Next Article in Journal
Synergistic Effects of SDS and Non-Ionic Surfactants on Ceramic Membrane Cleaning Performance Under Acidic Conditions
Previous Article in Journal
Study of Bacterial Elution from High-Efficiency Glass Fiber Filters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of Z-Scheme Heterojunction BiOCl/Bi2WO6 for Visible-Light Photocatalytic Degradation of Tetracycline Hydrochloride

by
Hetian Zhang
,
Zengying Zhu
,
Yajie Huang
,
Jiaxing Yu
and
Ming Li
*
College of Forestry, Northeast Forestry University, Harbin 150040, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Separations 2025, 12(5), 111; https://doi.org/10.3390/separations12050111
Submission received: 24 March 2025 / Revised: 24 April 2025 / Accepted: 25 April 2025 / Published: 28 April 2025
(This article belongs to the Special Issue Photocatalytic Degradation of Organic Pollutant in Wastewater)

Abstract

:
Tetracycline hydrochloride pollution poses a serious environmental threat; however, it is difficult to deal with by conventional methods. In this study, the Z-scheme BiOCl/Bi2WO6 composite was hydrothermally synthesized and evaluated for its ability to decompose tetracycline hydrochloride under visible light. The composite material was systematically characterized by XRD, SEM, TEM/HRTEM, XPS, FTIR, BET, PL, UV-Vis DRS, and EPR to analyze its structure, morphology, and optical/electrochemical properties. Characterization revealed that the composite featured a flower-ball structure with broader light absorption and higher solar energy efficiency. A narrow bandgap further facilitated charge separation, boosting photocatalytic performance. Among the synthesized materials, the 20% BiOCl/Bi2WO6 composite exhibited the best performance, removing 94% of tetracycline hydrochloride in 60 min, which was 5.2 times and 1.4 times higher than pure BiOCl and Bi2WO6, respectively. The rate constant was 10.8 times and 2.5 times higher than that of pure BiOCl and Bi2WO6. After five cycles, it maintained the 88.7% removal rate, with X-ray diffraction analysis confirming its structural stability and well mechanical properties. Electron paramagnetic resonance and radical scavenging experiments identified photogenerated holes (h+) and superoxide radicals (·O2) as the primary active species. This work highlights the fact that the prepared Z-scheme BiOCl/Bi2WO6 composite exhibited excellent photocatalytic performance in the degradation of tetracycline hydrochloride, demonstrating promising potential for practical applications.

Graphical Abstract

1. Introduction

In recent years, due to the overuse of antibiotics in disease treatment and livestock, antibiotic residues are ubiquitous in the water body, leading to severe water pollution [1]. In particular, tetracycline hydrochloride (TCH) is persistent and accumulative, posing significant threats to human health and ecosystems [2]. To deal with TCH pollution, various strategies have been developed, including adsorption [3,4], biodegradation [5,6], membrane separation [7,8], Fenton processes [9,10], and photocatalysis. Among these methods, photocatalytic technology has gained extensive attention due to its properties of cost-effectiveness, high efficiency, and eco-friendliness [11,12]. For example, evidence shows that potassium peroxymonosulfate activation of WO3/diatomite composites can effectively degrade TCH [13]; however, the great dependence on external oxidants significantly increases treatment costs. In addition, the rapid recombination of photogenerated carriers and the limited absorption of visible light significantly hinder the practical applications of photocatalysts [12]. Thus, the development of highly efficient photocatalysts is a matter of urgency.
Recently, bismuth-containing compounds have become a hotspot of research due to their unique properties, e.g., high photocatalytic performance, robust antimicrobial activity, non-toxicity, and strong absorption ability in the near-infrared region [11,14]. Bi2WO6 is an orthorhombic perovskite-structured material composed of (Bi2O2)2+ layers and perovskite-like octahedral (WO4)n2n− layers [15], which is widely used in hydrogen production, CO2 reduction, and the degradation of organic pollutants, due to its suitable bandgap, visible-light responsiveness, non-toxicity, and stability [16]. Wu et al. [17] enhanced hydrogen-production efficiency to 56.9 mol/g/h by adjusting the surface terminations to produce atomically thin two-dimensional layers of Bi2WO6. Jiang et al. [18] reported a fast photocatalytic conversion reaction of CO2 using three-dimensional hollow-structured Bi2WO6 synthesized via a solvothermal method.
The halide perovskite BiOCl is composed of alternating Bi2O2 and double Cl layers, forming a layered network structure that facilitates the migration of photogenerated electron-hole pairs [19]; however, the large bandgap (approximately 3.2–3.5 eV) limits absorption to UV light, resulting in a suboptimal photocatalytic performance [20]. To effectively utilize the performance advantages of BiOCl, various modification strategies have been proposed [21,22,23,24]. Yang et al. [25] synthesized S-doped hydroxylated BiOCl catalysts via a one-pot high-pressure reaction with the kinetic constant being 15 times that of BiOCl. Hu et al. [26] developed a recrystallization method to fabricate three-dimensional multilevel defect-rich BiOCl, exhibiting hexagonal prismatic morphology and containing numerous oxygen vacancies. Moreover, constructing heterostructures is an effective method of improving charge separation. By combining various semiconductor materials through certain methods, each material can leverage its advantages, enabling the heterostructure to meet the practical application requirements [27]. For example, compared to pure components, the successful construction of the Cd0.5Zn0.5S/Bi2WO6 S-scheme heterojunction significantly improved the separation and stability of photogenerated electron-hole pairs in degrading tetracycline and reducing Cr(VI) [28].
Former studies investigated Bi-based heterostructures for photocatalytic applications that rely on conventional type-I or type-II configurations [29], therefore limiting charge-separation efficiency. To improve charge separation and preserve redox potential, a Z-scheme BiOCl/Bi2WO6 heterostructure was synthesized in this study utilizing hydrothermal and deposition methods, supplemented by in situ growth techniques. This synthetic strategy offers an efficient and sustainable route to develop advanced photocatalysts due to its simplicity, cost-effectiveness, and eco-friendly nature. Comprehensive characterizations were conducted to measure the structure, morphology, optical, and electrochemical properties of materials using XRD, SEM, TEM/HRTEM, XPS, FTIR, BET, PL, UV-Vis DRS, and EPR techniques. The photocatalytic degradation of TCH under visible-light irradiation was systematically investigated to assess the influence of various parameters on the photocatalytic performance. The main active species involved in the photocatalytic processes were identified through radical scavenging experiments. This study also delves into the underlying mechanisms responsible for the enhancement of photocatalytic activity, providing insights into the development of more effective photocatalytic systems for environmental remediation.

