Next Article in Journal
Kinetic Modelling of Esterification and Transesterification Processes for Biodiesel Production Utilising Waste-Based Resource
Next Article in Special Issue
Visible-Light-Active Vanadium and Copper Co-Doped gCN Nanosheets with Double Direct Z-Scheme Heterojunctions for Photocatalytic Removal of Monocrotophos Pesticide in Water
Previous Article in Journal
NiFe Layered Double Hydroxide Electrocatalyst Prepared via an Electrochemical Deposition Method for the Oxygen Evolution Reaction
Previous Article in Special Issue
Natural Polymer-Based Iron Oxide (Fe3O4) Synthesis, Characterization and Its Application for 1-Amino-Nitrobenzene Degradation in Assistance with Oxidants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Indium Vanadium Oxide Nanosheet-Supported Nickel Iron Oxide Nanoplate Heterostructure for Synergistically Enhanced Photocatalytic Degradation of Tetracycline

1
Department of Physics, Acharya Nagarjuna University, Nagarjuna Nagar, Guntur 522 510, Andhra Pradesh, India
2
Department of Physics, Bapatla Engineering College, Bapatla 522 102, Andhra Pradesh, India
3
Department of Robotics Engineering, College of Mechanical and IT Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
4
School of Mechanical Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1471; https://doi.org/10.3390/catal12111471
Submission received: 17 October 2022 / Revised: 12 November 2022 / Accepted: 16 November 2022 / Published: 18 November 2022
(This article belongs to the Special Issue Synthesis and Photocatalytic Activity of Composite)

Abstract

:
Semiconductor-based heterogeneous photocatalytic oxidation processes have received considerable attention for the remediation of toxic pollutants. Herein, InVO4/NiFe2O4 nanocomposites were synthesized using a facile hydrothermal technique. Furthermore, various characterization results revealed the successful loading of NiFe2O4 nanoplates over InVO4 nanosheets, thereby signifying the formation of a heterostructure. The performance of the synthesized photocatalyst was tested for tetracycline (TC) antibiotic removal. The optimized InVO4/NiFe2O4 nanocomposite exhibits maximum photodegradation of TC molecules (96.68%) in 96 min; this is approximately 6.47 and 4.93 times higher than that observed when using NiFe2O4 and InVO4, respectively. The strong interaction between the InVO4 nanosheets and NiFe2O4 nanoplates can improve the visible-light absorption and hinder the recombination of charge carriers, further enhancing the photocatalytic performance. Moreover, hydroxyl radicals play a crucial role in the photodegradation of TC antibiotics.

1. Introduction

Water is an important resource for every living being on Earth [1,2]. Currently, the demand for clean water is increasing with the exponential growth of the population and the intensification of industrial development [3,4]. In recent years, the rapid growth of science and technology worldwide has promoted the establishment of factories in several sectors, such as semiconductors, microelectronics, and pharmaceuticals. During processing, most factories require substantial amounts of water and discharge wastewater containing harmful contaminants [5,6]. Among these, emerging pharmaceutical contaminants, such as tetracycline (TC) antibiotics, have been identified as one of the most important factors contributing to water pollution [7,8]. Therefore, owing to the extensive use of TC antibiotics and their low biodegradability, trace levels of TC molecules have been detected in various aquatic compartments [9,10]. Additionally, even limited amounts of TC antibiotics in freshwater may result in strains of resistant microorganisms in humans and other organisms [11,12]. Therefore, efficient and low-cost methods must be identified for water remediation [13,14].
Recently, the conservation of natural water resources and improvement of novel methodologies for wastewater treatment have emerged as key environmental issues. Numerous wastewater processing techniques are currently being utilized with variable degrees of efficiency [15,16]. Photocatalysis has been developed as a suitable technique owing to its economical, non-toxic, safe, and renewable characteristics [17,18]. Moreover, advanced oxidation processes are the most widely used techniques for treating toxic pollutants, such as TC antibiotics, which exhibit high chemical stability and low biodegradability [19,20]. Therefore, semiconductor-based heterogeneous photocatalytic oxidation processes under visible light have received considerable attention for the remediation of toxic pollutants [21,22]. Consequently, novel high-efficiency photocatalysts have attracted significant interest owing to their considerable potential for visible-light utilization [23,24].
Among the various semiconductor metal oxides, indium vanadate has promising properties and is increasingly being used in environmental and energy applications owing to its controllable shape and improved activity [25,26]. InVO4 has the advantages of low raw cost, nontoxicity, a visible bandgap (~2.2 eV), and excellent photocatalytic performance under visible light [27]. However, the photocatalytic degradation of antibiotics by InVO4 is limited because of the high recombination rate of photogenerated electrons (e) and holes (h+) [28]. To overcome its inherent shortcomings, the fabrication of InVO4-based heterostructured photocatalysts has been extensively analyzed, and this is considered a feasible and efficient strategy. Yao et al. constructed InVO4/TiO2 nanobelts using an electrospinning technique [29]. The as-prepared InVO4/TiO2 nanobelts exhibited excellent photocatalytic activity under solar-light irradiation, owing to their unique morphology and synergetic interactions. Furthermore, B12O15Cl6/InVO4 nanocomposites were successfully constructed using an in situ hydrothermal method [30]. The establishment of a Z-scheme among the individual components can significantly improve the separation of charge carriers and enhance the photocatalytic activity. Yuan et al. efficiently synthesized In2S3/InVO4 nanocomposites through a facile in situ ion-exchange technique [31]. After 60 min of visible-light irradiation, the prepared nanocomposite degraded TC by 71.4%. The formation of the heterostructure further improved the visible-light absorption ability along with interfacial interactions, thereby indicating photocatalytic activity.
Spinel-structured transition metal oxides, such as nickel ferrite, are abundant and promising minerals with a narrow bandgap (~1.8 eV). Recently, these materials have attracted considerable research interest owing to their unique characteristics [32]. Moreover, the oxidation of these materials has been recognized as a promising technique because of the simple and environmentally friendly reaction process [33]. Accordingly, the construction of an InVO4-based heterostructure that is incorporated with NiFe2O4 can be regarded as a potential strategy to facilitate charge separation at the interface, thereby enhancing the photocatalytic degradation of TC antibiotics.
Herein, InVO4 and NiFe2O4 are hydrothermally combined to construct a novel two-dimensional InVO4/NiFe2O4 heterostructure to remove TC antibiotics from the solution. The structure, morphology, chemical composition, and optical properties of the as-prepared photocatalysts are studied, and their photocatalytic activity is examined by degrading TC. Compared with the pure samples, the InVO4/NiFe2O4 heterostructure exhibits higher oxidation ability, faster kinetics, and higher stability owing to the lower radiative recombination rate of charge carriers. Furthermore, a possible charge transfer mechanism with enhanced photocatalytic activity is proposed. The present study provides a new and effective approach for achieving InVO4-based heterostructures, which could facilitate the preparation of similar types of materials.

