Next Article in Journal
Green Synthesis of Surface Modified Biochar for Simultaneous Removal of Steroidal Hormones and Heavy Metals from Wastewater: Optimisation by Central Composite Design
Previous Article in Journal
Provenance Indication of Rare Earth Elements in Lake Particulates from Environmentally Sensitive Regions
Previous Article in Special Issue
Synthesis of Sulfur-Doped Magnetic Iron Oxides for Efficient Removal of Lead from Aqueous Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Visible Light-Driven Photocatalytic Degradation of Tetracycline Using p-n Heterostructured Cr2O3/ZrO2 Nanocomposite

1
School of Civil Engineering and Architecture, Anhui Polytechnic University, Wuhu 241000, China
2
School of Chemical and Environmental Engineering, Anhui Polytechnic University, Wuhu 241000, China
3
Jiangsu Academy of Environmental Industry and Technology Corp, Nanjing 210036, China
*
Authors to whom correspondence should be addressed.
Water 2023, 15(20), 3702; https://doi.org/10.3390/w15203702
Submission received: 23 September 2023 / Revised: 17 October 2023 / Accepted: 19 October 2023 / Published: 23 October 2023
(This article belongs to the Special Issue Potential of Nanomaterials for Efficient Wastewater Treatment)

Abstract

:
Antibiotic pollution beyond the safety limits poses a significant threat to the environmental sustainability and human health which necessitates the development of efficient methods for reducing antibiotics in pharmaceutical wastewater. Photocatalysis is a proven technology which has drawn considerable attention in semiconductor photocatalysts. Our study aims to develop a highly efficient Cr2O3/ZrO2 photocatalyst for the degradation of tetracycline (TCL) under visible light. The synthesized catalyst was well characterized by XRD, HR-TEM-SAED, XPS, FT-IR, BET and UV-Vis-DRS methods. The effects of various parameters on photocatalytic degradation were evaluated in detail, showing that 97.1% of 50 mgL−1 tetracycline concentrations could be degraded within 120 min at pH 5 with a 0.1 gL−1 photocatalyst-loading concentration under visible light (300 W Xe lamp). The uniform distribution of spherical ZrO2 nanoparticles on the surface of the Cr2O3 nano-cubes efficiently reduced the recombination rate with an energy bandgap of 2.75 eV, which provided a faster photodegradation of tetracycline under visible light. In addition, a plausible degradation pathway and photoproducts generated during the photocatalytic degradation of TCL are proposed based on the LC-ESI/MS results, which suggested that efficient photodegradation was achieved during the visible light irradiation. Thus, our study reveals that the cost-effective Cr2O3-based photocatalyst with multi-reusability and efficient energy consumption could be an efficient photocatalyst for the rapid degradation of TCL during the wastewater treatment process.

1. Introduction

In the recent past, ecological imbalance has become a major environmental issue caused by the contamination of surface water due to rapid industrial development and the littering of household waste [1]. Several studies have demonstrated that chemical compounds released from pharmaceutical industries are known to be highly toxic and hazardous to aquatic ecosystems when they are left untreated [2,3]. In this context, tetracycline (TCL) belongs to the group of phenanthrene core antibiotics that are widely used to treat Gram-positive and Gram-negative bacteria, growth promoters for animals, intracellular mycoplasma, chlamydia, and rickettsia infections. Consequently, more than 50% of these antibiotics end up in the environment after they have been consumed by humans and animals [4]. The release of these antibiotic residues and their metabolites into the aquatic environment contributes to the rise in antibiotic resistance genes in the aquatic life and has had a significant influence on the environment. The high hydrophilicity of these compounds not only makes them biologically toxic for aquatic species but also enriches and transmits them through the food chain, resulting in serious health concerns for humans [5]. Due to the higher concentrations of these contaminates, clean water has also become scarce for humans and the environment [6]. Globally, the current unsafe situation has become a major concern, and it is expected to become worse as time goes on [7]. Thus, developing low-cost and efficient wastewater treatment technologies is urgently needed to combat this growing problem [8,9]. However, adsorption and coagulation processes are considered ineffective for conventional wastewater treatment systems due to their high operational costs and significant energy demands [10,11]. Furthermore, these existing water treatment methods are largely unfruitful due to high salinity and secondary pollutants [12].
The photocatalytic oxidation process is a well-known technology and a viable alternative to conventional wastewater treatment [13]. Through the usage of solar light as the source of energy and high-activity metal-oxide nanoparticles as the processing input, it is capable of degrading organic pollutants such as tetracycline in a sustainable manner [14]. Semiconductor photocatalysis which utilizes solar energy has drawn significant attention in the global scientific community due to its great potential in the search for better solutions to current environmental and energy issues [15]. This has led to the development of several nanomaterials like g-C3N4, FeAl2O4, Cu2O, Bi2WO6, CdS-N/ZnO, etc., for the photocatalytic degradation of various organic pollutants under visible light [16,17,18,19,20]. However, all these methods have limitations, such as high costs, slow degradation rates, UV-active materials, and reusability issues. Hence, in recent years, there has been increasing interest in the development of simple, cost-effective p-n heterojunction semiconductors that can be used as solar-harvesting photocatalysts to degrade TCL pollutants [21].
Following this, Cr2O3 is a widely recognized flagship p-type semiconducting material that can be used for photocatalysis enhancement. Furthermore, the strategic selection of Cr2O3 is based on a waste-to-resources conversion strategy, in which carcinogenic hexavalent chromium compounds that are utilized in various industrial applications can be converted into benign trivalent chromium oxide that can be used as a photocatalyst for environmental remediation. Thus, Cr2O3 has been envisioned as an alternative photoactive material through a modified method of coupling it with a stable metal oxide. For instance, Singh et al. proposed a new Fe3O4-Cr2O3 magnetic nanocomposite prepared using a simple wet chemical method for degradation of 4-chlorophenol under UV light [22]. Similarly, Ahmed et al. used sol–gel synthesis to prepare spherical mesoporous Cr2O3-TiO2 nanoparticles, and with UV light exposure, the photocatalyst exhibited 90.0% efficiency in removing methylene blue dye [23]. Consequently, due to its high rate of electron–hole recombination, Cr2O3 is widely employed as a co-catalyst in conjunction with other semiconducting materials. Hence, to overcome these drawbacks and make Cr2O3 a visible active photocatalyst, it must be coupled to a semiconducting material with a slightly higher bandgap [24]. As an n-type semiconductor, ZrO2 is becoming a widely used material in photocatalysis [25]. The unique properties of ZrO2 that include high refractive index, high thermal expansion, high optical transparency, chemical and photothermal stability, excellent corrosion resistance, and low thermal conductivity have led to its extensive use in photocatalysis in recent years [26]. However, its large bandgap of 3.87 eV and high electron–hole recombination rate makes ZrO2’s photocatalytic activity limited to the UV range [27]. Interestingly, several recent studies have demonstrated that the combination of various other metal oxides with ZrO2 to form a heterojunction would eventually increase its visible-light-active efficiency during the degradation process [28].
In response to these observations, we synthesized a stable and efficient heterostructure Cr2O3/ZrO2 catalyst via the coprecipitation method. The photocatalyst ensures ≥ 97.1% TCL degradation led by hydroxyl and superoxide radical species. The superior analytical performance and reusability of the proposed nanocomposite materials make them an attractive choice for environmental remediation applications in the antibiotic wastewater treatment process.

