W. and P.Q.; writing—original draft preparation, H.C.; writing—review and editing, Z.W. and F.S.; project administration, Z.W.; funding acquisition, B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 51927801.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to thank the staff at beamlines BL14B1 and BL20U1 and the experimental auxiliary system of Shanghai Synchrotron Radiation Facility (SSRF) for on-site assistance.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bacanlı, M.G. The two faces of antibiotics: An overview of the effects of antibiotic residues in foodstuffs. Arch. Toxicol. 2024, 98, 1717–1725. [Google Scholar] [CrossRef]
  2. Feng, D.; He, J.; Zheng, L.; Jiang, W.; Zhang, C.; Li, L.; Song, J.; Yao, W. Enhanced catalytic performance with Fe@α-Fe2O3 thin nanosheets by synergistic effect of photocatalysis and Fenton-like process. J. Phys. Chem. Solids 2021, 150, 109886. [Google Scholar] [CrossRef]
  3. Choi, Y.; He, H.; Dodd, M.C.; Lee, Y. Degradation kinetics of antibiotic resistance gene mecA of methicillin-resistant staphylococcus aureus (MRSA) during water disinfection with chlorine, ozone, and ultraviolet light. Environ. Sci. Technol. 2021, 55, 2541–2552. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, W.; Li, J.; Yao, Z.; Li, M. A review on the alternatives to antibiotics and the treatment of antibiotic pollution: Current development and future prospects. Sci. Total Environ. 2024, 926, 171757. [Google Scholar] [CrossRef] [PubMed]
  5. Xiao, J.; Lv, J.; Lu, Q. Building Fe2O3/MoO3 nanorod heterojunction enables better tetracycline photocatalysis. Mater. Lett. 2022, 311, 131580. [Google Scholar] [CrossRef]
  6. Moja, L.; Zanichelli, V.; Mertz, D.; Gandra, S.; Cappello, B.; Cooke, G.S.; Chuki, P.; Harbarth, S.; Pulcini, C.; Mendelson, M.; et al. WHO’s essential medicines and AWaRe: Recommendations on first- and second-choice antibiotics for empiric treatment of clinical infections. Clin. Microbiol. Infect. 2024, 30, S1–S51. [Google Scholar] [CrossRef]
  7. Zhu, L.; Song, W.; Liu, C.; Gu, W.; Zhao, M.; Zhao, Y. Efficient degradation of tetracycline hydrochloride by activated peroxymonosulfate with PdO/CuFe2O4/coal-bearing strata kaolinite composite. J. Phys. Chem. Solids 2024, 187, 111884. [Google Scholar] [CrossRef]
  8. Wang, B.; Guo, Y.; Li, Q.; Xin, C.; Tian, Y.; Zhang, W.; Yu, X. Design of porous ZrO2 with well-tuned band structures and strong visible-light harvesting via Zn doping for enhanced visible-light photocatalysis. Chem. Eng. J. 2024, 481, 148489. [Google Scholar] [CrossRef]
  9. He, N.; Yu, Z.; Yang, G.; Tan, Q.; Wang, J.; Chen, Y. Designing with A-site cation defects in LaFeO3: Removal of tetracycline hydrochloride in complex environments using photo-Fenton synergy. Chem. Eng. J. 2024, 484, 149613. [Google Scholar] [CrossRef]
  10. Khan, M.E.; Mohammad, A.; Ali, W.; Khan, A.U.; Hazmi, W.; Zakri, W.; Yoon, T. Excellent visible-light photocatalytic activity towards the degradation of tetracycline antibiotic and electrochemical sensing of hydrazine by SnO2–CdS nanostructures. J. Clean. Prod. 2022, 349, 131249. [Google Scholar] [CrossRef]
  11. Rathinam Thiruppathi Venkadajapathy, V.; Sivaperumal, S. Tailoring functional two-dimensional nanohybrids: A comprehensive approach for enhancing photocatalytic remediation. Ecotoxicol. Environ. Saf. 2024, 275, 116221. [Google Scholar] [CrossRef] [PubMed]
