Next Article in Journal
Zinc–Acetate–Amine Complexes as Precursors to ZnO and the Effect of the Amine on Nanoparticle Morphology, Size, and Photocatalytic Activity
Next Article in Special Issue
Fabrication of TaON/CdS Heterostructures for Enhanced Photocatalytic Hydrogen Evolution under Visible Light Irradiation
Previous Article in Journal
A Calix[8]arene-Based Catalyst for Suzuki–Miyaura Couplings with Reduced Pd_Leaching
Previous Article in Special Issue
Interface Engineering-Induced 1T-MoS2/NiS Heterostructure for Efficient Hydrogen Evolution Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Zn1−xCdxS Photocatalyst for Degradation of 2-CP and TC, Catalytic Mechanism

1
Key Laboratory of Marine Chemistry Theory and Technology of Ministry of Education, Ocean University of China, Qingdao 266100, China
2
Shandong Haijingtian Environmental Technology Corporation, Binzhou 256600, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1100; https://doi.org/10.3390/catal12101100
Submission received: 18 August 2022 / Revised: 13 September 2022 / Accepted: 21 September 2022 / Published: 23 September 2022

Abstract

:
Zn1−xCdxS catalysts with Zeolitic Imidazolate Framework-8 (ZIF-8) as the precursor were successfully prepared by ion exchange method, and the ability and electrochemical properties of a series of ZIF-8, ZnS and Zn1−xCdxS catalysts in photocatalytic degradation of 2-CP and TC were investigated. Doping of Cd ions was able to modulate the ZnS band gap width and improve the utilization of visible light by the photocatalyst. The nanocage catalysts with hollow structure of Zn1−xCdxS have better photocatalytic response. The removal of photocatalytic pollutants was up to 90% under optimal conditions. Using a Peroxymonosulfate (PMS)-assisted system to improve the degradation efficiency of 2-chlorophenol and tetracycline hydrochloride under visible light, we present a possible mechanism of Zn1−xCdxS as a photocatalyst for degradation in persistent pollutants and in PMS-assisted photocatalysis. Four active species, O2, h+, -OH, and SO4, can be generated to degrade 2-chlorophenol and tetracycline hydrochloride under PMS-assisted activation. Zn1−xCdxS nanocage with high activity and stability provides a feasible approach to catalytically remove persistent pollutants from aqueous solutions under visible light conditions.

1. Introduction

With the development of the world economy and the production and abuse of pesticides, deodorants, and antibiotic drugs being discharged into the environment, a large number of organic pollutants has been causing environmental pollution and water pollution [1]. These pollutants mainly include antibiotics (e.g., tetracycline hydrochloride, ciprofloxacin [2], etc.), chlorophenols (e.g., 2-chlorophenol, 4-chlorophenol, etc.), etc. Tetracycline hydrochloride [3,4], as an antibiotic that can be widely used, is often applied to prevent infections. Excessive use can cause resistance to microorganisms in the water and cause incalculable harm to the environment [5,6]. In contrast, 2-chlorophenol pollution is caused by the misuse of pesticides and other insecticides, which can cause incalculable harm to human health and daily life [7,8]. The removal of pollutants from water bodies has become an urgent issue at this stage.
At present, the treatment of pollutants in wastewater has been a great challenge. The most commonly used methods for the wastewater treatment containing pollutants are photocatalytic degradation [9], ion exchange [10], and adsorption [11,12]. Photocatalytic technology can convert pollutants in water into CO2 and H2O based on the rational use of visible light, without secondary pollution. A green, energy-saving and effective operation can effectively alleviate environmental pollution and the energy crisis [13,14]. However, the removal of pollutants by photocatalysis alone is more limited, and the removal of pollutants from water bodies by PMS-assisted photocatalysis [15,16]. This is due to the synergistic effect of photocatalysis and PMS on the activation of the photocatalyst and thus on the degradation in organic pollutants. PMS is able to capture photogenerated electrons, thus prolonging the complexation of electrons and catalyst, as well as increasing the efficiency of photocatalysis [17,18].
The core of photocatalysis is searching for a suitable and high performing catalyst [19]. Metal sulfides are widely used in photocatalysis due to their suitable band structure [20,21,22]. ZnS is an n-type semiconductor, which is widely used, simple and easy to be prepared, and has a high electron mobility [23,24]. Due to the band gap limitation of ZnS, it only responds well to UV light [25]. CdS has high visible light utilization [26,27]. Reddy et al. [28] synthesized CdS-ZnS core-shell nanoparticles by a two-step method, which expanded the surface area of the catalyst, increased the active sites of the catalyst, and significantly improved the degradation efficiency of MO. Liu et al. [29] synthesized ZnCdS solid solution by precipitation hydrothermal method, which has a suitable band gap compared with ZnS, can improve the utilization of visible light to a certain extent, and can separate free electrons and holes while preventing their complexation.
Metal–organic frameworks (MOFs) materials are porous materials composed of organic ligands and inorganic ions, which have gradually become the focus of scientists’ research, mainly due to their high specific surface area, high porosity and easy modification [30,31,32]. MOFs materials with hollow nanostructures can facilitate electron-hole migration, and thus have been widely studied. Chen et al. [33] successfully prepared ZnCdS hollow nanoframeworks with ZIF-8 as precursor, which greatly improved the separation and transfer of carriers and had superior performance in the field of photocatalytic hydrogen production. Bai et al. [34] synthesized Zn1−xCdxS /CdS heterojunctions via a low-temperature sulfidation method to effectively drive hydrogen production. Lin et al. [35] synthesized rhombic dodecahedral Cd0.5Zn0.5S that demonstrated the removal ability of organic pollutants.
In this paper, we have successfully prepared hollow nano Zn1−xCdxS materials by a two-step sulfidation and cation exchange method, using ZIF-8 as precursor, and investigated the removal of pollutants 2-CP and TC from water under visible light irradiation conditions, as well as under PMS-assisted conditions. Through a combination of capture experiments and kinetic fitting, we speculated on the possible reaction mechanism. This led us to propose a new method for the removal of pollutants from water.