2. Results and Discussion

2.1. X-ray Diffraction (XRD) Analysis

To understand the crystal structure and crystallinity of the InVO4/NiFe2O4 nanocomposites, the photocatalysts were characterized using XRD; the results are illustrated in Figure 1. From Figure 1a, the peaks identified at 18.62°, 25.62°, 31.04°, 33.04°, 35.14°, 46.96°, 56.47°, and 60.98° correspond to the (110), (021), (200), (112), (130), (222), (312), and (152) planes of InVO4 (JCPDS# 00-048-0898). The observed crystal planes match the standard results, thereby confirming the orthorhombic phase of InVO4 [34]. Furthermore, as shown in Figure 1b, the diffraction peaks at 30.08°, 35.44°, 43.26°, 57.38°, and 62.99° are in agreement with the (220), (311), (400), (511), and (440) planes of NiFe2O4 (JCPDS #00-074-2081) [35]. The XRD patterns of the InVO4/NiFe2O4 nanocomposites at different concentrations (Figure 1c–e) reveal the existence of both InVO4 (@) and NiFe2O4 ($). The loading of NiFe2O4 at different concentrations does not cause any deviations in the orthorhombic phase of InVO4. However, the intensified InVO4 diffraction patterns diminish with increasing NiFe2O4 concentration, confirming the successful formation of the InVO4/NiFe2O4 nanocomposite [36]. No additional peaks are observed in the XRD pattern, indicating that the pattern represents the pure phase, and no impurities are present.

2.2. Morphological Analysis

The morphology of a photocatalyst significantly influences its absorption of light, thereby determining its photocatalytic efficiency. Scanning electron microscopy (SEM), transmission electron microscopy (TEM), and high-resolution TEM (HRTEM) were used to examine the morphologies of InVO4, NiFe2O4, and InVO4/NiFe2O4 (IVNF-5.0) nanocomposites. As shown in Figure 2a,b, a sheet-like structure was observed for InVO4 through SEM analysis. SEM images of NiFe2O4 (Figure 2c,d) reveal a plate-like morphology with irregular shapes and sizes. Furthermore, Figure 2e,f clearly show that the IVNF-5.0 nanocomposite is composed of InVO4 nanosheets and NiFe2O4 nanoplates. Moreover, the NiFe2O4 nanoplates were successfully loaded onto the InVO4 nanosheets. The energy-dispersive spectroscopy (EDS) analysis of the IVNF-5.0 nanocomposite (Figure 2g) indicated the presence of individual components.
The TEM images illustrate the formation of InVO4 nanosheets (Figure 3a), NiFe2O4 nanoplates (Figure 3b), and the loading of NiFe2O4 nanoplates over InVO4 nanosheets (Figure 3c). Furthermore, the HRTEM image of IVNF-5.0 confirmed the effective loading of NiFe2O4 nanoplates on the InVO4 nanosheets (Figure 3d). As shown in Figure 3e,f, the lattice spacings of 0.265 and 0.245 nm were attributed to the (112) and (311) crystal planes of InVO4 and NiFe2O4, respectively [37,38]. Therefore, in terms of the position, the (311) plane of NiFe2O4 overlapped with the (112) crystal surface of InVO4, confirming the formation of the InVO4/NiFe2O4 heterostructure. The EDS color mapping of IVNF-5.0 (Figure 4) further confirmed the existence of In, V, Ni, Fe, and O.

2.3. X-ray Photoelectron Spectroscopy (XPS) Analysis

XPS was applied to survey the chemical state of each component of the InVO4/NiFe2O4 nanocomposite (Figure 5). In 3d, V 2p, O 1s, Fe 2p, and Ni 2p signals are clearly observed in the XPS general spectrum (Figure 5a) of the IVNF-5.0 photocatalyst. From Figure 5b, the XPS peak fitting of In 3d indicates that the signals at 444.72 and 452.29 eV are consistent with the +3 oxidation state of In 3d5/2 and In 3d3/2 energy levels, respectively [39]. As shown in Figure 5c, the XPS spectrum of V 2p exhibits two signals at 516.93 and 524.29 eV, which are assigned to the +5 oxidation state of V 2p3/2 and V 2p1/2 energy levels, respectively [40]. The asymmetric O 1s peak (Figure 5d) is divided into two peaks at 529.81 and 531.05 eV, corresponding to the oxygen species in the metal–oxygen bond [41]. The XPS spectrum of Fe 2p (Figure 5e) exhibits two significant signals at 710.85 and 724.16 eV, which are ascribed to Fe 2p3/2 and Fe 2p1/2, respectively, and are consistent with metal–oxygen formation [42]. Furthermore, the signals at 716.21 and 732.02 eV are attributed to the satellite signals of Fe3+ [43]. As shown in Figure 5f, the XPS spectrum of Ni 2p exhibits two signals at 855.22 and 873.31 eV, corresponding to Ni 2p3/2 and Ni 2p1/2, respectively [44]. The signals at 861.51 and 880.32 eV are assigned to the satellite signals of Ni2+ [45]. This analysis reveals a strong interaction between InVO4 and NiFe2O4 and confirms the formation of the nanocomposite.