2. Materials and Methods

2.1. Synthesis of Bi2WO6, BiOCl, and BiOCl/Bi2WO6 Composites

Firstly, 2 mmol of Bi(NO3)3·5H2O and 1 mmol of Na2WO4·2H2O were dissolved into deionized water and transferred to a 100 mL Teflon-lined stainless-steel autoclave. The mixture underwent a hydrothermal reaction at 180 °C for 12 h to obtain Bi2WO6.
Next, 2 mmol of KCl and 1 mmol of Bi(NO3)3·5H2O were dissolved in 40 mL of deionized water and stirred for 1 h. Different amounts of Bi2WO6 were added to the above-mentioned solution and stirred for another 1 h. Then, the solution was transferred into a 100 mL Teflon-lined autoclave to react at 180 °C for 18 h to obtain BiOCl/Bi2WO6. The mass proportions of BiOCl in the BiOCl/Bi2WO6 composites were 10%, 20%, 30%, and 40%, denoted as 10-BOC/BWO, 20-BOC/BWO, 30-BOC/BWO, and 40-BOC/BWO, respectively. Pure BiOCl was synthesized by omitting Bi2WO6 in these steps. Additionally, to confirm the improved performance results from true heterojunction formation, a physical mixture of BiOCl and Bi2WO6 with 20% BiOCl by mass (20-BOC+BWO) was synthesized under the same conditions.

2.2. Characterization of the Synthesized Samples

The absorbance of the samples was analyzed using an ultraviolet-visible spectrophotometer (UV-5100, Yuanxi Instruments Co., Ltd., Shanghai, China). The crystal phase structure was characterized by powder X-ray diffraction (XRD, TD-3500, Dandong Tongda Technology Co., Ltd., Dandong, China) with a scan range from 10° to 80°. Morphological features were examined by a field-emission scanning electron microscope (SEM, Gemini 300, Carl Zeiss AG, Oberkochen, Germany). Transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) were performed to assess microcrystalline structures (JEM-2100F, JEOL Ltd., Tokyo, Japan). X-ray photoelectron spectroscopy (XPS) analysis was conducted using a Thermo ESCALAB 250Xi (Thermo Fisher Scientific Inc., Waltham, MA, USA). Surface functional groups of BiOCl, Bi2WO6, and BiOCl/Bi2WO6 were analyzed with Fourier transform infrared spectroscopy (FT-IR, Nicolet Nexus 670, Thermo Nicolet Corporation, Waltham, MA, USA). Brunauer–Emmett–Teller (BET) tests for nitrogen adsorption–desorption isotherms were conducted by a Quantachrome Autosorb-iQ. Photoluminescence (PL) spectra were recorded to measure the recombination of photo-generated electron-hole pairs using a fluorescence spectrophotometer (PL, FLS980, Edinburgh Instruments Ltd., Livingston, UK). The optical-absorption properties were measured using a UV-vis diffuse reflectance spectrometer (UV-vis DRS, UV3600, Shimadzu Corporation, Kyoto, Japan).

2.3. Photocatalytic Degradation Experiment

The degradation experiment was carried out using a xenon light source (PLS-SXE300+, Beijing Perfectlight Technology Co., Ltd., Beijing, China) to simulate sunlight. The experiments were conducted in a photoreactor with a circulating water-cooling system to maintain a constant temperature of 25 °C. First, 100 mg of the photocatalyst was ultrasonically dispersed in 100 mL of TCH solution (20 mg/L). The solution was stirred in the dark for 30 min to achieve adsorption-desorption equilibrium before light exposure. During the photocatalytic degradation under simulated sunlight (350 W xenon lamp, λ > 420 nm), samples were collected every 10 min and filtered through a 0.22 μm membrane to measure TCH concentration using a UV-visible spectrophotometer (357 nm). The remaining TCH solution containing photocatalyst was centrifuged to collect powders for 4 repetitions of the photocatalytic experiment in order to facilitate an assessment of the stability of the photocatalyst.
Electron paramagnetic resonance spectroscopy (EPR, Bruker A300, Bruker Corporation, Karlsruhe, Germany) was utilized to measure the oxidative active species mainly involved in the reaction. The active species superoxide radicals (·O2), photogenerated holes (h+), and hydroxyl radicals (·OH) were quenched by adding p-benzoquinone (BQ), ethylenediaminetetraacetic acid disodium salt (EDTA-2Na), and isopropanol (IPA), respectively, to investigate the photocatalytic mechanism. Moreover, Cl and HCO3 were added to examine the influence of co-ions on TCH degradation.