  12. Zhou, Y.; Zhu, Z.; Ma, X.; Wu, J.; Lv, X.; Qi, F.; He, P.; Qi, Y.; Lv, C.; Pan, W. Research progress in β-Fe2O3-based catalysts: An overview of the synthesis, conversion and applications. J. Environ. Chem. Eng. 2023, 11, 110864. [Google Scholar] [CrossRef]
  13. Wang, W.; Li, G.; Xia, D.; An, T.; Zhao, H.; Wong, P.K. Photocatalytic nanomaterials for solar-driven bacterial inactivation: Recent progress and challenges. Environ. Sci. Nano 2017, 4, 782–799. [Google Scholar] [CrossRef]
  14. Wang, X.; Lin, S.; Cui, N.; Qi, K.; Liu, S.-Y.; Khan, I. Synthesis of ZnWO4/NiWO4 photocatalysts and their application in tetracycline hydrochloride degradation and antibacterial activities. J. Taiwan Inst. Chem. Eng. 2024, 157, 105408. [Google Scholar] [CrossRef]
  15. Gu, C.; Li, C.; Minezawa, N.; Okazaki, S.; Yamaguchi, K.; Suzuki, K. Multi-stimuli-responsive polymer degradation by polyoxometalate photocatalysis and chloride ions. Nanoscale 2024, 16, 8013–8019. [Google Scholar] [CrossRef] [PubMed]
  16. Li, X.; Li, Z.; Xing, Z.; Song, Z.; Ye, B.; Wang, Z.; Wu, Q. UV-LED/P25-based photocatalysis for effective degradation of isothiazolone biocide. Front. Environ. Sci. Eng. 2020, 15, 85. [Google Scholar] [CrossRef]
  17. Carbuloni, C.F.; Savoia, J.E.; Santos, J.S.P.; Pereira, C.A.A.; Marques, R.G.; Ribeiro, V.A.S.; Ferrari, A.M. Degradation of metformin in water by TiO2–ZrO2 photocatalysis. J. Environ. Manag. 2020, 262, 110347. [Google Scholar] [CrossRef]
  18. Dhull, P.; Sudhaik, A.; Sharma, V.; Raizada, P.; Hasija, V.; Gupta, N.; Ahamad, T.; Nguyen, V.-H.; Kim, A.; Shokouhimehr, M.; et al. An overview on InVO4-based photocatalysts: Electronic properties, synthesis, enhancement strategies, and photocatalytic applications. Mol. Catal. 2023, 539, 113013. [Google Scholar] [CrossRef]
  19. Yang, R.; Zhang, Y.; Fan, Y.; Wang, R.; Zhu, R.; Tang, Y.; Yin, Z.; Zeng, Z. InVO4-based photocatalysts for energy and environmental applications. Chem. Eng. J. 2021, 428, 131145. [Google Scholar] [CrossRef]
  20. Li, J.; Zheng, M.; Wei, F.; Dong, C.; Xiu, Z.; Mu, W.; Zhou, X.; Ding, Y.; Han, X. Fe doped InVO4 nanosheets with rich surface oxygen vacancies for enhanced electrochemical nitrogen fixation. Chem. Eng. J. 2021, 431, 133383. [Google Scholar] [CrossRef]
  21. Chaison, J.; Wetchakun, K.; Wetchakun, N. Investigation of the physical, optical, and photocatalytic properties of CeO2/Fe-doped InVO4 composite. J. Phys. Chem. Solids 2017, 111, 95–103. [Google Scholar] [CrossRef]
  22. Jin, L.; Liu, H.; Ye, L.; Huang, Y.; Liu, X.; Huang, D. Revealing the primary role of the V4+/V5+ cycle in InVO4 catalysts for promoting the photo-Fenton reaction. Environ. Sci. Nano 2023, 11, 942–950. [Google Scholar] [CrossRef]
  23. Zuo, G.; Ma, S.; Yin, Z.; Chen, W.; Wang, Y.; Ji, Q.; Xian, Q.; Yang, S.; He, H. Z-Scheme modulated charge transfer on InVO4@ZnIn2S4 for durable overall water splitting. Small 2023, 19, 2207031. [Google Scholar] [CrossRef]
  24. An, X.; Chen, Y.; Ao, M.; Jin, Y.; Zhan, L.; Yu, B.; Wu, Z.; Jiang, P. Sequential photocatalytic degradation of organophosphorus pesticides and recovery of orthophosphate by biochar/α-Fe2O3/MgO composite: A new enhanced strategy for reducing the impacts of organophosphorus from wastewater. Chem. Eng. J. 2022, 435, 135087. [Google Scholar] [CrossRef]