2. Results and Discussion

2.1. Characterization of the Synthesized Zn1−xCdxS

As can be seen in Figure 1a, the synthesized Zn1−xCdxS catalyst with ZIF-8 as the precursor has a very distinctive dodecahedral structure with an average particle size of 800 nm. The SEM images of Zn1−xCdxS from Figure 1c,d show that the dodecahedral structure of the ZIF-8 precursor did not change significantly after sulphidation and Cd doping. The SEM images of the ZIF-8 precursors showed that the dodecahedral structure did not change significantly after sulphidation and Cd doping, but a clear hollow structure was formed inside. The modification of the ZIF-8 precursor did not destroy the complete dodecahedral structure and retained the prismatic frame. The elemental distribution of Zn1−xCdxS is also shown in Figure 1e, where the elements Zn, Cd and S are uniformly distributed in the catalyst. In Figure 2 the SEM images were analysed using image analysis software and a certain amount of particles were selected to measure their average initial size, giving an average particle size of approximately 800 nm for the Zn1−xCdxS catalyst.
In this work, ZIF-8 was synthesized according to the method reported in the literature by standing at room temperature, and according to the XRD image Figure 3a, the synthesized ZIF-8 has obvious diffraction peaks and a more consistent structure with the simulated ZIF-8, indicating the successful preparation of ZIF-8 crystals with high crystallinity. The X-ray diffraction pattern after ion exchange is shown in Figure 3b, where the position of the diffraction peak on the (1,1,1) crystal plane is gradually shifted to a lower angle as the doping of Cd2+ increases, indicating that the introduction of Cd2+ is not a simple physical mixing, but is introduced into the ZnS lattice.
The light absorption performance determines the light utilization and catalytic activity of the photocatalyst. The light absorption properties of ZnS and Zn0.8Cd0.2S were investigated using UV-vis diffuse reflectance spectroscopy (UV-vis DRS). According to Figure 4, the ZnS catalyst with ZIF-8 catalyst as precursor has a strong absorption peak at 360 nm, which is consistent with what was previously reported [36]. The visible light absorption of Zn0.8Cd0.2S is enhanced due to the introduction of Cd2+. The bandgap energy (Eg) of the catalyst can be calculated by the Kubelka–Munk method using the α(hv) = A(Eg)n/2, where α, h, v and A denote the absorption coefficient, Planck’s constant, optical frequency and constant, respectively [37]. The calculation shows that the bandgap energy of ZnS is 3.60 eV, while that of Zn0.8Cd0.2S is 2.88 eV.
The N2 adsorption–desorption isotherms were used to analyze the specific surface area, pore volume and pore size distribution of the samples. It can be seen that the catalysts all exhibit a similar shape to the Type I isotherms. The specific surface areas of ZIF-8, ZnS and Zn0.8Cd0.2S measured by the Barrett–Joyner–Halenda (BJH) method were 1035.7082 m2/g, 261.6770 m2/g and 576.2332 m2/g, respectively, and the pore volumes were 0.4900 cm3/g, 0.2816 cm3/g, 0.1160 cm3/g. The larger specific surface area and pore volume of Zn0.8Cd0.2S compared to ZnS is due to the larger ionic radius of Cd2+ compared to Zn2+, which is caused by the ion exchange [33]. Both the increase in specific surface area and pore volume facilitate increased adsorption of pollutant molecules, thus promoting photocatalytic degradation reactions.

2.2. Photocatalytic Performance of Photocatalysts

The performance of the catalysts was assessed by their effectiveness in removing persistent pollutants from aqueous solutions. Then, 2-CP (50 mg/L) was removed from aqueous solutions with an initial pH of six using catalysts containing different Cd contents. The experimental results in Figure 5a show that light irradiation has little effect on the degradation in 2-CP when no catalyst is added. The best degradation effect of 90% was achieved with the addition of a catalyst with a Cd doping ratio of 20% under the same experimental conditions. In contrast, the adsorption effect of the catalyst on pollutants decreases with the increase in the amount of Cd introduced into the catalyst. The aqueous solution containing the contaminant TC (40 mg/L) was removed by Zn1−xCdxS catalysts with different Cd-containing ratios at an initial pH = 6. The experimental results Figure 5b show that without the addition of Zn1−xCdxS catalyst the removal efficiency for TC is almost zero, while with the addition of catalyst the best degradation can reach 87%. This indicates that the Zn1−xCdxS catalyst has a high efficiency in utilizing visible light and can promote photoelectron transfer.
The initial pH of the solution containing the persistent pollutants was adjusted with 1 M hydrochloric acid. Figure 5c,d shows that both the degradation effect and the adsorption effect decreased significantly at pH = 2 and pH = 4. The possible reason for this is that the structure of the hollow nanocatalyst may be damaged under strongly acidic conditions, thus affecting the degradation in the pollutants. The reported performance of the photocatalysts applied for the degradation in TC and 2-CP is presented in Table 1 and Table 2. Compared with the photocatalyst materials listed in the tables, it reveals that the Zn1−xCdxS catalyst synthesized herein has excellent visible light catalytic activity for the degradation in TC and 2-CP organic pollutants.
The two contaminants were added to the aqueous solution simultaneously at the same levels as when the experiments were conducted separately, and the aqueous solution was scanned at full wavelength using a UV-Vis spectrometer. As shown in Figure 6, the levels of both pollutants in the aqueous solution showed a decreasing trend with time after the light was switched on. This indicates that the Zn1−xCdxS catalyst showed a superior removal efficiency for TC when both persistent pollutants were present in the aqueous solution at the same time.

2.3. Kinetic Studies

Both pollutants were optimally degraded when the initial pH of the solution = 6, and the catalysts showed the best photocatalytic performance for each pollutant. Under such optimum initial conditions, a linear fit was performed for the removal of the persistent pollutants from the solution, as shown in Figure 7a,b, and the results of the fit were consistent with a pseudo-first order kinetic model. For both contaminants, a 20% Cd doping rate resulted in a maximum degradation rate of 0.03706 min−1 for 2-CP and 0.00776 min−1 for TC. As shown in Figure 7c,d, when a 1M hydrochloric acid solution was used to adjust the initial pH of the solution, the fitted results were consistent with the pseudo-first order kinetic model, with the maximum degradation rate at pH = 6 when on Zn0.8Cd0.2S.

