Next Article in Journal
Photocatalytic Nanofiltration Membrane Using Zr-MOF/GO Nanocomposite with High-Flux and Anti-Fouling Properties
Next Article in Special Issue
Facile Synthesis of Porous Hexapod Ag@AgCl Dual Catalysts for In Situ SERS Monitoring of 4-Nitrothiophenol Reduction
Previous Article in Journal
Recent Progresses on Metal Halide Perovskite-Based Material as Potential Photocatalyst
Previous Article in Special Issue
Pt-Amorphous Barium Aluminum Oxide/Carbon Catalysts for an Enhanced Methanol Electrooxidation Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reduced Graphene Oxide/ZnIn2S4 Nanocomposite Photocatalyst with Enhanced Photocatalytic Performance for the Degradation of Naproxen under Visible Light Irradiation

1
Key Laboratory of Yangtze River Water Environment, Ministry of Education, Shanghai 200092, China
2
College of Environmental Science and Engineering, Tongji University, Shanghai 200092, China
3
Shanghai Institute of Pollution Control and Ecological Security, Shanghai 200092, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(6), 710; https://doi.org/10.3390/catal10060710
Submission received: 4 June 2020 / Revised: 19 June 2020 / Accepted: 22 June 2020 / Published: 24 June 2020

Abstract

:
The development of photocatalysts with visible light response is of great significance to cope with energy crisis and environmental remediation. In this study, a visible light-driven photocatalyst reduced graphene oxide/ZnIn2S4 (rGO/ZIS) was prepared by a facile one-pot hydrothermal method. The photocatalyst was used for the degradation of naproxen under visible light illumination and it exhibited remarkably degradation efficiency (nearly 99% within 60 min). The improved photocatalytic degradation performance can be attributed to the enhancement of light adsorption capacity and effective separation of photoinduced electron–hole pairs. The reactive species quenching experiments and EPR measurements demonstrated that superoxide radical (•O2) and hole (h+) play a dominant role in the photocatalytic degradation reactions. In addition, the degradation intermediates were identified and the degradation pathway was suggested.

1. Introduction

Over the last few decades, pharmaceuticals and personal care products (PPCPs)—an emerging class of refractory organic pollutants—have been ubiquitously detected in wastewater treatment plant effluents, surface water, ground water, and even some drinking water [1]. It has become a serious environmental and human health concern because of the potential risks posed by PPPCs which include acute and chronic toxicity, the occurrence of antibiotic resistance genes, etc. [2]. Naproxen (NPX), as one of representative nonsteroidal anti-inflammatory drug (NSAID), has been widely used in the treatment of rheumatoid arthritis due to its analgesic and antipyretic properties [3,4]. It has been frequently detected in various water environment, and its concentration ranges from ng/L to µg/L. Although the concentration of NPX is relatively low, long-term accumulation and enrichment in water and soil will have a serious impact on aquatic organisms and ecosystem [5]. In addition, some researchers have found that people who ingest trace NPX have a higher risk of heart attack or bladder cancer [6,7]. However conventional water and waste treatment processes are insufficient for the removal of NPX. Therefore, highly effective treatment methods for the elimination of NPX are urgently needed to develop.
Among different approaches for the removal of PPCPs, Photocatalytic degradation has become a promising method by virtue of its high efficiency and energy-saving. However, the traditional photocatalyst (such as TiO2 and ZnO) with wide band gap can only exploit ultraviolet region light which accounts for only 4% of the total solar spectrum, thus greatly restricts their practical applications [8]. Therefore, in order to make full use of solar energy, it is necessary to develop visible light driven photocatalysts.
Recently, ZnIn2S4, as an important 2D layered ternary chalcogenide semiconductor, has attracted increasing attention because of its unique electronic and optical characteristics [8,9]. After Lei’s group synthetized visible-light-driven photocatalyst ZnIn2S4 in 2003, researches regarding ternary metal chalcogenide ZnIn2S4 are increasingly reported [10]. It has been a promising photocatalyst used in contaminants degradation [11,12,13], hydrogen evolution [14,15,16], and CO2 reduction [17] due to its appropriate band edges, eco-friendly property, excellent photoactivity, as well as chemical stability. Reduced graphene oxide (rGO), a two-dimensional (2D) sp2 carbon allotrope, has attracted a wide range of interests owing to its outstanding features such as large theoretical specific surface area, good electrical conductivity as well as excellent chemical stability [18]. The high electron conductivity can hinder the recombination of photoinduced electron–hole pairs, thus improving photocatalytic activity.
Therefore, the combination of ZnIn2S4 and RGO is expected to further improve the photocatalytic performance of ZnIn2S4. On one hand, RGO can act as a good catalyst support, on the other hand, it can be used as a favorable acceptor and mediator of photogenerated electrons [19]. Most of studies have demonstrated that the addition of RGO can improve the performance of composite photocatalysts [19,20,21,22]. Xie et al. synthesized nano-sized rGO/CuInS2 photocatalyst via a simple one-step solvothermal method [21]. The results demonstrated that the photocatalytic efficiency of the rGO/CuInS2 was far superior to that of pure rGO/CuInS2 for the 2-nitrophenol removal. Li et al. synthesized graphene coated ZnIn2S4 microspheres—with enhanced interfacial contact [22]. The results shown that the large interfacial area and speedy charge carriers transfer route are of great significance for the heterojunction materials to facilitate photoinduced charge separation.
Herein, we synthesized reduced GO/ZnIn2S4 (rGO/ZIS) nanocomposite photocatalyst by a simple one-step hydrothermal method. The morphologies and optical-electrical properties were characterized. The photocatalytic activities of rGO/ZIS was evaluated by photocatalytic degradation of naproxen under visible light. In addition, the degradation intermediates were identified by LC/MS/MS, the possible degradation pathway was proposed and the degradation mechanism was elucidated.