2.4. Optical Absorption Analysis

The ability to absorb light in the visible region is a critical property of photocatalysts. Herein, the responses of the prepared samples to light were determined using diffuse reflectance spectroscopy (DRS). The absorption properties for visible light and the bandgaps of InVO4, NiFe2O4, and InVO4/NiFe2O4 nanocomposites are presented in Figure 6. As shown in Figure 6a, both InVO4 and NiFe2O4 displayed maximum absorbance with band edges at approximately 579 and 680 nm, respectively [46,47]. Moreover, compared with pure InVO4, the absorption performance of InVO4/NiFe2O4 nanocomposites for visible light was significantly improved toward higher wavelength regions. Therefore, the InVO4/NiFe2O4 nanocomposites demonstrate a good response in the visible region, in contrast to the pure samples. This confirms the construction of a heterostructure, which is beneficial for enhanced photocatalytic activity [48]. Furthermore, the bandgaps of InVO4, NiFe2O4, and InVO4/NiFe2O4 nanocomposites were evaluated based on the corresponding Tauc plots (Figure 6b–d). After extrapolation, the bandgaps of InVO4, NiFe2O4, IVNF-2.5, IVNF-5.0, and IVNF-7.5 were measured as 2.142, 1.824, 1.992, 1.938, and 1.852 eV, respectively. A reduction was observed in the bandgap between the pure InVO4 and the InVO4/NiFe2O4 nanocomposites. Consequently, the visible-light absorption capacity of the InVO4/NiFe2O4 nanocomposites is better than that of the pure samples.

2.5. Photoluminescence (PL) Analysis

The PL intensity of each sample was tested at the same excitation wavelength (325 nm). The PL spectra of InVO4, NiFe2O4, and InVO4/NiFe2O4 nanocomposites are shown in Figure 7. For InVO4, a strong PL emission peak was observed at approximately 586 nm, which is attributed to the recombination of charge carriers [49]. Furthermore, strong PL emission peaks were observed for NiFe2O4 at approximately 465 and 589 nm, which were caused by oxygen vacancies and structural defects [50]. Compared to the pure samples, the PL intensities of the InVO4/NiFe2O4 nanocomposites were significantly reduced, and the IVNF-5.0 sample exhibited the lowest PL intensity. Therefore, the InVO4/NiFe2O4 nanocomposites display a relatively high ability to separate the photogenerated charge carriers in addition to enhanced photocatalytic performance.

2.6. Photocatalytic Performance

The photodegradation efficiency of the as-synthesized photocatalysts was examined for TC removal under visible light. Before irradiation with visible light, the reactive solution was stirred for 30 min in the dark to ensure that the TC solution was evenly distributed on the photocatalyst surface, and the adsorption process reached equilibrium. The absorbance spectra for the degradation of TC over InVO4/NiFe2O4 nanocomposites were recorded every 12 min over a time period of 96 min; the results are presented in Figure 8a,b. The degradation efficiencies of the pure NiFe2O4 and InVO4 samples were 35.26% and 40.97%, respectively. Furthermore, the degradation efficiencies of IVNF-2.5, IVNF-5.0, and IVNF-7.5 were 74.72%, 96.68%, and 81.27%, respectively. Therefore, upon visible-light irradiation, the order of increasing photocatalytic degradation performance is given by: IVNF-5.0 > IVNF-7.5 > IVNF-2.5 > InVO4 > NiFe2O4.
To further evaluate the photocatalytic degradation of TC by the InVO4/NiFe2O4 nanocomposite, a pseudo-first-order kinetic model was considered. The results of kinetic modeling are presented in Figure 9a,b. The kinetic rate constants of NiFe2O4, InVO4, IVNF-2.5, IVNF-5.0, and IVNF-7.5 were calculated to be 0.0045, 0.0059, 0.0132, 0.0291, and 0.0161 min−1, respectively, quantitatively confirming that the IVNF-5.0 photocatalyst is superior to the other samples.
Moreover, the rate constant of IVNF-5.0 was 6.47 times that of NiFe2O4 and 4.93 times that of InVO4. Therefore, the IVNF-5.0 nanocomposite is demonstrated to be a better photocatalyst than the other prepared samples; it significantly enhances the separation ability for photogenerated carriers and retains a strong redox ability, thereby promoting the TC degradation efficiency. The catalytic removal of TC in the present work was compared with that previously reported in the literature, as shown in Table 1. The data reveal the improved photocatalytic activity of the InVO4/NiFe2O4 nanocomposites.
The influence of catalyst loading is an important parameter in determining the degradation of TC by the IVNF-5.0 photocatalyst. The results obtained using various catalyst loadings are shown in Figure 10a. With the gradual increase in catalyst loading, the TC degradation efficiency increases, and the time required for the degradation to attain equilibrium decreases. When the loading of IVNF-5.0 is 10 mg, the TC degradation efficiency is only 78.54%. Furthermore, when the loading of IVNF-5.0 is increased to 15 mg, the TC degradation effect is optimal, with an efficiency of approximately 96.68%. However, when the loading of IVNF-5.0 is further increased to 20 mg, the TC degradation effect is decreased, and its efficiency is approximately 62.33%. As the catalyst loading concentration continues to increase, the rate of degradation begins to decrease at a particular limit because of the decreased diffusion of light, which hinders the generation of radical species [56]. Furthermore, the stability of the InVO4/NiFe2O4 nanocomposite was examined under the optimal experimental conditions for various practical applications. A stability test was conducted with the optimized sample IVNF-5.0 for four cycles under visible light; the results are depicted in Figure 10b. After the IVNF-5.0 photocatalyst degraded TC for 96 min, the photocatalyst was recycled, and the abovementioned process was repeated four times. After four cycles, the TC degradation efficiency reached 93.49%, indicating that the IVNF-5.0 photocatalyst can be effectively reused and has good stability.
Additionally, radical quenching experiments were conducted to identify the influence of free radicals on photodegradation; the results are shown in Figure 11. Herein, BQ, IPA, and TEA were used as scavengers for superoxide (•O2) and hydroxyl (•OH) radicals and h+, respectively. As shown in Figure 11, when BQ is added to the reactive solution, TC removal over the IVNF-5.0 photocatalyst decreases considerably to 86.75%. After the addition of TEA, TC removal noticeably decreases to 74.61%. After adding IPA, TC removal is further decreased to 41.29%. Based on these results, •OH radicals are considered to play a more dominant role than •O2 radicals and h+ in photocatalytic reactions. Furthermore, to elaborate on the generation and separation efficiency of photoexcited electron–hole pairs, transient photocurrent (I–t) tests were conducted. The I–t responses of NiFe2O4, InVO4, and InVO4/NiFe2O4 nanocomposites under visible light are presented in Figure 11b. The IVNF-5.0 nanocomposite exhibited the strongest response among all samples. Moreover, the I–t response intensity of the IVNF-5.0 nanocomposite was maintained despite several on–off cycles of visible-light irradiation, which demonstrates its excellent photochemical stability.

