Next Article in Journal
Review on Preparation of Perovskite Solar Cells by Pulsed Laser Deposition
Previous Article in Journal
Molten Bismuth–Bismuth/Zinc Oxide Composites for High-Temperature Thermal Energy Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of Zn0.5Cd0.5S/Bi4O5Br2 Heterojunction for Enhanced Photocatalytic Degradation of Tetracycline Hydrochloride

1
School of Materials and Chemistry, Southwest University of Science and Technology, Mianyang 621010, China
2
State Key Laboratory of Environmental-Friendly Energy Materials, Southwest University of Science and Technology, Mianyang 621010, China
*
Author to whom correspondence should be addressed.
Inorganics 2024, 12(5), 127; https://doi.org/10.3390/inorganics12050127
Submission received: 23 March 2024 / Revised: 16 April 2024 / Accepted: 19 April 2024 / Published: 24 April 2024
(This article belongs to the Section Inorganic Materials)

Abstract

:
The development of efficient catalysts with visible light response for the removal of pollutants in an aqueous environment has been a hotspot in the field of photocatalysis research. A Zn0.5Cd0.5S (ZCS) nanoparticle/Bi4O5Br2 ultra-thin nanosheet heterojunction was constructed by ultrasound-assisted solvothermal method. The morphology, structure, and optoelectronic properties of the composite were characterized by scanning electron microscopy, X-ray diffraction, X-ray photoelectron spectroscopy, and UV–vis diffuse reflectance spectra. Under simulated visible light illumination, the photocatalytic performance was evaluated through degradation of tetracycline hydrochloride. Results show that the degradation effect by the optimum ZCS/Bi4O5Br2 catalyst is superior to pure materials with the kinetic constant that is 1.7 and 9.6 times higher than those of Bi4O5Br2 and ZCS, and also has better stability and reusability. Trapping experiments and electron paramagnetic resonance tests find that free radicals in the photocatalytic system are superoxide radicals and holes. This work provides a referable idea for the development of more efficient and recyclable photocatalysts.

Graphical Abstract

1. Introduction

Tetracycline hydrochloride (TCH) is widely used in medicine and agriculture as a broad-spectrum antibiotic. It is stable and difficult to degrade naturally, and its overuse poses a serious threat to the environment and human beings [1,2]. The increasing amounts of various types of organic dyes discharged from industrial wastewater also pose a great challenge for water purification [3]. Conventional treatment technologies have difficulty removing pollutants in an efficient and environmentally friendly manner. In recent years, photocatalytic technology has emerged as a potential strategy for degrading hazardous pollutants [4,5]. Semiconductor photocatalytic materials have a strong redox capacity and are capable of converting most organic pollutants into H2O, CO2, and other small inorganic molecules for purifying harmful substances. Therefore, developing a new efficient photocatalyst is particularly important.
Bismuth oxyhalide BiOX (X = Cl, Br, I) materials have become a research hotspot in the field of photocatalysis due to their unique layered structure and superior photocatalytic performance [6]. However, their positive conduction band position (CB) and poor carrier separation ability limit the practical applications [7]. Density-functional-transfer (DFT) calculations of BiOBr show that its valence band is hybridized by O 2p and Br 4p orbitals, and the conduction band consists of Bi 6p orbitals; thus, its energy band position and band gap width can be changed by adjusting the ratio of Bi, O, and Br [8]. Bi4O5Br2 constructed by this bismuth-rich strategy retains the 2D layered structure of BiOBr but has a narrower band gap, stronger photogenerated carrier separation efficiency, and better photocatalytic performance [9,10]. However, the photocatalytic role of pure Bi4O5Br2 is limited, with low visible light utilization efficiency and high complexation rate of a photogenerated electron–hole pair [11,12]. Therefore, modifications are needed to improve the catalytic activity, including changing the morphology [13], constructing heterojunctions [14,15,16], doping [17,18], and introducing surface defects [19,20]. Selecting ideal semiconductors with suitable energy band structure to construct heterojunctions with excellent photocatalytic performance is of great significance. For example, Zhang et al. constructed In2O3/Bi4O5Br2 S-scheme heterojunctions with interfacial oxygen vacancies using a hydrothermal and chemical precipitation method. The obtained catalyst was capable of degrading tetracycline efficiently [21]. Qian et al. synthesized a 2D/2D Bi4O5Br2/Bi2WO6 Z-scheme heterojunction by chemical co-precipitation and hydrothermal method, which was effective for the stable removal of ciprofloxacin [22]. Therefore, rationally designing heterojunction materials is an effective strategy for improving the photocatalytic activity of Bi4O5Br2 while preserving the redox capacity.
In recent years, ZnxCd1−xS solid solution with excellent photocatalytic properties has been recognized as a promising semiconductor material. One of the highlights is that its forbidden bandwidth and optical response range can be flexibly tuned by varying the molar ratio of Zn2+ and Cd2+ [23]. Among the series of zinc cadmium sulfide solid solutions, ZCS has been the most extensively researched and widely used for photocatalytic hydrogen precipitation [24,25], degradation of pollutants [26,27], and carbon dioxide reduction [28]. However, pure ZCS has some limitations, such as low efficiency of photogenerated electron–hole pair separation and photocorrosion [29]. These defects can be compensated by constructing heterojunctions with other semiconductors to improve the photocatalytic activity.
Herein, a series of ZCS/Bi4O5Br2 heterojunctions were constructed by using a facile ultrasound-assisted solvothermal method. The prepared composite catalysts exhibit the following properties: (1) The ZCS/Bi4O5Br2 heterojunction forms a built-in electric field and accelerates the transport of photogenerated carriers at the interface. (2) Bi4O5Br2 serves as an oxidation semiconductor to capture the holes of ZCS and strengthen the photocatalytic stability, thereby slightly mitigating the photocorrosion of ZCS nanoparticles. The photocatalytic performance of the material was optimized by appropriately adjusting the ratio of ZCS and Bi4O5Br2, which showed an obvious removal effect for several pollutants (TCH, OTC, RhB, and MB). The stability and reusability were demonstrated by cycling tests, and possible TCH degradation pathways and mechanistic predictions were suggested.

