Next Article in Journal
Structural Characterization, and Antioxidant, Hypoglycemic and Immunomodulatory Activity of Exopolysaccharide from Sanghuangporus sanghuang JM-1
Previous Article in Journal
Negative Effects of Occurrence of Mycotoxins in Animal Feed and Biological Methods of Their Detoxification: A Review
Previous Article in Special Issue
Thermal Exfoliation and Phosphorus Doping in Graphitic Carbon Nitride for Efficient Photocatalytic Hydrogen Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Extended Interfacial Charge Transference in CoFe2O4/WO3 Nanocomposites for the Photocatalytic Degradation of Tetracycline Antibiotics

1
Yunnan Provincial Key Laboratory of Entomological Biopharmaceutical R&D, College of Pharmacy, Dali University, Dali 671000, China
2
Asset and Laboratory Management Division, Dali University, Dali 671000, China
3
Department of Chemistry, Abdul Wali Khan University, Mardan 23200, Pakistan
4
UNESCO-UNISA Africa Chair in Nanosciences and Nanotechnology, College of Graduate Studies, University of South Africa, Muckleneuk Ridge, P.O. Box 392, Pretoria 0002, South Africa
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(19), 4561; https://doi.org/10.3390/molecules29194561
Submission received: 22 July 2024 / Revised: 19 September 2024 / Accepted: 19 September 2024 / Published: 25 September 2024

Abstract

:
The large-scale utilization of antibiotics has opened a separate chapter of pollution with the generation of reactive drug-resistant bacteria. To deal with this, in this work, different mass ratios of CoFe2O4/WO3 nanocomposites were prepared following an in situ growth method using the precursors of WO3 and CoFe2O4. The structure, morphology, and optical properties of the nanocomposite photocatalysts were scrutinized by X-ray diffraction (XRD), UV-visible diffuse reflectance spectra (UV-Vis DRS), photoluminescence spectrum (PL), etc. The experimental data signified that the loading of CoFe2O4 obviously changed the optical properties of WO3. The photocatalytic performance of CoFe2O4/WO3 composites was investigated by considering tetracycline as a potential pollutant. The outcome of the analyzed data exposed that the CoFe2O4/WO3 composite with a mass ratio of 5% had the best degradation performance for tetracycline eradication under the solar light, and a degradation efficiency of 77% was achieved in 20 min. The monitored degradation efficiency of the optimized photocatalyst was 45% higher compared with the degradation efficiency of 32% for pure WO3. Capturing experiments and tests revealed that hydroxyl radical (·OH) and hole (h+) were the primary eradicators of the target pollutant. This study demonstrates that a proper mass of CoFe2O4 can significantly push WO3 for enhanced eradication of waterborne pollutants.

1. Introduction

Antibiotics are a class of secondary metabolites, also known as antimicrobial materials, produced by microorganisms (bacteria, fungi, and actinomycetes) or synthesized synthetically. They are widely used in medicines, aquaculture, animal husbandry, and other fields to treat fatal diseases caused by microbial infections [1,2]. The widespread use of antibiotics, although providing benefits, has resulted in the introduction of antibiotic accumulation in water bodies, causing water pollution that largely affects the stability of ecosystems [3]. It has caused serious environmental issues, and antibiotics pose ecological risks by affecting microbial diversity and plant growth and development, with enormous effects on human health [4]. In recent years, the potential risk of a new tetracycline antibiotic to the environment and human health has attracted widespread attention as it is a typical broad-spectrum antibiotic and widely used in health care centers, aquaculture, and animal husbandry [5,6]. Therefore, finding efficient methods to control the mass of tetracycline in water bodies has become a current research hotspot, which has far-reaching practical impact in the protection of ecological environments.
Photocatalytic degradation of tetracycline has many advantages as it is a potential method to solve environmental problems. This method uses light energy to decompose organic pollutants without additional chemical reagents and is environmentally friendly [7]. Photocatalytic technology, as an advanced oxidation method, is based on the principle of solar energy conversion into chemical energy with the aid of oxidation realized in the presence of oxygen reactive species such as ·OH, ·O2, and 1O2 produced during the photocatalytic process. The role of the photocatalyst in the reaction process is to direct the reactants (organic pollutants) to undergo a chain oxidation-reduction reaction towards the degradation of pollutants into CO2, H2O, and some inorganic ions, while ensuring that it itself remains unchanged before and after the reaction [8,9,10]. Under light conditions, excitation of the photocatalyst generates electrons (e) in the conduction band and holes (h+) in the valence band [11]. The excited e-h+ combine with O2 and H2O to form free radicals to completely oxidize and degrade the pollutants [12,13,14]. It should be noted that the photogenerated carriers formed by the excitation of the material should be properly separated under the action of an internal electric field for migration to the catalyst surface. The produced electrons and holes chemically react with O2 and H2O/OH to give birth to superoxide radical (·O2) and ·OH/H2O2 free radicals, which effectively remove all the toxic materials present in water bodies [15,16]. The recombination of e and h+ in this process leads to a decrease in the active electron and hole lifetime, which reduces the degradation efficiency of catalysts [17,18,19,20,21].
The forbidden bandwidth of WO3 is adjustable in the range of 2.5–3.5 eV, while the potential of its valence band is about 3.2 V. This potential endows a strong oxidation ability to effectively remove many toxic materials [22,23,24]. However, WO3 still has the disadvantages of high electron–hole pair recombination rates and narrow light absorption range, which greatly restrict its practical application [25,26,27]. Cobalt ferrate (CoFe2O4) as a typical spinel-type ferrite has widely been used in photocatalysis owing to its permanent magnetism, saturation magnetization strength, magnetostriction coefficient, and relatively large magnetic crystal anisotropy constant. Its physico-chemical stability is relatively good, and it has been widely applied in capacitor energy storage, advanced oxidation, biomedicine, and many other aspects [28,29,30]. CoFe2O4 has a band gap around 1~3 eV, and its ferromagnetism boosts rapid separation from liquid–solid systems under external magnetic fields [31]. CoFe2O4 nanomaterials can fully utilize visible light to achieve high light utilization. Therefore, loading a certain amount of CoFe2O4 on the WO3 surface will greatly improve its photocatalytic performance.
In this paper, WO3 photocatalyst was prepared by a hydrothermal method and loaded with different masses of CoFe2O4 to obtain CoFe2O4/WO3 composite photocatalysts. The photocatalysts were checked under solar light to degrade tetracycline antibiotic as a potential pollutant in drinking water bodies.