3. Results and Discussion

3.1. Characterization of Catalysts

XRD analysis (Figure S1) confirmed the successful synthesis of Bi2WO6 and BiOCl; the characteristic diffraction peaks at 2θ values corresponded to orthorhombic Bi2WO6 (JPCDS card No.79-2381) and tetragonal BiOCl (JPCDS card No.85-0861). No significant shifts in the diffraction peaks of the BiOCl/Bi2WO6 composites were observed, indicating that the presence of BiOCl did not cause lattice distortion in Bi2WO6 [30]. Peaks became more prominent as the BiOCl content increased, confirming the successful integration. However, the intensity of the (002) and (003) planes of BiOCl was reduced in the composites, suggesting that Bi2WO6 suppressed BiOCl crystal growth [30].
SEM images (Figure 1) show that BiOCl consists of irregular, dispersed nanosheets, while Bi2WO6 forms flower-like structures with voids [31]. The composite shows BiOCl nanosheets covering the Bi2WO6 surface and maintaining a similar flower-like morphology, confirming successful formation of composite. TEM and HRTEM images (Figure 2a–e) show irregular nanosized particles; these images aligned with the SEM results. As magnification increased (Figure 2b,c), particle details and crystal boundaries became clearer. The fast Fourier transform analysis (Figure 2d,e) confirmed the presence of clear lattice fringes in the BiOCl/Bi2WO6, reflecting the high crystallinity and ordered crystal structure. The crystal planes in regions I–IV indicated the presence of both BiOCl and Bi2WO6 and further confirmed the successful synthesis of the BiOCl/Bi2WO6 composite. Energy-dispersive X-ray spectroscopy (EDX) (Figure 2f) revealed the presence of Bi, W, Cl, and O elements in the composites.
XPS spectra (Figure S2a) further confirmed the presence of Bi, W, Cl, and O elements in the BiOCl/Bi2WO6 composite. The spectra were calibrated using the C1s peak at 284.8 eV (Figure S2b) [32]. High-resolution Bi 4f spectra (Figure S2c) indicated chemical-state changes in Bi during composite formation, while O 1s spectra (Figure S2d) showed peaks for W-O, Bi-O, and adsorbed oxygen [33,34]. The Cl 2p spectra (Figure S2e) revealed peaks of Cl 2p1/2 and Cl 2p3/2, with a shift in the Cl 2p peak of the BiOCl/Bi2WO6 heterojunction. This shift indicated a strong interaction between Cl and WO42− interactions and an increase in binding energy due to electron loss in BiOCl [35]. W 4f spectra (Figure S2f) confirmed the presence of W6+; moreover, the shifts of binding energy indicated electron accumulation in Bi2WO6 and an increase in surface-charge density. These changes confirmed electron transfer from BiOCl to Bi2WO6, establishing an internal electric field and successful heterojunction formation [36].
As shown in Figure S3, the FTIR spectra of BiOCl, Bi2WO6, and BiOCl/Bi2WO6 reveal functional group changes before and after composite formation. The peaks at 520 cm−1 and 1382 cm−1 in BiOCl represent the Bi-O vibrations and Bi-Cl stretching, respectively [37]. In Bi2WO6 and BiOCl/Bi2WO6, the peaks at 573 cm−1 and 731 cm−1, as well as 818 cm−1 and 1383 cm−1, represent Bi-O, W-O, and W-O-W bonds, respectively [38,39].
Figure S4a displays nitrogen adsorption–desorption isotherms, revealing type-IV isotherms for Bi2WO6 and BiOCl/Bi2WO6, characterized by H3 hysteresis loops [40,41]. In contrast, BiOCl exhibits a type-III isotherm, typical of non-porous or microporous materials with minimal surface area. Pore-size distribution (Figure S4b–d), determined by the Barrett–Joyner–Halenda (BJH) and density functional theory (DFT) methods, shows predominant pore sizes of 9–15 nm for Bi2WO6 and BiOCl/Bi2WO6. As summarized in Table S1, BiOCl/Bi2WO6 exhibits the highest specific surface area, pore volume, and pore size, indicating that it is favorable for pollutant adsorption [42]. The optical-absorption characteristics of the prepared samples were analyzed using UV-Vis diffuse reflectance spectroscopy. Figure S5a shows that BiOCl has an optical-absorption edge at 370 nm, whereas Bi2WO6 and BiOCl/Bi2WO6 display light-absorption edges between 440 and 460 nm.
The bandgap (Eg) of photocatalysts is determined by the Kubelka–Munk equation:
Ah υ 1 / n = A h υ E g
where α is the absorption coefficient, h is the Planck constant, υ is the photon frequency, Eg is the bandgap, A is a constant, and n is 1/2 and 2 for direct and indirect bandgap semiconductors, respectively. Both pure BiOCl and Bi2WO6 are classified as indirect bandgap semiconductors, meaning n is 2 [30]. As illustrated in Figure S5b, the Eg of BiOCl, Bi2WO6, and BiOCl/Bi2WO6 are 3.29 eV, 2.56 eV, and 2.51 eV, respectively. The reduced Eg of the composite material facilitates carrier migration within the BiOCl/Bi2WO6 heterojunction, enhancing light-absorption efficiency and photocatalytic activity [43].
PL spectroscopy (Figure 3) was employed to evaluate the separation efficiency of photo-generated electron-hole pairs. The BiOCl/Bi2WO6 composite shows lower fluorescence intensity compared to BiOCl, suggesting that the heterojunction structure effectively suppresses the recombination of electron holes. The enhanced carrier separation prolongs their lifetime and improves charge-transfer efficiency, ultimately enhancing photocatalytic performance. However, the observed reduction in PL intensity for the BiOCl/Bi2WO6 composite is relatively modest, implying that single PL may not be sufficient to fully validate the improvement in carrier-separation efficiency.