2. Results and Discussion

2.1. Material Characterization

The samples were characterized by X-ray diffraction (XRD), as displayed in Figure 1. The diffraction peaks at 24.23°, 29.58°, 31.82°, and 45.65° are observed in pure Bi4O5Br2, corresponding to the (112), (11-3), (020), and (422) crystal planes of the monoclinic phase (JCPDS No. 37-0699), respectively [30]. The peaks between cubic CdS and ZnS appearing at 27.44°, 45.49°, and 53.67° belong to the (111), (220), and (311) crystallographic planes of ZCS, respectively, which indicates the successful synthesis of ZCS [31,32]. The main diffraction peaks of as-prepared ZCS/Bi4O5Br2 composites clearly correspond to Bi4O5Br2 and ZCS without the appearance of other impurity peaks. With the gradual introduction of ZCS, the peak intensity of ZCS in the composite catalysts was strengthened, while the crystal surface strength of Bi4O5Br2 was weakened and then increased. This represented the gradual loading of ZCS to Bi4O5Br2, and the materials were successfully composited.
The microscopic morphology and structure of ZCS, Bi4O5Br2, and ZCS/Bi4O5Br2 were observed by scanning electron microscopy (SEM), transmission electron microscopy (TEM), and high-resolution TEM (HRTEM). Figure 2a shows the morphology of ZCS, which consists of small, aggregated nano-microspheres. The pure Bi4O5Br2 exhibits irregular stacked nanosheets with relatively smooth surfaces, as shown in Figure 2b. A scanning image of 1.5-ZCS/Bi4O5Br2 composite (Figure 2c) shows that ZCS nanoparticles are uniformly loaded on Bi4O5Br2 sheet structures, and the surfaces become rough. EDS and elemental mapping images (Figure 3) confirm the uniformly distributed elements of Zn, Cd, S, Bi, Br, and O in 1.5-ZCS/Bi4O5Br2.
The morphology of nanoparticles and ultra-thin nanosheets was similarly detected in TEM images (Figure 4a), with ZCS nanoparticles dispersed on the surface of Bi4O5Br2 nanosheets. The material was further analyzed by HRTEM (Figure 4b,c), the lattice fringes reveal that the lattice spacing of 0.302 nm corresponds to the (11-3) crystal plane of Bi4O5Br2, and 0.335 nm matches with the (111) crystal plane of ZCS from the boxed section. The result is also matched with XRD, further demonstrating the successful synthesis of the ZCS/Bi4O5Br2 heterojunction interface.
The elemental compositions and valence states were analyzed using X-ray photoelectron spectroscopy (XPS). The full spectrum of 1.5-ZCS/Bi4O5Br2 shown in Figure 5a reveals the presence of Bi, O, Br, Zn, Cd, and S. Figure 5b displays that the two characteristic peaks Zn 2p1/2 and Zn 2p3/2 of the Zn 2p spectrum in 1.5-ZCS/Bi4O5Br2 are attributed to 1044.64 and 1021.61 eV, respectively, proving the presence of Zn2+ [33]. The Cd 3d peaks in 1.5-ZCS/Bi4O5Br2 at 411.38 and 404.37 eV, belonging to Cd 3d3/2 and Cd 3d5/2 of Cd2+ (Figure 5c), respectively [34]. In the Bi 4f spectrum shown in Figure 5d, Bi 4f5/2 and Bi 4f7/2 are attributed to 163.78 and 158.48 eV, respectively, indicating that Bi element exists in a +3 valence state [11]. Meanwhile, the S 2p peaks of S2− are observed from the spectrum with binding energies at 161.82 and 160.98 eV belonging to S 2p1/2 and S 2p3/2, respectively [35]. The result is consistent with the S 2p peaks in pure ZCS. The peaks appearing at 68.80 and 67.80 eV in Figure 5e correspond to Br 3d3/2 and Br 3d5/2, respectively, confirming the existence of Br [36]. Additionally, the O 1s peaks (Figure 5f) can be fitted to three peaks at 532.25, 530.76, and 529.26 eV related to Bi-O bond and surface-adsorbed oxygen [37]. Compared with pure Bi4O5Br2 or ZCS, the binding energies of Bi, Br, and O from 1.5-ZCS/Bi4O5Br2 are slightly negatively shifted possibly due to the escape of electrons, and the elements involving Zn, Cd, and S are positively shifted due to the increase of the surrounding electron density, indicating the electronic interaction between Bi4O5Br2 and ZCS [38].
The N2 adsorption–desorption isotherms and pore size distributions of the corresponding samples are presented in Figure 6. The curves of Bi4O5Br2 and 1.5-ZCS/Bi4O5Br2 are clearly consistent with type IV isotherm, indicating the presence of mesoporous structure of the materials. Evidently, 1.5-ZCS/Bi4O5Br2 (68.7 m2/g, 0.232 cm3/g) has a higher specific surface area and pore volume than Bi4O5Br2 (24.6 m2/g, 0.167 cm3/g). The result demonstrates that the introduction of ZCS can significantly increase the specific surface area of Bi4O5Br2, thus providing more catalytically active sites with enhanced adsorption capacity and photocatalytic performance.

2.2. Optoelectronic Properties

The light absorption capacities and energy band structures of ZCS, Bi4O5Br2, and ZCS/Bi4O5Br2 were revealed by UV–vis diffuse reflectance spectroscopy (DRS). Figure 7a shows the diffuse reflection spectra of the obtained samples in the region of 200–800 nm. The absorption edges of ZCS and Bi4O5Br2 are located around 519 and 538 nm, respectively. Compared with pure Bi4O5Br2, the composite photocatalysts exhibit enhanced light absorption intensity and a significantly red-shifted characteristic of the absorption edges, indicating that the combination of ZCS and Bi4O5Br2 can expand the absorption range of sunlight [39]. Results demonstrates that the effective complexation of ZCS with Bi4O5Br2 can improve the absorption capacity and utilization of visible light. The energy bandgap (Eg) values of Bi4O5Br2 (2.65 eV) and ZCS (2.43 eV) are calculated in Figure 7b according to the Kubelka–Munk formula [40]:
( α hv ) = A ( hv E g ) n / 2
The bandgap value of 1.5-ZCS/Bi4O5Br2 is slightly reduced, which is roughly at 2.38 eV. The narrow band gap made it easier for electrons to jump, and the photocatalytic activity was enhanced. The flat band potential positions −0.49 V (vs. Ag/AgCl) of Bi4O5Br2 and −0.88 V (vs. Ag/AgCl) of ZCS are obtained from the tangent intercepts of Mott–Schottky curves, respectively (Figure 7c,d). Converted to standard hydrogen electrode potentials according to the following formula [26]:
E NHE =   E Ag / AgCl + 0.198
The flat band potentials for Bi4O5Br2 and ZCS are −0.29 and −0.68 V (vs. NHE), respectively. The tangent slopes of M–S curves are both positive, indicating that ZCS and Bi4O5Br2 are typical n-type semiconductors [41]. For n-type semiconductors, the flat band is generally 0.1 or 0.2 V higher than the conduction band (CB) potential, so the CB positions of Bi4O5Br2 and ZCS are −0.49 and −0.88 V (vs. NHE), respectively. Finally, applying the following equation [42]:
E VB = E CB + E g
The valence band (VB) potentials of Bi4O5Br2 and ZCS are calculated to be 2.16 and 1.55 V, respectively.
The separation efficiency of photogenerated carriers in semiconductors was examined by photoluminescence (PL) spectroscopy. PL spectra of Bi4O5Br2, ZCS, and 1.5-ZCS/Bi4O5Br2 heterojunction measured at the excitation light wavelength of 325 nm are shown in Figure 8a. The fluorescence intensity of 1.5-ZCS/Bi4O5Br2 is weaker than that of uncompounded photocatalysts, illustrating that the possibility of its photogenerated carrier recombination becomes significantly lower [43]. The transfer and separation efficiency of light-induced carriers were further evaluated by impedance tests. Figure 8b displays the electrical impedance spectra (EIS) of Bi4O5Br2, ZCS, and 1.5-ZCS/Bi4O5Br2 heterojunction. The impedance arc radius of 1.5-ZCS/Bi4O5Br2 is obviously smaller than that of ZCS and Bi4O5Br2, which suggests that the introduction of ZCS can reduce the resistance of the material. Evidently, ZCS/Bi4O5Br2 composite catalysts showed low charge migration resistance and fast electron transport.