2. Materials and Methods

2.1. Synthesis of Cr2O3 Nano-Cubes

Firstly, to synthesize pure chromium oxide cubes, chromium (III) nitrate nonahydrate (≥99.99%; Sigma Aldrich, Shanghai, China) was used as the precursor. In the initial process, the Cr starting material (0.1 M) was mixed constantly for about 0.5 h, followed by the dropwise addition of ammonia (0.3 M) to control the pH 6.0, to obtain a green-colored precipitate. Then, the precipitate was separated and washed several times with deionized water in order to remove any soluble impurities. After obtaining chromium hydroxide precipitate, it was dried for 24 h at 90 °C followed by calcination at 900 °C for 5 h to obtain the chromium oxide cubes with an intense green color.

2.2. Preparation of Cr2O3-ZrO2 Nanocomposite

To prepare the Cr2O3/ZrO2 nanocomposite, the co-precipitation approach was employed. Initially, zirconium (IV) tetra-butoxide (≥99.99%; Sigma Aldrich, Shanghai, China) precursor was mixed with 15 mL solution of H2O2 (30% w/v), followed by stirring for 0.5 h, resulting in a white-colored colloidal peroxo complex of Zirconium. To the above suspension, 50 mL of distilled water and 0.25 g of previously synthesized Cr2O3 cubes were added, and the suspension was stirred for 24 h. Further, the complete suspension was then transferred to a Teflon-lined autoclave maintained at 100 °C for 12 h. After bringing the autoclave to room temperature, the catalyst was washed several times with water, dried at 80 °C, followed by calcination at 500 °C for 2 h at a ramp rate of 5 °C/min. The final powder was denoted as a Cr2O3/ZrO2 nanocomposite. For comparison, the pure ZrO2 nanoparticles were synthesized by following the same above-mentioned method without addition of Cr2O3 cubes.

2.3. Instrumentations

Several instruments have been used to characterize the as-synthesized photocatalytic materials. The wide angel XRD pattern for the synthesized photocatalyst materials were recorded using a BRUKER D8 Advance X-ray diffractometer (Karlsruhe, Germany) with 2θ ranges from 10° to 90°. The structural pattern and surface morphology of the prepared nanocomposite materials was studied using HR-TEM-SAED (Thermo Fisher Talos F200S G2 electron microscopes; Pleasanton, CA, USA). The energy band gap measurements for the prepared nanocomposites were calculated using a UV-Visible spectrophotometer equipped with diffused reflectance integrating sphere (DRS) (Shimadzu UV-360; Kyoto, Japan). An X-ray photoelectron spectrometer (XPS) (Thermo Scientific K-AlphaX; Pleasanton, CA, USA) with monochromatic Al Kα radiation (150 W, 15 kV, and 1486.74 eV) was used to study the oxidation states and surface composition of the prepared photocatalyst materials. FT-IR spectral measurements were carried out using a Thermo Scientific (Pleasanton, CA, USA) Nicolet iS10 model. BET analysis was carried out using a Kubo-X1000 high-performance micropore analyzer (Bbbm024; Shenzhen, China).

2.4. Photocatalytic Methodology

Photocatalytic activity of the Cr2O3/ZrO2 catalyst was assessed in terms of degradation of TCL. The experiment was conducted in a laboratory-scale photocatalytic chamber equipped with a 300 W Xe lamp (simulated solar light). Initially, 100 mg of the photocatalyst was suspended in 100 mL of an aqueous solution containing 50 mgL−1 TCL. To eliminate the background reductions (adsorption/desorption) in TCL drug concentrations, the tubes containing TCL solutions and photocatalyst material were stirred in dark conditions for 20 min. The degradation efficiency of the samples was measured by collecting 2.5 mL aliquots at fixed intervals during irradiation. Before HPLC analysis, 2 mL of the sample was centrifuged, then filtered through 0.2 mm syringe filters. To monitor the decrease in TCL concentration, the withdrawn samples were analyzed using HPLC at its characteristic absorbance of 354 nm with an UV detector, while acetonitrile (30% v/v) and water (70% v/v) were used as the mobile phase at a flow rate of 1 mL min−1. Meanwhile, the HPLC that connected to a tandem mass spectrophotometer (LC-ESI/MS/MS) was used to identify the degradation products during degradation. The ESI method was used in identifying the mass, helium gas was utilized at a flow rate ~1 mL min−1, and a 16 V of fragment voltage was maintained during analysis. We followed the same procedure under visible light irradiation to optimize various physiochemical parameters that may influence degradation efficiency.