3.2. Photocatalytic Activity

In Figure 4a, the TCH solution containing photocatalyst was equilibrated in the dark for 30 min to achieve adsorption–desorption equilibrium. The dark treatment led to variations in the initial (t = 0) concentrations across samples, which can be attributed to the difference in the ability of the composites to absorb the TCH pollution during the treatment. As shown in Figure S4 and Table S1, BiOCl/Bi2WO6 composites possessed a higher surface area and pore volume compared to the pure BiOCl and Bi2WO6, leading to greater TCH adsorption at equilibrium. After irradiation for 60 min, the concentration of TCH decreased with time. The BiOCl/Bi2WO6 composite, particularly 20-BOC/BWO, exhibited superior degradation efficiency compared to pure BiOCl and Bi2WO6. The removal efficiencies of BiOCl and Bi2WO6 were 18.2% and 66.8%, respectively, while the removal efficiencies of BiOCl/Bi2WO6 composites ranged from 83.1% (10-BOC/BWO) to 94.0% (20-BOC/BWO). Notably, the photocatalytic TCH degradation rate of 20-BOC/BWO showed significant improvement. Additionally, the degradation efficiency of 20-BOC+BWO was only 73.4%, which was significantly lower than that of 20-BOC/BWO (94.0%), demonstrating that the performance enhancement stems from genuine heterojunction formation. The apparent quantum yield (AQY) serves as a critical tool for evaluating the efficiency of photon utilization in photocatalytic processes [44,45], defined as the ratio of the number of degraded molecules to the number of incident photons. All experiments were performed under the same visible-light irradiation source and intensity; the number of incident photons remained constant. Therefore, the differences in TCH degradation performance among the photocatalysts directly reflect their relative AQY values. The 20% BiOCl/Bi2WO6 composite exhibited the highest degradation efficiency, indicating its superior photocatalytic capability compared to the other synthesized materials, particularly pure BiOCl and Bi2WO6.
Kinetic analysis reveals that TCH degradation follows a pseudo-first-order model:
ln(C0/Ct) = kt
where C0, Ct, k, and t represent the concentration of TCH (mg/L) at initial and t time, the apparent reaction-rate constant, and time (min), respectively.
Figure 4b shows the pseudo-first-order kinetic fitting curves and apparent rate constants for each sample under visible light. In particular, the reaction-rate constant of the BiOCl/Bi2WO6 composite was significantly higher than that of pure BiOCl and Bi2WO6. Specifically, the degradation rate reached 0.0312 min−1 when the proportion of BiOCl in the composite was 20%, which was 2.5 times and 10.8 times higher than that of pure Bi2WO6 and BiOCl, respectively. These results confirmed that the introduction of BiOCl enhanced the photocatalytic performance of the BiOCl/Bi2WO6 composite.
Figure S6a,b illustrates the influence of initial TCH concentrations on the photocatalytic degradation performance of the 20-BOC/BWO composite material. As shown in Figure S6a, the pollutant-removal rate decreased as initial TCH concentration increased. Figure S6b demonstrates that the apparent rate constants declined as the pollutant concentration increased. Figure S6c demonstrates that photocatalytic efficiency increased with catalyst mass; when the dosage of catalyst was 0.1 g, maximum degradation efficiencies achieved 93%; however, the efficiency declined when the dosage of catalyst exceeded 0.1 g. Excessive catalyst dosage can cause turbidity, blocking light paths and reducing light absorption [46]. Additionally, increased catalyst content may lead to particle aggregation, decreasing the effective reaction surface area. The pseudo-first-order kinetic curves and apparent rate constants in Figure S6d show that the rate constants initially increased and then decreased with catalyst dosage. According to the previous studies, this behavior may be attributed to factors such as light scattering and reduced light penetration, which inhibited the effective utilization of the photocatalyst surface [47,48].
To further assess photocatalytic performance, 100 mg of 20-BOC/BWO was applied to degrade 100 mL of congo red (CR, 100 mg/L), rhodamine B (RhB, 50 mg/L), and TCH (20 mg/L) under simulated visible light. As shown in Figure S6e, 99.7% CR was degraded within 20 min, highlighting its high efficiency. After 40 min, the degradation rate of RhB also reached 99.7%, while the degradation rate of TCH increased to 93.9% over 60 min. Overall, 20-BOC/BWO composite demonstrated excellent versatility to degrade CR, RhB, and TCH. Figure S6f illustrates the influence of anions in the solution on the photocatalytic degradation of TCH. The results indicate that the addition of Cl, SO42−, and NO3 reduced the degradation rate; moreover, the inhibition strengthened as the concentration of these anions increased.
To evaluate the stability of the 20-BOC/BWO composite photocatalyst, TCH degradation experiments were conducted over five consecutive cycles under identical conditions. As shown in Figure S7a, despite a slight decrease in efficiency after the five cycles, the degradation rate remained 88%, showing strong stability and mechanical robustness. Figure S7b presents the XRD results of the 20-BOC/BWO composite catalyst before and after the cycling experiments. The positions and intensities of the diffraction peaks were unchanged, confirming that the crystal structure remained stable after five cycles. These results further demonstrate that the 20-BOC/BWO composite catalyst is suitable for repeated use in photocatalytic applications.

3.3. Mechanism of Enhanced Photoactivity

EPR experiments further identified h+ and ·O2 as the dominant active species in the BOC/BWO photocatalytic system. Figure 5a shows that the TEMPO-h+ signal peak intensity decreased under illumination, indicating increased h+ generation due to effective electron-hole pair separation [43]. In Figure 5b, characteristic quartet signals for DMPO-·O2 adducts appeared only under illumination, reinforcing the fact that h+ and ·O2 were key active species in the degradation process.
Radical scavenging experiments were conducted to elucidate the photocatalytic degradation mechanism of TCH and identify the primary active species involved. IPA, BQ, and EDTA-2Na were utilized as scavengers for ·OH, ·O2, and h+, respectively. As shown in Figure 6a, IPA, BQ, and EDTA-2Na decreased the photocatalytic degradation efficiencies by 7.1%, 57.2%, and 88.4%, respectively, indicating that ·OH played a minor role. Conversely, the removal of h+ and ·O2 significantly hindered photocatalytic degradation, showing that these active species were vital and played a key role in the degradation of TCH.
The valence band (VB) and conduction band (CB) potentials of the photocatalysts are calculated as Equation (3) [49]:
Eg = EVB − ECB
where EVB and ECB are the potentials of VB and CB, respectively.
According to the XPS valence band spectra of the BiOCl and Bi2WO6 photocatalysts in Figure 6b, the EVB values of p-type BiOCl and n-type Bi2WO6 are 2.12 eV and 2.28 eV, respectively. The Eg of BiOCl and Bi2WO6 are 3.29 eV and 2.56 eV, resulting in ECB values for BiOCl and Bi2WO6 of −1.17 eV and −0.28 eV, respectively.
To explain the enhanced photocatalytic capabilities of the binary system, two mechanisms are proposed in Figure 7a,b: a traditional type-II mechanism and a direct Z-scheme mechanism. As shown in Figure 7a, the ECB of BiOCl is more negative than that of Bi2WO6, while the EVB of Bi2WO6 is more positive than that of BiOCl. In the traditional mechanism, photo-induced electrons from the CB of BiOCl transfer to Bi2WO6. However, the reductive capacity of electrons in the CB of Bi2WO6 is weaker than the O2/·O2 potential (−0.33 eV), thus preventing ·O2 radical generation [50]. This contradicts the trapping experiments and EPR results that clearly indicate the formation of ·O2 radicals. Therefore, the traditional type-II mechanism cannot accurately explain the observed photocatalytic activity, demonstrating that it is inappropriate to adopt this mechanism.
In contrast, the direct Z-scheme model illustrated in Figure 7b provides a clearer explanation for the enhanced photocatalytic activity in decomposing TCH [51]. In this mechanism, photo-generated electrons transfer from the CB of BiOCl to the VB of Bi2WO6, effectively separating electrons and holes, generating the active radicals necessary for TCH degradation. These radicals, characterized by their high reactivity, drive a series of photoredox reactions. The detailed reactions are outlined in Equations (4)–(8).
BiOCl/Bi2WO6 + hν → BiOCl (h+ + e)/Bi2WO6 (h+ + e)
BiOCl (h+)/Bi2WO6 (e) → BiOCl/Bi2WO6 (recombination)
Bi2WO6 (h+) + TCH → degradation products
BiOCl (e) + O2 → ·O2
·O2 + TCH → degradation products