2.3. Photocatalytic Performance

The photocatalytic performance of ZCS, Bi4O5Br2, and ZCS/Bi4O5Br2 was evaluated by degrading TCH. The adsorption–desorption equilibrium experiment was first conducted, as shown in Figure S1. After stirring in the dark for 30 min, the concentration of TCH remains essentially constant, indicating that equilibrium had been reached. In Figure 9a, there is almost no degradation that occurred without the presence of the photocatalyst, proving that TCH is very stable in the natural environment. The degradation rate by pure Bi4O5Br2 is 65.8% within 60 min of irradiation, while the degradation rate of ZCS is 14.4%. However, the photocatalytic activity of ZCS/Bi4O5Br2 heterojunctions improved significantly. Among them, 1.5-ZCS/Bi4O5Br2 shows the highest degradation rate of 84.5%, which is 1.3 and 5.9 times that of Bi4O5Br2 and ZCS, respectively. In Figure 9b,c, the removal process of TCH is further analyzed applying the pseudo-first-order kinetic model. It is calculated that 1.5-ZCS/Bi4O5Br2 (0.02189 min−1) has the highest apparent rate constant, which is about 1.7 and 9.6 times higher than that of pure Bi4O5Br2 (0.01297 min−1) and ZCS (0.00229 min−1), respectively. The above findings corroborate that the effective combination of ZCS and Bi4O5Br2 can improve the photocatalytic degradation performance. From the UV–vis absorption spectra (Figure 9d), it can be observed that the intensity of the main characteristic peak of TCH diminishes with time, proving the occurrence of degradation. The reported photocatalytic materials for degrading TCH are displayed in Table 1, and it can be concluded from the comparison that ZCS/Bi4O5Br2 catalyst has advantages in degrading TCH. Further, the suitability of 1.5-ZCS/Bi4O5Br2 for different pollutants was further investigated. Figure 9e demonstrates that the degradation efficiency of OTC, RhB, and MB can reach 83.3%, 97.7%, and 97.3%, respectively, implying its potential for the removal of contaminants.
In practice application, the situation was complex and variable. Thus, the effect of different conditions on the degradation of ZCS/Bi4O5Br2 heterojunction was explored, including the initial pH and concentration of TCH solution, as well as the catalyst dosage. As shown in Figure S2a, 1.5-ZCS/Bi4O5Br2 catalyst has the best photodegradation ability at a neutral initial pH value, and also maintains favorable TCH removal under different pH gradients, indicating a wider pH application range. Figure S2b shows the removal rate of TCH with different catalyst dosages. The dosage increases from 0.2 to 0.5 g/L, and the degradation rate increases from 78.0% to 86.0%. More pollutants can be adsorbed to the catalyst surface to participate in the photoreaction, and more reactive sites can be provided. The differences in TCH removal effects were not significant when 1.5-ZCS/Bi4O5Br2 dosage was 0.4 and 0.5 g/L, respectively. Therefore, 0.4 g/L was the optimal catalyst dosage. Figure S2c shows the degradation efficiency at different TCH concentrations. The concentrations range from 10 to 40 mg/L, and the degradation rates are 87.3%, 84.5%, 80.2%, and 75.5%, respectively. As the pollutant concentration increased, the removal rate decreased, which was due to the accumulation of substrates and intermediates masking some active sites on the surface of catalyst. And the active radicals produced by a certain amount of catalyst were also limited. Considering the practical application and economic value, 20 mg/L of TCH solution was selected for the following experiments.
The stability and reusability of the photocatalyst are also an important evaluation index. The solution after each round of degradation was filtered off the supernatant, and the catalyst was collected by centrifugation, washed, and dried for the next round of photodegradation experiments. As illustrated in Figure 10a, the photocatalytic activity of 1.5-ZCS/Bi4O5Br2 decreases slightly after five cycles of degradation experiments but is still superior to that of the pure ZCS and Bi4O5Br2. The result indicates that the catalyst has a long service life. Peak structures of XRD pattern (Figure 10b) before and after cycling remain basically the same, which also demonstrates the stability of its crystalline phase structure.

2.4. Analysis of Possible Degradation Pathways

The intermediate products in the degradation of TCH were identified using LC-MS. On the basis of the mass spectrometry m/z results (Figure S3), the corresponding intermediates and two possible degradation pathways are deduced. The strong peak at m/z = 445 corresponds to the mass-to-charge ratio of TCH. In pathway I, the introduction of a hydroxyl group produces the intermediate P1 (m/z = 459), followed by N-demethylation and hydroxylation occurring with the detection of P2 (m/z = 417). P2 dissociates to P3 (m/z = 274) via ring-opening and reduction reaction. Furthermore, P2 is dehydrated to produce P4 (m/z = 399) and then degrades to P5 (m/z = 338) by ring-opening, demethylation, and hydroxylation reactions. In pathway II, TCH is degraded to the intermediate P6 (m/z = 391) via deamination reaction and hydroxylation, and then by a ring-opening reaction to produce P7 (m/z = 234), which undergoes a dehydration process and other reactions to produce P8 (m/z = 217). These intermediate substances may be further degraded to obtain smaller masses of organic molecules (P9–P11, m/z = 194, 171, and 114), finally mineralizing to CO2 and H2O. The postulated degradation pathways are shown in Figure 11.