3. Results and Discussion

3.1. Characterization

To determine the crystal structure and phase orientation of the prepared nanomaterials, XRD analysis was performed. Figure 1 illustrates the wide range of XRD patterns of Cr2O3, ZrO2, and Cr2O3/ZrO2 NCs materials as well as their JCPDS card patterns. The XRD diffraction pattern of the synthesized Cr2O3 nanoparticles observed at respective 2θ values and the corresponding indexed planes are 2θ = 24.80 (0 1 2), 33.82 (1 0 4), 35.80 (1 1 0), 41.17 (0 0 6), 42.73 (1 1 3), 45.53 (2 0 2) 51.43 (0 2 4), 56.30 (1 1 6), 59.41 (0 1 8), 62.80 (2 1 4), and 64.95 (3 0 0). These reflection planes indicated that the synthesized nanoparticles have a rhombohedral phase with an R3c space group, in accordance with the JCPDS Card No. 85-0869. In the same manner, the diffraction peaks of the pure ZrO2 NPs reveal the peaks with the following values 2θ = 28.8 (−11.1), 32.01 (−11.1), 50.10 (12.2), and 61.76 (−13.1) indicating that the ZrO2 is in monoclinic phase (JCPDS Card No. 37-1484). As a result of coupling the ZrO2 with Cr2O3, no significant change in the peak position of the Cr2O3 lattice planes was observed, and both the rhombohedral and monoclinic distribution were clearly visible in the Cr2O3-ZrO2 nanocomposite, indicating that ZrO2 had no effect on the Cr2O3 matrix. This observation confirms that ZrO2 was deposited only on the Cr2O3 NP surface, possibly because Zr4+ has a larger ionic radius than Cr3+. Furthermore, peaks corresponding to ZrO2 and Cr2O3 are clearly observed on the nanocomposite, indicating that no other detectable impurities are present. In addition, the Debye–Scherrer equation D = Kλ/β cos θ was applied to further determine the average crystalline size for the nanocomposite, in which β represents the full width at half maximum (FWHM) of the diffraction peak, θ indicates the angle between the incident and diffracted beams, λ specifies the wavelength of X-ray beam (1.540 Å for CuKα), and K is the shape factor (0.9). The crystalline size of the prepared composite has been carried out using the major intense peaks and their average size has been calculated as mentioned above. The crystalline size of the Cr2O3, ZrO2, and Cr2O3/ZrO2 nanocomposite was found to be around 30.91, 25.84, and 27.50, respectively.
The light absorption properties of the as-prepared catalysts were investigated using diffuse reflectance UV-Vis spectroscopy (UV-Vis-DRS). The spectral results indicated that Cr2O3 enhanced ZrO2 absorption in the visible region, with a slight redshift in the absorption edge. This indicates that the Cr2O3/ZrO2 heterostructure photocatalysts have effective activity in the visible range. Due to its octahedral geometrical preferences, the absorption spectra for pure Cr2O3 NPs exhibits three major peaks. This is the reason for the curved wave pattern in Tauc’s plot of the corresponding band gap energy of the Cr2O3 NPs. The band gap energy of the nanocomposite (Cr2O3/ZrO2) was calculated using Tauc’s plot using the formula: αhν = A (hν − Eg) n. Where A is constant, Eg represents band gap energy, the exponent n = ½ for direct band gap transition, hν is photon energy, and α is the attenuation constant. According to the results, the energy band gaps of pristine Cr2O3, ZrO2, and Cr2O3/ZrO2 are found to be 2.12, 3.81, and 2.75 eV, respectively (Figure 2a). As a result of these increased absorptions as well as the reduced band gap energy in ZrO2, the photocatalyst obtained higher efficiency under the visible light illumination.
The presence of multiple functional groups in pure Cr2O3, ZrO2, and Cr2O3/ZrO2 nanocomposites was confirmed by FT-IR analysis (Figure 2b). From the results, it shows that the strong vibration at 540 cm−1 was assigned to pristine Cr2O3 that indicates the presence of a Cr-O bond, while the strong vibration at 620 cm−1 was ascribed to the Cr2O3 crystalline peak. Similarly, the strong vibration peak at 530 cm−1 could be assigned to the Zr-O bond in pristine ZrO2. As a result of the intervening peak at 530 cm−1 of ZrO2 in Cr2O3/ZrO2 nanocomposites, a peak broadening was observed for the twin peaks (540 cm−1 and 620 cm−1) of Cr2O3. This further supports the formation of Cr2O3/ZrO2 nanocomposites. A sharp band that appeared on all three peaks at 1620 cm−1 may be related to the bending mode of hydroxyl groups from the surface-adsorbed molecules [29,30,31]. As a result of the formation of heterostructures, atmospheric CO2 will react with the surface hydroxyl groups of Cr2O3/ZrO2, which will result in the consumption of hydroxyl groups at vibrational band at 3450 cm−1. These FTIR results are in complete agreement with our XRD results.
The morphological characteristics of the synthesized composite were investigated through HR-TEM. The results clearly evidenced that Cr2O3 are in nano-cubes (Figure 3a), while ZrO2 are in slightly agglomerated particles (Figure 3b). It is evidenced that there is a formation of a Cr2O3/ZrO2 heterojunction during the coprecipitation process as it can be seen from the images that the Cr2O3 nano-cubes are clearly visible and the ZrO2 nanoparticles are agglomerated with equal distribution under Cr2O3 nano-cubes (Figure 3c). In addition, the average particle size of Cr2O3, ZrO2, and Cr2O3/ZrO2 NCs was calculated, and the corresponding values are about 36.5 nm, 27.5 nm, and 30.5 nm, respectively [10,32]. Additionally, the high-magnification TEM images clearly revealed that the lattice fringes of Cr2O3 and ZrO2 with d-spacing values are about 0.332 and 0.237 nm (Figure 3d,e). Likewise, the SAED analysis of Cr2O3/ZrO2 composite confirms the microcrystalline phase and the indexed planes match the p-XRD analysis (Figure 3f). With these morphological features, it could be suggested that better adsorption and diffusion of TCL analytes are achievable, which enables effective removal under visible light.
Surface composition, purity, and elemental analysis were carried out by X-ray photoelectron spectroscopy (XPS) for the as-prepared composite catalyst. From the results, it can be seen from the survey scan spectrum of the Cr2O3/ZrO2 nanocomposite which showed all the elements of the composite and confirms their uniform distribution in the composite (Figure 4a). Accordingly, the Cr2O3/ZrO2 nanocomposite exhibited high-intensity peaks at 180.32 eV, 580.25 eV, and 530.87 eV, ascribed to Zr 3d, Cr 2P, and O 1s, respectively. It is worth mentioning that the absence of the N 1s peak suggests that the NO3− ion has been successfully removed from the precursor Cr(NO3)9H2O during the synthesis and calcination process. The high-resolution Cr 2p spectrum (Figure 4b) shows two well-resolved peaks at 574.6 and 584.7 eV corresponding to Cr 2p3/2 and Cr 2p1/2 peaks, respectively. In addition, the energy separation of 9.8 eV between the 2p3/2 and 2p1/2 peaks indicated that the Cr 2p orbital contains chromium in a trivalent state (Cr3+). Likewise, the deconvoluted Zr 3d orbital (Figure 4c) shows distinct states of Zr 3d3/2 and Zr 3d5/2, with binding energies of 198.8 eV and 182.1 eV, respectively, indicating an oxidation state of Zr4+. Furthermore, the O 1s high resolution spectrum (Figure 4d) displays a former peak at 529.6 eV, that is assigned to the lattice oxygen of the composite, while the other intense peak at 532.3 eV could be ascribed to the OH species that result from the chemisorbed water [33,34,35,36]. According to the XPS results, the atomic percentages of the nanocomposite elements are found to be chromium (Cr3+)—25.5%, zirconium (Zr4+)—12.3%, oxygen (O)—41.7% and carbon (C)—20.5%, confirming the uniform distribution of the elements and the formation of a heterostructure between them.
In addition, the BET and pore size distribution analysis were conducted on the prepared samples to determine the surface area and pore characteristics (Figure 5). The surface area of the bare Cr2O3 was found to be about 23.83 m2/g with a pore volume and pore diameter of 0.033 cm3/g and 4.05 nm, respectively. Similarly, the BET surface area of pristine ZrO2 nanoparticles was found to be about 14.74 m2/g with a pore volume and diameter of 0.0012 cm3/g and 2.90 nm, respectively. However, the Cr2O3/ZrO2 nanocomposite exhibits an increased surface area of 38.43 m2/g as well as a mid-way pore volume of 0.025 cm3/g and pore diameter of 4.35 nm. The increased surface area and pore properties of the Cr2O3/ZrO2 nanocomposite confirm that ZrO2 nanoparticles have been dispersed onto the surface of Cr2O3 nanoparticles rather than intercalated into the lattice (Figure 5).