4. Conclusions

In this study, Z-scheme heterojunction BiOCl/Bi2WO6 was successfully synthesized via a hydrothermal method and utilized to decompose TCH. The results indicated that the BiOCl/Bi2WO6 composite material exhibited a significantly higher carrier-migration rate and visible-light-absorption ability, greatly outperforming the pure Bi2WO6 and BiOCl in the degradation of TCH. EPR and radical scavenging tests showed that h+ and·O2 are the key active species in the photocatalytic degradation of TCH. This study presents the Z-scheme BiOCl/Bi2WO6 composite as a promising and efficient photocatalyst for TCH degradation, offering great potential for practical applications. Future research can explore its performance in complex water bodies and investigate large-scale production for environmental cleanliness.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/separations12050111/s1, Table S1. Pore structure parameters of the prepared samples; Figure S1. XRD patterns of Bi2WO6, BiOCl, X-BOC/BWO (X = 10, 20, 30, and 40); Figure S2. XPS survey spectra of BiOCl, Bi2WO6, and BiOCl/Bi2WO6: (a) full spectrum, (b) C 1s, (c) Bi 4f, (d) O1s, (e) Cl 2p, and (f) W 4f; Figure S3. FTIR spectra of BiOCl, Bi2WO6, and BiOCl/Bi2WO6; Figure S4. Nitrogen adsorption-desorption isotherms (a), BJH adsorption pore size distribution curves (b), BJH desorption pore size distribution curves (c), and DFT pore size differential distribution diagram (d) of BiOCl, Bi2WO6, and BiOCl/Bi2WO6; Figure S5. The UV-vis spectra (a) and the band gap energy (b) of the as-prepared photocatalysts; Figure S6. Photocatalytic degradation curves and pseudo-first-order kinetic fitting curves of 20-BOC/BWO at different TCH concentrations (a,b), catalyst dosages (c,d); different pollutants (e), and coexisting ions (f); Figure S7. (a) The cyclic stability test of the 20-BOC/BWO, and (b) the XRD of the used and fresh 20-BOC/BWO.

Author Contributions

Methodology, H.Z., Y.H. and J.Y.; Investigation, Y.H.; Data curation, J.Y.; Writing—original draft, H.Z. and Z.Z.; Writing—review & editing, M.L.; Supervision, M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Undergraduate Training Programs for Innovations (202310225109) and Natural Science Foundation of Heilongjiang Province of China (LH2023D003).