2.5. Photocatalytic Mechanism

During the photocatalysis process, organic pollutants were degraded through the generation of active species including ·O2, h+, and ·OH by the catalysts. Corresponding scavengers such as para-benzoquinone (BQ, ·O2 scavenger), ethylenediaminetetraacetic acid disodium salt (EDTA-2Na, h+ scavenger), and isopropanol (IPA, ·OH scavenger) were added during the degradation of TCH to identify the main active substances by trapping experiments. As depicted in Figure 12a, the degradation rates decrease from 84.5% to 44.5% (·O2), 61.6% (h+), and 83.5% (·OH), indicating that ·O2 and h+ are the most important active radicals. The EPR tests further confirmed the production of ·O2 and h+ in the photocatalytic system. In Figure 12b, the signal of DMPO-·O2 could not be recognized under dark condition for 1.5-ZCS/Bi4O5Br2. The EPR peak intensity increases after 15 min of light, and the signal is obviously enhanced. The signal generated under dark condition belongs to the capture agent TEMPO, and the intensity of the EPR peak decreases after providing light instead, precisely because the generation of h+ neutralizes the signal from TEMPO, making the signal weaker (Figure 12c).
The CB and VB positions of ZCS are −0.88 and 1.55 V, while Bi4O5Br2 are −0.49 and 2.16 V, respectively. Both semiconductors have more negative potentials and can produce ·O2 (O2/·O2, −0.33 V vs. NHE) [48]. The redox capacity of Bi4O5Br2 enables H2O molecules to produce ·OH (H2O/·OH, 1.99 V vs. NHE) [49], but ·OH (OH/·OH, 2.27 V vs. NHE) derived from OH cannot be generated [50]. Thus, ·OH had minimal participation in the degradation of TCH, which matched the trapping experimental results. Under visible light excitation, photogenerated electrons and holes were produced in the VB and CB of ZCS and Bi4O5Br2, respectively. If the photocatalytic mechanism followed the traditional type II heterojunction transfer theory, ·O2 would not be generated due to the more positive CB of ZCS than ·O2 (O2/·O2), and the VB of Bi4O5Br2 would be more negative than ·OH (·OH/H2O) or ·OH (·OH/OH) and unable to produce ·OH, which was contrary to the trapping experiments. Combining the above analyses, the transfer mechanism could be hypothesized in Figure 13. When the two semiconductors contacted closely, the electrons of ZCS spontaneously flowed to Bi4O5Br2 until the Fermi energy levels reached equilibrium. Similarly, electron consumption accelerates the transfer of photogenerated holes in the VB of Bi4O5Br2 to that of ZCS with a higher potential. At the heterojunction interface, Bi4O5Br2 was negatively charged and ZCS was positively charged, thus forming an internal electric field. The recombination of electrons in the CB of Bi4O5Br2 and holes in the VB of ZCS under Coulomb force and photoexcitation promoted photogenerated charge separation. Therefore, the soluble oxygen could be reduced to ·O2 with the presence of photogenerated electrons in the CB of ZCS, and the remaining highly oxidized h+ in the VB of Bi4O5Br2 was directly involved in the degradation of pollutants. The whole process is summarized as follows:
ZCS / Bi 4 O 5 Br 2 + h v     ZCS / Bi 4 O 5 Br 2 ( e , h + )
e ( Bi 4 O 5 Br 2 ) + h + ZCS   Recombination
e ( ZCS ) + O 2 · O 2
h + ( Bi 4 O 5 Br 2 ) + H 2 O   · OH + h +
· O 2 + h + + · OH + TCH CO 2 + H 2 O + Small   molecules

3. Experimental Section

3.1. Preparation of Zn0.5Cd0.5S

Zn0.5Cd0.5S was easily prepared by using the solvothermal method. Quantities of 0.3296 g of Zn(CH3COO)2·2H2O and 0.3998 g of Cd(CH3COO)2·2H2O were completely dissolved by ultrasonication with 40 mL of ethylene glycol solution, then Na2S solution was added dropwise according to the molar ratio of Zn: Cd: S = 1: 1: 3 and stirred for 4 h. Next, the mixed solution was poured into a 100 mL Teflon-lined autoclave at 160 °C for 4 h. The sample was washed with ethanol and deionized water repeatedly by centrifugation after natural cooling and dried under vacuum at 60 °C overnight to obtain ZCS nanoparticles.

3.2. Preparation of Bi4O5Br2

Bi4O5Br2 nanosheets were prepared according to a previous report with minor modifications [51]. Specifically, 0.7277 g of Bi(NO3)3·5H2O was added into 15 mL of ethylene glycol solution, and 0.1786 g of KBr was added into 10 mL of deionized water, dissolved by sonication for 0.5 h. Thereafter, the KBr solution was slowly dripped into Bi(NO3)3 solution, and the white colloid appeared gradually. The pH value was adjusted to about 9 by using dilute ammonia and continued stirring. Finally, the homogeneously mixed solution of Bi4O5Br2 precursor was kept at 140 °C for 12 h. The reaction product was washed and dried to obtain milky white Bi4O5Br2 powder.

3.3. Preparation of Zn0.5Cd0.5S/Bi4O5Br2

A simple ultrasound-assisted solvothermal method was used to synthesize Zn0.5Cd0.5S/Bi4O5Br2 composites. A total of 0.1000 g of Bi4O5Br2 was dispersed ultrasonically in 20 mL of ethanol to form a homogeneous solution. Next, ZCS was added at the mass ratio X:1 wt.% (X = 0.25, 0.50, 1.0, 1.5, 2.0, which was the mass ratio of ZCS and Bi4O5Br2) and sonicated for 1 h, followed by stirring for 2 h. The solutions were transferred into a 50 mL Teflon-lined autoclave, and 100 °C heat was employed for 10 h. After cooling, the composites were washed with ethanol and deionized water to obtain the final substance named X-ZCS/Bi4O5Br2 (X = 0.25, 0.50, 1.0, 1.5, 2.0). The whole preparation process is displayed in Scheme 1.

3.4. Characterization

Specific characterizations are detailed in the Supplementary Materials.

3.5. Photocatalytic Performance Measurements

The photocatalytic performance was evaluated by using TCH (20 mg/L), OTC, RhB, and MB (10 mg/L) as target pollutants, pH = 7 for all contaminant solutions, and a 300 W xenon lamp as a simulated light source. About 20 mg of catalyst was weighed and dispersed homogeneously into 50 mL of contaminant solution and stirred under dark condition for 0.5 h to reach adsorption–desorption equilibrium. Then, 4 mL of suspension was extracted at 15 min intervals after irradiation and centrifuged to test the maximum absorbance at 357 nm of the obtained supernatant with a Shimadzu UV-1900 spectrophotometer. The degradation rate is calculated via the following equation:
Removal   rate % = 1     C t / C 0   ×   100 %
where C0 is the initial concentration of the pollutant solution, and Ct is the concentration at a certain degradation time t (min).