2.4. Effect of Temperature

The reaction temperature was screened under the basis of the optimum degradation conditions listed in Section 3.2. As shown in Figure 8, the photocatalyst Zn1−xCdxS showed a decreasing trend in the degradation in both organic pollutants with increasing temperature at a dosage of 50 mg and pH = 6. This may be due to the fact that the dissolved oxygen in the organic pollutant solution declines as the temperature increases, resulting in a downward concentration of superoxide radicals leading to a decrease in the degradation performance of the pollutants [48].

2.5. Photoelectrochemical Analysis

The transient photocurrent test of the catalyst sample allows the study of the broadband response of the catalyst. It can be seen from Figure 9 that the current intensity changes significantly at the moment the light is turned on. The strong photocurrent indicates that the doping of Cd promotes the carrier transfer and separation, and the strong current intensity of Zn1−xCdxS is due to the change of band gap width by the doping of Cd, which improves the utilization of visible light.

2.6. Degradation Properties of Zn1−xCdxS-Activated PMS

The optimum photocatalyst Zn0.8Cd0.2S was selected based on the results of the previous experiments to investigate the degradation in TC and 2-CP with different PMS dosages. In this section, different concentrations of PMS were set up for the degradation in TC and 2-CP. As shown in Figure 10a, b, the degradation efficiency of 2-CP increased from 79% to 90% with the increase in PMS dosage, and the best degradation effect and the maximum reaction rate were achieved when the PMS dosage was 0.02 mM. The degradation efficiency of 2-CP increased from 79% to 90%, and the best degradation effect and maximum reaction rate was achieved at 0.02 mM PMS. According to Figure 10c,d, the degradation efficiency of TC increased and then decreased with increasing PMS dosage, and the best degradation effect was achieved at 0.02 mM. The graph shows that when the dosage is more than 0.02 mM, the degradation process is more obviously inhibited. This may be due to the fact that the excess amount of PMS obstructs other types of reactive oxygen radicals (ROS) from generating. As the amount of photocatalyst is constant and therefore the amount of PMS that can be activated is constant, an excess of PMS may not be activated and may cause an increase in turbidity in the solution, thus preventing light from reaching the solution, resulting in poor light utilization [49].

2.7. Possible Photocatalytic Mechanism

The capture experiments can help to understand the main active species in the photocatalytic degradation process. The scavengers added to the solution containing persistent pollutants were isopropyl alcohol, p-benzoquinone and EDTA-2Na to capture hydroxyl radicals (•OH), superoxide radicals (•O2) and vacancies (h+), respectively. The addition of BQ to the solution containing TC and to the solution containing 2-CP, respectively, greatly reduced the degradation rate of both contaminants. The addition of IPA and EDTA-2Na to the solution containing contaminants did not significantly hinder the degradation effect. Thus, the experiments can prove that superoxide radicals are the main active species for the degradation in TC and 2-CP (Figure 11).
•O2, •OH, h+ are the main reactive oxygen radicals during the mineralization of TC molecules and 2-CP molecules. During the course of the experiment, the organic pollutant molecules are first adsorbed on the surface of the catalyst and when the light starts, the electrons are first excited from the valence band to the conduction band. O2 continues to be adsorbed on the surface, forming O2 with electrons, which is then further reduced to H2O2 and -OH. In addition, the h+ accumulated on the VB can also be used for the degradation in pollutant molecules as seen in Figure 12. The •O2 play a major role in the degradation of organic pollutants throughout the photocatalytic reaction process. In contrast, h+ and •OH play an auxiliary role in the degradation in organic pollutants. The introduction of Cd2+ enhances the light absorption, accelerates charge separation and transfer, enhances the adsorption of oxygen and produces more highly active substances, which is conducive to enhancing the activity of photocatalytic degradation in organic pollutants.
From this, we propose a possible photocatalytic reaction, which is as follows:
Zn ( 1 x ) CdxS + hv e + h +
O 2 + e O 2
O 2 + 2 H + 2 O H
e + H S O 5 O H + S O 4
O 2 + T C Product
O H + T C Product
h + + T C Product
S O 4 + T C Product
O 2 + 2 C P Product
h + + 2 C P Product
O H + 2 C P Product
S O 4 + 2 C P Product

2.8. Photocatalyst Stability Testing

To investigate the stability of the optimized catalyst, the photocatalyst material synthesized for the degradation in persistent pollutants was recovered from solution, washed three times with anhydrous ethanol and deionized water and then dried in an oven, and will be repeated for photocatalytic degradation experiments. Figure 13a,b show the stability and recoverability of the photocatalysts implanted. The photocatalytic experiments were repeated three times en route and there was no significant change in the degradation efficiency for the pollutants (TC and 2-CP). This indicates that under suitable conditions, the prepared photocatalysts are highly stable and can be repeated for photocatalytic degradation experiments, which is particularly important in the field of photocatalytic removal of contaminants from aqueous solutions.

3. Materials and Method

3.1. Chemicals

Zinc acetate dihydrate (Zn(CH3COO)2·2H2O,AR), Cadmium acetate dehydrate (Cd(CH3COO)2·2H2O,AR), 2-Chlorophenol (2-CP, 99%), Tetracycline hydrochloride (TC, 99%), thioacetamide (TAA, 98%), Ammonium persulfate ((NH4)2S2O8,AR), 1-methylimidazole (98%) and 2-methylimidazole (98%, C4H6N2) were purchased from Aladdin Industries (Shanghai, China). Methanol (AR), ethanol (AR), ethanediol (AR), p-benzoquinone (BQ), isopropanol (IPA, AR), EDTA-2Na were purchased from Sinopharm Chemical Regent (Shanghai, China). All of the materials were used without further purification.