2. Results and Discussion

2.1. Phase Analysis

The XRD data in Figure 1 show the characteristic diffraction peaks of WO3 at 2 θ values of 22.98°, 23.62, 24.24°, 34.08°, and 49.82°, which correspond to the (002), (020), (200), (202), and (140) crystal planes in the standard card JCPDS:43-1035. The results indicate monoclinic crystalinity of WO3 [32]. The introduction of CoFe2O4 basically does not affect the position of the diffraction peaks of WO3, proving that it does not change the crystalline structure of WO3 [33]. Individual diffraction peaks of CoFe2O4 are masked by the diffraction peaks of WO3. At the same time, no other obvious diffraction peaks are detected, showing that the samples are highly crystalline.

2.2. Morphological Analysis

Figure 2a,b show the SEM images of 5%CoFe2O4/WO3. It can be understood that the surface of photocatalysts is rough and is made of countless nanosheets and fine particles with visible voids, which are conducive to photocatalysis. The nanosheet surface of WO3 is tightly attached and bonded with CoFe2O4, which proves that CoFe2O4/WO3 composites are successfully prepared. Figure 2c demonstrates that all the constituent elements are present in the composite, as shown in the EDS energy spectrum. However, the elements of CoFe2O4 are not detected, which may be due to its low content. The presence of the C element is attributed to some low boiling point volatile organic compounds in the air adsorbed on the sample surface, and the Al element is introduced from the test carrier. No other impurity signals are found, indicating that the composite material does not contain impurities. Figure 2d shows that all four elements, Co, Fe, O, and W, are present in the CoFe2O4/WO3 composite and indirectly indicates that the absence of CoFe2O4 in the EDS spectra is due to its low content.

2.3. Energy Band Structure

The optical properties of WO3, CoFe2O4, and 5%CoFe2O4/WO3 were characterized by UV-Vis diffuse reflectance, as presented in Figure 3a,b. WO3 and 5%CoFe2O4/WO3 samples absorb UV-visible light in the range of 220–800 nm. The absorption intensity of 5%CoFe2O4/WO3 is higher than that of WO3 in the wavelength range of 220–690 nm, and the absorption edges of samples appear to be red shifted, indicating that the absorbance capacity of 5%CoFe2O4/WO3 is better than that of WO3, and the absorption performance of the composites in the UV-visible region has been improved to some extent [34].
The Tauc-plot equation is as follows [35,36]:
αhν = A(hν − Eg)n/2
Eg = hν
Here, α, A, hν, and Eg are the absorption coefficient, constant, photon energy, and band gap energy, respectively. The value of n is related to the type of semiconductor when the material is a direct transition semiconductor, the value of “n” is 1. When the material is an indirect transition semiconductor, the value of “n” is 4. According to the above equation, a plot of “hν” corresponding to “(αhν)2/n” was drawn. The tangents intersect the x-axis at 2.97 eV, 2.80 eV, and 1.58 eV, corresponding to the forbidden band widths (Eg) of WO3, 5%CoFe2O4/WO3, and CoFe2O4 [37,38]. From the experimental data, the band gap energy of the prepared 5%CoFe2O4/WO3 is relatively low compared with WO3, as evidenced by the fact that 5%CoFe2O4/WO3 is more easily excited by visible light with a broadened spectral response range. Thus, the introduction of CoFe2O4 leads to the reduction of the band gap energy of WO3, and stronger light absorption may be one of the reasons for its higher photocatalytic efficiency.
To measure the flat-band potentials, the impedance–potential method was used for WO3, 5%CoFe2O4/WO3, and CoFe2O4. The measured data obtained were inverted by the Mott-Schottky theory. The obtained Mott–Schottky curves for the catalysts of WO3, 5%CoFe2O4/WO3, and CoFe2O4 are provided in Figure 3c. The slopes of the curves are all greater than 0, signifying that the samples are all n-type semiconductors. The intercepts of the tangent lines on the x-axis of the Mott–Schottky curves for the catalysts of WO3, 5%CoFe2O4/WO3, and CoFe2O4 are −0.55 V, −0.67 V, and −0.60 V, respectively (vs. SCE). The value of flat-band potentials (Efb) of the material was calculated according to Equation (3) [39]:
E 0 = E f b + R T F
Here, E0 is the intercept of the tangent line of the Mott–Schottky curve, R is the gas constant, T is the temperature, and F is the Faraday’s constant. The calculated Efb of WO3, 5%CoFe2O4/WO3, and CoFe2O4 are −0.58 V, −0.70 V, and −0.63 V, respectively.
Normally, it is generally accepted that the value of the flat-band potential of an n-type semiconductor is approximately equal to the value of the conduction-band potential [40,41]. The valence band potential of the sample was calculated according to Equation (4):
ECB = EVB − Eg
Here, ECB, EVB, and Eg represent the CB potential, VB potential, and band gap energy, respectively. As demonstrated in Figure 3b, the Eg of WO3, 5%CoFe2O4/WO3, and CoFe2O4 are 2.97 eV, 2.80 eV, and 1.58 eV, respectively. From the values of CB potential and Eg, the VB potentials of WO3, 5%CoFe2O4/WO3, and CoFe2O4 are 2.39 V, 2.10 V, and 0.95 V, respectively. Based on the above experimental results, the visualized energy band structure is shown in Figure 3d.

2.4. BET Surface Area Analysis

The specific surface areas, pore diameters, and pore volumes of CoFe2O4, WO3, and 5%CoFe2O4/WO3 are displayed in Table 1. The specific surface area of 5%CoFe2O4/WO3 is smaller than WO3, which may be attributed to the arrival of CoFe2O4 in the nanocomposite. In general, the smaller the specific surface area of the catalyst, the fewer the active sites. Accordingly, the contact between the catalyst and the pollutant reduces, which is not conducive to the improvement of the degradation efficiency of the catalyst [42,43].
The nitrogen adsorption and desorption diagrams of CoFe2O4, WO3, and 5%CoFe2O4/WO3 are shown in Figure 4a. WO3 and 5%CoFe2O4/WO3 have similar shapes; both have desorption hysteresis loops in the range of P/P0 = 0.2~1.0, which suggests that the materials have mesoporous structures [44]. The adsorption capacity of the three materials increases with an increase in P/P0, which is relatively low in a certain range. When P/P0 = 0.9, the adsorption capacity has a large increase. Overall, the adsorption capacity of the three materials (CoFe2O4 > WO3 > 5%CoFe2O4/WO3) is in accordance with the results of the specific surface area test. Figure 4b shows that the pore volume is reduced significantly after loading CoFe2O4, indicating that the reduction of composite pore volume is mainly attributed to CoFe2O4. The decrease in specific surface area and pore volume may be due to the close interfacial contact between CoFe2O4 and WO3 [45].