3.2. Catalytic Degradation of TCL and Pathway of the Cr2O3-ZrO2 Composite

The photocatalytic degradation of TCL via the as-prepared Cr2O3-ZrO2 composite has been carried out on a laboratory-designed photoreactor (Figure 6). Firstly, the effect of pH, in which three different values of pH 5 (acidic), pH 7 (neutral), and pH 9 (basic) were investigated to understand its influence on TCL degradation. The results indicated that all the pH levels showed higher efficiency in TCL degradation while the acidic pH showed the highest efficiency in TCL degradation after 120 min under visible light. The results indicated that the basic pH has significantly reduced the TCL degradation because at the basic pH the catalyst surface would be negatively charged, hence the unstable negatively charged TCL molecules would easily repel by the catalyst, thus the degradation has been decreased (Figure 6a). However, the acidic pH is more favorable for TCL degradation in which the surface attraction of the TCL molecule and the positively charged catalyst surface are higher to provide active sites and generate highly reactive organic species to degrade the TCL molecule under visible light [37,38]. To demonstrate the surface charge of the Cr2O3/ZrO2 nanocomposite, the zeta potential analysis at various pH levels was measured to elucidate the isoelectric point of the photocatalyst, and, as shown in Figure S1, the composite has the isoelectric point at pH 6.22. The enhanced photocatalytic degradation of TCL at pH 5.0 is attributed to the strong electrostatic interaction of the positively charged surface of the Cr2O3/ZrO2 photocatalyst, with the anionic tetracycline drug molecule. Beyond the isoelectric point (>6.22), the decline in photocatalytic drug degradation efficiency is attributed to the electrostatic repulsion between the negatively charged photocatalyst surface and the anionic tetracycline drug molecules.
In addition, the effect of catalyst load on the degradation of TCL has also been analyzed during photodegradation. From the results, it indicated that in the dark, there is no significant reduction in TCL concentration due to physical adsorption on the catalyst. Meanwhile, the degradation efficiency has been increased with the increase in the amount of the catalyst from 0.25 g/L to 1.0 g/L. However, subsequently, the efficiency has declined with the increase in catalyst load to 2.0 g/L (Figure 6b). This could be because with the increase in the catalyst load after the optimum ratio, the viscosity of the suspension would increase and subsequently decreases the light penetration onto the catalyst. This could lead to a reduction in light availability for the active sites of the catalyst, which could significantly decrease the degradation efficiency [39]. Thus, it was found that 1.0 g/L of the catalyst was the optimum catalyst load for the enhanced TCL degradation under visible light.
From the above results, it is observed that at the optimum conditions the composite exhibited about 97.2% of TCL degradation in 120 min. It is worth mentioning that the pure Cr2O3 and ZrO2 showed insignificant degradation efficiency in TCL degradation (Figure 6c,d). This could be because the energy band gap of Cr2O3 is too low which leads to the absorption of the visible light more easily to generate charge carriers, however, the recombination rate is significantly higher. Similarly, ZrO2 has a slightly higher energy band gap, which is insufficient to generate photo-induced charge carriers. Thus, the composite has obtained an optimum energy band gap in which the photogenerated charge carriers would easily transfer to generate active radical species like OH•− and O2•− to degrade the TCL to a lower molecule.
Furthermore, the plausible degradation pathway of TCL by the Cr2O3/ZrO2 nanocomposite at the optimized conditions was investigated by LC-ESI/MS analysis (Figure 7). The results showed that there could reasonably be three sites to be attacked by the reactive oxygen species (OH•− or O2•−). Initially, the hydroxyl radical attacks the amine group and breaks the bond to form DP-1 (m/z = 430); similarly, in the other way, it could be attacked by the super oxide anion to form DP-2 (m/z = 490). Furthermore, the degradation products undergo effective attack by the hydroxyl radical to form hydroxylated compounds (DP-3, DP-4, and DP-5). Eventually, after formation of hydroxylated compounds, they further undergo an effective attack by both the hydroxyl and superoxide anion radicals to form ring opening structures (DP-8 to DP-12), which would further mineralize to generate smaller mineral acids, H2O and CO2.