3.2. Synthesis of ZIF-8

ZIF-8 was synthesized by improving on the previously reported synthesis method [50,51]. Zn(CH3COO)2·2H2O and 1-methylimidazole, 2-methylimidazole were dissolved in 20 mL of anhydrous methanol, sonicated until completely dissolved, then noted as solution A and B, respectively. Solution A was poured into solution B with rapid stirring to mix thoroughly and stirred slowly at room temperature for 24 h. The white solid obtained by centrifugation was washed three times with anhydrous ethanol and dried overnight in a drying oven at 60 °C.

3.3. Synthesis of Zn1−xCdxS Nanocages

Zn1−xCdxS nanocages are synthesized by improving on the previous work [52]. ZIF-8 was dissolved in 40 mL of anhydrous methanol and 500 mg of thioacetamide (TAA) was dissolved in 10 mL of anhydrous ethanol and stirred separately until completely dissolved. The two liquids were mixed and stirred slowly and continuously at 50 °C for 24 h. A certain amount of Cd(CH3COO)2·2H2O was taken and dissolved in ethylene glycol to dissolve completely. The above two solutions were mixed and stirred for 30 min and then transferred to a 100 mL PTFE-lined hydrothermal reactor and kept at 160 °C for 4 h. The resulting samples were separated by centrifugation, washed three times with anhydrous ethanol and dried in a drying oven at 60 °C overnight. The amount of Cd(CH3COO)2·2H2O was added to obtain different Zn1-xCdxS catalysts (x = 0.2, 0.4 and 0.6). The preparation process is shown in Scheme 1.

3.4. Photocatalytic Activity Measurements

2-chlorophenol (2-CP) and tetracycline hydrochloride (TC) were selected as organic pollutants. The photocatalytic degradation experiments were carried out in a 100 mL reaction vessel containing 100 mL of the organic pollutant solution (50 mg/L of 2-CP or 40 mg/L of TC). After adding 50 mg of photocatalyst to the organic contaminant solution and stirring for 30 min in the dark until adsorption–desorption equilibrium, the contaminants were irradiated with visible light at a cut-off wavelength of 420 nm using a 300 W xenon lamp (Magnesium Rexon MC-PF300, Beijing, China). Samples were taken at 15 min intervals and filtered through a 0.45 μm filter to remove catalyst particles from the solution, and the absorbance of 2-chlorophenol and tetracycline hydrochloride were measured at 274 nm and 356 nm, respectively, using a UV spectrophotometer (PERSEE, TU-1950, Beijing, China), with the method of scanning from 200 nm to 450 nm.
The photocatalytic degradation efficiency can be calculated by Equation (13), and C0 and Ct denote the initial concentration of the pollutant and the concentration after a reaction period, respectively.
D e g r a d a t i o n   r a t e : D % = C 0 C t C 0 × 100 %

3.5. PMS Activation Photocatalytic Reaction Test

For photocatalytic reactions using PMS activation, 50 mg of catalyst samples were added to the reactor (100 mL, 50 mg/L of 2-CP or 40 mg/L of TC), stirred for 0.5 h in the dark to reach adsorption–desorption equilibrium, and then 0.01 mM, 0.02 mM and 0.3 mM PMS oxidants were added to the reactor. A sample of 4 mL was taken at 15 min intervals and the catalyst was separated from the solution using a 0.45 μm needle filter and tested as in 2.4 above.

3.6. Free radical Capture Test

Under the same conditions as the photocatalytic degradation in pollutants, p-benzoquinone (BQ), isopropanol (IPA), and EDTA-2Na were used as trapping agents for superoxide radicals (•O2), hydroxyl radicals (•OH), and holes (h+). For the cycling experiments of photocatalysts, the photocatalysts were centrifuged and washed with deionized water and continued to be used for photocatalytic degradation cycle experiments.

3.7. Electrochemical Testing

Photocurrent tests were performed on an electrochemical workstation with a standard triple electrode. The electrolyte was 0.5 M H2SO4, the counter electrode was a platinum wire, the reference electrode was an Ag/AgCl electrode, and the working electrode was a catalyst sample coated on an ITO electrode. The working electrode was prepared by taking 5 mg of sample dissolved in 150 μL of anhydrous ethanol, adding 15 μL of naphthol to the solution and sonicating. The sonication finished solution was coated on the ITO glass with an area of 10 mm ×15 mm and the coating area was 10 mm × 10 mm.

4. Conclusions

In summary, Zn1−xCdxS is excellent for the degradation in pollutants under UV-Vis conditions, but has a limiting effect on pH adjustment. When the pH was close to neutral, the catalyst showed good adsorption and photocatalytic degradation, achieving over 90% removal of the pollutant within 120 min. At the same time, some inhibition of the photocatalytic process occurs with increasing temperature. In photocurrent response tests it was shown that the doping of Cd broadened the absorption of light and had a higher migration rate of electrons. It was shown that •O2 was the main reactive radical. We therefore propose a possible electron transfer mechanism and reaction mechanism. Zn1−xCdxS also has a certain activation ability for PMS and shows a certain promotion effect for the removal of organic pollutants when PMS is dosed at 0.02 mM in solutions containing organic pollutants.