2.5. Photodegradation Study

The dark adsorption equilibrium time for the photocatalytic degradation experiments was 30 min, and the degradation experiments were conducted after the adsorption equilibrium. Experiments were steered to investigate the effect of different CoFe2O4 loadings on the photocatalytic performance, and the results are displayed in Figure 5a. CoFe2O4/WO3 with CoFe2O4 percentage content of 2.5%, 5%, 7.5%, and 10% were synthesized for the degradation of tetracycline. The experimental results show that the introduction of CoFe2O4 can significantly improve the photocatalytic degradation activity of TC by WO3, and the best degradation performance is achieved when the compound ratio of CoFe2O4 is 5%. After that, the degradation efficiency of tetracycline gradually decreases with an increase in the loading ratio, which may be attributed to the excessive occupation of the active sites on the catalyst surface by the loaded CoFe2O4. The experimental results also show that hydrogen peroxide alone does not have a better degradation efficiency for TC.
Regarding the stability of the materials, we compared the XRD patterns of the optimized photocatalyst with the best catalytic performance before and after the degradation experiments to analyze whether its crystalline phase was changed, as shown in Figure 5b. The positions and intensities of the characteristic peaks of the XRD patterns before and after the reaction are almost unchanged, indicating that the composites are stable.
Potassium persulfate (KPS), ascorbic acid (AA), n-butanol (n-BA), and sodium oxalate (SO) were added to the solution as trapping agents for e, ·O2, ·OH, and h+ to validate the experimental mechanism, while maintaining the same conditions for the photodegradation experiments [46,47]. At the end of dark adsorption, the samples were added with 1 mL of capturing agent and irradiated under solar light for 100 min. The results in Figure 5c show that the degradation efficiency of the catalyst decreases to some extent after the addition of the capture agent. The experimental results show that e, ·O2, ·OH, and h+ all play certain roles in the degradation of tetracycline, while ·OH and h+ are the main active substances that play a role in the degradation.
At the maximum excitation wavelength, fluorescence spectroscopy was measured, as shown in Figure 5d. It is well known that a higher fluorescence intensity represents a higher rate of electron–hole combination [48]. Comparing the fluorescence intensity of 5%CoFe2O4/WO3 and WO3, the fluorescence intensity of 5%CoFe2O4/WO3 < WO3, which indicates that the introduction of CoFe2O4 inhibits the recombination of e and h+ to a certain extent. As a result, the lifetimes of the photogenerated carriers of 5%CoFe2O4/WO3 are prolonged, which improves their photocatalytic performance.
To investigate the effect of CoFe2O4 loading on the charge transfer efficiency of the catalysts, electrochemical impedance spectroscopy (EIS) was measured in a 0.1 mol/L KCl solution containing 5 mM K3[Fe(CN)6]/K4[Fe(CN)6]. As shown in Figure 5e, it can be seen that 5%CoFe2O4/WO3 has a lower arc radius than WO3 in the high-frequency region of the EIS Nyquist plot, so 5%CoFe2O4/WO3 has a higher carrier transfer efficiency [49].
To further understand the effective separation efficiency of the photogenerated electron–hole pairs of the samples, we performed photocurrent experiments in NaSO4 (0.1 mol/L) electrolyte solution to detect the intensity of the photocurrent generated by the materials. The interval between the light switched on and switched off was 20 s, and the measurements were completed in 300 s. Largely, the stronger the photocurrent, the more efficient the separation of electrons and holes, and the better the performance of the photocatalyst [50,51]. From Figure 5f, it can be seen that the photocurrent response of 5%CoFe2O4/WO3 is stronger than that of WO3 and CoFe2O4 throughout the testing time. This signifies that the photogenerated electron–hole recombination rate is reduced after loading CoFe2O4 compared to that of the pure WO3, which is conducive to the enhancement of the photoelectrochemical performance of WO3.

2.6. Photocatalytic Mechanism

Based on the results of the free radical trapping experiments, energy band structure analysis, and photocatalytic degradation efficiency, we propose a possible mechanism for the photocatalytic degradation of tetracycline. On one hand, compared with the CoFe2O4 pure phase material, the photocatalytic degradation efficiency is significantly increased after composite with WO3. As shown in Figure 5a, the degradation efficiency increases from 39% to 77%, demonstrating an increase of 38%. Combined with the results of the capture experiments and energy band structure analysis, it suggests that the oxidation reaction may be dominated by holes in VB with a higher oxidation potential. In other words, oxidation occurs in the VB of WO3 with a more positive potential. On the other hand, the photocatalytic degradation efficiency of the composites is also significantly improved with the introduction of 5%CoFe2O4 into the materials compared with WO3. As shown in Figure 5a, the degradation efficiency increases from 45% to 77%, showing an increase of 32%. Also, combining the results of the trapping experiments and energy band structure analysis, we believe that the photogenerated hole and electron pairs of the composites are separated more efficiently, which results in more photogenerated holes and electrons participating in the degradation reaction. Meanwhile, this conjecture is also verified by the results of photoluminescence, photocurrent, and AC impedance experiments. Coincidentally, this scenario fits well with the S-scheme heterojunction building theory. When irradiated by solar light, CoFe2O4 and WO3 materials are excited when the energy of the incident light is larger than their band gap energies. As a result, the electrons in VB jump to the CB, generating high electron density in the CB and photogenerated h+ in VB of the two components [52]. As shown in Figure 6, when the two semiconductor materials are compounded together, the contact surface and difference in the Fermi energy levels drive the electron transfer from the CoFe2O4 surface with a higher reduction potential to the WO3 surface, which leads to the formation of a built-in electric field. The direction of the electric field is from the reduced semiconductor CoFe2O4 to the oxidized semiconductor WO3 [53]. At the same time, due to the transfer of electrons, the depletion of electrons in the reduced semiconductor and the accumulation of electrons in the oxidized semiconductor cause, respectively, upward and downward bending of the energy bands near the contact surfaces of CoFe2O4 and WO3 [54]. Under the combined effects of the spatial influence from such energy band bending, the mutual repulsion of the homogeneous charges by the Coulomb force, and the electron driven by the built-in electric field, the excited electrons can be transferred from the downward-bended CB of WO3 to the upward-bended VB of CoFe2O4 to combine with holes in the VB band of CoFe2O4, while the reverse transfer of electrons from the CB of CoFe2O4 to the CB of WO3 cannot occur [55]. The construction of this heterojunction is successful in allowing the WO3 to retain its high oxidizing activity at the expense of its lower reducing activity. Meanwhile, the composite material better spatially realizes the effective separation of photogenerated carriers [56]. The effectively separated photogenerated electrons react with O2 to generate superoxide radical (·O2) near CB with a more negative potential, while the generated h+ react with H2O or OH to generate hydroxyl radical (·OH) near VB with a more positive potential. These generated radicals fully degrade the tetracycline pollutants.