2.7. Photocatalytic Mechanism

The reaction mechanism of the InVO4/NiFe2O4 nanocomposite under visible-light irradiation is shown in Figure 12. Moreover, the valence band (VB) and conduction band (CB) edge potentials of InVO4 and NiFe2O4 were projected using the Butler–Ginley equation to further understand the migration of charge carriers after heterostructure formation [57]. The susceptibility (χ) values of InVO4 and NiFe2O4 are approximately 5.753 and 5.836 eV, respectively. Therefore, InVO4 has VB and CB potentials of +2.324 and +0.182 eV, respectively, whereas NiFe2O4 has VB and CB potentials of +2.248 and +0.424 eV, respectively. During photocatalysis, visible-light irradiation results in the formation of h+ and e, which are transferred to the CB and VB of the InVO4/NiFe2O4 heterostructure. Therefore, e on the CB of InVO4 are transferred to the CB of NiFe2O4. Simultaneously, h+ in the VB of InVO4 are delivered to the VB of NiFe2O4. Furthermore, e in the CB of NiFe2O4 can effectively generate •OH radicals because of their lower negative reduction potential, and they are used for the degradation of TC molecules. The photoinduced h+ in the VB of NiFe2O4 is employed for the exchange of a water molecule and is thereafter converted to •OH radicals in the visible region. Therefore, •OH radicals are the strongest oxidants for TC removal. Therefore, a heterojunction can significantly detach electron–hole pairs and produce highly active species on the surface, which enhances the photocatalytic degradation ability.

3. Materials and Methods

3.1. Preparation of InVO4 Nanosheets

InVO4 powder was prepared using a hydrothermal method. Initially, NH4VO3 and InCl3.4H2O were added to 60 mL of deionized water at an equal molar ratio (0.1 M) and vigorously stirred for 60 min until dissolution. Subsequently, HNO3 solution was added dropwise to maintain the pH below 2. The solution was magnetically stirred for 30 min, poured into a Teflon-lined stainless-steel autoclave, heat-treated at 180 °C for 18 h, and naturally cooled to ambient temperature. The resulting products were centrifuged thrice with deionized water and ethanol and dried overnight in an oven at 60 °C.

3.2. Preparation of NiFe2O4 Nanoplates

NiFe2O4 powder was prepared using a calcination technique. Initially, FeCl3.6H2O and NiCl2.6H2O were liquefied in 100 mL of deionized water at a 1:2 molar ratio and stirred for 30 min. The obtained red-brown product was washed thrice and calcinated in a muffle furnace at 500 °C for 3 h.

3.3. Synthesis of InVO4/NiFe2O4 Nanocomposites

Herein, InVO4/NiFe2O4 nanocomposites were obtained using a hydrothermal technique. Typically, 100 mg of the prepared InVO4 powder and different concentrations of NiFe2O4 (2.5, 5.0, and 7.5 mg) powder were separately transferred to 30 mL of deionized water under stirring for 30 min. Different concentrations of NiFe2O4 were then added to the InVO4 solution and stirred vigorously for 1 h. Subsequently, the solution was poured into a Teflon-lined stainless-steel autoclave and heat-treated at 150 °C for 12 h. After being naturally cooled to ambient temperature, the collected products were centrifuged thrice and dried overnight in an oven at 60 °C. For simplicity, the nanocomposites were denoted as IVNF-2.5, IVNF-5.0, and IVNF-7.5. A detailed description of the reagents, characterization, and photodegradation tests can be found in the Supplementary Material.

4. Conclusions

Herein, novel InVO4/NiFe2O4 nanocomposites were successfully prepared using a facile hydrothermal method. XRD analysis confirmed the orthorhombic phase of InVO4 and spinel structure of NiFe2O4. Morphological analysis confirmed the strong interactions between the InVO4 nanosheets and NiFe2O4 nanoplates. From the DRS analysis, a reduction was observed in the bandgap between the pure InVO4 and InVO4/NiFe2O4 nanocomposites, thereby indicating improved visible-light absorption than that of the pure samples. Furthermore, the optimized IVNF-5.0 nanocomposite was demonstrated to be an efficient photocatalyst for the degradation of TC (96.68%) in 96 min under visible light. The InVO4/NiFe2O4 photocatalyst exhibited good stability and reusability after four cycles. •OH radicals play a more dominant role than •O2 radicals and h+ in the photocatalytic reaction. Therefore, the synergetic effect between the InVO4 nanosheets and NiFe2O4 nanoplates significantly enhances the separation ability for the photogenerated carriers and retains their strong redox ability, thereby promoting the TC degradation efficiency.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12111471/s1, Figure S1. XRD pattern of IVNF-5.0 nancomposite (a) before and (b) after photocatalytic activity; Figure S2. SEM images of IVNF-5.0 nancomposite (a) before and (b) after photocatalytic activity; Figure S3. TEM images of IVNF-5.0 nancomposite (a) before and (b) after photocatalytic activity.

Author Contributions

Conceptualization, methodology, formal analysis, and writing—original draft preparation: N.S.; Conceptualization, funding, supervision, and review: V.A.; Conceptualization, methodology, data acquisition, formal analysis, and writing—original draft preparation: R.K.; supervision and review: D.-Y.L.; supervision, funding, and review: J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Korean Government’s National Research Foundation (NRF) (2020R1A2C1012439), Republic of Korea.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors have no conflicts of interest to declare that are relevant to the content of this article.