4. Conclusions

In this work, a ZCS/Bi4O5Br2 heterojunction was successfully prepared. The better degradation properties of optimized ZCS/Bi4O5Br2 were attributed to the formation of an electric field inside the heterojunction, realizing the fast separation of carriers at the interface between ZCS and Bi4O5Br2 while also preserving their redox capabilities. Compared with ZCS (14.3%) and Bi4O5Br2 (65.8%), 1.5-ZCS/Bi4O5Br2 could degrade TCH up to 84.5% and exhibited excellent stability, as indicated by recycling experiments. Here, ·O2 and h+ played a dominant role during the elimination process. The suitability for other pollutants’ degradation was validated, and the results showed that it also had a better removal effect on OTC (83.3%), RhB (97.7%), and MB (97.3%). Therefore, this study provides a possible strategy for the modification of bismuth-rich materials and the preparation of efficient photocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics12050127/s1, Figure S1: the adsorption curve of 1.5-ZCS/Bi4O5Br2 under dark; Figure S2: effect of (a) initial pH of TCH solution, (b) catalyst dosage, and (c) pollutant concentration; Figure S3: LC-MS spectra of TCH degradation process during different irradiation times: (a) 0 min, (b) 30 min, (c) 60 min. In addition, specific characterization methods are placed in the supplementary materials.