3. Experimental Section

3.1. Synthesis of Materials

All the purchased reagents/chemicals for the experiment were of analytical grade and used as such without further refinement. Ferric nitrate ninhydrate, sodium hydroxide, and potassium persulfate were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Sodium tungstate and cobalt nitrate hexahydrate were purchased from Shanghai Mclean Biochemical Technology Co. Ethanol was purchased from Yunnan Yang Lin Industrial Development Zone, Shan Dian Pharmaceutical Co. Tetracycline was purchased from Shanghai Aladdin Biochemical Technology Co. N-butanol was purchased from the Luoyang City Chemical Reagent Factory. Sodium oxalate was purchased from Tianjin Hengxing Chemical Reagent Manufacturing Co. Ascorbic acid was purchased from Shanghai Titan Scientific Co (Shanghai, China). Nitric acid was purchased from Chongqing Chuandong Chemical Industry (Group) Co. (Chongqing, China), and deionized water (conductivity 18.25 μs/cm) was homemade in the laboratory.
WO3 preparation: For the preparation, 0.3299 g Na2WO4·2H2O was accurately weighed and dissolved in 15 mL of deionized water, and 10 mL of HNO3 (0.25 mol/L) was added to it with the help of a measuring cylinder. The mixture was stirred for 10 min and then transferred to a 100 mL sealed reactor for hydrothermal synthesis and heated for 12 h in an oven at 180 °C. The cooled mixture was then washed three times with deionized water and anhydrous ethanol. The prepared photocatalyst was dried at 60 °C. After grinding, it was put in a muffle furnace and calcined at 500 °C for 2 h (20 °C/min).
Preparation of CoFe2O4: For the preparation, 1.0812 g Fe(NO3)3·9H2O and 0.5826 g Co(NO3)2·6H2O were accurately weighed with an analytical balance and dissolved in 50 mL of deionized water under magnetic stirring until the solution was transparent. The pH of the mixture was regulated to 10 with NaOH (1M) solution. After stirring for 10 min, it was poured into a 100 mL sealed reactor for hydrothermal synthesis and heated for 10 h at 160 °C. After that, it was cooled down to room temperature and then washed three times each with water and ethanol. The dried photocatalysts were grinded and calcined in a muffle furnace at 600 °C for 3 h (10 °C/min temperature rise).
CoFe2O4/WO3 preparation: For the preparations, 2.13 mL, 4.3 mL, 6.4 mL, 8.5 mL, and 12.8 mL of Fe(NO3)3·9H2O (0.01 mol/L) solution were prepared separately and, respectively, mixed with 1.07 mL, 2.13 mL, 3.2 mL, 4.3 mL, and 6.4 mL solutions of Co(NO3)2·6H2O under stirring. After adding about 100 mg WO3 to each solution, the mixtures were heated for 10 h at 160 °C in a sealed reactor for hydrothermal synthesis. After that, each sample was cooled down to room temperature and then washed three times each with water and ethanol. The dried photocatalysts were grinded and calcined in a muffle furnace at 600 °C for 3 h (10 °C/min temperature rise).

3.2. Characterization

The crystal structures of photocatalysts were evaluated by means of an X-ray powder diffractometer (XRD, DX-2700BH, Dandong Haoyuan Instrument Co., Ltd., Liaoning, China) with a CuKα radiation source. The microstructure of samples was obtained by scanning electron microscopy (SEM, JSM 7600), and the elemental composition of the material was characterized by EDS. The light absorption characters were measured with the help of a UV-2600i UV-Vis spectrophotometer (UV-vis, Shimadzu, Suzhou, China) and an integrating sphere assembly. Fluorescence intensity was detected at the sample excitation wavelength by a fluorescence spectrophotometer model (Spectrofluorometer-FS5 (PL, Edinburgh Instruments Ltd., Edinburgh, UK). The specific surface area of the materials was analyzed using a fully automated high-performance physical adsorption tester (BET, BSP-PS4T).

3.3. Photocatalytic Activity

A 10 mg sample was weighed into a beaker and mixed with 50 mL of tetracycline solution (20 mg·L−1). Dark adsorption experiments were carried out after 2 min of ultrasonication (30 min to reach equilibrium between adsorption and desorption), and the absorbance of about 4 mL sample was measured after the saturation of adsorption.
Photocatalytic experiments were carried out with WO3, CoFe2O4, and CoFe2O4/WO3 composites with different loadings using a photocatalytic device with a Xenon lamp. Tetracycline was used as a model pollutant. Next, 10 mg of photocatalyst and 50 mL of tetracycline solution (20 mg/L) were added to a beaker and sonicated for 2 min. The mixture was kept under a Xenon lamp at a distance of 20 cm, and the experiments were carried out after the dark adsorption equilibrium was reached. Before the photocatalytic reaction was started, 15 µL of hydrogen peroxide was added to the system, and the photocatalytic experiment was extended to 20 min. Next, 4 mL of the reaction solution was taken at the end of the reaction to detect the absorbance in the wavelength range of 200–600 nm. The degradation rate of tetracycline was calculated by Equation (5):
D % = A 0 A A 0 = C 0 C C 0 × 100 %
Here, D is the degradation efficiency, A0 and A are the absorbances, and C0 and C are, respectively, the concentrations of the stocked tetracycline solution and the solution irritated under light for a certain period.
To detect the active species majorly responsible for the removal of tetracycline, several different quenching agents such as ascorbic acid, potassium persulfate, n-butanol, and sodium oxalate were introduced to the photocatalytic system to trap, respectively, ·O2, e, ·OH, and h+. The photocatalytic reactions were conducted in the same manner as discussed for the measurement of the photocatalytic activity.
At the end of the photocatalytic experiments, the optimized photocatalyst with the best degradation efficiency was filtered and separated. The sample was dried at 80 °C in an oven and subjected to XRD inspection. The post-reaction XRD patterns of the samples were compared with the pre-reaction XRD patterns to evaluate the catalyst stability.