References

  1. Velempini, T.; Prabakaran, E.; Pillay, K. Recent developments in the use of metal oxides for photocatalytic degradation of pharmaceutical pollutants in water—A review. Mater. Today Chem. 2021, 19, 100380. [Google Scholar] [CrossRef]
  2. Subhiksha, V.; Kokilavani, S.; Sudheer Khan, S. Recent advances in degradation of organic pollutant in aqueous solutions using bismuth based photocatalysts: A review. Chemosphere 2022, 290, 133228. [Google Scholar] [CrossRef] [PubMed]
  3. Zia, J.; Fatima, F.; Riaz, U. A comprehensive review on the photocatalytic activity of polythiophene-based nanocomposites against degradation of organic pollutants. Catal. Sci. Technol. 2021, 11, 6630–6648. [Google Scholar] [CrossRef]
  4. Zuo, W.; Zhang, L.; Zhang, Z.; Tang, S.; Sun, Y.; Huang, H.; Yu, Y. Degradation of organic pollutants by intimately coupling photocatalytic materials with microbes: A review. Crit. Rev. Biotechnol. 2021, 41, 273–299. [Google Scholar] [CrossRef] [PubMed]
  5. Ateia, M.; Alalm, M.G.; Awfa, D.; Johnson, M.S.; Yoshimura, C. Modeling the degradation and disinfection of water pollutants by photocatalysts and composites: A critical review. Sci. Total Environ. 2020, 698, 134197. [Google Scholar] [CrossRef]
  6. Gusain, R.; Gupta, K.; Joshi, P.; Khatri, O.P. Adsorptive removal and photocatalytic degradation of organic pollutants using metal oxides and their composites: A comprehensive review. Adv. Colloid Interface Sci. 2019, 272, 102009. [Google Scholar] [CrossRef]
  7. Abdurahman, M.H.; Abdullah, A.Z.; Shoparwe, N.F. A comprehensive review on sonocatalytic, photocatalytic, and sonophotocatalytic processes for the degradation of antibiotics in water: Synergistic mechanism and degradation pathway. Chem. Eng. J. 2021, 413, 127412. [Google Scholar] [CrossRef]
  8. Chen, G.; Yu, Y.; Liang, L.; Duan, X.; Li, R.; Lu, X.; Yan, B.; Li, N.; Wang, S. Remediation of antibiotic wastewater by coupled photocatalytic and persulfate oxidation system: A critical review. J. Hazard. Mater. 2021, 408, 124461. [Google Scholar] [CrossRef]
  9. Abbasnia, A.; Zarei, A.; Yeganeh, M.; Sobhi, H.R.; Gholami, M.; Esrafili, A. Removal of tetracycline antibiotics by adsorption and photocatalytic-degradation processes in aqueous solutions using metal organic frameworks (MOFs): A systematic review. Inorg. Chem. Commun. 2022, 145, 109959. [Google Scholar] [CrossRef]
  10. Dai, Y.; Liu, M.; Li, J.; Yang, S.; Sun, Y.; Sun, Q.; Wang, W.; Lu, L.; Zhang, K.; Xu, J.; et al. A review on pollution situation and treatment methods of tetracycline in groundwater. Sep. Sci. Technol. 2020, 55, 1005–1021. [Google Scholar] [CrossRef]
  11. Koutavarapu, R.; Jang, W.Y.; Rao, M.C.; Arumugam, M.; Shim, J. Novel BiVO4-nanosheet-supported MoS2-nanoflake-heterostructure with synergistic enhanced photocatalytic removal of tetracycline under visible light irradiation. Chemosphere 2022, 305, 135465. [Google Scholar] [CrossRef]
  12. 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]
  13. Ahmed, S.; Rasul, M.G.; Brown, R.; Hashib, M.A. Influence of parameters on the heterogeneous photocatalytic degradation of pesticides and phenolic contaminants in wastewater: A short review. J. Environ. Manag. 2011, 92, 311–330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Parul; Kaur, K.; Badru, R.; Singh, P.P.; Kaushal, S. Photodegradation of organic pollutants using heterojunctions: A review. J. Environ. Chem. Eng. 2020, 8, 103666. [Google Scholar] [CrossRef]
  15. Sharma, K.; Dutta, V.; Sharma, S.; Raizada, P.; Hosseini-Bandegharaei, A.; Thakur, P.; Singh, P. Recent advances in enhanced photocatalytic activity of bismuth oxyhalides for efficient photocatalysis of organic pollutants in water: A review. J. Ind. Eng. Chem. 2019, 78, 1–20. [Google Scholar] [CrossRef]
  16. Fauzi, A.A.; Jalil, A.A.; Hassan, N.S.; Aziz, F.F.A.; Azami, M.S.; Hussain, I.; Saravanan, R.; Vo, D.V.N. A critical review on relationship of CeO2-based photocatalyst towards mechanistic degradation of organic pollutant. Chemosphere 2022, 286, 131651. [Google Scholar] [CrossRef]
  17. Sudhaik, A.; Raizada, P.; Rangabhashiyam, S.; Singh, A.; Nguyen, V.-H.; Van Le, Q.; Khan, A.A.P.; Hu, C.; Huang, C.-W.; Ahamad, T.; et al. Copper sulfides based photocatalysts for degradation of environmental pollution hazards: A review on the recent catalyst design concepts and future perspectives. Surf. Interf. 2022, 33, 102182. [Google Scholar] [CrossRef]
  18. Qin, K.; Zhao, Q.; Yu, H.; Xia, X.; Li, J.; He, S.; Wei, L.; An, T. A review of bismuth-based photocatalysts for antibiotic degradation: Insight into the photocatalytic degradation performance, pathways and relevant mechanisms. Environ. Res. 2021, 199, 111360. [Google Scholar] [CrossRef]
  19. Silvestri, S.; Fajardo, A.R.; Iglesias, B.A. Supported porphyrins for the photocatalytic degradation of organic contaminants in water: A review. Environ. Chem. Lett. 2022, 20, 731–771. [Google Scholar] [CrossRef]