  25. Ivanov, N.P.; Dran’kov, A.N.; Shichalin, O.O.; Lembikov, A.O.; Buravlev, I.Y.; Mayorov, V.Y.; Balanov, M.I.; Rogachev, K.A.; Kaspruk, G.D.; Pisarev, S.M.; et al. Composite magnetic sorbents based on magnetic Fe3O4 coated by Zn and Al layered double hydroxide for U(VI) removal from aqueous media. J. Radioanal. Nucl. Chem. 2024, 333, 1213–1230. [Google Scholar] [CrossRef]
  26. Balybina, V.A.; Dran’kov, A.N.; Shichalin, O.O.; Savel’eva, N.Y.; Kokorina, N.G.; Kuular, Z.C.; Ivanov, N.P.; Krasitskaya, S.G.; Ivanets, A.I.; Papynov, E.K. Mesoporous layered double hydroxides: Synthesis for high effective uranium ions sorption from seawater and salt solutions on nanocomposite functional materials. J. Compos. Sci. 2023, 7, 458. [Google Scholar] [CrossRef]
  27. Tamtam, M.R.; Koutavarapu, R.; Shim, J. InVO4 nanosheets decorated with ZnWO4 nanorods: A novel composite and its enhanced photocatalytic performance under solar light. Environ. Res. 2023, 227, 115735. [Google Scholar] [CrossRef]
  28. Ding, W.; Lin, X.; Ma, G.; Lu, Q. Designed formation of InVO4/CeVO4 hollow nanobelts with Z-scheme charge transfer: Synergistically boosting visible-light-driven photocatalytic degradation of tetracycline. J. Environ. Chem. Eng. 2020, 8, 104588. [Google Scholar] [CrossRef]
  29. Latif, M.; Hammad Aziz, M.; Shaheen, F.; Mansoor Ali, S.; Asif, M.; Huang, Q. Enhanced photocatalysis activity of Co0.5Mg0.5Fe2O4/rGO nanocomposites for tetracycline antibiotic degradation. Mater. Lett. 2023, 360, 135756. [Google Scholar] [CrossRef]
  30. Zhang, H.; Xie, L.; Liang, Z.; Huang, C.; Wang, H.; Hu, J.; Jiang, Z.; Song, F. Tandem catalysis for enhanced CO oxidation over the Bi–Au–SiO2 interface. Nucl. Sci. Tech. 2023, 34, 108. [Google Scholar] [CrossRef]
  31. Wang, X.; Lin, X.; Wu, X.; Lynch, I. Z-scheme Fe@Fe2O3/BiOBr heterojunction with efficient carrier separation for enhanced heterogeneous photo-Fenton activity of tetracycline degradation: Fe2+ regeneration, mechanism insight and toxicity evaluation. Environ. Res. 2024, 252, 118396. [Google Scholar] [CrossRef] [PubMed]
  32. Hu, C.; He, J.; Liang, J.; Lin, T.; Liu, Q. Heterogeneous photo-Fenton catalyst α-Fe2O3@g-C3N4@NH2-MIL-101(Fe) with dual Z-Scheme heterojunction for degradation of tetracycline. Environ. Res. 2023, 231, 116313. [Google Scholar] [CrossRef] [PubMed]
  33. Guo, T.; Wang, K.; Zhang, G.; Wu, X. A novel α-Fe2O3@g-C3N4 catalyst: Synthesis derived from Fe-based MOF and its superior photo-Fenton performance. Appl. Surf. Sci. 2019, 469, 331–339. [Google Scholar] [CrossRef]
  34. Li, M.; Liu, H.; Pang, S.; Yan, P.; Liu, M.; Ding, M.; Zhang, B. Facile fabrication of three-dimensional fusiform-like α-Fe2O3 for enhanced photocatalytic performance. Nanomaterials 2021, 11, 2650. [Google Scholar] [CrossRef] [PubMed]
  35. Keerthana, S.P.; Yuvakkumar, R.; Ravi, G.; Kumar, P.; Elshikh, M.S.; Alkhamis, H.H.; Alrefaei, A.F.; Velauthapillai, D. A strategy to enhance the photocatalytic efficiency of α-Fe2O3. Chemosphere 2021, 270, 129498. [Google Scholar] [CrossRef] [PubMed]
  36. Lejbini, M.B.; Sangpour, P.; Tajabadi, F.; Kolahi, A. Photocatalytic upgrading α-Fe2O3 nanoparticles by incorporating MoS2/rGO nanosheets. Mater. Res. Express 2020, 7, 055011. [Google Scholar] [CrossRef]
  37. Niu, Y.; Li, M.; Jia, X.; Shi, Z.; Liu, H.; Zhang, X. Structures and photocatalytic activity of α-Fe2O3@TiO2 core-shell nanoparticles. Solid State Commun. 2022, 345, 114683. [Google Scholar] [CrossRef]