3.4. Photoelectrochemical Test

Ten mg of photocatalyst was dispersed in a mixture of 5 mL each of ethanol and ethylene glycol under sonication for 30 min followed by vigorous stirring for 4 h to form a homogenous suspension. Next, 20 µL of the formed suspension was evenly coated on ITO conductive glass, followed by natural air drying for 30 min. The open-circuit potential was measured in a 0.1 mol-L−1 Na2SO4 solution. To deeply investigate the separation efficiency of photogenerated e-h+ pairs, the instantaneous photocurrent response curves (I-t) of the samples were obtained on the electrochemical workstation of the CHI-660D system using 0.1 of mol∙L−1 Na2SO4 as the electrolyte. Electrochemical impedance spectroscopy (EIS) was performed in a 0.1 mol-L−1 KCl solution containing 5 mmol·L−1 K3(Fe(CN)6]/K4(Fe(CN)6)(1:1).

4. Conclusions

In this study, WO3, CoFe2O4, and CoFe2O4/WO3 composite nanomaterials with different mass ratios were prepared by loading CoFe2O4 on the surface of WO3 using the in situ growth method. The prepared materials were characterized using XRD, SEM, and BET, which proved that the prepared materials were in accordance with the relevant study. The results analyzed by UV-Vis DRS, PL, EIS, and AC impedance showed that the photogenerated carriers produced by 5%CoFe2O4/WO3 subjected to excitation had longer lifetimes, and the composite had a stronger spectral response in the UV-visible region. From the experimental observations, it was concluded that in the case of sample dosages of 10 mg and 20 mg/L of tetracycline, the photoreaction of the 5%CoFe2O4/WO3 sample in 20 min had the highest degradation rate as the degradation efficiency reached 77%. This demonstrated that the introduction of CoFe2O4 enhanced the optical properties of WO3, and CoFe2O4/WO3 was an excellent photocatalyst. We believe that the higher photocatalytic performance of the composites is attributed to the S-scheme heterojunction, which allows the composites to achieve an efficient spatial separation of photogenerated electrons and holes without sacrificing stronger oxidizing properties.