  20. He, X.; Kai, T.; Ding, P. Heterojunction photocatalysts for degradation of the tetracycline antibiotic: A review. Environ. Chem. Lett. 2021, 19, 4563–4601. [Google Scholar] [CrossRef]
  21. Hu, X.; Ye, Y.; Dong, W.; Huang, Y.; Zhu, M. Quantitative evaluation of structure-activity relationships in heterogeneous photocatalytic oxidation towards organic contaminants. Appl. Catal. B 2022, 309, 121238. [Google Scholar] [CrossRef]
  22. Calvete, M.J.F.; Piccirillo, G.; Vinagreiro, C.S.; Pereira, M.M. Hybrid materials for heterogeneous photocatalytic degradation of antibiotics. Coord. Chem. Rev. 2019, 395, 63–85. [Google Scholar] [CrossRef]
  23. Zhou, X.; Chen, Y.; Li, C.; Zhang, L.; Zhang, X.; Ning, X.; Zhan, L.; Luo, J. Construction of LaNiO3 nanoparticles modified g-C3N4 nanosheets for enhancing visible light photocatalytic activity towards tetracycline degradation. Sep. Pur. Technol. 2019, 211, 179–188. [Google Scholar] [CrossRef]
  24. Khanh Huyen, N.T.; Pham, T.-D.; Dieu Cam, N.T.; Van Quan, P.; Van Noi, N.; Hanh, N.T.; Thanh Tung, M.H.; Dao, V.-D. Fabrication of titanium doped BiVO4 as a novel visible light driven photocatalyst for degradation of residual tetracycline pollutant. Ceram. Int. 2021, 47, 34253–34259. [Google Scholar] [CrossRef]
  25. 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. 2022, 428, 131145. [Google Scholar] [CrossRef]
  26. Van Thuan, D.; Nguyen, T.L.; Pham Thi, H.H.; Thanh, N.T.; Ghotekar, S.; Sharma, A.K.; Binh, M.T.; Nga, T.T.; Pham, T.-D.; Cam, D.P. Development of Indium vanadate and Silver deposited on graphitic carbon nitride ternary heterojunction for advanced photocatalytic degradation of residual antibiotics in aqueous environment. Opt. Mater. 2022, 123, 111885. [Google Scholar] [CrossRef]
  27. Sreeram, N.; Aruna, V.; Koutavarapu, R.; Lee, D.-Y.; Shim, J. Visible-light-driven indium vanadium oxide nanosheets supported bismuth tungsten oxide nanoflakes heterostructure as an efficient photocatalyst for the tetracycline degradation. Chemosphere 2022, 299, 134477. [Google Scholar] [CrossRef]
  28. Wang, Z.; Wang, J.; Pan, Y.; Liu, F.; Lai, Y.; Li, J.; Jiang, L. Preparation and characterization of a novel and recyclable InVO4/ZnFe2O4 composite for methylene blue removal by adsorption and visible-light photocatalytic degradation. Appl. Surf. Sci. 2020, 501, 144006. [Google Scholar] [CrossRef]
  29. Yao, L.; Guo, E.; Wei, M.; Wang, Q.; Lu, Q. Electrospun Mesoporous InVO4/TiO2 Nanobelts with Enhanced Photocatalytic Properties. Photochem. Photobiol. 2019, 95, 1122–1130. [Google Scholar] [CrossRef]
  30. Yao, J.; Huang, L.; Li, Y.; Liu, J.; Liu, J.; Shu, S.; Huang, L.; Zhang, Z. Facile synthesizing Z-scheme Bi12O15Cl6/InVO4 heterojunction to effectively degrade pollutants and antibacterial under light-emitting diode light. J. Colloid Interf. Sci. 2022, 627, 224–237. [Google Scholar] [CrossRef]
  31. Yuan, X.; Jiang, L.; Liang, J.; Pan, Y.; Zhang, J.; Wang, H.; Leng, L.; Wu, Z.; Guan, R.; Zeng, G. In-situ synthesis of 3D microsphere-like In2S3/InVO4 heterojunction with efficient photocatalytic activity for tetracycline degradation under visible light irradiation. Chem. Eng. J. 2019, 356, 371–381. [Google Scholar] [CrossRef]
  32. Gebreslassie, G.; Bharali, P.; Chandra, U.; Sergawie, A.; Baruah, P.K.; Das, M.R.; Alemayehu, E. Hydrothermal Synthesis of g-C3N4/NiFe2O4 Nanocomposite and Its Enhanced Photocatalytic Activity. Appl. Organomet. Chem. 2019, 33, e5002. [Google Scholar] [CrossRef]
  33. Mishra, P.; Behera, A.; Kandi, D.; Parida, K. Facile construction of a novel NiFe2O4@P-doped g-C3N4 nanocomposite with enhanced visible-light-driven photocatalytic activity. Nanoscale Adv. 2019, 1, 1864–1879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Wang, C.; Yang, B.; Liu, H.; Xia, F.; Xiao, J. Potentiometric ammonia sensor with InVO4 sensing electrode. Sens. Actuators B 2020, 316, 128140. [Google Scholar] [CrossRef]
  35. Atla, R.; Shaik, B.; Oh, T.H. NiFe2O4 nanoplates decorated on MoS2 nanosheets as an effective visible light-driven heterostructure photocatalyst for the degradation of methyl orange. J. Mater. Sci. Mater. Electron. 2022, 1–14. [Google Scholar] [CrossRef]
  36. Shi, Z.; Lan, L.; Li, Y.; Yang, Y.; Zhang, Q.; Wu, J.; Zhang, G.; Zhao, X. Co3O4/TiO2 Nanocomposite Formation Leads to Improvement in Ultraviolet–Visible-Infrared-Driven Thermocatalytic Activity Due to Photoactivation and Photocatalysis–Thermocatalysis Synergetic Effect. ACS Sustain. Chem. Eng. 2018, 6, 16503–16514. [Google Scholar] [CrossRef]
  37. Hafeez, H.Y.; Lakhera, S.K.; Shankar, M.V.; Neppolian, B. Synergetic improvement in charge carrier transport and light harvesting over ternary InVO4-g-C3N4/rGO hybrid nanocomposite for hydrogen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 7530–7540. [Google Scholar] [CrossRef]
  38. Zhao, X.; Zhang, Y.-L.; Wang, X.-X.; Shi, H.-L.; Wang, W.-Z.; Cao, M.-S. Enhanced microwave absorption properties of NiFe2O4 nanocrystal deposited reduced graphene oxides. J. Mater. Sci. Mater. Electron. 2016, 27, 11518–11523. [Google Scholar] [CrossRef]