2. Results and Discussion

2.1. Characterization

The morphology and microstructure of the composites were studied by SEM and TEM. As can be seen in Figure 1a, pure ZnIn2S4 shows an irregular morphology formed by self-assembly of sheet structure, while homogeneous spherical particles were observed in rGO/ZIS-1 (Figure 1b,c), which is probably because the addition of GO enables ZnIn2S4 to be fixed on its surface and self-assembled uniformly in all directions. Furthermore, Figure 1d–f shows the HRTEM of rGO/ZIS-1, which exhibits a nanoscale layered structure, the interplanar distances of 0.33 nm should be assigned to the (101) plane of ZnIn2S4. The microstructure and compositional distribution of the composite photocatalyst were further investigated by EDS mapping. The EDS elemental mapping (Figure 1g) clearly revealed that the elements of S, In, and Zn were uniformly distributed in a spherical particle located in the center of analyzed area, however C and O elements were homogeneously scattered in the surrounding space, confirming the successful combination of ZnIn2S4 and graphene oxide.
The XRD patterns of as-prepared samples were investigated and shown in Figure 2a. For the pure ZnIn2S4, all the characteristic peaks could be indexed to a hexagonal phase of ZnIn2S4 (JCPDS no. 65–2023) [8]. Graphene oxide (GO) shows a characteristic peak at 2θ = 10.8° and a weak peak at 2θ = 42°, which can be ascribed to the (002) and (100) planes respectively [23], however, no obvious diffraction peak of GO was observed in rGO/ZIS, indicating the reduction and/or pyrolysis of GO during the hydrothermal reaction [23,24]. The peak positions of rGO/ZIS stay unchanged, implying that the incorporation of GO in the composite hardly alter the crystal phase of rGO/ZIS, which may due to the high dispersion or low content of rGO [21]. In addition, however, the FT-IR spectra can further confirm the existence of RGO in the nanocomposite, as shown in Figure 2b. The peak at 1048, 1250, and 1727 cm−1 displayed in GO can be ascribed to the stretching vibration bands of oxygen-containing functional groups including C–O, C–OH, and C=O [25]. for ZnIn2S4, two peaks at 1619 and 1396 cm−1 are attributed to the water molecules and hydroxyl groups absorbed on surface [18]. rGO/ZIS nanocomposites show peaks at 1048 and 1620 cm−1 corresponding to the C–O stretching vibrations and skeletal vibration, respectively. the intensity of some peaks weakened or disappeared in the FT-IR spectrum of ZIS/rGO further indicating the partial reduction of graphene oxide to rGO.
The analysis of chemical composition and corresponding valence states in composite was carried out using XPS measurements (Figure 3). Figure 3a shows the XPS spectra for a survey scan of rGO/ZIS-1, the result demonstrated that the main elements of the composite are Zn, In, S, C, and O, respectively. The peaks shown in Figure 3b centered at 1022.18 and 1045.32 eV can be ascribed to Zn 2p3/2 and Zn 2p1/2. The two symmetric peaks at 452.56 eV and 445 eV were assigned to In 3d3/2 and In 3d5/2, respectively [8].