Author Contributions

S.D. conducted catalyst synthesis, characterization, and the activity test. J.D. contributed to data analysis. Y.Y. discussed the photocatalytic mechanism. A.Z. and K.Q. conceived the project and wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (Grants 22268003, 52272287).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lodhi, A.F.; Zhang, Y.; Adil, M.; Deng, Y. Antibiotic discovery: Combining isolation chip (iChip) technology and co-culture technique. Appl. Microbiol. Biotechnol. 2018, 102, 7333–7341. [Google Scholar] [CrossRef]
  2. Zinner, S.H. Antibiotic use: Present and future. New Microbiol. 2007, 30, 321–325. [Google Scholar] [PubMed]
  3. Wei, Z.; Liu, J.; Shangguan, W. A review on photocatalysis in antibiotic wastewater: Pollutant degradation and hydrogen production. Chin. J. Catal. 2020, 41, 1440–1450. [Google Scholar] [CrossRef]
  4. Kumar, L.; Ragunathan, V.; Chugh, M.; Bharadvaja, N. Nanomaterials for remediation of contaminants: A review. Environ. Chem. Lett. 2021, 19, 3139–3163. [Google Scholar] [CrossRef]
  5. Peng, X.; Luo, W.; Wu, J.; Hu, F.; Hu, Y.; Xu, L.; Xu, G.; Jian, Y.; Dai, H. Carbon quantum dots decorated heteroatom co-doped core-shell Fe0@POCN for degradation of tetracycline via multiply synergistic mechanisms. Chemosphere 2021, 268, 128806. [Google Scholar] [CrossRef] [PubMed]
  6. Zheng, X.; Liu, Y.; Liu, X.; Li, Q.; Zheng, Y. A novel PVDF-TiO2@g-C3N4 composite electrospun fiber for efficient photocatalytic degradation of tetracycline under visible light irradiation. Ecotoxicol. Environ. Saf. 2021, 210, 111866. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, S.; Xia, Y.; Yan, G.; Chen, M.; Wang, X.; Wu, L.; Liang, R. PDI bridged MIL-125(Ti)-NH2 heterojunction with frustrated Lewis pairs: A promising photocatalyst for Cr(VI) reduction and antibacterial application. Appl. Catal. B 2022, 317, 121798. [Google Scholar] [CrossRef]
  8. Song, J.; Zhang, J.; Zada, A.; Ma, Y.; Qi, K. CoFe2O4/NiFe2O4 S-scheme composite for photocatalytic decomposition of antibiotic contaminants. Ceram. Int. 2023, 49, 12327–12333. [Google Scholar] [CrossRef]
  9. Zheng, K.; Chen, J.; Gao, X.; Cao, X.; Wu, S.; Su, J. Photocatalytic degradation of tetracycline by Phosphorus-doped carbon nitride tube combined with peroxydisulfate under visible light irradiation. Water Sci. Technol. 2021, 84, 1919–1929. [Google Scholar] [CrossRef]
  10. Dharman, R.K.; Oh, T.H. Fabrication of g-C3N4@N-doped Bi2MoO6 heterostructure for enhanced visible-light-driven photocatalytic degradation of tetracycline pollutant. Chemosphere 2023, 338, 139513. [Google Scholar] [CrossRef]
  11. Zhao, Y.; Li, Y.; Sun, L. Recent advances in photocatalytic decomposition of water and pollutants for sustainable application. Chemosphere 2021, 276, 130201. [Google Scholar] [CrossRef]
  12. Jiang, X.; Zhou, Q.; Lian, Y. Efficient Photocatalytic Degradation of Tetracycline on the MnFe2O4/BGA Composite under Visible Light. Int. J. Mol. Sci. 2023, 24, 9378. [Google Scholar] [CrossRef] [PubMed]
  13. Ma, Y.; Peng, Q.; Sun, M.; Zuo, N.; Mominou, N.; Li, S.; Jing, C.; Wang, L. Photocatalytic oxidation degradation of tetracycline over La/Co@TiO2 nanospheres under visible light. Environ. Res. 2022, 215, 114297. [Google Scholar] [CrossRef] [PubMed]
  14. Sreeram, N.; Aruna, V.; Koutavarapu, R.; Lee, D.Y.; Rao, M.C.; Shim, J. Fabrication of InVO4/SnWO4 heterostructured photocatalyst for efficient photocatalytic degradation of tetracycline under visible light. Environ. Res. 2023, 220, 115191. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, E.Z.; Wang, Y.; Xiao, D. Polymer Nanocomposites for Photocatalytic Degradation and Photoinduced Utilizations of Azo-Dyes. Polymers 2021, 13, 1215. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, X.; Liu, X.; Zhu, L.; Tao, X.; Wang, X. One-step fabrication of novel MIL-53(Fe, Al) for synergistic adsorption-photocatalytic degradation of tetracycline. Chemosphere 2022, 291, 133032. [Google Scholar] [CrossRef]
  17. Lalliansanga; Tiwari, D.; Lee, S.M.; Kim, D.J. Photocatalytic degradation of amoxicillin and tetracycline by template synthesized nano-structured Ce3+@TiO2 thin film catalyst. Environ. Res. 2022, 210, 112914. [Google Scholar] [CrossRef]
  18. Pandi, K.; Preeyanghaa, M.; Vinesh, V.; Madhavan, J.; Neppolian, B. Complete photocatalytic degradation of tetracycline by carbon doped TiO2 supported with stable metal nitrate hydroxide. Environ. Res. 2022, 207, 112188. [Google Scholar] [CrossRef]
  19. Guan, X.-H.; Qu, P.; Guan, X.; Wang, G.-S. Hydrothermal synthesis of hierarchical CuS/ZnS nanocomposites and their photocatalytic and microwave absorption properties. RSC Adv. 2014, 4, 15579–15585. [Google Scholar] [CrossRef]
  20. Guan, X.-H.; Yang, L.; Guan, X.; Wang, G.-S. Synthesis of a flower-like CuS/ZnS nanocomposite decorated on reduced graphene oxide and its photocatalytic performance. RSC Adv. 2015, 5, 36185–36191. [Google Scholar] [CrossRef]
  21. Simonsen, M.E.; Jensen, H.; Li, Z.; Søgaard, E.G. Surface properties and photocatalytic activity of nanocrystalline titania films. J. Photochem. Photobiol. A 2008, 200, 192–200. [Google Scholar] [CrossRef]
  22. Xiang, Q.; Meng, G.F.; Zhao, H.B.; Zhang, Y.; Li, H.; Ma, W.J.; Xu, J.Q. Au Nanoparticle Modified WO3 Nanorods with Their Enhanced Properties for Photocatalysis and Gas Sensing. J. Phys. Chem. C 2010, 114, 2049–2055. [Google Scholar] [CrossRef]
  23. Teoh, L.G.; Shieh, J.; Lai, W.H.; Hung, I.M.; Hon, M.H. Structure and optical properties of mesoporous tungsten oxide. J. Alloys Compd. 2005, 396, 251–254. [Google Scholar] [CrossRef]
  24. Su, J.; Feng, X.; Sloppy, J.D.; Guo, L.; Grimes, C.A. Vertically Aligned WO3 Nanowire Arrays Grown Directly on Transparent Conducting Oxide Coated Glass: Synthesis and Photoelectrochemical Properties. Nano Lett. 2011, 11, 203–208. [Google Scholar] [CrossRef] [PubMed]