  39. Dang, M.T.; Lefebvre, J.; Wuest, J.D. Recycling Indium Tin Oxide (ITO) Electrodes Used in Thin-Film Devices with Adjacent Hole-Transport Layers of Metal Oxides. ACS Sustain. Chem. Eng. 2015, 3, 3373–3381. [Google Scholar] [CrossRef]
  40. Yang, H.; Wang, P.; Zhang, J.; Zhang, L.; Yan, J. Microwave hydrothermal synthesis of SbVO4 nanospheres as anode materials for sodium ion batteries. Ionics 2020, 26, 1267–1273. [Google Scholar] [CrossRef]
  41. McInnes, A.; Plant, S.R.; Ornelas, I.M.; Palmer, R.E.; Wijayantha, K.G.U. Enhanced photoelectrochemical water splitting using oxidized mass-selected Ti nanoclusters on metal oxide photoelectrodes. Sustain. Energy Fuels 2017, 1, 336–344. [Google Scholar] [CrossRef] [Green Version]
  42. Rajan, A.; Sharma, M.; Sahu, N.K. Assessing magnetic and inductive thermal properties of various surfactants functionalised Fe3O4 nanoparticles for hyperthermia. Sci. Rep. 2020, 10, 15045. [Google Scholar] [CrossRef] [PubMed]
  43. Li, L.; Ma, P.; Hussain, S.; Jia, L.; Lin, D.; Yin, X.; Lin, Y.; Cheng, Z.; Wang, L. FeS2/carbon hybrids on carbon cloth: A highly efficient and stable counter electrode for dye-sensitized solar cells. Sustain. Energy Fuels 2019, 3, 1749–1756. [Google Scholar] [CrossRef]
  44. Cheng, M.; Fan, H.; Song, Y.; Cui, Y.; Wang, R. Interconnected hierarchical NiCo2O4 microspheres as high-performance electrode materials for supercapacitors. Dalton Trans. 2017, 46, 9201–9209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Gao, Y.; Mi, L.; Wei, W.; Cui, S.; Zheng, Z.; Hou, H.; Chen, W. Double Metal Ions Synergistic Effect in Hierarchical Multiple Sulfide Microflowers for Enhanced Supercapacitor Performance. ACS Appl. Mater. Interf. 2015, 7, 4311–4319. [Google Scholar] [CrossRef] [PubMed]
  46. 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]
  47. Sankaranarayanan, R.; Shailajha, S.; Mubina, M.S.K.; Anilkumar, C.P. Effect of Ni2+ and Fe3+ Ion Concentrations on Structural, Optical, Magnetic, and Impedance Response of NiFe2O4 Nanoparticles Prepared by Sol-Gel Process. J. Supercond. Novel Magn. 2020, 33, 3631–3642. [Google Scholar] [CrossRef]
  48. Kite, S.V.; Kadam, A.N.; Sathe, D.J.; Patil, S.; Mali, S.S.; Hong, C.K.; Lee, S.W.; Garadkar, K.M. Nanostructured TiO2 Sensitized with MoS2 Nanoflowers for Enhanced Photodegradation Efficiency toward Methyl Orange. ACS Omega 2021, 6, 17071–17085. [Google Scholar] [CrossRef]
  49. Faisal, M.; Iqbal, A.; Adam, F.; Jothiramalingam, R. Effect of Cu doping on the photocatalytic activity of InVO4 for hazardous dye photodegradation under LED light and its mechanism. Water Sci. Technol. 2021, 84, 576–595. [Google Scholar] [CrossRef]
  50. Babu, B.; Koutavarapu, R.; Shim, J.; Yoo, K. SnO2 quantum dots decorated NiFe2O4 nanoplates: 0D/2D heterojunction for enhanced visible-light-driven photocatalysis. Mater. Sci. Semicond. Process. 2020, 107, 104834. [Google Scholar] [CrossRef]
  51. Zhang, X.; Zhang, J.; Yu, J.; Zhang, Y.; Cui, Z.; Sun, Y.; Hou, B. Fabrication of InVO4/AgVO3 heterojunctions with enhanced photocatalytic antifouling efficiency under visible-light. Appl. Catal. B 2018, 220, 57–66. [Google Scholar] [CrossRef]
  52. Shen, J.; Yang, H.; Feng, Y.; Cai, Q.; Shen, Q. Synthesis of 3D hierarchical porous TiO2/InVO4 nanocomposites with enhanced visible-light photocatalytic properties. Solid State Sci. 2014, 32, 8–12. [Google Scholar] [CrossRef]
  53. Sandra, C.D.; Alberto, E.G.; Ricardo, G. Heterojunction formation on InVO4/N-TiO2 with enhanced visible light photocatalytic activity for reduction of 4-NP. Mater. Sci. Semicond. Process. 2019, 89, 201–211. [Google Scholar] [CrossRef]
  54. Zhang, X.; Zhang, J.; Yu, J.; Zhang, Y.; Yu, F.; Jia, L.; Tan, Y.; Zhu, Y.; Hou, B. Enhancement in the photocatalytic antifouling efficiency over cherimoya-like InVO4/BiVO4 with a new vanadium source. J. Colloid Interface Sci. 2019, 533, 358–368. [Google Scholar] [CrossRef]
  55. Ma, D.; Zhang, Y.; Gao, M.; Xin, Y.; Wu, J.; Bao, N. RGO/InVO4 hollowed-out nanofibers: Electrospinning synthesis and its application in photocatalysis. Appl. Surf. Sci. 2015, 353, 118–126. [Google Scholar] [CrossRef]
  56. Rahman, K.H.; Kar, A.K. Influence of catalyst loading on photocatalytic degradation efficiency of CTAB-assisted TiO2 photocatalyst towards methylene blue dye solution. Bull. Mater. Sci. 2022, 45, 18. [Google Scholar] [CrossRef]
  57. Grubač, Z.; Katić, J.; Metikoš-Huković, M. Energy-Band Structure as Basis for Semiconductor n-Bi2S3/n-Bi2O3 Photocatalyst Design. J. Electrochem. Soc. 2019, 166, H433–H437. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of (a) InVO4, (b) NiFe2O4, (c) IVNF-2.5, (d) IVNF-5.0, and (e) IVNF-7.5 nanocomposites.
Figure 1. X-ray diffraction patterns of (a) InVO4, (b) NiFe2O4, (c) IVNF-2.5, (d) IVNF-5.0, and (e) IVNF-7.5 nanocomposites.
Catalysts 12 01471 g001
Figure 2. Scanning electron microscopy images of (a,b) InVO4 nanosheets, (c,d) NiFe2O4 nanoplates, (e,f) IVNF-5.0 nanocomposite, and (g) energy-dispersive spectroscopy (EDS) analysis of the IVNF-5.0 nanocomposite.