3.3. Photocatalysis Degradation Mechanism

To understand the possible reactive oxygen species involved in the degradation, a series of radical trapping and quantification experiments was carried out. The results indicated that after the addition of isopropyl alcohol (OH•− scavenger) the degradation efficiency of TCL was significantly reduced, while benzoquinone (O2•− scavenger) has less influence than isopropyl alcohol on the TCL degradation. Furthermore, the addition of triethanolamine (h+ scavenger) had a slightly lower influence on the TCL degradation (Figure 8a). These phenomena were further supported by the radical quantification results (Figure 8b,c) (Figure S2) and confirmed that the OH•− and O2•− species generation was increased with the increase in irradiation time and produced more species to be involved in TCL degradation.
Furthermore, the photocatalytic degradation mechanism of TCL by the Cr2O3/ZrO2 composite was proposed and is shown in Figure 9. It is feasible to propose a traditional heterojunction-type photocatalytic mechanism for Cr2O3 and ZrO2 since these semiconductors can form a unique p-n heterojunction structure. Upon visible light illumination, it is possible for electrons from the valence band (VB) of Cr2O3 to become excited to the conduction band (CB) of Cr2O3, creating holes. Meanwhile, the matched energy between ZrO2 and Cr2O3 allows for the migration of photogenerated electrons from Cr2O3 CB to ZrO2 CB, which produces superoxide radicals via further reaction with the dissolved oxygen. In the similar way, the holes in the VB of ZrO2 can be relocated to the VB of Cr2O3, allowing for the formation of hydroxyl radicals by the photochemical reactions. As a result, the probability of electron/hole recombination is considerably reduced, since electrons concentrate primarily on the CB of ZrO2 while holes stay in the VB of Cr2O3. As a result of the reduced electron/hole recombination and the formation of reactive oxygen species, tetracycline drug molecules are mineralized into CO2, H2O, and other non-toxic compounds.
In addition, the recycle test has also been performed by centrifuging the used catalyst at 5000 rpm for 5 min and collected, followed by drying at 60 °C before using it in the next cycle. From the results, as shown in Figure 10a, the proposed nanocomposite materials can undergo six successive cycles without losing their photocatalytic efficiency. In addition, the XRD results reveals an unaltered crystallinity along with retention of the structural characteristics of the photocatalyst even after six cycles of usage (Figure 10b). The superior performance of the proposed photocatalyst has been compared with similar type of photocatalysts reported in the literature [32,33,34,35,36,37,38,39], and the observations are described in Table 1.

4. Conclusions

A novel Cr2O3–ZrO2 nanocomposite has been successfully synthesized for the efficient degradation of the tetracycline antibiotic under visible light. Our results indicated that the ZrO2 nanoparticles were uniformly distributed on the Cr2O3 cubes which facilitates a strong interaction to form a heterojunction which further leads to improving the charge separation rate and reducing the recombination rate. Furthermore, the composite showed enhanced degradation efficiency towards TCL, which was about 97.1% after 120 min of visible light irradiation at the optimal conditions (pH = 5, catalyst load = 0.1 gL−1, and TCL concentration = 50 mgL−1). The radical trapping and the active species scavenging experimental results revealed that the OH•−, O2•−, and photogenerated holes are the main species involved in the TCL degradation. Thus, our study suggests new insights into the preparation of Cr- and Zr-based metal oxide heterojunctions with efficient visible light harvesting capabilities for degradation of antibiotic pollutants under visible light during pharmaceutical wastewater treatment processes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w15203702/s1, Figure S1: Zeta potential vs. pH for Cr2O3-ZrO2 nanocomposite; Figure S2: (a) Reaction pathway between terepthalic acid and hydroxyl radical with the formation of fluorescent 2-hydroxy terepthalic acid (b) Reaction pathway between NBT and superoxide radicals with the formation of formazan.

Author Contributions

Conceptualization, X.W. and S.N.; methodology, X.W. and S.N.; validation, X.W. and P.C.; resources, J.L. and B.L.; writing—original draft preparation, X.W.; writing—review and editing, X.W., X.Y. and S.N.; visualization, X.W.; funding acquisition, X.W. and S.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Anhui Polytechnic University Introduction of Talents Research Start-up Fund (Grant No: 2023YQQ011), Natural Science Key Foundation of Educational Commission of Anhui Province (Grant No. 2022AH050989) and Scientific Research Project of Anhui Polytechnic University (Grant No. Xjky2022170).