In addition, the binding energies of S 2p3/2 and S 2p1/2 in Figure 3c are located at 161.67 eV and 162.87 eV, respectively, which is associated with S2− in the composite [26].Figure 3e shows the high-resolution spectrum of C 1s in rGO/ZIS-1, the binding energy at 284.6, 286.5, 287.5, and 289 eV were corresponding to C=C/C–C, C–O, C=O, and O–C=O, respectively [27]. Compared with GO (Figure S1), the intensity of oxygen containing functional groups in rGO/ZIS-1 is significantly reduced, indicating that the GO was partially reduced to rGO during the hydrothermal reaction. In addition, Figure S2 and Table S1 displayed the results of N2 adsorption-desorption isotherm and BJH pore size distribution of as-prepared samples. All the samples showed a type IV isotherm and a H3 hysteresis loop, which indicated that the pores in composites are mainly formed by self-assembly of slit-like lamellar structures. Compared with GO and ZnIn2S4, the specific surface area and pore volume of rGO/ZIS-1 were improved, which facilitated the adsorption and degradation of pollutants by providing more surface-active sites.
The optical properties of the as-prepared samples was studied by the UV–vis DRS analysis, and the results were shown in Figure 4a. ZnIn2S4 had a high absorbance in the whole UV–vis light spectrum, and the absorbance edge is around 700 nm. The light absorbance of rGO/ZIS-1enhanced as compared with ZnIn2S4. Furthermore, the band gap energy (Eg) of the samples can be estimated using the equation
α h ν = A h ν E g n / 2
where α, h, ν, A, and Eg represents the absorption coefficient, Planck constant, light frequency, proportional constant and band gap energy, respectively. The (αhν)2 vs. hν-profiles for GO, ZnIn2S4 and rGO/ZIS-1 were depicted in Figure 4b. The band gap energy of samples is determined by extrapolation method when (αhν)2 is equal to 0. The values of band gap energy of ZnIn2S4 and rGO/ZIS-1 is calculated to be 1.93 eV and 1.88 eV. Compared with the energy band of ZnIn2S4, the energy band of rGO/ZIS-1 is slightly narrowed, which indicated that the presence of GO does not significantly affect the energy band structure of ZnIn2S4. The valance band (VB) and conduction band (CB) edge potential of rGO/ZIS-1can be determined by the equation
E V B = X E e + 0.5 E g
where EVB is VB edge potential, X, Ee, and Eg represents Mulliken’s electronegativity, the energy of free electrons on the hydrogen scale (~4.5 eV) and band gap energy, respectively [28]. The X value of pure ZnIn2S4 is estimated to be about 4.86, thus the EVB and ECB were calculated to be +1.3 and −0.58 eV, respectively. PL analysis was further carried out to study the recombination rate of the photogenerated electron and holes. Figure 4c shows the PL emission intensities of as-prepared samples centered at around 465 nm, which is in the order of ZnIn2S4 > rGO/ZIS-1 > GO. It is well known that the stronger the PL emission intensity is, the higher the recombination rate and possibility of the photogenerated carriers are [29]. The emission intensity of rGO/ZIS-1 reduced evidently when compared with ZnIn2S4, indicating that the incorporation of GO into the ZnIn2S4 clearly inhibit the recombination rate of photogenerated electron–hole pairs, and thus enhance the photocatalytic performance. In addition, the interface charge separation efficiency was measured by electrochemical impedance spectra (EIS). In general, the smaller curvature diameter of the EIS Nyquist plot means the smaller interfacial electron transfer resistance [8]. It can be seen in Figure 4d that the curvature diameter of rGO/ZIS-1 is smaller than ZnIn2S4, which suggested that the rGO/ZIS-1 could improve the photocatalytic activity by reducing the interfacial charge transfer resistance and then facilitating the separation and migration of charges at the interface.