Author Contributions

Conceptualization, methodology, validation, formal analysis, investigation, data curation, visualization and writing—original draft preparation, L.L.; resources, writing—review and editing, supervision, project administration and funding acquisition, J.S. and B.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the financial support received from the basic research project of Sichuan Province for Science and Technology Development (No. 2019YJ0355), Outstanding Youth Science and Technology Talents Program of Sichuan Province (No. 19JCQN0085), the Project of State Key Laboratory of Environment-friendly Energy Materials, Southwest University of Science and Technology (project No. 22fksy18).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lu, Z.Y.; Ma, Y.L.; Zhang, J.T.; Fan, N.S.; Huang, B.C.; Jin, R.C. A critical review of antibiotic removal strategies: Performance and mechanisms. J. Water Process Eng. 2020, 38, 101681. [Google Scholar] [CrossRef]
  2. Yang, J.; Chen, Y.; Xu, P.; Li, Y.; Jia, X.; Song, H. Fabrication of compressible and underwater superoleophobic carbon/g-C3N4 aerogel for wastewater purification. Mater. Lett. 2019, 254, 210–213. [Google Scholar] [CrossRef]
  3. Al-Buriahi, A.K.; Al-Gheethi, A.A.; Kumar, P.S.; Mohamed, R.M.S.R.; Yusof, H.; Alshalif, A.F.; Khalifa, N.A. Elimination of rhodamine B from textile wastewater using nanoparticle photocatalysts: A review for sustainable approaches. Chemosphere 2022, 287, 132162. [Google Scholar] [CrossRef]
  4. Wang, J.; Zhuan, R. Degradation of antibiotics by advanced oxidation processes: An overview. Sci. Total Environ. 2020, 701, 135023. [Google Scholar] [CrossRef]
  5. Sharma, K.; Dutta, V.; Sharma, S.; Raizada, P.; Hosseini-Bandegharaei, A.; Thakur, P.; Singh, P. Recent advances in enhanced photocatalytic activity of bismuth oxyhalides for efficient photocatalysis of organic pollutants in water: A review. J. Ind. Eng. Chem. 2019, 78, 1–20. [Google Scholar] [CrossRef]
  6. Miao, Z.; Wang, Q.; Zhang, Y.; Meng, L.; Wang, X. In situ construction of S-scheme AgBr/BiOBr heterojunction with surface oxygen vacancy for boosting photocatalytic CO2 reduction with H2O. Appl. Catal. B Environ. 2022, 301, 120802. [Google Scholar] [CrossRef]
  7. Rashid, J.; Abbas, A.; Chang, L.C.; Iqbal, A.; Haq, I.U.; Rehman, A.; Awan, S.U.; Arshad, M.; Rafique, M.; Barakat, M.A. Butterfly cluster like lamellar BiOBr/TiO2 nanocomposite for enhanced sunlight photocatalytic mineralization of aqueous ciprofloxacin. Sci. Total Environ. 2019, 665, 668–677. [Google Scholar] [CrossRef]
  8. Zang, J.; Chen, C.; Chen, X.; Yang, Y. Bi4O5Br2 nanosheets modified by carbon quantum dots: Efficient BPA degradation induced by enhanced absorption and Z-scheme charge transfer. J. Alloys Compd. 2023, 935, 167988. [Google Scholar] [CrossRef]
  9. Bai, Y.; Chen, T.; Wang, P.; Wang, L.; Ye, L. Bismuth-rich Bi4O5X2 (X = Br, and I) nanosheets with dominant {101} facets exposure for photocatalytic H2 evolution. Chem. Eng. J. 2016, 304, 454–460. [Google Scholar] [CrossRef]
  10. Mao, D.; Ding, S.; Meng, L.; Dai, Y.; Sun, C.; Yang, S.; He, H. One-pot microemulsion-mediated synthesis of Bi-rich Bi4O5Br2 with controllable morphologies and excellent visible-light photocatalytic removal of pollutants. Appl. Catal. B Environ. 2017, 207, 153–165. [Google Scholar] [CrossRef]
  11. Zhao, X.; You, Y.; Huang, S.; Wu, Y.; Ma, Y.; Zhang, G.; Zhang, Z. Z-scheme photocatalytic production of hydrogen peroxide over Bi4O5Br2/g-C3N4 heterostructure under visible light. Appl. Catal. B Environ. 2020, 278, 119251. [Google Scholar] [CrossRef]
  12. Ma, Y.; Qian, X.; Arif, M.; Xia, J.; Fan, H.; Luo, J.; He, G.; Chen, H. Z-scheme Bi4O5Br2/MIL-88B(Fe) heterojunction for boosting visible light catalytic oxidation of tetracycline hydrochloride. Appl. Surf. Sci. 2023, 611, 155667. [Google Scholar] [CrossRef]
  13. Jin, X.; Lv, C.; Zhou, X.; Xie, H.; Sun, S.; Liu, Y.; Meng, Q.; Chen, G. A bismuth rich hollow Bi4O5Br2 photocatalyst enables dramatic CO2 reduction activity. Nano Energy 2019, 64, 103955. [Google Scholar] [CrossRef]
  14. Chang, F.; Li, S.; Shi, Z.; Qi, Y.; Liu, D.; Liu, X.; Chen, S. Boosted photocatalytic NO removal performance by S-scheme hierarchical composites WO3/Bi4O5Br2 prepared through a facile ball-milling protocol. Sep. Purif. Technol. 2021, 278, 119662. [Google Scholar] [CrossRef]
  15. Chang, F.; Yan, W.; Wang, X.; Peng, S.; Li, S.; Hu, X. Strengthened photocatalytic removal of bisphenol a by robust 3D hierarchical n-p heterojunctions Bi4O5Br2-MnO2 via boosting oxidative radicals generation. Chem. Eng. J. 2022, 428, 131223. [Google Scholar] [CrossRef]
  16. Chen, Y.; Zhao, C.; Ma, S.; Xing, P.; Hu, X.; Wu, Y.; He, Y. Fabrication of a Z-scheme AgBr/Bi4O5Br2 nanocomposite and its high efficiency in photocatalytic N2 fixation and dye degradation. Inorg. Chem. Front. 2019, 6, 3083–3092. [Google Scholar] [CrossRef]
  17. Weng, P.; Cai, Q.; Wu, H.; Zhang, L.; Wu, K.; Guo, J. Facile synthesis of flower-like CQDs/S-Bi4O5Br2 composites as a highly efficient visible-light response photocatalyst for ciprofloxacin degradation. J. Mater. Sci. 2022, 57, 1977–1993. [Google Scholar] [CrossRef]
  18. Wu, D.; Wu, C.; Tan, H. Facile synthesis of novel I-doped Bi4O5Br2 nanosheets with enhanced visible light photocatalytic activity. J. Mater. Sci. Mater. Electron. 2018, 29, 11090–11095. [Google Scholar] [CrossRef]
  19. Jia, X.; Shen, Z.; Han, Q.; Bi, H. Rod-like Bi4O5I2/Bi4O5Br2 step-scheme heterostructure with oxygen vacancies synthesized by calcining the solid solution containing organic group. Chin. J. Catal. 2022, 43, 288–302. [Google Scholar] [CrossRef]
  20. Li, P.; Cao, W.; Zhu, Y.; Teng, Q.; Peng, L.; Jiang, C.; Feng, C.; Wang, Y. NaOH-induced formation of 3D flower-sphere BiOBr/Bi4O5Br2 with proper-oxygen vacancies via in-situ self-template phase transformation method for antibiotic photodegradation. Sci. Total Environ. 2020, 715, 136809. [Google Scholar] [CrossRef]
  21. Zhang, X.; Zha, X.; Luo, Y.; Liu, T.; Chen, G.; He, X. In2O3 Nanoparticle/Bi4O5Br2 Nanosheet S-Scheme Heterojunctions with Interfacial Oxygen Vacancies for Photocatalytic Degradation of Tetracycline. ACS Appl. Nano Mater. 2023, 6, 11877–11887. [Google Scholar] [CrossRef]
  22. Qian, X.; Ma, Y.; Arif, M.; Xia, J.; He, G.; Chen, H. Construction of 2D/2D Bi4O5Br2/Bi2WO6 Z-scheme heterojunction for highly efficient photodegradation of ciprofloxacin under visible light. Sep. Purif. Technol. 2023, 316, 123794. [Google Scholar] [CrossRef]