Data Availability Statement

Data is contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Iskandar, K.; Murugaiyan, J.; Hammoudi Halat, D.; Hage, S.E.; Chibabhai, V.; Adukkadukkam, S.; Roques, C.; Molinier, L.; Salameh, P.; Van Dongen, M. Antibiotic discovery and resistance: The chase and the race. Antibiotics 2022, 11, 182. [Google Scholar] [CrossRef]
  2. Hanna, N.; Tamhankar, A.J.; Stålsby Lundborg, C. Antibiotic concentrations and antibiotic resistance in aquatic environments of the WHO Western Pacific and South-East Asia regions: A systematic review and probabilistic environmental hazard assessment. Lancet Planet. Health 2023, 7, e45–e54. [Google Scholar] [CrossRef] [PubMed]
  3. Haciosmanoglu, G.G.; Mejias, C.; Martin, J.; Santos, J.L.; Aparicio, I.; Alonso, E. Antibiotic adsorption by natural and modified clay minerals as designer adsorbents for wastewater treatment: A comprehensive review. J. Environ. Manag. 2022, 317, 115397. [Google Scholar] [CrossRef]
  4. Rathod, M.; Haldar, S.; Basha, S. Nanocrystalline cellulose for removal of tetracycline hydrochloride from water via biosorption: Equilibrium, kinetic and thermodynamic studies. Ecol. Eng. 2015, 84, 240–249. [Google Scholar] [CrossRef]
  5. Li, S.J.; Peng, L.; Yang, C.G.; Song, S.X.; Xu, Y.F. Cometabolic biodegradation of antibiotics by ammonia oxidizing microorganisms during wastewater treatment processes. J. Environ. Manag. 2022, 305, 114336. [Google Scholar] [CrossRef] [PubMed]
  6. Bueno, I.; He, H.; Kinsley, A.C.; Ziemann, S.J.; Degn, L.R.; Nault, A.J.; Beaudoin, A.L.; Singer, R.S.; Wammer, K.H.; Arnold, W.A. Biodegradation, photolysis, and sorption of antibiotics in aquatic environments: A scoping review. Sci. Total Environ. 2023, 897, 165301. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, N.; Peng, L.; Gu, Y.; Liang, C.Z.; Pott, R.W.M.; Xu, Y.F. Insights into biodegradation of antibiotics during the biofilm-based wastewater treatment processes. J. Clean Prod. 2023, 393, 136321. [Google Scholar] [CrossRef]
  8. Zhu, Y.J.; Wang, Y.Y.; Zhou, S.; Jiang, X.X.; Ma, X.; Liu, C. Robust performance of a membrane bioreactor for removing antibiotic resistance genes exposed to antibiotics: Role of membrane foulants. Water Res. 2018, 130, 139–150. [Google Scholar] [CrossRef]
  9. Gu, W.Y.; Huang, X.Y.; Tian, Y.H.; Cao, M.; Zhou, L.; Zhou, Y.; Lu, J.; Lei, J.Y.; Zhou, Y.B.; Wang, L.Z.; et al. High-efficiency adsorption of tetracycline by cooperation of carbon and iron in a magnetic Fe/porous carbon hybrid with effective Fenton regeneration. Appl. Surf. Sci. 2021, 538, 147813. [Google Scholar] [CrossRef]
  10. Yang, Q.X.; Yan, Y.; Yang, X.F.; Liao, G.Y.; He, J.; Wang, D.S. The effect of complexation with metal ions on tetracycline degradation by Fe2+/3+ and Ru3+ activated peroxymonosulfate. Chem. Eng. J. 2022, 429, 132178. [Google Scholar] [CrossRef]
  11. Ajiboye, T.O.; Oyewo, O.A.; Onwudiwe, D.C. The performance of bismuth-based compounds in photocatalytic applications. Surf. Interfaces 2021, 23, 100927. [Google Scholar] [CrossRef]
  12. Wang, C.C.; Yan, R.Y.; Cai, M.J.; Liu, Y.P.; Li, S.J. A novel organic/inorganic S-scheme heterostructure of TCPP/Bi12O17Cl2 for boosting photodegradation of tetracycline hydrochloride: Kinetic, degradation mechanism, and toxic assessment. Appl. Surf. Sci. 2023, 610, 155346. [Google Scholar] [CrossRef]
  13. Nguyen, T.L.; Tran, H.H.; Huynh, T.C.; Le, K.T.; Cao, T.M.; Pham, V.V. Harnessing potassium peroxymonosulfate activation of WO3/diatomite composites for efficient photocatalytic degradation of tetracycline. RSC Adv. 2024, 14, 25019–25030. [Google Scholar] [CrossRef]
  14. Xu, M.; Yang, J.K.; Sun, C.Y.; Liu, L.; Cui, Y.; Liang, B. Performance enhancement strategies of bi-based photocatalysts: A review on recent progress. Chem. Eng. J. 2020, 389, 124402. [Google Scholar] [CrossRef]
  15. Chawla, H.; Chandra, A.; Ingole, P.P.; Garg, S. Recent advancements in enhancement of photocatalytic activity using bismuth-based metal oxides Bi2MO6 (M = W, Mo, Cr) for environmental remediation and clean energy production. J. Ind. Eng. Chem. 2021, 95, 1–15. [Google Scholar] [CrossRef]
  16. Chen, T.; Liu, L.Z.; Hu, C.; Huang, H.W. Recent advances on Bi2WO6-based photocatalysts for environmental and energy applications. Chin. J. Catal. 2021, 42, 1413–1438. [Google Scholar] [CrossRef]
  17. Wu, S.J.; Sun, J.G.; Li, Q.; Hood, Z.D.; Yang, S.; Su, T.M.; Peng, R.; Wu, Z.; Sun, W.W.; Kent, P.R.C.; et al. Effects of surface terminations of 2D Bi2WO6 on photocatalytic hydrogen evolution from water splitting. ACS Appl. Mater. Interfaces 2020, 12, 20067–20074. [Google Scholar] [CrossRef]
  18. Jiang, Z.Y.; Liang, X.Z.; Zheng, H.L.; Liu, Y.Y.; Wang, Z.Y.; Wang, P.; Zhang, X.Y.; Qin, X.Y.; Dai, Y.; Whangbo, M.H.; et al. Photocatalytic reduction of CO2 to methanol by three-dimensional hollow structures of Bi2WO6 quantum dots. Appl. Catal. B-Environ. 2017, 219, 209–215. [Google Scholar] [CrossRef]
  19. Song, S.J.; Xing, Z.P.; Zhao, H.A.; Li, Z.Z.; Zhou, W. Recent advances in bismuth-based photocatalysts: Environment and energy applications. Green Energy Environ. 2023, 8, 1232–1264. [Google Scholar] [CrossRef]
  20. Yu, G.L.; Sun, Q.F.; Yang, Y.; Chen, S.; Long, Y.N.; Li, Y.F.; Ge, S.Y.; Zheng, D. BiOCl-based composites for photocatalytic degradation of antibiotics: A review of synthesis method, modification, factors affecting photodegradation and toxicity assessment. J. Alloy. Compd. 2024, 981, 173733. [Google Scholar] [CrossRef]
  21. Sánchez-Rodríguez, D.; Jasso-Salcedo, A.B.; Hedin, N.; Church, T.L.; Aizpuru, A.; Escobar-Barrios, V.A. Semiconducting nanocrystalline bismuth oxychloride (BiOCl) for photocatalytic reduction of CO2. Catalysts 2020, 10, 998. [Google Scholar] [CrossRef]
  22. Cao, J.Y.; Li, J.J.; Chu, W.; Cen, W.L. Facile synthesis of Mn-doped BiOCl for metronidazole photodegradation: Optimization, degradation pathway, and mechanism. Chem. Eng. J. 2020, 400, 125813. [Google Scholar] [CrossRef]
  23. Wang, C.Y.; Zhang, Y.J.; Wang, W.K.; Pei, D.N.; Huang, G.X.; Chen, J.J.; Zhang, X.; Yu, H.Q. Enhanced photocatalytic degradation of bisphenol A by Co-doped BiOCl nanosheets under visible light irradiation. Appl. Catal. B-Environ. 2018, 221, 320–328. [Google Scholar] [CrossRef]
  24. Cao, J.Y.; Cen, W.L.; Jing, Y.; Du, Z.X.; Chu, W.; Li, J.J. P-doped BiOCl for visible light photodegradation of tetracycline: An insight from experiment and calculation. Chem. Eng. J. 2022, 435, 134683. [Google Scholar] [CrossRef]
  25. Yang, B.; He, L.P.; Guo, Y.; Zhao, T.; Shi, Z.Z.; Liu, C.X.; Yang, M.; Yang, Q.; Sun, S.D.; Cui, J. Enhanced catalysis for degradation of antibiotic by hydroxyl-functionalized S-doped BiOCl with high capacity of local spatial charge separation. Colloid Surf. A-Physicochem. Eng. Asp. 2023, 669, 131448. [Google Scholar] [CrossRef]
  26. Hu, L.J.; Ding, Z.C.; Yan, F.; Li, K.; Feng, L.; Wang, H.Q. Construction of hexagonal prism-like defective BiOCl hierarchitecture for photocatalytic degradation of tetracycline hydrochloride. Nanomaterials 2022, 12, 2700. [Google Scholar] [CrossRef]
  27. He, X.H.; Kai, T.H.; Ding, P. Heterojunction photocatalysts for degradation of the tetracycline antibiotic: A review. Environ. Chem. Lett. 2021, 19, 4563–4601. [Google Scholar] [CrossRef]
  28. Li, S.J.; Cai, M.J.; Liu, Y.P.; Wang, C.C.; Yan, R.Y.; Chen, X.B. Constructing Cd0.5Zn0.5S/Bi2WO6 S-scheme heterojunction for boosted photocatalytic antibiotic oxidation and Cr(VI) reduction. Adv. Powder Mater. 2023, 2, 100073. [Google Scholar] [CrossRef]
  29. Fenelon, E.; Bui, D.P.; Tran, H.H.; You, S.J.; Wang, Y.F.; Cao, T.M.; Van Pham, V. Straightforward synthesis of SnO2/Bi2S3/BiOCl-Bi24O31Cl10 composites for drastically enhancing rhodamine B photocatalytic degradation under visible light. ACS Omega 2020, 5, 20438–20449. [Google Scholar] [CrossRef]