Data Availability Statement

Data are available from the corresponding author on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dong, Y.; Xu, D.; Zhang, J.; Wang, Q.; Pang, S.; Zhang, G.; Campos, L.C.; Lv, L.; Liu, X.; Gao, W.; et al. Enhanced antibiotic wastewater degradation by intimately coupled B-Bi3O4Cl photocatalysis and biodegradation reactor: Elucidating degradation principle systematically. J. Hazard. Mater. 2023, 445, 130364. [Google Scholar] [CrossRef] [PubMed]
  2. Yu, Y.; Hu, X.; Li, M.; Fang, J.; Leng, C.; Zhu, X.; Xu, W.; Qin, J.; Yao, L.; Liu, Z.; et al. Constructing mesoporous Zr-doped SiO2 onto efficient Z-scheme TiO2/g-C3N4 heterojunction for antibiotic degradation via adsorption-photocatalysis and mechanism insight. Environ. Res. 2022, 214, 114189. [Google Scholar] [CrossRef] [PubMed]
  3. Li, J.; Li, Y.; Zhu, M.; Mei, Q.; Tang, X.; Wu, Y.; Yue, S.; Tang, Y.; Wang, Q. Constructing aloe-emodin/FeOOH organic-inorganic heterojunction for synergetic photocatalysis-Fenton eliminating antibiotic pollutants. J. Environ. Chem. Eng. 2023, 11, 109775. [Google Scholar] [CrossRef]
  4. Sharma, M.; Mandal, M.K.; Pandey, S.; Kumar, R.; Dubey, K.K. Visible-Light-Driven Photocatalytic Degradation of Tetracycline Using Heterostructured Cu2O-TiO2 Nanotubes, Kinetics, and Toxicity Evaluation of Degraded Products on Cell Lines. ACS Omega 2022, 7, 33572–33586. [Google Scholar] [CrossRef] [PubMed]
  5. Wu, S.; Hu, H.; Lin, Y.; Zhang, J.; Hu, Y.H. Visible light photocatalytic degradation of tetracycline over TiO2. Chem. Eng. J. 2020, 382, 122842. [Google Scholar] [CrossRef]
  6. Li, C.; Tian, Q.; Zhang, Y.; Li, Y.; Yang, X.; Zheng, H.; Chen, L.; Li, F. Sequential combination of photocatalysis and microalgae technology for promoting the degradation and detoxification of typical antibiotics. Water Res. 2022, 210, 117985. [Google Scholar] [CrossRef] [PubMed]
  7. Ji, B.; Zhang, J.; Zhang, C.; Li, N.; Zhao, T.; Chen, F.; Hu, L.; Zhang, S.; Wang, Z. Vertically Aligned ZnO@ZnS Nanorod Chip with Improved Photocatalytic Activity for Antibiotics Degradation. ACS Appl. Nano Mater. 2018, 1, 793–799. [Google Scholar] [CrossRef]
  8. Naraginti, S.; Yu, Y.Y.; Fang, Z.; Yong, Y.C. Novel tetrahedral Ag3PO4@N-rGO for photocatalytic detoxification of sulfamethoxazole: Process optimization, transformation pathways and biotoxicity assessment. Chem. Eng. J. 2019, 375, 122035. [Google Scholar] [CrossRef]
  9. Tan, R.; Wang, Y.; Jin, Z.; Zhang, P.; Luo, H.; Liu, D.; Mamba, B.B.; Kuvarega, A.T.; Gui, J. Preparation of carbon-coated brookite@anatase TiO2 heterophase junction nanocables with enhanced photocatalytic performance. Photochem. Photobiol. Sci. 2020, 19, 966–975. [Google Scholar] [CrossRef]
  10. Sompalli, N.K.; Mohanty, A.; Mohan, A.M.; Deivasigamani, P. Heterojunction Cr2O3-Ag2O nanocomposite decorated porous polymer monoliths a new class of visible light fast responsive heterogeneous photocatalysts for pollutant clean-up. J. Environ. Chem. Eng. 2021, 9, 104846. [Google Scholar] [CrossRef]
  11. Sompalli, N.K.; Das, A.; De, S.S.; Mohan, A.M.; Deivasigamani, P. Mesoporous monolith designs of mixed phased titania codoped Sm3+/Er3+ composites: A super responsive visible light photocatalysts for organic pollutant clean-up. Appl. Surf. Sci. 2020, 504, 144350. [Google Scholar] [CrossRef]
  12. Yang, X.; Wang, D. Photocatalysis: From Fundamental Principles to Materials and Applications. ACS Appl. Energy Mater. 2018, 1, 6657–6693. [Google Scholar] [CrossRef]
  13. Wen, Y.; Feng, M.; Zhang, P.; Zhou, H.-C.; Sharma, V.K.; Ma, X. Metal Organic Frameworks (MOFs) as Photocatalysts for the Degradation of Agricultural Pollutants in Water. ACS ES&T Eng. 2021, 1, 804–826. [Google Scholar] [CrossRef]
  14. Li, Y.; Chen, F.; Wang, Y.; Tang, N.; He, R. Semiconductor Photocatalysis for Water Purification; Elsevier Inc.: Amsterdam, The Netherlands, 2019. [Google Scholar] [CrossRef]
  15. Elaheh, K.; Naimeh, S. The Application of Photocatalytic Materials for Efficient Air Purification; Elsevier Inc.: Amsterdam, The Netherlands, 2020. [Google Scholar] [CrossRef]
  16. Oluwole, A.O.; Olatunji, O.S. Photocatalytic degradation of tetracycline in aqueous systems under visible light irridiation using needle-like SnO2 nanoparticles anchored on exfoliated g-C3N4. Environ. Sci. Eur. 2022, 34, 5. [Google Scholar] [CrossRef]
  17. Zhou, X.; Chen, X.; Han, W.; Han, Y.; Guo, M.; Peng, Z.; Fan, Z.; Shi, Y.; Wan, S. Tetracycline Removal by Hercynite-Biochar from the Co-Pyrolysis of Red Mud-Steel Slag-Sludge. Nanomaterials 2022, 12, 2595. [Google Scholar] [CrossRef] [PubMed]
  18. Zhou, Q.; Hong, P.; Shi, X.; Li, Y.; Yao, K.; Zhang, W.; Wang, C.; He, J.; Zhang, K.; Kong, L. Efficient degradation of tetracycline by a novel nanoconfinement structure Cu2O/Cu@MXene composite. J. Hazard. Mater. 2023, 448, 130995. [Google Scholar] [CrossRef] [PubMed]
  19. Jia, K.; Liu, G.; Lang, D.N.; Chen, S.F.; Yang, C.; Wu, R.L.; Wang, W.; Wang, J.D. Fast photodegradation of antibiotics and dyes by an anionic surfactant-aided CdS/ZnO nanodispersion. New J. Chem. 2022, 46, 11303–11314. [Google Scholar] [CrossRef]
  20. He, Y.; Wang, D.; Li, X.; Fu, Q.; Yin, L.; Yang, Q.; Chen, H. Photocatalytic degradation of tetracycline by metal-organic frameworks modified with Bi2WO6 nanosheet under direct sunlight. Chemosphere 2021, 284, 131386. [Google Scholar] [CrossRef]
  21. 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]