  23. Li, Y.; Sun, B.; Lin, H.; Ruan, Q.; Geng, Y.; Liu, J.; Wang, H.; Yang, Y.; Wang, L.; Tam, K.C. Efficient visible-light induced H2 evolution from T-CdxZn1−xS/defective MoS2 nano-hybrid with both bulk twinning homojunctions and interfacial heterostructures. Appl. Catal. B Environ. 2020, 267, 118702. [Google Scholar] [CrossRef]
  24. Song, J.; Sun, R.; Chen, Y.; Sun, D.; Li, X. l-Cysteine assisted synthesis of Zn0.5Cd0.5S solid solution with different morphology, crystal structure and performance for H2 evolution. Int. J. Hydrogen Energy 2018, 43, 18220–18231. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Lu, D.; Li, H.; Kondamareddy, K.K.; Wang, H.; Zhang, B.; Wang, J.; Wu, Q.; Zeng, Y.; Zhang, X.; et al. Enhanced visible Light-Driven photocatalytic hydrogen evolution and stability for noble Metal-Free MoS2/Zn0.5Cd0.5S heterostructures with W/Z phase junctions. Appl. Surf. Sci. 2022, 586, 152770. [Google Scholar] [CrossRef]
  26. Dong, W.; Cai, T.; Wang, L.; Liu, C.; Chen, H.; Li, W.; Liu, Y.; Xia, X. Optimized removal of antibiotics over Cd0.5Zn0.5S/NiCo-LDH: Constructing a homojunctions-heterojunctions composite photocatalyst. J. Environ. Chem. Eng. 2022, 10, 108624. [Google Scholar] [CrossRef]
  27. Qi, S.; Wang, D.; Zhao, Y.; Xu, H. Core–shell g-C3N4@Zn0.5Cd0.5S heterojunction photocatalysts with high photocatalytic activity for the degradation of organic dyes. J. Mater. Sci. Mater. Electron. 2019, 30, 5284–5296. [Google Scholar] [CrossRef]
  28. Madhusudan, P.; Wageh, S.; Al-Ghamdi, A.A.; Zhang, J.; Cheng, B.; Yu, Y. Graphene-Zn0.5Cd0.5S nanocomposite with enhanced visible-light photocatalytic CO2 reduction activity. Appl. Surf. Sci. 2020, 506, 144683. [Google Scholar] [CrossRef]
  29. Qiu, J.; Zhang, X.-F.; Zhang, X.; Feng, Y.; Li, Y.; Yang, L.; Lu, H.; Yao, J. Constructing Cd0.5Zn0.5S@ZIF-8 nanocomposites through self-assembly strategy to enhance Cr(VI) photocatalytic reduction. J. Hazard. Mater. 2018, 349, 234–241. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, X.; Zhang, Y.; Miao, W.; Zhang, X.; Shi, Y.; Tang, Z.; Shi, H. Insight into the pathway, improvement of performance and photocatalytic mechanism of active carbon/Bi4O5Br2 composite for cefixime and rhodamine B removal. J. Photochem. Photobiol. Chem. 2023, 444, 114892. [Google Scholar] [CrossRef]
  31. Zhou, X.; Wu, J.; Xiao, Y.; Jiang, Y.; Zhang, W.; Liu, Y.; Liu, Z.; Zhang, J. Boosting photoelectron transport in Zn0.5Cd0.5S/Sn3O4 heterostructure through close interface contact for enhancing photocatalytic H2 generation and degradation of tetracycline hydrochloride. Sep. Purif. Technol. 2023, 311, 123243. [Google Scholar] [CrossRef]
  32. Xiao, Y.; Jiang, Y.; Zhou, E.; Zhang, W.; Liu, Y.; Zhang, J.; Wu, X.; Qi, Q.; Liu, Z. In-suit fabricating an efficient electronic transport channels via S-scheme polyaniline/Cd0.5Zn0.5S heterojunction for rapid removal of tetracycline hydrochloride and hydrogen production. J. Mater. Sci. Technol. 2023, 153, 205–218. [Google Scholar] [CrossRef]
  33. Li, A.; Pang, H.; Li, P.; Zhang, N.; Chen, G.; Meng, X.; Liu, M.; Liu, X.; Ma, R.; Ye, J. Insights into the critical dual-effect of acid treatment on ZnxCd1−xS for enhanced photocatalytic production of syngas under visible light. Appl. Catal. B Environ. 2021, 288, 119976. [Google Scholar] [CrossRef]
  34. Madhusudan, P.; Shi, R.; Xiang, S.; Jin, M.; Chandrashekar, B.N.; Wang, J.; Wang, W.; Peng, O.; Amini, A.; Cheng, C. Construction of highly efficient Z-scheme ZnxCd1−xS/Au@g-C3N4 ternary heterojunction composite for visible-light-driven photocatalytic reduction of CO2 to solar fuel. Appl. Catal. B Environ. 2021, 282, 119600. [Google Scholar] [CrossRef]
  35. Cao, W.; Jiang, C.; Chen, C.; Zhou, H.; Wang, Y. A novel Z-scheme CdS/Bi4O5Br2 heterostructure with mechanism analysis: Enhanced photocatalytic performance. J. Alloys Compd. 2021, 861, 158554. [Google Scholar] [CrossRef]
  36. Ye, L.; Jin, X.; Liu, C.; Ding, C.; Xie, H.; Chu, K.H.; Wong, P.K. Thickness-ultrathin and bismuth-rich strategies for BiOBr to enhance photoreduction of CO2 into solar fuels. Appl. Catal. B Environ. 2016, 187, 281–290. [Google Scholar] [CrossRef]
  37. Meng, S.; Bi, Y.; Yan, T.; Zhang, Y.; Wu, T.; Shao, Y.; Wei, D.; Du, B. Room-temperature fabrication of bismuth oxybromide/oxyiodide photocatalyst and efficient degradation of phenolic pollutants under visible light. J. Hazard. Mater. 2018, 358, 20–32. [Google Scholar] [CrossRef] [PubMed]
  38. Cai, M.; Liu, Y.; Wang, C.; Lin, W.; Li, S. Novel Cd0.5Zn0.5S/Bi2MoO6 S-scheme heterojunction for boosting the photodegradation of antibiotic enrofloxacin: Degradation pathway, mechanism and toxicity assessment. Sep. Purif. Technol. 2023, 304, 122401. [Google Scholar] [CrossRef]
  39. Xia, D.; Hu, L.; Wang, Y.; Xu, B.; Liao, Y.; He, C.; Ye, L.; Liang, X.; Ye, Y.; Shu, D. Immobilization of facet-engineered Ag3PO4 on mesoporous Al2O3 for efficient industrial waste gas purification with indoor LED illumination. Appl. Catal. B Environ. 2019, 256, 117811. [Google Scholar] [CrossRef]
  40. Guo, F.; Huang, X.; Chen, Z.; Ren, H.; Li, M.; Chen, L. MoS2 nanosheets anchored on porous ZnSnO3 cubes as an efficient visible-light-driven composite photocatalyst for the degradation of tetracycline and mechanism insight. J. Hazard. Mater. 2020, 390, 122158. [Google Scholar] [CrossRef]
  41. Pei, C.Y.; Chen, Y.G.; Wang, L.; Chen, W.; Huang, G.B. Step-scheme WO3/CdIn2S4 hybrid system with high visible light activity for tetracycline hydrochloride photodegradation. Appl. Surf. Sci. 2021, 535, 147682. [Google Scholar] [CrossRef]
  42. Lai, C.; Zhang, M.; Li, B.; Huang, D.; Zeng, G.; Qin, L.; Liu, X.; Yi, H.; Cheng, M.; Li, L.; et al. Fabrication of CuS/BiVO4 (0 4 0) binary heterojunction photocatalysts with enhanced photocatalytic activity for Ciprofloxacin degradation and mechanism insight. Chem. Eng. J. 2019, 358, 891–902. [Google Scholar] [CrossRef]
  43. Xu, Y.; You, Y.; Huang, H.; Guo, Y.; Zhang, Y. Bi4NbO8Cl {001} nanosheets coupled with g-C3N4 as 2D/2D heterojunction for photocatalytic degradation and CO2 reduction. J. Hazard. Mater. 2020, 381, 121159. [Google Scholar] [CrossRef] [PubMed]
  44. Xu, Y.; Ma, Y.; Ji, H.; Huang, S.; Xie, M.; Zhao, Y.; Xu, H.; Li, H. Enhanced long-wavelength light utilization with polyaniline/bismuth-rich bismuth oxyhalide composite towards photocatalytic degradation of antibiotics. J. Colloid Interface Sci. 2019, 537, 101–111. [Google Scholar] [CrossRef]