  25. Rong, J.; Zhang, T.; Qiu, F.; Rong, X.; Zhu, X.; Zhang, X. Preparation of hierarchical micro/nanostructured Bi2S3-WO3 composites for enhanced photocatalytic performance. J. Alloys Compd. 2016, 685, 812–819. [Google Scholar] [CrossRef]
  26. Khampuanbut, A.; Santalelat, S.; Pankiew, A.; Channei, D.; Pornsuwan, S.; Faungnawakij, K.; Phanichphant, S.; Inceesungvorn, B. Visible-light-driven WO3/BiOBr heterojunction photocatalysts for oxidative coupling of amines to imines: Energy band alignment and mechanistic insight. J. Colloid Interface Sci. 2020, 560, 213–224. [Google Scholar] [CrossRef]
  27. Gu, X.; Lin, S.; Qi, K.; Yan, Y.; Li, R.; Popkov, V.; Almjasheva, O. Application of tungsten oxide and its composites in photocatalysis. Sep. Purif. Technol. 2024, 345, 127299. [Google Scholar] [CrossRef]
  28. Abdolmohammad-Zadeh, H.; Rahimpour, E. CoFe2O4 nano-particles functionalized with 8-hydroxyquinoline for dispersive solid-phase micro-extraction and direct fluorometric monitoring of aluminum in human serum and water samples. Anal. Chim. Acta 2015, 881, 54–64. [Google Scholar] [CrossRef]
  29. Sonia, S.; Kumar, P.; Kumar, A. Multifunctional CoFe2O4/ZnO nanocomposites: Probing magnetic and photocatalytic properties. Nanotechnology 2023, 35, 145705. [Google Scholar] [CrossRef]
  30. Rezaeivala, Z.; Imanparast, A.; Mohammadi, Z.; Najafabad, B.K.; Sazgarnia, A. The multimodal effect of Photothermal/Photodynamic/Chemo therapies mediated by Au-CoFe2O4@Spiky nanostructure adjacent to mitoxantrone on breast cancer cells. Photodiagn. Photodyn. Ther. 2023, 41, 103269. [Google Scholar] [CrossRef]
  31. Ajay, K.; Manisha, C.; Arush, S.; Manita, T.; Amit, K.; Deepak, P.; Lakhveer, S. Robust visible light active PANI/LaFeO3/CoFe2O4 ternary heterojunction for the photo-degradation and mineralization of pharmaceutical effluent: Clozapine. Photodiagn. Photodyn. Ther. 2021, 9, 106159. [Google Scholar]
  32. Naciri, Y.; Hsini, A.; Bouziani, A.; Tanji, K.; El Ibrahimi, B.; Ghazzal, M.N.; Bakiz, B.; Albourine, A.; Benlhachemi, A.; Navío, J.A.; et al. Z-scheme WO3/PANI heterojunctions with enhanced photocatalytic activity under visible light: A depth experimental and DFT studies. Chemosphere 2022, 292, 133468. [Google Scholar] [CrossRef] [PubMed]
  33. Renukadevi, S.; Jeyakumari, A.P. A one-pot microwave irradiation route to synthesis of CoFe2O4-g-C3N4 heterojunction catalysts for high visible light photocatalytic activity: Exploration of efficiency and stability. Diamond Relat. Mater. 2020, 109, 108012. [Google Scholar] [CrossRef]
  34. Liu, S.; Dong, S.; Hao, Y.; Qi, K.; Peng, A. Gd-doped ZnFe2O4 multi-shell microspheres for enhancing photocatalytic H2 production or antibiotic degradation. J. Rare Earths 2024, in press. [Google Scholar] [CrossRef]
  35. Academic MedicineLing, Y.; Dai, Y. Direct Z-scheme hierarchical WO3/BiOBr with enhanced photocatalytic degradation performance under visible light. Appl. Surf. Sci. 2020, 509, 145201. [Google Scholar]
  36. Wang, Z.; Liu, R.; Zhang, J.; Dai, K. S-scheme Porous g-C3N4/Ag2MoO4 Heterojunction Composite for CO2 Photoreduction. Chin. J. Struct. Chem. 2022, 41, 2206015–2206022. [Google Scholar]
  37. Ehsan, M.F.; Fazal, A.; Hamid, S.; Arfan, M.; Khan, I.; Usman, M.; Shafiee, A.; Ashiq, M.N. CoFe2O4 decorated g-C3N4 nanosheets: New insights into superoxide anion mediated photomineralization of methylene blue. J. Environ. Chem. Eng. 2020, 8, 104556. [Google Scholar] [CrossRef]
  38. Palanisamy, G.; Bhuvaneswari, K.; Bharathi, G.; Pazhanivel, T.; Grace, A.N.; Pasha, S.K.K. Construction of magnetically recoverable ZnS–WO3–CoFe2O4 nanohybrid enriched photocatalyst for the degradation of MB dye under visible light irradiation. Chemosphere 2021, 273, 129687. [Google Scholar] [CrossRef]
  39. Wang, X.; Liu, S.; Lin, S.; Qi, K.; Yan, Y.; Ma, Y. Visible Light Motivated the Photocatalytic Degradation of P-Nitrophenol by Ca2+-Doped AgInS2. Molecules 2024, 29, 361. [Google Scholar] [CrossRef]
  40. Huang, J.; Shang, Q.; Huang, Y.; Tang, F.; Zhang, Q.; Liu, Q.; Jiang, S.; Hu, F.; Liu, W.; Luo, Y.; et al. Oxyhydroxide Nanosheets with Highly Efficient Electron–Hole Pair Separation for Hydrogen Evolution. Angew. Chem. Int. Ed. 2016, 55, 2137–2141. [Google Scholar] [CrossRef]
  41. Liu, Y.; Zhang, Q.; Yang, B.; Xu, J.; Yan, Y.; He, J. Synthesis of Zn2+ doped AgInxSy sub-microspheres and its visible light photocatalytic activity. J. Mater. Sci. Mater. Electron. 2019, 30, 15257–15266. [Google Scholar] [CrossRef]
  42. Cao, X.; Chen, W.; Zhao, P.; Yang, Y.; Yu, D.G. Electrospun Porous Nanofibers: Pore-Forming Mechanisms and Applications for Photocatalytic Degradation of Organic Pollutants in Wastewater. Polymers 2022, 14, 3990. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, Z.; Xue, X.; Chen, X. A novel g-C3N4 nanosheet/Ag3PO4/α-Bi2O3 ternary dual Z-scheme heterojunction with increased light absorption and expanded specific surface area for efficient photocatalytic removal of TC. Dalton Trans. 2022, 51, 8015–8027. [Google Scholar] [CrossRef] [PubMed]
  44. Hayat, A.; Rahman, M.U.; Khan, I.; Khan, J.; Sohail, M.; Yasmeen, H.; Liu, S.; Qi, K.; Lv, W. Conjugated Electron Donor–Acceptor Hybrid Polymeric Carbon Nitride as a Photocatalyst for CO2 Reduction. Molecules 2019, 24, 1779. [Google Scholar] [CrossRef] [PubMed]
  45. Xu, Y.; Xu, H.-Y.; Shan, L.-W.; Liu, Y.; Cao, M.-C.; Jin, L.-G.; Dong, L.-M. Photocatalysis Meets Piezoelectricity in a Type-I Oxygen Vacancy-Rich BaTiO3/BiOBr Heterojunction: Mechanism Insights from Characterizations to DFT Calculations. Inorg. Chem. 2024, 63, 6500–6513. [Google Scholar] [CrossRef]
  46. Zhang, J.; Zhao, Y.; Zhang, K.; Zada, A.; Qi, K. Sonocatalytic degradation of tetracycline hydrochloride with CoFe2O4/g-C3N4 composite. Ultrason. Sonochem. 2023, 94, 106325. [Google Scholar] [CrossRef]