  22. Singh, K.K.; Senapati, K.K.; Borgohain, C.; Sarma, K.C. Newly developed Fe3O4–Cr2O3 magnetic nanocomposite for photocatalytic decomposition of 4-chlorophenol in water. J. Environ. Sci. 2017, 52, 333–340. [Google Scholar] [CrossRef]
  23. Ahmed, M.A.; Abou-Gamra, Z.M.; Salem, A.M. Photocatalytic degradation of methylene blue dye over novel spherical mesoporous Cr2O3/TiO2 nanoparticles prepared by sol-gel using octadecylamine template. J. Environ. Chem. Eng. 2017, 5, 4251–4261. [Google Scholar] [CrossRef]
  24. Mohanapandian, K.; Krishnan, A. Synthesis, structural, morphological and optical properties of Cu2+ doped Cr2O3 nanoparticles. Int. J. Adv. Eng. Technol. E 2016, 7, 273–279. Available online: http://www.technicaljournalsonline.com/ijeat/VOL%20VII/IJAET%20VOL%20VII%20ISSUE%20II%20APRIL%20JUNE%202016/20167250.pdf (accessed on 16 October 2023).
  25. Rani, V.; Sharma, A.; Kumar, A.; Singh, P.; Thakur, S.; Singh, A.; Van Le, Q.; Nguyen, V.H.; Raizada, P. ZrO2-Based Photocatalysts for Wastewater Treatment: From Novel Modification Strategies to Mechanistic Insights. Catalysts 2022, 12, 1418. [Google Scholar] [CrossRef]
  26. Agorku, E.S.; Kuvarega, A.T.; Mamba, B.B.; Pandey, A.C.; Mishra, A.K. Enhanced visible-light photocatalytic activity of multi-elements-doped ZrO2 for degradation of indigo carmine. J. Rare Earths 2015, 33, 498–506. [Google Scholar] [CrossRef]
  27. Kadam, A.; Dhabbe, R.; Gophane, A.; Sathe, T.; Garadkar, K. Template free synthesis of ZnO/Ag2O nanocomposites as a highly efficient visible active photocatalyst for detoxification of methyl orange. J. Photochem. Photobiol. B Biol. 2016, 154, 24–33. [Google Scholar] [CrossRef] [PubMed]
  28. Aldeen, E.M.S.; Jalil, A.A.; Mim, R.S.; Alhebshi, A.; Hassan, N.S.; Saravanan, R. Altered zirconium dioxide based photocatalyst for enhancement of organic pollutants degradation: A review. Chemosphere 2022, 304, 135349. [Google Scholar] [CrossRef] [PubMed]
  29. Yazdi, M.N.; Yamini, Y.; Asiabi, H. Multiwall carbon nanotube-zirconium oxide nanocomposite hollow fiber solid phase microextraction for determination of polyaromatic hydrocarbons in water, coffee and tea samples. J. Chromatogr. A 2018, 1554, 8–15. [Google Scholar] [CrossRef]
  30. Ivanova, T.; Gesheva, K.; Cziraki, A.; Szekeres, A.; Vlaikova, E. Structural transformations and their relation to the optoelectronic properties of chromium oxide thin films. J. Phys. Conf. Ser. 2008, 113, 012030. [Google Scholar] [CrossRef]
  31. Ocaa, M. Nanosized Cr2O3 hydrate spherical particles prepared by the urea method. J. Eur. Ceram. Soc. 2001, 21, 931–939. [Google Scholar] [CrossRef]
  32. Jagadeesan, D.; Sompalli, N.K.; Mohan, A.M.; Rao, C.V.S.B.; Nagarajan, S.; Deivasigamani, P. ZrO2–Ag2O nanocomposites encrusted porous polymer monoliths as high-performance visible light photocatalysts for the fast degradation of pharmaceutical pollutants. Photochem. Photobiol. Sci. 2022, 21, 1273–1286. [Google Scholar] [CrossRef]
  33. Kadari, A.; Schemme, T.; Kadri, D.; Wollschläger, J. XPS and morphological properties of Cr2O3 thin films grown by thermal evaporation method. Results Phys. 2017, 7, 3124–3129. [Google Scholar] [CrossRef]
  34. Cao, Z.; Zuo, C. Cr2O3/carbon nanosheet composite with enhanced performance for lithium ion batteries. RSC Adv. 2017, 7, 40243–40248. [Google Scholar] [CrossRef]
  35. Liu, J.; Liao, M.; Imura, M.; Tanaka, A.; Iwai, H.; Koide, Y. Low on-resistance diamond field effect transistor with high-k ZrO2 as dielectric. Sci. Rep. 2014, 4, 2–6. [Google Scholar] [CrossRef]
  36. Luo, J.; Luo, X.; Hu, C.; Crittenden, J.C.; Qu, J. Zirconia (ZrO2) Embedded in Carbon Nanowires via Electrospinning for Efficient Arsenic Removal from Water Combined with DFT Studies. ACS Appl. Mater. Interfaces 2016, 8, 18912–18921. [Google Scholar] [CrossRef] [PubMed]
  37. Divakaran, K.; Baishnisha, A.; Balakumar, V.; Perumal, K.N.; Meenakshi, C.; Kannan, R.S. Photocatalytic degradation of tetracycline under visible light using TiO2@ sulfur doped carbon nitride nanocomposite synthesized via in-situ method. J. Environ. Chem. Eng. 2021, 9, 105560. [Google Scholar] [CrossRef]
  38. McGrady, J.; Yamashita, S.; Kano, S.; Yang, H.; Abe, H. Charge transfer across the Cr2O3, Fe2O3, and ZrO2 oxide/water interface: A pulse radiolysis study. Radiat. Phys. Chem. 2021, 180, 109240. [Google Scholar] [CrossRef]
  39. Safari, G.; Hoseini, M.; Seyedsalehi, M.; Kamani, H.; Jaafari, J.; Mahvi, A. Photocatalytic degradation of tetracycline using nanosized titanium dioxide in aqueous solution. Int. J. Environ. Sci. Technol. 2015, 12, 603–616. [Google Scholar] [CrossRef]
  40. Linh, N.X.D.; Hanh, N.T.; Cuong, L.M.; Huong, N.T.; Ha, N.T.T.; Trinh, T.D.; Van Noi, N.; Cam, N.T.D.; Pham, T.D. Facile Fabrication of α-Fe2O3/g-C3N4 Z-Scheme Heterojunction for Novel Degradation of Residual Tetracycline. Top. Catal. 2023, 66, 139–148. [Google Scholar] [CrossRef]
  41. Li, X.; Xiong, J.; Gao, X.; Ma, J.; Chen, Z.; Kang, B.; Liu, J.; Li, H.; Feng, Z.; Huang, J. Novel BP/BiOBr S-scheme nano-heterojunction for enhanced visible-light photocatalytic tetracycline removal and oxygen evolution activity. J. Hazard. Mater. 2020, 387, 121690. [Google Scholar] [CrossRef]
  42. 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]