  45. Shi, Y.; Wang, L.; Dong, S.; Miao, X.; Zhang, M.; Sun, K.; Zhang, Y.; Cao, Z.; Sun, J. Wool-ball-like BiOBr@ZnFe-MOF composites for degradation organic pollutant under visible-light: Synthesis, performance, characterization and mechanism. Opt. Mater. 2022, 131, 112580. [Google Scholar] [CrossRef]
  46. Luangwanta, T.; Chachvalvutikul, A.; Kaowphong, S. Facile synthesis and enhanced photocatalytic activity of a novel FeVO4/Bi4O5Br2 heterojunction photocatalyst through step-scheme charge transfer mechanism. Colloids Surf. Physicochem. Eng. Asp. 2021, 627, 127217. [Google Scholar] [CrossRef]
  47. Ma, X.; Chen, K.; Niu, B.; Li, Y.; Wang, L.; Huang, J.; She, H.; Wang, Q. Preparation of BiOCl0.9I0.1/β-Bi2O3 composite for degradation of tetracycline hydrochloride under simulated sunlight. Chin. J. Catal. 2020, 41, 1535–1543. [Google Scholar] [CrossRef]
  48. Li, X.; Sun, Y.; Xiong, T.; Jiang, G.; Zhang, Y.; Wu, Z.; Dong, F. Activation of amorphous bismuth oxide via plasmonic Bi metal for efficient visible-light photocatalysis. J. Catal. 2017, 352, 102–112. [Google Scholar] [CrossRef]
  49. Wang, W.; Li, Z.; Wu, K.; Dai, G.; Chen, Q.; Zhou, L.; Zheng, J.; Ma, L.; Li, G.; Wang, W.; et al. Novel Ag-bridged dual Z-scheme g-C3N4/BiOI/AgI plasmonic heterojunction: Exceptional photocatalytic activity towards tetracycline and the mechanism insight. J. Environ. Sci. 2023, 131, 123–140. [Google Scholar] [CrossRef]
  50. Fei, Q.; Yin, H.; Yuan, C.; Zhang, Y.; Zhao, Q.; Lv, H.; Zhang, Y.; Zhang, Y. Visible-light-driven AgI/Bi4O5I2 S-scheme heterojunction for efficient tetracycline hydrochloride removal: Mechanism and degradation pathway. Chemosphere 2023, 337, 139326. [Google Scholar] [CrossRef]
  51. Wang, Y.; Xiao, X.; Ding, T.; Lu, M. Mixed 1T-2H phase MoSe2 as interfacial charge-transfer-bridge to boosting photocatalytic activity of dual Z-scheme AgI/1T-2H MoSe2/Bi4O5Br2 heterojunction. J. Alloys Compd. 2021, 875, 160092. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) ZCS, Bi4O5Br2, and (b) ZCS/Bi4O5Br2 composites.
Figure 1. XRD patterns of (a) ZCS, Bi4O5Br2, and (b) ZCS/Bi4O5Br2 composites.
Inorganics 12 00127 g001
Figure 2. SEM images of (a) ZCS, (b) Bi4O5Br2, and (c) 1.5-ZCS/Bi4O5Br2.
Figure 2. SEM images of (a) ZCS, (b) Bi4O5Br2, and (c) 1.5-ZCS/Bi4O5Br2.
Inorganics 12 00127 g002
Figure 3. EDS elemental mappings of 1.5-ZCS/Bi4O5Br2.
Figure 3. EDS elemental mappings of 1.5-ZCS/Bi4O5Br2.
Inorganics 12 00127 g003
Figure 4. (a) TEM and (b,c) HRTEM images of 1.5-ZCS/Bi4O5Br2.
Figure 4. (a) TEM and (b,c) HRTEM images of 1.5-ZCS/Bi4O5Br2.
Inorganics 12 00127 g004
Figure 5. XPS spectra of ZCS, Bi4O5Br2, and 1.5-ZCS/Bi4O5Br2: (a) full spectra, (b) Zn 2p, (c) Cd 3d, (d) Bi 4f, (e) Br 3d, and (f) O 1s.
Figure 5. XPS spectra of ZCS, Bi4O5Br2, and 1.5-ZCS/Bi4O5Br2: (a) full spectra, (b) Zn 2p, (c) Cd 3d, (d) Bi 4f, (e) Br 3d, and (f) O 1s.
Inorganics 12 00127 g005
Figure 6. (a) N2 adsorption–desorption isotherms and (b) pore size distributions (inset) of Bi4O5Br2 and 1.5-ZCS/Bi4O5Br2.
Figure 6. (a) N2 adsorption–desorption isotherms and (b) pore size distributions (inset) of Bi4O5Br2 and 1.5-ZCS/Bi4O5Br2.
Inorganics 12 00127 g006
Figure 7. (a) UV–vis DRS of obtained samples, (b) Tauc plots of Bi4O5Br2, ZCS, and 1.5-ZCS/Bi4O5Br2, (c,d) Mott–Schottky plots of ZCS and Bi4O5Br2.
Figure 7. (a) UV–vis DRS of obtained samples, (b) Tauc plots of Bi4O5Br2, ZCS, and 1.5-ZCS/Bi4O5Br2, (c,d) Mott–Schottky plots of ZCS and Bi4O5Br2.
Inorganics 12 00127 g007
Figure 8. (a) PL spectra and (b) EIS Nyquist plots of Bi4O5Br2, ZCS, and 1.5-ZCS/Bi4O5Br2.
Figure 8. (a) PL spectra and (b) EIS Nyquist plots of Bi4O5Br2, ZCS, and 1.5-ZCS/Bi4O5Br2.
Inorganics 12 00127 g008
Figure 9. (a) Photodegradation efficiencies of TCH by different samples, (b) pseudo-first-order degradation kinetics curves, (c) apparent rate constants of corresponding catalysts, (d) UV–vis spectrum changes of TCH, and (e) photodegradation efficiencies of OTC, RhB, and MB.
Figure 9. (a) Photodegradation efficiencies of TCH by different samples, (b) pseudo-first-order degradation kinetics curves, (c) apparent rate constants of corresponding catalysts, (d) UV–vis spectrum changes of TCH, and (e) photodegradation efficiencies of OTC, RhB, and MB.
Inorganics 12 00127 g009
Figure 10. (a) Cycling experiments of 1.5-ZCS/Bi4O5Br2, (b) XRD plots before and after cycling.
Figure 10. (a) Cycling experiments of 1.5-ZCS/Bi4O5Br2, (b) XRD plots before and after cycling.
Inorganics 12 00127 g010
Figure 11. Proposed TCH degradation pathways by 1.5-ZCS/Bi4O5Br2.
Figure 11. Proposed TCH degradation pathways by 1.5-ZCS/Bi4O5Br2.
Inorganics 12 00127 g011
Figure 12. (a) Trapping experiments for the photodegradation of TCH with 1.5-ZCS/Bi4O5Br2, (b,c) EPR spectra of DMPO-·O2 and TEMPO-h+.
Figure 12. (a) Trapping experiments for the photodegradation of TCH with 1.5-ZCS/Bi4O5Br2, (b,c) EPR spectra of DMPO-·O2 and TEMPO-h+.
Inorganics 12 00127 g012
Figure 13. Schematic diagram of possible photocatalytic mechanism.
Figure 13. Schematic diagram of possible photocatalytic mechanism.
Inorganics 12 00127 g013
Scheme 1. Schematic illustration of the preparation of ZCS/Bi4O5Br2 composites.
Scheme 1. Schematic illustration of the preparation of ZCS/Bi4O5Br2 composites.
Inorganics 12 00127 sch001
Table 1. Comparison of degradation rates of TCH by different catalysts.
Table 1. Comparison of degradation rates of TCH by different catalysts.
CatalystDosage
(g/L)
Concentration
(mg/L)
Time (min)Degradation Rate (%)Reference
PANI/Bi4O5Br20.42024085.7[44]
BiOBr@ZnFe-MOF0.1559079.2[45]
Bi4O5I2/Bi4O5Br20.42012090.2[19]
FeVO4/Bi4O5Br20.52015088[46]
BiOCl0.9I0.1/β-Bi2O30.42012082.4[47]
Zn0.5Cd0.5S/Bi4O5Br20.4206084.5This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Luo, L.; Shen, J.; Jin, B. Construction of Zn0.5Cd0.5S/Bi4O5Br2 Heterojunction for Enhanced Photocatalytic Degradation of Tetracycline Hydrochloride. Inorganics 2024, 12, 127. https://doi.org/10.3390/inorganics12050127

AMA Style

Luo L, Shen J, Jin B. Construction of Zn0.5Cd0.5S/Bi4O5Br2 Heterojunction for Enhanced Photocatalytic Degradation of Tetracycline Hydrochloride. Inorganics. 2024; 12(5):127. https://doi.org/10.3390/inorganics12050127

Chicago/Turabian Style

Luo, Lan, Juan Shen, and Bo Jin. 2024. "Construction of Zn0.5Cd0.5S/Bi4O5Br2 Heterojunction for Enhanced Photocatalytic Degradation of Tetracycline Hydrochloride" Inorganics 12, no. 5: 127. https://doi.org/10.3390/inorganics12050127

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