Author Contributions

J.T.: Conceptualization, Investigation, Data curation, Writing-original draft; G.W.: Validation, Data curation; Z.W.: Validation, Data curation; H.S.: Validation, Data curation; L.L.: Investigation, Validation; C.L.: Supervision, Validation, Writing—review and editing; J.B.: Methodology, Supervision, Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the 2021 China Ministry of Education University-Industry Collaborative Education Project under grant number KGRCYJY2021011.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, R.; Yang, Z.-H.; Wang, Q.-P.; Bai, Y.; Liu, J.-B.; Zheng, Y.; Zhang, Y.-R.; Xiong, W.-P.; Ahmad, K.; Fan, C.-Z. Rapid startup of thermophilic anaerobic digester to remove tetracycline and sulfonamides resistance genes from sewage sludge. Sci. Total Environ. 2018, 612, 788–798. [Google Scholar] [CrossRef]
  2. Tozar, T.; Boni, M.; Staicu, A.; Pascu, M.L. Optical Characterization of Ciprofloxacin Photolytic Degradation by UV-Pulsed Laser Radiation. Molecules 2021, 26, 2324. [Google Scholar] [CrossRef]
  3. Liu, X.; Huang, L.; Wu, X.; Wang, Z.; Dong, G.; Wang, C.; Liu, Y.; Wang, L. Bi2Zr2O7 nanoparticles synthesized by soft-templated sol-gel methods for visible-light-driven catalytic degradation of tetracycline. Chemosphere 2018, 210, 424–432. [Google Scholar] [CrossRef]
  4. Sun, S.; Wang, Y.; Zhou, L.; Wang, X.; Kang, C. Enhanced degradation mechanism of tetracycline by MnO2 with the presence of organic acids. Chemosphere 2022, 286, 131606. [Google Scholar] [CrossRef]
  5. Wang, X.; Zhu, G.; Wang, C.; Niu, Y. Effective degradation of tetracycline by organic-inorganic hybrid materials induced by triethylenediamine. Environ. Res. 2021, 198, 111253. [Google Scholar] [CrossRef]
  6. Zhang, Q.; Xu, J.; Ma, X.; Xu, J.; Yun, Z.; Zuo, Q.; Wang, L. A novel Fe-based bi-MOFs material for photocatalytic degradation of tetracycline: Performance, mechanism and toxicity assessment. J. Water Process Eng. 2021, 44, 102364. [Google Scholar] [CrossRef]
  7. Channei, D.; Nakaruk, A.; Khanitchaidecha, W.; Jannoey, P.; Phanichphant, S. Hybrid high-porosity rice straw infused with BiVO4 nanoparticles for efficient 2-chlorophenol degradation. Int. J. Appl. Ceram. Technol. 2019, 16, 1060–1068. [Google Scholar] [CrossRef]
  8. Yasmeen, H.; Zada, A.; Ali, S.; Khan, I.; Ali, W.; Khan, W.; Khan, M.; Anwar, N.; Ali, A.; Huerta-Flores, A.M.; et al. Visible light-excited surface plasmon resonance charge transfer significantly improves the photocatalytic activities of ZnO semiconductor for pollutants degradation. J. Chin. Chem. Soc. 2020, 67, 1611–1617. [Google Scholar] [CrossRef]
  9. Wu, S.; Hu, H.; Lin, Y.; Zhang, J.; Hu, Y.H. Visible light photocatalytic degradation of tetracycline over TiO2. Chem. Eng. J. 2020, 382, 122842. [Google Scholar] [CrossRef]
  10. Lim, S.J.; Kim, T.-H. Combined treatment of swine wastewater by electron beam irradiation and ion-exchange biological reactor system. Sep. Purif. Technol. 2015, 146, 42–49. [Google Scholar] [CrossRef]
  11. Shu, J.; Wang, Z.; Huang, Y.; Huang, N.; Ren, C.; Zhang, W. Adsorption removal of Congo red from aqueous solution by polyhedral Cu2O nanoparticles: Kinetics, isotherms, thermodynamics and mechanism analysis. J. Alloys Compd. 2015, 633, 338–346. [Google Scholar] [CrossRef]
  12. Lei, C.; Zhu, X.; Zhu, B.; Jiang, C.; Le, Y.; Yu, J. Superb adsorption capacity of hierarchical calcined Ni/Mg/Al layered double hydroxides for Congo red and Cr(VI) ions. J. Hazard. Mater. 2017, 321, 801–811. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, H.; Wu, Y.; Feng, M.; Tu, W.; Xiao, T.; Xiong, T.; Ang, H.; Yuan, X.; Chew, J.W. Visible-light-driven removal of tetracycline antibiotics and reclamation of hydrogen energy from natural water matrices and wastewater by polymeric carbon nitride foam. Water Res. 2018, 144, 215–225. [Google Scholar] [CrossRef] [PubMed]
  14. Cabrera-Reina, A.; Martínez-Piernas, A.B.; Bertakis, Y.; Xekoukoulotakis, N.P.; Agüera, A.; Sánchez Pérez, J.A. TiO2 photocatalysis under natural solar radiation for the degradation of the carbapenem antibiotics imipenem and meropenem in aqueous solutions at pilot plant scale. Water Res. 2019, 166, 115037. [Google Scholar] [CrossRef]
  15. Li, C.-J.; Zhao, R.; Peng, M.-Q.; Gong, X.-L.; Xia, M.; Li, K.; Huang, Z.-B.; Xia, D.-S. Mechanism study on denitration by new PMS modified bamboo charcoal bifunctional photocatalyst. Chem. Eng. J. 2017, 316, 544–552. [Google Scholar] [CrossRef]
  16. Oh, W.-D.; Ng, C.-Z.; Ng, S.-L.; Lim, J.-W.; Leong, K.-H. Rapid degradation of organics by peroxymonosulfate activated with ferric ions embedded in graphitic carbon nitride. Sep. Purif. Technol. 2020, 230, 115852. [Google Scholar] [CrossRef]
  17. Shao, H.; Zhao, X.; Wang, Y.; Mao, R.; Wang, Y.; Qiao, M.; Zhao, S.; Zhu, Y. Synergetic activation of peroxymonosulfate by Co3O4 modified g-C3N4 for enhanced degradation of diclofenac sodium under visible light irradiation. Appl. Catal. B Environ. 2017, 218, 810–818. [Google Scholar] [CrossRef]