  47. Cui, Q.; Gu, X.; Zhao, Y.; Qi, K.; Yan, Y. S-scheme CuInS2/ZnS heterojunctions for the visible light-driven photocatalytic degradation of tetracycline antibiotic drugs. J. Taiwan Inst. Chem. Eng. 2023, 142, 104679. [Google Scholar] [CrossRef]
  48. Li, W.; Zhuang, C.; Li, Y.; Gao, C.; Jiang, W.; Sun, Z.; Qi, K. Anchoring ultra-small TiO2 quantum dots onto ultra-thin and large-sized Mxene nanosheets for highly efficient photocatalytic water splitting. Ceram. Int. 2021, 47, 21769–21776. [Google Scholar] [CrossRef]
  49. Li, B.; Xu, H.-Y.; Chi, G.-H.-N.; Dong, L.-M.; Shan, L.-W.; Jin, L.-G.; Zhuang, Y.-L.; Cao, M.-C.; He, X.-L.; Qi, S.-Y. Integrating the confinement effect and bimetallic cycles in a hierarchical Co3O4@Co3O4/Fe3O4 yolk-shell nanoreactor for peroxymonosulfate activation enhancement. Chem. Eng. J. 2024, 483, 149403. [Google Scholar] [CrossRef]
  50. Mohammad, W.K.; Reda, M.M.; Detlef, W.B. Controlled synthesis of Ag2O/g-C3N4 heterostructures using soft and hard templates for efficient and enhanced visible-light degradation of ciprofloxacin. Ceram. Int. 2021, 47, 31073–31083. [Google Scholar]
  51. Li, F.; Liao, B.; Shen, J.; Ke, J.; Zhang, R.; Wang, Y.; Niu, Y. Enhancing Photocatalytic Activities for Sustainable Hydrogen Evolution on Structurally Matched CuInS2/ZnIn2S4 Heterojunctions. Molecules 2024, 29, 2447. [Google Scholar] [CrossRef] [PubMed]
  52. Albukhari, S.M.; Ismail, A.A. Construction of cobalt ferrite nanoparticles anchored on mesoporous WO3 for accelerated photoreduction of Cr(VI). J. Alloys Compd. 2023, 967, 171788. [Google Scholar] [CrossRef]
  53. Qiao, H.; Du, R.; Zhou, S.; Wang, Q.; Ren, J.; Wang, D.; Li, H. Efficient Charge Carriers Separation and Transfer Driven by Interface Electric Field in FeS2@ZnIn2S4 Heterojunction Boost Hydrogen Evolution. Molecules 2024, 29, 4269. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, J.; Zhao, Y.; Qi, K.; Liu, S.-y. CuInS2 quantum-dot-modified g-C3N4 S-scheme heterojunction photocatalyst for hydrogen production and tetracycline degradation. J. Mater. Sci. Technol. 2024, 172, 145–155. [Google Scholar] [CrossRef]
  55. Fu, J.; Xu, Q.; Low, J.; Jiang, C.; Yu, J. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H2-production photocatalyst. Appl. Catal. B 2019, 243, 556–565. [Google Scholar] [CrossRef]
  56. Zhang, J.; Bifulco, A.; Amato, P.; Imparato, C.; Qi, K. Copper indium sulfide quantum dots in photocatalysis. J. Colloid Interface Sci. 2023, 638, 193–219. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of WO3, CoFe2O4, and CoFe2O4/WO3 samples.
Figure 1. XRD patterns of WO3, CoFe2O4, and CoFe2O4/WO3 samples.
Molecules 29 04561 g001
Figure 2. (a,b) SEM images, (c) EDS energy spectra, and (d) elemental mapping analysis of Co, Fe, O, and W for the sample of 5%CoFe2O4/WO3.
Figure 2. (a,b) SEM images, (c) EDS energy spectra, and (d) elemental mapping analysis of Co, Fe, O, and W for the sample of 5%CoFe2O4/WO3.
Molecules 29 04561 g002
Figure 3. (a) UV-visible diffuse reflectance spectra; (b) forbidden bandwidth; (c) plot of flat-band potential; (d) energy band structure of the samples.
Figure 3. (a) UV-visible diffuse reflectance spectra; (b) forbidden bandwidth; (c) plot of flat-band potential; (d) energy band structure of the samples.
Molecules 29 04561 g003
Figure 4. (a) Nitrogen adsorption and desorption diagrams of CoFe2O4, WO3, and 5%CoFe2O4/WO3 samples; (b) pore size distributions.
Figure 4. (a) Nitrogen adsorption and desorption diagrams of CoFe2O4, WO3, and 5%CoFe2O4/WO3 samples; (b) pore size distributions.
Molecules 29 04561 g004
Figure 5. (a) Photocatalytic degradation performance of tetracycline for samples; (b) XRD patterns of 5%CoFe2O4/WO3 before and after reaction; (c) Effect of different capture agents on the performance of photocatalytic degradation of tetracycline with 5%CoFe2O4/WO3; (d) fluorescence spectroscopy; (e) alternating current impedance spectra; and (f) photocurrent response.
Figure 5. (a) Photocatalytic degradation performance of tetracycline for samples; (b) XRD patterns of 5%CoFe2O4/WO3 before and after reaction; (c) Effect of different capture agents on the performance of photocatalytic degradation of tetracycline with 5%CoFe2O4/WO3; (d) fluorescence spectroscopy; (e) alternating current impedance spectra; and (f) photocurrent response.
Molecules 29 04561 g005
Figure 6. Reaction mechanism of the photocatalytic degradation of tetracycline by CoFe2O4/WO3.
Figure 6. Reaction mechanism of the photocatalytic degradation of tetracycline by CoFe2O4/WO3.
Molecules 29 04561 g006
Table 1. Sample-specific surface area, pore diameter, and pore volume test results.
Table 1. Sample-specific surface area, pore diameter, and pore volume test results.
SampleBET Specific Surface Area (m2/g)Pore Size (nm)Pore Volume (mL/g)
CoFe2O427.520252.11060.4280
WO310.375722.26590.0686
5%CoFe2O4/WO34.776620.09000.0308
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

Dong, S.; Dai, J.; Yang, Y.; Zada, A.; Qi, K. Extended Interfacial Charge Transference in CoFe2O4/WO3 Nanocomposites for the Photocatalytic Degradation of Tetracycline Antibiotics. Molecules 2024, 29, 4561. https://doi.org/10.3390/molecules29194561

AMA Style

Dong S, Dai J, Yang Y, Zada A, Qi K. Extended Interfacial Charge Transference in CoFe2O4/WO3 Nanocomposites for the Photocatalytic Degradation of Tetracycline Antibiotics. Molecules. 2024; 29(19):4561. https://doi.org/10.3390/molecules29194561

Chicago/Turabian Style

Dong, Suiying, Jiafu Dai, Ying Yang, Amir Zada, and Kezhen Qi. 2024. "Extended Interfacial Charge Transference in CoFe2O4/WO3 Nanocomposites for the Photocatalytic Degradation of Tetracycline Antibiotics" Molecules 29, no. 19: 4561. https://doi.org/10.3390/molecules29194561

Article Metrics

Back to TopTop