  43. Pan, T.; Chen, D.; Xu, W.; Fang, J.; Wu, S.; Liu, Z.; Wu, K.; Fang, Z. Anionic polyacrylamide-assisted construction of thin 2D-2D WO3/g-C3N4 Step-scheme heterojunction for enhanced tetracycline degradation under visible light irradiation. J. Hazard. Mater. 2020, 393, 122366. [Google Scholar] [CrossRef]
  44. Jiang, H.; Wang, Q.; Chen, P.; Zheng, H.; Shi, J.; Shu, H.; Liu, Y. Photocatalytic degradation of tetracycline by using a regenerable (Bi)BiOBr/rGO composite. J. Clean. Prod. 2022, 339, 130771. [Google Scholar] [CrossRef]
  45. Phuong, D.M.; Duong, T.A.; Huong, N.T.; Khoa, N.V.; Hanh, N.T.; Phuong, N.M.; Pham, T.D.; Trang, H.T.; Van Noi, N. Enhancement of visible light photocatalytic removal of residual tetracycline by Ni doped WO3 nano structures. Inorg. Chem. Commun. 2023, 157, 111329. [Google Scholar] [CrossRef]
  46. Fu, Q.; Meng, Y.; Yao, Y.; Shen, H.; Xie, B.; Ni, Z.; Xia, S. Construction of facet orientation-supported Z-scheme heterojunction of BiVO4 (110)-Fe2O3 and its photocatalytic degradation of tetracycline. J. Environ. Chem. Eng. 2023, 11, 111060. [Google Scholar] [CrossRef]
  47. Siddhardhan, E.V.; Surender, S.; Arumanayagam, T. Degradation of tetracycline drug in aquatic environment by visible light active CuS/CdS photocatalyst. Inorg. Chem. Commun. 2023, 147, 110244. [Google Scholar] [CrossRef]
Figure 1. p-XRD pattern of pure Cr2O3, ZrO2, and Cr2O3-ZrO2 nanocomposites.
Figure 1. p-XRD pattern of pure Cr2O3, ZrO2, and Cr2O3-ZrO2 nanocomposites.
Water 15 03702 g001
Figure 2. (a) Tauc plot (inset: UV-vis absorption spectra) and (b) FT-IR spectra of pure Cr2O3, ZrO2 and Cr2O3–ZrO2 nanocomposites.
Figure 2. (a) Tauc plot (inset: UV-vis absorption spectra) and (b) FT-IR spectra of pure Cr2O3, ZrO2 and Cr2O3–ZrO2 nanocomposites.
Water 15 03702 g002
Figure 3. TEM images of (a) Cr2O3, (b) ZrO2, and (c) Cr2O3–ZrO2; (d,e) HR-TEM images and (f) SAED pattern of Cr2O3–ZrO2 nanocomposites.
Figure 3. TEM images of (a) Cr2O3, (b) ZrO2, and (c) Cr2O3–ZrO2; (d,e) HR-TEM images and (f) SAED pattern of Cr2O3–ZrO2 nanocomposites.
Water 15 03702 g003
Figure 4. (a) XPS survey spectra of Cr2O3/ZrO2 nanocomposite, and (bd) high resolution deconvoluted XPS spectra for Cr 2p, Zr 3d, and O 1s orbitals.
Figure 4. (a) XPS survey spectra of Cr2O3/ZrO2 nanocomposite, and (bd) high resolution deconvoluted XPS spectra for Cr 2p, Zr 3d, and O 1s orbitals.
Water 15 03702 g004
Figure 5. BET analysis spectra of Cr2O3, ZrO2, and Cr2O3–ZrO2 nanocomposites.
Figure 5. BET analysis spectra of Cr2O3, ZrO2, and Cr2O3–ZrO2 nanocomposites.
Water 15 03702 g005
Figure 6. Influence of, (a) solution pH, (b) photocatalyst quantity, (c,d) degradation kinetics of TCL molecules under optimized experimental conditions.
Figure 6. Influence of, (a) solution pH, (b) photocatalyst quantity, (c,d) degradation kinetics of TCL molecules under optimized experimental conditions.
Water 15 03702 g006
Figure 7. Plausible photocatalytic degradation pathway of TCL by Cr2O3/ZrO2.
Figure 7. Plausible photocatalytic degradation pathway of TCL by Cr2O3/ZrO2.
Water 15 03702 g007
Figure 8. Radical quantification of (a) O2•− and (b) OH•−; (c) photocatalytic efficiency of nanocomposite in the presence of different trapping agents.
Figure 8. Radical quantification of (a) O2•− and (b) OH•−; (c) photocatalytic efficiency of nanocomposite in the presence of different trapping agents.
Water 15 03702 g008
Figure 9. Photocatalytic mechanism of Cr2O3–ZrO2 nanocomposite under visible light.
Figure 9. Photocatalytic mechanism of Cr2O3–ZrO2 nanocomposite under visible light.
Water 15 03702 g009
Figure 10. (a) Reusability studies and (b) XRD analysis of Cr2O3–ZrO2 nanocomposite under optimized experimental conditions.
Figure 10. (a) Reusability studies and (b) XRD analysis of Cr2O3–ZrO2 nanocomposite under optimized experimental conditions.
Water 15 03702 g010
Table 1. Comparison of literature reports on the degradation of tetracycline drug with the proposed method.
Table 1. Comparison of literature reports on the degradation of tetracycline drug with the proposed method.
S. No.PhotocatalystLight SourceDegradation TimeDegradation (%)Catalyst Amount (mg)Drug Conc. (mg/L)Ref.
1.α-Fe2O3/g-C3N4 NCsVisible (32 mW/cm2, W)180 min95.05010[40]
2.Black Phosphorus/BiOBr NCsVisible (300 W/cm2, Xe)90 min79.010050[41]
3.BiWO6Visible (300 W/cm2, Xe)180 min79.73020[42]
4.WO3/g-C3N4Visible (300 W/cm2, Xe)60 min78.05080[43]
5.(Bi)BiOBr/rGO NCsVisible (300 W/cm2, Xe)140 min98.05020[44]
6.Ni-WO3LED (35W)105 min76.05010[45]
7.BiVO4/Fe2O3Visible (300 W/cm2, Xe)60 min91.53015[46]
8.CuS/CdS NCsSolar Light 50 min90.05020[47]
9.Cr2O3/ZrO2Visible (300 W/cm2, Xe)120 min97.110050Present Work
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

Wei, X.; Naraginti, S.; Chen, P.; Li, J.; Yang, X.; Li, B. Visible Light-Driven Photocatalytic Degradation of Tetracycline Using p-n Heterostructured Cr2O3/ZrO2 Nanocomposite. Water 2023, 15, 3702. https://doi.org/10.3390/w15203702

AMA Style

Wei X, Naraginti S, Chen P, Li J, Yang X, Li B. Visible Light-Driven Photocatalytic Degradation of Tetracycline Using p-n Heterostructured Cr2O3/ZrO2 Nanocomposite. Water. 2023; 15(20):3702. https://doi.org/10.3390/w15203702

Chicago/Turabian Style

Wei, Xueyu, Saraschandra Naraginti, Pengli Chen, Jiyuan Li, Xiaofan Yang, and Buwei Li. 2023. "Visible Light-Driven Photocatalytic Degradation of Tetracycline Using p-n Heterostructured Cr2O3/ZrO2 Nanocomposite" Water 15, no. 20: 3702. https://doi.org/10.3390/w15203702

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