Figure 2. Scanning electron microscopy images of (a,b) InVO4 nanosheets, (c,d) NiFe2O4 nanoplates, (e,f) IVNF-5.0 nanocomposite, and (g) energy-dispersive spectroscopy (EDS) analysis of the IVNF-5.0 nanocomposite.
Catalysts 12 01471 g002
Figure 3. Transmission electron microscopy (TEM) images of (a) InVO4, (b) NiFe2O4, and (c) IVNF-5.0; (d) high-resolution TEM image; (e) lattice fringes and (f) selected area electron diffraction patterns of the IVNF-5.0 nanocomposite.
Figure 3. Transmission electron microscopy (TEM) images of (a) InVO4, (b) NiFe2O4, and (c) IVNF-5.0; (d) high-resolution TEM image; (e) lattice fringes and (f) selected area electron diffraction patterns of the IVNF-5.0 nanocomposite.
Catalysts 12 01471 g003
Figure 4. EDS elemental color mapping of the IVNF-5.0 nanocomposite. (a) Mapping region, (b) In, (c) V, (d) Ni, (e) Fe, and (f) O.
Figure 4. EDS elemental color mapping of the IVNF-5.0 nanocomposite. (a) Mapping region, (b) In, (c) V, (d) Ni, (e) Fe, and (f) O.
Catalysts 12 01471 g004
Figure 5. (a) Survey, (b) In 3d, (c) V 2p, (d) O 1s, (e) Fe 2p, and (f) Ni 2p high resolution X-ray photoelectron spectra of the IVNF-5.0 nanocomposite.
Figure 5. (a) Survey, (b) In 3d, (c) V 2p, (d) O 1s, (e) Fe 2p, and (f) Ni 2p high resolution X-ray photoelectron spectra of the IVNF-5.0 nanocomposite.
Catalysts 12 01471 g005
Figure 6. (a) Optical absorption spectra and (bd) Tauc plots of the synthesized samples.
Figure 6. (a) Optical absorption spectra and (bd) Tauc plots of the synthesized samples.
Catalysts 12 01471 g006
Figure 7. Photoluminescence spectra of the as-prepared samples.
Figure 7. Photoluminescence spectra of the as-prepared samples.
Catalysts 12 01471 g007
Figure 8. (a) Absorbance spectra for the degradation of tetracycline (TC) over the IVNF-5.0 nanocomposite and (b) degradation efficiency of the as-prepared samples.
Figure 8. (a) Absorbance spectra for the degradation of tetracycline (TC) over the IVNF-5.0 nanocomposite and (b) degradation efficiency of the as-prepared samples.
Catalysts 12 01471 g008
Figure 9. (a) Pseudo-first-order kinetics and (b) kinetic rate constants of the as-prepared catalysts.
Figure 9. (a) Pseudo-first-order kinetics and (b) kinetic rate constants of the as-prepared catalysts.
Catalysts 12 01471 g009
Figure 10. (a) Catalyst loading and (b) stability test over the IVNF-5.0 nanocomposite for the degradation of TC.
Figure 10. (a) Catalyst loading and (b) stability test over the IVNF-5.0 nanocomposite for the degradation of TC.
Catalysts 12 01471 g010
Figure 11. (a) Radical quenching experiment for the degradation of TC over IVNF-5.0 nanocomposite, and (b) I–t curves of the samples under the ON–OFF states.
Figure 11. (a) Radical quenching experiment for the degradation of TC over IVNF-5.0 nanocomposite, and (b) I–t curves of the samples under the ON–OFF states.
Catalysts 12 01471 g011
Figure 12. Photocatalytic degradation mechanism of TC molecules by the InVO4/NiFe2O4 nanocomposite under visible light.
Figure 12. Photocatalytic degradation mechanism of TC molecules by the InVO4/NiFe2O4 nanocomposite under visible light.
Catalysts 12 01471 g012
Table 1. A comparison of photocatalytic removal of pollutants under visible light using InVO4-based nanocomposites with recent examples from the literature.
Table 1. A comparison of photocatalytic removal of pollutants under visible light using InVO4-based nanocomposites with recent examples from the literature.
PhotocatalystSynthesis MethodCatalyst Loading mg/mLRemoval DyeRemoval Time (min)Removal Efficiency (%)Ref.
InVO4/B12O15Cl6Hydrothermal0.50TC9078.85[30]
InVO4/CeVO4Electrospinning0.40TC9092.40[46]
InVO4/AgVO3Ion exchange0.25RhB20099.99[51]
InVO4/TiO2Hydrothermal0.50RhB18055.30[52]
InVO4/ZnFe2O4Hydrothermal0.20MB24088.74[28]
InVO4/N-TiO2Sol-gel0.75NP9089.00[53]
In2S3/InVO4Anion-exchange0.50TC6071.40[31]
InVO4/BiVO4Hydrothermal0.50RhB28099.78[54]
rGO/InVO4Electrospinning0.50RhB12093.00[55]
InVO4/NiFe2O4Hydrothermal0.15TC9696.68Present study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sreeram, N.; Aruna, V.; Koutavarapu, R.; Lee, D.-Y.; Shim, J. Novel Indium Vanadium Oxide Nanosheet-Supported Nickel Iron Oxide Nanoplate Heterostructure for Synergistically Enhanced Photocatalytic Degradation of Tetracycline. Catalysts 2022, 12, 1471. https://doi.org/10.3390/catal12111471

AMA Style

Sreeram N, Aruna V, Koutavarapu R, Lee D-Y, Shim J. Novel Indium Vanadium Oxide Nanosheet-Supported Nickel Iron Oxide Nanoplate Heterostructure for Synergistically Enhanced Photocatalytic Degradation of Tetracycline. Catalysts. 2022; 12(11):1471. https://doi.org/10.3390/catal12111471

Chicago/Turabian Style

Sreeram, N., V. Aruna, Ravindranadh Koutavarapu, Dong-Yeon Lee, and Jaesool Shim. 2022. "Novel Indium Vanadium Oxide Nanosheet-Supported Nickel Iron Oxide Nanoplate Heterostructure for Synergistically Enhanced Photocatalytic Degradation of Tetracycline" Catalysts 12, no. 11: 1471. https://doi.org/10.3390/catal12111471

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