  30. Cheng, Y.; Guo, Z.; Wu, C.; Zuo, H.; Liu, J.; Chang, X.; Yan, Q. Construction of BiOCl/Bi2WO6 Z-scheme heterojunction with close interfacial contact using CNT as electron medium. Colloid Surf. A-Physicochem. Eng. Asp. 2024, 681, 132847. [Google Scholar] [CrossRef]
  31. Zheng, Y.; Sun, Y. Construction of a flower-like S-scheme Bi2WO6/BiOCl nano-heterojunction with enhanced visible-light photocatalytic properties. New J. Chem. 2022, 46, 21418–21427. [Google Scholar] [CrossRef]
  32. Xu, M.; Lu, M.; Yang, Y.; Ai, L.; Fan, H.; Guo, N.; Wang, L. Efficient degradation of pollutants by Bi2WO6/BiOCl heterojunction activated peroxymonosulfate: Performance and mechanism. J. Environ. Chem. Eng. 2024, 12, 112156. [Google Scholar] [CrossRef]
  33. Sun, Z.H.; Guo, J.J.; Zhu, S.M.; Mao, L.; Ma, J.; Zhang, D. A high-performance Bi2WO6-graphene photocatalyst for visible light-induced H2 and O2 generation. Nanoscale 2014, 6, 2186–2193. [Google Scholar] [CrossRef]
  34. Zhou, H.R.; Wen, Z.P.; Liu, J.; Ke, J.; Duan, X.G.; Wang, S.B. Z-scheme plasmonic Ag decorated WO3/Bi2WO6 hybrids for enhanced photocatalytic abatement of chlorinated-VOCs under solar light irradiation. Appl. Catal. B-Environ. 2019, 242, 76–84. [Google Scholar] [CrossRef]
  35. Ma, Y.; Chen, Z.; Qu, D.; Shi, J. Synthesis of chemically bonded BiOCl@Bi2WO6 microspheres with exposed (020) Bi2WO6 facets and their enhanced photocatalytic activities under visible light irradiation. Appl. Surf. Sci. 2016, 361, 63–71. [Google Scholar] [CrossRef]
  36. Zhang, L.Y.; Zhang, J.J.; Yu, H.G.; Yu, J.G. Emerging S-scheme photocatalyst. Adv. Mater. 2022, 34, 2107668. [Google Scholar] [CrossRef]
  37. Song, J.M.; Mao, C.J.; Niu, H.L.; Shen, Y.H.; Zhang, S.Y. Hierarchical structured bismuth oxychlorides: Self-assembly from nanoplates to nanoflowers via a solvothermal route and their photocatalytic properties. Crystengcomm 2010, 12, 3875–3881. [Google Scholar] [CrossRef]
  38. Zhong, X.; Wu, W.T.; Jie, H.N.; Tang, W.Y.; Chen, D.Y.; Ruan, T.; Bai, H.P. Degradation of norfloxacin by copper-doped Bi2WO6-induced sulfate radical-based visible light-Fenton reaction. RSC Adv. 2020, 10, 38024–38032. [Google Scholar] [CrossRef]
  39. Kan, L.; Chang, C.; Wang, Q.; Wang, X. Glycol assisted splitting BiOIO3 into plasmonic bismuth coupled with BiOI co-modified Bi2WO6 (BiOI/Bi/Bi2WO6) to form indirect Z-scheme heterojunction for efficient photocatalytic degradation of BPA. Sep. Purif. Technol. 2022, 297, 121537. [Google Scholar] [CrossRef]
  40. Sing, K.S.W. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Recommendations 1984). Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  41. Fu, R.R.; Zeng, X.Q.; Ma, L.; Gao, S.M.; Wang, Q.Y.; Wang, Z.Y.; Huang, B.B.; Dai, Y.; Lu, J. Enhanced photocatalytic and photoelectrochemical activities of reduced TiO2-x/BiOCl heterojunctions. J. Power Sources 2016, 312, 12–22. [Google Scholar] [CrossRef]
  42. Wu, J.; Zhang, W.; Tian, Z.; Zhao, Y.; Shen, Z. Facile fabrication of Bi2WO6/BiOCl hierarchical structure as adsorbents for methylene blue dye removal. Mater. Res. Express 2019, 6, 055034. [Google Scholar] [CrossRef]
  43. Yang, J.; Yang, L.; Fang, M.; Li, L.; Fu, F.; Xu, H.; Li, M.; Fan, X. A compact Z-scheme heterojunction of BiOCl/Bi2WO6 for efficiently photocatalytic degradation of gaseous toluene. J. Colloid Interface Sci. 2023, 631, 44–54. [Google Scholar] [CrossRef]
  44. Pham, M.-T.; Tran, H.-H.; Nguyen, T.-M.T.; Bui, D.-P.; Huang, Y.; Cao, J.; You, S.-J.; Van Viet, P.; Nam, V.H.; Wang, Y.-F. Revealing DeNOx and DeVOC reactions via the study of the surface and bandstructure of ZnSn(OH)6 photocatalysts. Acta Mater. 2021, 215, 117068. [Google Scholar] [CrossRef]
  45. Pham, M.T.; Nguyen, T.M.T.; Bui, D.P.; Wang, Y.F.; Tran, H.H.; You, S.J. Enhancing quantum efficiency at Ag/g-C3N4 interfaces for rapid removal of nitric oxide under visible light. Sustain. Chem. Pharm. 2022, 25, 9. [Google Scholar] [CrossRef]
  46. Zhang, C.; Li, Y.; Li, M.Q.; Shuai, D.M.; Zhou, X.Y.; Xiong, X.Y.; Wang, C.; Hu, Q. Continuous photocatalysis via photo-charging and dark-discharging for sustainable environmental remediation: Performance, mechanism, and influencing factors. J. Hazard. Mater. 2021, 420, 126607. [Google Scholar] [CrossRef]
  47. Bognár, S.; Putnik, P.; Maksimovic, I.; Velebit, B.; Putnik-Delic, M.; Merkulov, D.S. Sustainable removal of tolperisone from waters by application of photocatalysis, nanotechnology, and chemometrics: Quantification, environmental toxicity, and degradation optimization. Nanomaterials 2022, 12, 4199. [Google Scholar] [CrossRef]
  48. Jovanović, D.; Orčić, D.; Šojić Merkulov, D.; Despotović, V.; Finčur, N. Study of factors impacting the light-driven removal of ciprofloxacin (Ciprocinal®) and identification of degradation products using LC–ESI–MS2. J. Photochem. Photobiol. A-Chem. 2025, 460, 116119. [Google Scholar] [CrossRef]
  49. Ren, H.T.; Qi, F.; Labidi, A.; Zhao, J.J.; Wang, H.; Xin, Y.; Luo, J.M.; Wang, C.Y. Chemically bonded carbon quantum dots/Bi2WO6 S-scheme heterojunction for boosted photocatalytic antibiotic degradation: Interfacial engineering and mechanism insight. Appl. Catal. B-Environ. 2023, 330, 122587. [Google Scholar] [CrossRef]
  50. Guo, Y.; Ao, Y.H.; Wang, P.F.; Wang, C. Mediator-free direct dual-Z-scheme Bi2S3/BiVO4/MgIn2S4 composite photocatalysts with enhanced visible-light-driven performance towards carbamazepine degradation. Appl. Catal. B-Environ. 2019, 254, 479–490. [Google Scholar] [CrossRef]
  51. Yuan, Y.; Guo, R.T.; Hong, L.F.; Ji, X.Y.; Lin, Z.D.; Li, Z.S.; Pan, W.G. A review of metal oxide-based Z-scheme heterojunction photocatalysts: Actualities and developments. Mater. Today Energy 2021, 21, 100829. [Google Scholar] [CrossRef]
Figure 1. SEM images of (ad) BiOCl, (eh) Bi2WO6, and (il) BiOCl/Bi2WO6.
Figure 1. SEM images of (ad) BiOCl, (eh) Bi2WO6, and (il) BiOCl/Bi2WO6.
Separations 12 00111 g001aSeparations 12 00111 g001b
Figure 2. (ac) TEM images, (d) HRTEM image, (e) fast Fourier transform lattice stripe enlargements, and (f) EDX spectra of BiOCl/Bi2WO6.
Figure 2. (ac) TEM images, (d) HRTEM image, (e) fast Fourier transform lattice stripe enlargements, and (f) EDX spectra of BiOCl/Bi2WO6.
Separations 12 00111 g002aSeparations 12 00111 g002b
Figure 3. PL spectra of BiOCl, Bi2WO6, and BiOCl/Bi2WO6.
Figure 3. PL spectra of BiOCl, Bi2WO6, and BiOCl/Bi2WO6.
Separations 12 00111 g003
Figure 4. (a) Photocatalytic degradation curves of TCH, (b) pseudo-first-order kinetic fitting curves for Bi2WO6, BiOCl, X-BOC/BWO (X = 10, 20, 30, and 40), and 20-BOC/BWO.
Figure 4. (a) Photocatalytic degradation curves of TCH, (b) pseudo-first-order kinetic fitting curves for Bi2WO6, BiOCl, X-BOC/BWO (X = 10, 20, 30, and 40), and 20-BOC/BWO.
Separations 12 00111 g004
Figure 5. EPR signals of (a) TEMPO-h+ and (b) DMPO- · O 2 for 20-BOC/BWO.
Figure 5. EPR signals of (a) TEMPO-h+ and (b) DMPO- · O 2 for 20-BOC/BWO.
Separations 12 00111 g005
Figure 6. (a) Effect of different scavengers on the photodegradation of TCH by 20-BOC/BWO and (b) XPS valence band spectra of BiOCl and Bi2WO6 photocatalysts.
Figure 6. (a) Effect of different scavengers on the photodegradation of TCH by 20-BOC/BWO and (b) XPS valence band spectra of BiOCl and Bi2WO6 photocatalysts.
Separations 12 00111 g006
Figure 7. The charge-transfer mechanism of BiOCl/Bi2WO6 under visible-light irradiation: (a) a traditional type-II mechanism and (b) a direct Z-scheme mechanism.
Figure 7. The charge-transfer mechanism of BiOCl/Bi2WO6 under visible-light irradiation: (a) a traditional type-II mechanism and (b) a direct Z-scheme mechanism.
Separations 12 00111 g007
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