  38. Yogesh, K.; Rohit, K.; Pankaj, R.; Aftab Aslam Parwaz, K.; Arachana, S.; Quyet Van, L.; Van Huy, N.; Rangabhashiyam, S.; Sourbh, T.; Pardeep, S. Current status of Hematite (α-Fe2O3) based Z-scheme photocatalytic systems for environmental and energy applications. J. Environ. Chem. Eng. 2022, 10, 107427. [Google Scholar] [CrossRef]
  39. Solanki, R.; Jarosova, M.; Al-Azzawi, W.K.; Machek, P.; Alsultany, F.H.; Khalaji, A.D.; Al Mashhadani, Z.I. Sonochemical-assisted synthesis of α-Fe2O3 nanoparticles and their photocatalytic activity toward methylene blue and methyl orange dyes. Appl. Phys. A 2022, 128, 741. [Google Scholar] [CrossRef]
  40. Wang, Y. Study on Synthesis and Application of Ferric Oxide and Its Complex. Master’s Thesis, Yangzhou University, Yangzhou, China, 2012. [Google Scholar]
  41. Kumar, A.; Prajapati, P.K.; Pal, U.; Jain, S.L. Ternary rGO/InVO4/Fe2O3 Z-scheme heterostructured photocatalyst for CO2 reduction under visible light irradiation. ACS Sustain. Chem. Eng. 2018, 6, 8201–8211. [Google Scholar] [CrossRef]
  42. Lian, P.; Qin, A.; Liu, Z.; Ma, H.; Liao, L.; Zhang, K.; Li, N. Facile synthesis to porous TiO2 nanostructures at low temperature for efficient visible-light degradation of tetracycline. Nanomaterials 2024, 14, 943. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, Y.; Gao, J.; Xin, X.; Wang, L.; Li, H.; Zheng, X.; Jiang, Y. Immobilization laccase on heterophase TiO2 microsphere as a photo-enzyme integrated catalyst for emerging contaminants degradation under visible light. Appl. Mater. Today 2020, 21, 100810. [Google Scholar] [CrossRef]
  44. Li, Y.; Zhang, H.; Zhang, D.; Yao, S.; Dong, S.; Chen, Q.; Fan, F.; Jia, H.; Dong, M. Construction of Bi2WO6/g-C3N4 Z-scheme heterojunction and its enhanced photocatalytic degradation of tetracycline with persulfate under solar light. Molecules 2024, 29, 1169. [Google Scholar] [CrossRef]
  45. Ji, L.; Li, K.; Gao, M.; Lu, L.; Liang, Q.; Lan, H.; Lu, W.; Zhang, W.; Zhang, Y. Elucidating tetracycline degradation in a coupled system of photocatalysis combined with bacillus cereus. J. Water Process Eng. 2023, 57, 104616. [Google Scholar] [CrossRef]
  46. Chen, L.; Xu, B.; Jin, M.; Chen, L.; Yi, G.; Xing, B.; Zhang, Y.; Wu, Y.; Li, Z. Excellent photocatalysis of Bi2WO6 structured with oxygen vacancies in degradation of tetracycline. J. Mol. Struct. 2023, 1278, 134911. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the synthesis steps of InVO4/α-Fe2O3.
Figure 1. Schematic diagram of the synthesis steps of InVO4/α-Fe2O3.
Nanomaterials 14 01441 g001
Figure 2. XRD diffraction patterns of InVO4, α-Fe2O3, InVO4/0.25α-Fe2O3, and *InVO4/0.25α-Fe2O3 (InVO4/0.25α-Fe2O3 after degradation once) in the range of (a) 2θ = 5–35° and (b) 2θ = 11.8–12.5°.
Figure 2. XRD diffraction patterns of InVO4, α-Fe2O3, InVO4/0.25α-Fe2O3, and *InVO4/0.25α-Fe2O3 (InVO4/0.25α-Fe2O3 after degradation once) in the range of (a) 2θ = 5–35° and (b) 2θ = 11.8–12.5°.
Nanomaterials 14 01441 g002
Figure 3. (a) N2 adsorption–desorption isotherm and (b) the pore size distribution curves of the samples.
Figure 3. (a) N2 adsorption–desorption isotherm and (b) the pore size distribution curves of the samples.
Nanomaterials 14 01441 g003
Figure 4. SEM images of (a) α-Fe2O3, (b) InVO4, and (c) InVO4/0.25α-Fe2O3 complex samples.