  18. Gong, Y.; Zhao, X.; Zhang, H.; Yang, B.; Xiao, K.; Guo, T.; Zhang, J.; Shao, H.; Wang, Y.; Yu, G. MOF-derived nitrogen doped carbon modified g-C3N4 heterostructure composite with enhanced photocatalytic activity for bisphenol A degradation with peroxymonosulfate under visible light irradiation. Appl. Catal. B Environ. 2018, 233, 35–45. [Google Scholar] [CrossRef]
  19. Feng, C.; Chen, Z.; Jing, J.; Sun, M.; Tong, H.; Hou, J. Band structure and enhanced photocatalytic degradation performance of Mg-doped CdS nanorods. Phys. B Condens. Matter 2020, 594, 412363. [Google Scholar] [CrossRef]
  20. Yan, S.C.; Li, Z.S.; Zou, Z.G. Photodegradation Performance of g-C3N4 Fabricated by Directly Heating Melamine. Langmuir 2009, 25, 10397–10401. [Google Scholar] [CrossRef]
  21. Bo, T.; Wang, Y.; Wang, J.; Zhao, Z.; Zhang, J.; Zheng, K.; Lin, T.; Zhang, B.; Shao, L. Photocatalytic H2 evolution on CdS modified with partially crystallized MoS2 under visible light irradiation. Chem. Phys. Lett. 2020, 746, 137305. [Google Scholar] [CrossRef]
  22. Su, J.; Yu, S.; Xu, M.; Guo, Y.; Sun, X.; Fan, Y.; Zhang, Z.; Yan, J.; Zhao, W. Enhanced visible light photocatalytic performances of few-layer MoS2@TiO2 hollow spheres heterostructures. Mater. Res. Bull. 2020, 130, 110936. [Google Scholar] [CrossRef]
  23. Ethiraj, A.S.; Hebalkar, N.; Sainkar, S.R.; Urban, J.; Kulkarni, S.K. Synthesis and Investigation of ZnS Nanoparticles Adsorbed On Functionalised Silica Particles. Surf. Eng. 2004, 20, 367–372. [Google Scholar] [CrossRef]
  24. Mahajan, J.; Jeevanandam, P. A facile thermal decomposition approach for the synthesis of SiO2@ZnS core-shell nanoparticles and their application as effective adsorbent for the removal of congo red. Mater. Today Commun. 2021, 26, 102085. [Google Scholar] [CrossRef]
  25. Guo, J.; Khan, S.; Cho, S.-H.; Kim, J. ZnS nanoparticles as new additive for polyethersulfone membrane in humic acid filtration. J. Ind. Eng. Chem. 2019, 79, 71–78. [Google Scholar] [CrossRef]
  26. Ruan, D.; Fujitsuka, M.; Majima, T. Exfoliated Mo2C nanosheets hybridized on CdS with fast electron transfer for efficient photocatalytic H2 production under visible light irradiation. Appl. Catal. B Environ. 2020, 264, 118541. [Google Scholar] [CrossRef]
  27. Zhong, W.; Wu, X.; Wang, P.; Fan, J.; Yu, H. Homojunction CdS Photocatalysts with a Massive S2—Adsorbed Surface Phase: One-Step Facile Synthesis and High H2-Evolution Performance. ACS Sustain. Chem. Eng. 2020, 8, 543–551. [Google Scholar] [CrossRef]
  28. Reddy, C.V.; Shim, J.; Cho, M. Synthesis, structural, optical and photocatalytic properties of CdS/ZnS core/shell nanoparticles. J. Phys. Chem. Solids 2017, 103, 209–217. [Google Scholar] [CrossRef]
  29. Liu, M.; Wang, L.; Lu, G.; Yao, X.; Guo, L. Twins in Cd1−xZnxS solid solution: Highly efficient photocatalyst for hydrogen generation from water. Energy Environ. Sci. 2011, 4, 1372–1378. [Google Scholar] [CrossRef]
  30. Wang, D.; Li, Z. Iron-based metal–organic frameworks (MOFs) for visible-light-induced photocatalysis. Res. Chem. Intermed. 2017, 43, 5169–5186. [Google Scholar] [CrossRef]
  31. Zhang, X.; Wang, J.; Dong, X.-X.; Lv, Y.-K. Functionalized metal-organic frameworks for photocatalytic degradation of organic pollutants in environment. Chemosphere 2020, 242, 125144. [Google Scholar] [CrossRef] [PubMed]
  32. Jiang, W.; Li, Z.; Liu, C.; Wang, D.; Yan, G.; Liu, B.; Che, G. Enhanced visible-light-induced photocatalytic degradation of tetracycline using BiOI/MIL-125(Ti) composite photocatalyst. J. Alloys Compd. 2021, 854, 157166. [Google Scholar] [CrossRef]
  33. Chen, J.; Chen, J.; Li, Y. Hollow ZnCdS dodecahedral cages for highly efficient visible-light-driven hydrogen generation. J. Mater. Chem. A 2017, 5, 24116–24125. [Google Scholar] [CrossRef]
  34. Bai, T.; Shi, X.; Liu, M.; Huang, H.; Yu, M.-H.; Zhang, J.; Bu, X.-H. A metal–organic framework-derived Zn1−xCdxS/CdS heterojunction for efficient visible light-driven photocatalytic hydrogen production. Dalton Trans. 2021, 50, 6064–6070. [Google Scholar] [CrossRef] [PubMed]
  35. Wu, L.; Fu, H.; Wei, Q.; Zhao, Q.; Wang, P.; Wang, C.-C. Porous Cd0.5Zn0.5S nanocages derived from ZIF-8: Boosted photocatalytic performances under LED-visible light. Environ. Sci. Pollut. Res. 2021, 28, 5218–5230. [Google Scholar] [CrossRef] [PubMed]
  36. Song, H.; Wang, N.; Meng, H.; Han, Y.; Wu, J.; Xu, J.; Xu, Y.; Zhang, X.; Sun, T. A facile synthesis of a ZIF-derived ZnS/ZnIn2S4 heterojunction and enhanced photocatalytic hydrogen evolution. Dalton Trans. 2020, 49, 10816–10823. [Google Scholar] [CrossRef]
  37. Luca, V.; Djajanti, S.; Howe, R.F. Structural and Electronic Properties of Sol−Gel Titanium Oxides Studied by X-ray Absorption Spectroscopy. J. Phys. Chem. B 1998, 102, 10650–10657. [Google Scholar] [CrossRef]