Zhang, H.; Zhu, Z.; Huang, Y.; Yu, J.; Li, M. Construction of Z-Scheme Heterojunction BiOCl/Bi2WO6 for Visible-Light Photocatalytic Degradation of Tetracycline Hydrochloride. Separations 2025, 12, 111. https://doi.org/10.3390/separations12050111

AMA Style

Zhang H, Zhu Z, Huang Y, Yu J, Li M. Construction of Z-Scheme Heterojunction BiOCl/Bi2WO6 for Visible-Light Photocatalytic Degradation of Tetracycline Hydrochloride. Separations. 2025; 12(5):111. https://doi.org/10.3390/separations12050111

Chicago/Turabian Style

Zhang, Hetian, Zengying Zhu, Yajie Huang, Jiaxing Yu, and Ming Li. 2025. "Construction of Z-Scheme Heterojunction BiOCl/Bi2WO6 for Visible-Light Photocatalytic Degradation of Tetracycline Hydrochloride" Separations 12, no. 5: 111. https://doi.org/10.3390/separations12050111

APA Style

Zhang, H., Zhu, Z., Huang, Y., Yu, J., & Li, M. (2025). Construction of Z-Scheme Heterojunction BiOCl/Bi2WO6 for Visible-Light Photocatalytic Degradation of Tetracycline Hydrochloride. Separations, 12(5), 111. https://doi.org/10.3390/separations12050111

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