Figure 4. SEM images of (a) α-Fe2O3, (b) InVO4, and (c) InVO4/0.25α-Fe2O3 complex samples.
Nanomaterials 14 01441 g004
Figure 5. (a) TEM and (b) HRTEM images of InVO4/0.25α-Fe2O3 complex samples.
Figure 5. (a) TEM and (b) HRTEM images of InVO4/0.25α-Fe2O3 complex samples.
Nanomaterials 14 01441 g005
Figure 6. (a) EDS elemental mapping of InVO4/0.25α-Fe2O3 complexes and (b) the corresponding elemental mapping for In, V, O and Fe.
Figure 6. (a) EDS elemental mapping of InVO4/0.25α-Fe2O3 complexes and (b) the corresponding elemental mapping for In, V, O and Fe.
Nanomaterials 14 01441 g006
Figure 7. High-resolution XPS spectra of InVO4/0.25α-Fe2O3 complexes. (a) XPS survey spectra, (b) V 2p, (c) O 1s, (d) In 3d, and (e) Fe 2p.
Figure 7. High-resolution XPS spectra of InVO4/0.25α-Fe2O3 complexes. (a) XPS survey spectra, (b) V 2p, (c) O 1s, (d) In 3d, and (e) Fe 2p.
Nanomaterials 14 01441 g007
Figure 8. (a) Ultraviolet–visible absorption spectra of InVO4, α-Fe2O3, and InVO4/0.25α-Fe2O3 and the tauc plot for the band gap determination of (b) InVO4, (c) α-Fe2O3, and (d) InVO4/0.25α-Fe2O3.
Figure 8. (a) Ultraviolet–visible absorption spectra of InVO4, α-Fe2O3, and InVO4/0.25α-Fe2O3 and the tauc plot for the band gap determination of (b) InVO4, (c) α-Fe2O3, and (d) InVO4/0.25α-Fe2O3.
Nanomaterials 14 01441 g008
Figure 9. (a) The effect of irradiation time on the degradation efficiency of different photocatalysts (InVO4, α-Fe2O3, and InVO4/α-Fe2O3). (b) The effect of irradiation time on the degradation efficiency of different TC initial concentrations. (c) Initial pH effect of experimental conditions on the degradation of TC.
Figure 9. (a) The effect of irradiation time on the degradation efficiency of different photocatalysts (InVO4, α-Fe2O3, and InVO4/α-Fe2O3). (b) The effect of irradiation time on the degradation efficiency of different TC initial concentrations. (c) Initial pH effect of experimental conditions on the degradation of TC.
Nanomaterials 14 01441 g009
Figure 10. (a) XRD patterns of the InVO4/0.25α-Fe2O3 photocatalyst before and after the degradation of TC. (b) Reusability of the InVO4/0.25α-Fe2O3 photocatalyst.
Figure 10. (a) XRD patterns of the InVO4/0.25α-Fe2O3 photocatalyst before and after the degradation of TC. (b) Reusability of the InVO4/0.25α-Fe2O3 photocatalyst.
Nanomaterials 14 01441 g010
Figure 11. (a) Trapping experiments of the active species. (b) Possible mechanism of TC degradation by the InVO4/0.25α-Fe2O3 composite photocatalyst.
Figure 11. (a) Trapping experiments of the active species. (b) Possible mechanism of TC degradation by the InVO4/0.25α-Fe2O3 composite photocatalyst.
Nanomaterials 14 01441 g011
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chang, H.; Wang, Y.; Qiao, P.; Sun, B.; Wang, Z.; Song, F. Formulating InVO4/α-Fe2O3 Heterojunction Composites for Photocatalytic Tetracycline Hydrochloride Degradation. Nanomaterials 2024, 14, 1441. https://doi.org/10.3390/nano14171441

AMA Style

Chang H, Wang Y, Qiao P, Sun B, Wang Z, Song F. Formulating InVO4/α-Fe2O3 Heterojunction Composites for Photocatalytic Tetracycline Hydrochloride Degradation. Nanomaterials. 2024; 14(17):1441. https://doi.org/10.3390/nano14171441

Chicago/Turabian Style

Chang, Haoxu, Yayang Wang, Panzhe Qiao, Bo Sun, Zhengbang Wang, and Fei Song. 2024. "Formulating InVO4/α-Fe2O3 Heterojunction Composites for Photocatalytic Tetracycline Hydrochloride Degradation" Nanomaterials 14, no. 17: 1441. https://doi.org/10.3390/nano14171441

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