  38. Chen, P.; Dong, N.; Zhang, J.; Wang, W.; Tan, F.; Wang, X.; Qiao, X.; Keung Wong, P. Investigation on visible-light photocatalytic performance and mechanism of zinc peroxide for tetracycline degradation and Escherichia coli inactivation. J. Colloid Interface Sci. 2022, 624, 137–149. [Google Scholar] [CrossRef]
  39. He, Q.; Ge, M. Visible-light activation of peroxydisulfate by magnetic BiOBr/MnFe2O4 nanocomposite toward degradation of tetracycline. J. Mater. Sci. Mater. Electron. 2022, 33, 5859–5877. [Google Scholar] [CrossRef]
  40. Wang, H.; Liao, B.; Lu, T.; Ai, Y.; Liu, G. Enhanced visible-light photocatalytic degradation of tetracycline by a novel hollow BiOCl@CeO2 heterostructured microspheres: Structural characterization and reaction mechanism. J. Hazard. Mater. 2020, 385, 121552. [Google Scholar] [CrossRef]
  41. Kameli, S.; Mehrizad, A. Ultrasound-assisted Synthesis of Ag-ZnS/rGO and its Utilization in Photocatalytic Degradation of Tetracycline Under Visible Light Irradiation. Photochem. Photobiol. 2019, 95, 512–521. [Google Scholar] [CrossRef] [PubMed]
  42. Zhu, B.; Jiang, G.; Kong, C.; Sun, J.; Liu, F.; Wang, Y.; Zhao, C.; Liu, C. Photocatalytic degradation of organic pollutants in water by N-doping ZnS with Zn vacancy: Enhancement mechanism of visible light response and electron flow promotion. Environ. Sci. Pollut. Res. 2022, 29, 58716–58729. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, L.; Wang, J.; Zhang, Y.; Zhou, B.; Tan, P.; Pan, J. Construction of S-scheme BiOCl/CdS composite for enhanced photocatalytic degradation of antibiotic. J. Mater. Sci. Mater. Electron. 2022, 33, 13303–13315. [Google Scholar] [CrossRef]
  44. Barakat, M.A.; Kumar, R.; Almeelbi, T.; Al-Mur, B.A.; Eniola, J.O. Sustainable visible light photocatalytic scavenging of the noxious organic pollutant using recyclable and reusable polyaniline coupled WO3/WS2 nanohybrid. J. Clean. Prod. 2022, 330, 129942. [Google Scholar] [CrossRef]
  45. Alafif, Z.O.; Anjum, M.; Kumar, R.; Abdelbasir, S.M.; Barakat, M.A. Synthesis of CuO–GO/TiO2 visible light photocatalyst for 2-chlorophenol degradation, pretreatment of dairy wastewater and aerobic digestion. Appl. Nanosci. 2019, 9, 579–591. [Google Scholar] [CrossRef]
  46. Ombaka, L.M.; McGettrick, J.D.; Oseghe, E.O.; Al-Madanat, O.; Rieck genannt Best, F.; Msagati, T.A.M.; Davies, M.L.; Bredow, T.; Bahnemann, D.W. Photocatalytic H2 production and degradation of aqueous 2-chlorophenol over B/N-graphene-coated CuO/TiO2: A DFT, experimental and mechanistic investigation. J. Environ. Manag. 2022, 311, 114822. [Google Scholar] [CrossRef]
  47. Hassan, N.S.; Jalil, A.A.; Aziz, F.F.A.; Fauzi, A.A.; Azami, M.S.; Jusoh, N.W.C. Tailoring the Silica Amount in Stabilizing the Tetragonal Phase of Zirconia for Enhanced Photodegradation of 2-Chlorophenol. Top. Catal. 2020, 63, 1145–1156. [Google Scholar] [CrossRef]
  48. Saadati, F.; Keramati, N.; Ghazi, M.M. Influence of parameters on the photocatalytic degradation of tetracycline in wastewater: A review. Crit. Rev. Environ. Sci. Technol. 2016, 46, 757–782. [Google Scholar] [CrossRef]
  49. Ji, Q.; Cheng, X.; Wu, Y.; Xiang, W.; He, H.; Xu, Z.; Xu, C.; Qi, C.; Li, S.; Zhang, L.; et al. Visible light absorption by perylene diimide for synergistic persulfate activation towards efficient photodegradation of bisphenol A. Appl. Catal. B Environ. 2021, 282, 119579. [Google Scholar] [CrossRef]
  50. Jiang, Z.; Sun, H.; Qin, Z.; Jiao, X.; Chen, D. Synthesis of novel ZnS nanocages utilizing ZIF-8 polyhedral template. Chem. Commun. 2012, 48, 3620–3622. [Google Scholar] [CrossRef]
  51. Zhan, W.-w.; Kuang, Q.; Zhou, J.-z.; Kong, X.-j.; Xie, Z.-x.; Zheng, L.-s. Semiconductor@Metal–Organic Framework Core–Shell Heterostructures: A Case of ZnO@ZIF-8 Nanorods with Selective Photoelectrochemical Response. J. Am. Chem. Soc. 2013, 135, 1926–1933. [Google Scholar] [CrossRef] [PubMed]
  52. Qiu, B.; Zhu, Q.; Xing, M.; Zhang, J. A robust and efficient catalyst of CdxZn1−xSe motivated by CoP for photocatalytic hydrogen evolution under sunlight irradiation. Chem. Commun. 2017, 53, 897–900. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a,b) SEM images of ZIF-8; (c,d) SEM images of Zn1−xCdxS; (e) corresponding EDS mapping of Zn1−xCdxS.
Figure 1. (a,b) SEM images of ZIF-8; (c,d) SEM images of Zn1−xCdxS; (e) corresponding EDS mapping of Zn1−xCdxS.
Catalysts 12 01100 g001
Figure 2. Particle size distribution of prepared Zn1-xCdxS catalyst.
Figure 2. Particle size distribution of prepared Zn1-xCdxS catalyst.
Catalysts 12 01100 g002
Figure 3. XRD of (a) ZIF-8 and ZnS, and (b) different metal ratios of Zn1-xCdxS.
Figure 3. XRD of (a) ZIF-8 and ZnS, and (b) different metal ratios of Zn1-xCdxS.
Catalysts 12 01100 g003
Figure 4. (a) UV-vis DRS of ZnS and Zn0.8Cd0.2S (b) Plot of transformed Kubelka-Munk function versus photon energy for ZnS and Zn0.8Cd0.2S.
Figure 4. (a) UV-vis DRS of ZnS and Zn0.8Cd0.2S (b) Plot of transformed Kubelka-Munk function versus photon energy for ZnS and Zn0.8Cd0.2S.
Catalysts 12 01100 g004
Figure 5. Photocatalytic removal curves under visible light irradiation. (a) Degradation in 2-CP; (b) Degradation in TC; (c) Degradation in 2-CP, HCl adjust Ph; (d) Degradation in 2-CP,HCl adjust pH.
Figure 5. Photocatalytic removal curves under visible light irradiation. (a) Degradation in 2-CP; (b) Degradation in TC; (c) Degradation in 2-CP, HCl adjust Ph; (d) Degradation in 2-CP,HCl adjust pH.
Catalysts 12 01100 g005
Figure 6. UV-spectra of TC and 2-CP degradation over time on stream.
Figure 6. UV-spectra of TC and 2-CP degradation over time on stream.
Catalysts 12 01100 g006
Figure 7. The Kinetic fitting curves. (a) The Kinetic fittings of 2-CP. (b) The Kinetic fittings of TC. (c) The Kinetic fittings of 2-CP, HCl adjust pH. (d) The Kinetic fittings of TC, HCl adjust pH.
Figure 7. The Kinetic fitting curves. (a) The Kinetic fittings of 2-CP. (b) The Kinetic fittings of TC. (c) The Kinetic fittings of 2-CP, HCl adjust pH. (d) The Kinetic fittings of TC, HCl adjust pH.
Catalysts 12 01100 g007
Figure 8. (a) Effect of temperature on the degradation in organic pollutant 2-CP (b) Effect of temperature on the degradation in organic pollutant TC.
Figure 8. (a) Effect of temperature on the degradation in organic pollutant 2-CP (b) Effect of temperature on the degradation in organic pollutant TC.
Catalysts 12 01100 g008
Figure 9. Transient photocurrent responses of catalysts.
Figure 9. Transient photocurrent responses of catalysts.
Catalysts 12 01100 g009
Figure 10. Photocatalytic contaminant removal curves and kinetic fitting curves for ZIF-8 activated PMS. (a) Degradation and (b) The kinetic fittings of 2-CP. (c) Degradation and (d) The kinetic fittings of TC.
Figure 10. Photocatalytic contaminant removal curves and kinetic fitting curves for ZIF-8 activated PMS. (a) Degradation and (b) The kinetic fittings of 2-CP. (c) Degradation and (d) The kinetic fittings of TC.
Catalysts 12 01100 g010
Figure 11. Scavenger experiments.
Figure 11. Scavenger experiments.
Catalysts 12 01100 g011
Figure 12. Possible photocatalytic mechanism of photocatalytic TC degradation and 2-CP degradation.
Figure 12. Possible photocatalytic mechanism of photocatalytic TC degradation and 2-CP degradation.
Catalysts 12 01100 g012
Figure 13. Photo-stability and recyclability of the optimized catalyst. (a) Application to 2-CP de-gradability reusability trials. (b) Application to TC degradability reusability trials.
Figure 13. Photo-stability and recyclability of the optimized catalyst. (a) Application to 2-CP de-gradability reusability trials. (b) Application to TC degradability reusability trials.
Catalysts 12 01100 g013
Scheme 1. Schematic diagram of hollow Zn1-xCdxS nanocage preparation.
Scheme 1. Schematic diagram of hollow Zn1-xCdxS nanocage preparation.
Catalysts 12 01100 sch001
Table 1. Comparison of the performance of different photocatalysts for TC degradation.
Table 1. Comparison of the performance of different photocatalysts for TC degradation.
PhotocatalystsConcentration of Catalyst (g/L)Concentration of TC (mg/L)Light SourceIrradiation Time (h)Degradation Rate (%)Reference
Zn1−xCdxS0.550300 W Xenon lamp
(>400 nm)
291This work
ZnO20.150300 W Xenon lamp
(>400 nm)
483.7[38]
BiOBr/MnFe2O4120simulated sunlight lamp1.576.5[39]
BiOCl@CeO20.510300 W Xenon lamp
(>400 nm)
292[40]
Ag-ZnS/rGO1.2510300 W Xenon lamp
(>400 nm)
290[41]
N-ZnS0.310300 W Xenon lamp
(>400 nm)
2.590[42]
BiOCl/CdS610300 W Xenon lamp
(>400 nm)
1.793[43]
Table 2. Comparison of the performance of different photocatalysts for the degradation in 2-CP.
Table 2. Comparison of the performance of different photocatalysts for the degradation in 2-CP.
PhotocatalystsConcentration of Catalyst (g/L)Concentration of 2-CP (mg/L)Light SourceIrradiation Time (h)Degradation Rate (%)Reference
Zn1-xCdxS0.550300W Xenon lamp
(>400 nm)
290This work
WO3/WS2/PANI0.535300 W Xenon lamp
(>400 nm)
391[44]
CuO-GO/TiO20.510104W cool white lamp3.585[45]
B/N-graphene-coated Cu/TiO20.2551000W Xenon lamp (<550 nm)1.392[46]
SiO2/ZrO20.37510400W metal halide lamp
(>400 nm)
492[47]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tan, J.; Wei, G.; Wang, Z.; Su, H.; Liu, L.; Li, C.; Bian, J. Application of Zn1−xCdxS Photocatalyst for Degradation of 2-CP and TC, Catalytic Mechanism. Catalysts 2022, 12, 1100. https://doi.org/10.3390/catal12101100

AMA Style

Tan J, Wei G, Wang Z, Su H, Liu L, Li C, Bian J. Application of Zn1−xCdxS Photocatalyst for Degradation of 2-CP and TC, Catalytic Mechanism. Catalysts. 2022; 12(10):1100. https://doi.org/10.3390/catal12101100

Chicago/Turabian Style

Tan, Jingxin, Guoqiang Wei, Zhen Wang, Hui Su, Lingtao Liu, Chunhu Li, and Junjie Bian. 2022. "Application of Zn1−xCdxS Photocatalyst for Degradation of 2-CP and TC, Catalytic Mechanism" Catalysts 12, no. 10: 1100. https://doi.org/10.3390/catal12101100

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