2.2. Photocatalytic Performance

The photocatalytic activities of rGO/ZIS composites were tested by the degradation of NPX under visible light irradiation and the results were displayed in Figure 5 as can be seen from Figure 5a, in the scenario of dark adsorption and self-degradation of NPX, NPX is hardly degraded or removed. Compared with this, all the rGO/ZIS composites showed a remarkable removal efficiency for NPX within 60 min. Furthermore, the photodegradation rate can be calculated by the pseudo first-order kinetic equation
ln c t / c 0 = k t
where k represents apparent rate constant, t is the irradiation time, C0 and Ct are the initial and t time concentration of pollutants. rGO/ZIS-1 demonstrated the fastest degradation rate (99% within 60 min) and the highest k value than that of other samples. The k value corresponding to the degradation of NPX was 0.082 min−1, about 1.3 times higher than that of ZnIn2S4 alone (Figure 5b), which indicated that the introduction of GO made a good contribution to the promotion of NPX degradation efficiency. However, appropriate GO concentration is very important to improve the photocatalytic activity of rGO/ZIS composites. Therefore, rGO/ZIS-1 was chosen as the optimal photocatalyst for the following experiments. As shown in Figure 5c, with the decrease of initial concentration of NPX, the degradation rate of NPX decreased gradually, which is due to the ample active sites and reactive species has a quantitative superiority over NPX with low concentration. By the similar token, the higher concentration of photocatalyst will produce higher concentration of active sites and reactive species, which will accelerate the degradation rate of NPX (Figure 5d).
Because of the widespread existence of natural organic matter (NOM) in environmental water, it will affect the photocatalytic degradation to a certain extent. As a major constitute of NOM, the effect of humic acid (HA) with different concentration on the photodegradation of NPX was investigated. It can be seen from Figure 5e, the degradation rate of NPX decreased gradually with the increase of HA concentration, On the one hand, it can be attributed to the competitive effect of HA on reactive species, on the other hand, it may be due to the shielding effect of HA on irradiation light. It is well known that pH not only affects the surface charges of catalyst and ionization of organic compounds, but also influences their interactions. Therefore, the effect of initial pH (pHini = 4.04, 6.01, 7.99, and 9.97) on the photocatalytic degradation performance of NPX was carried out. The pKa value of NPX is 4.15, thus NPX existed as a neutral molecule when pH < 4.15 and deprotonated to anionic form NPX when pH > 4.15. The measured isoelectric point of rGO/ZIS-1 was about 6.95 (Figure S3), hence when pH > 6.95, the electrostatic repulsion between negatively charged surface of rGO/ZIS-1and NPX prevented them from approaching each other, resulting in the reduction of degradation rate. In addition, rGO/ZIS-1 was used for the removal of diclofenac to testify its universality and the results were shown in Figure S4. rGO/ZIS-1 exhibited the best photocatalytic performance than that of other as-prepared photocatalysts and the removal efficiency reached 99.4% within 60 min. The rate constant k is 5.6 times higher than that of pure ZnIn2S4.

2.3. Photocatalytic Mechanism

Generally, hydroxyl radical (•OH), superoxide radical (•O2) and hole (h+) are the main three reactive species that are responsible for the photocatalytic degradation of organic pollutants. In order to estimate their contribution during the photocatalytic process of NPX, the reactive species quenching experiments were carried out using different scavengers. As shown in Figure 6a, ethylenediamine tetra acetic acid disodium salt (EDTA-2Na, 1.0 mM), isopropanol (IPA, 1.0 mM), p-benzoquinone (BQ, 1.0 mM), and furfuryl alcohol (FFA, 1.0 mM) were chosen as the scavengers for h+, •OH, •O2, and 1O2 respectively. No obvious effect on the NPX photodegradation efficiency was observed when the IPA and FFA were added as quenching agents respectively, suggesting the •OH and 1O2 were not the dominant reactive species in the experiments of NPX-photodegradation. However, the removal of NPX was greatly inhibited in the presence of EDTA-2Na or BQ, and the NPX degradation efficiency decreases from 96.44% to 35.46% and 60.19%, respectively (Figure 6b). Which indicated that the h+ and •O2 might play the key roles in the NPX photocatalytic degradation process. To further affirm the above results, EPR measurements were conducted to in suit determine the •OH and •O2 radicals produced on the photocatalysts surface as the results shown in Figure 6c, no EPR signal was detected in the absence of light irradiation. By comparison, when the light is on, both ZnIn2S4 and rGO/ZIS-1 exhibited typical 1:1:1:1 spectra of DMPO−•O2 adducts, and the signal of rGO/ZIS-1 is stronger than that of ZnIn2S4, implying the enhancement of •O2 generation. However, no signal of DMPO-•OH adducts was observed, indicating the absence of •OH during the NPX degradation process, which is consistent with the reactive species quenching experiments in Figure 6a based on the above results, h+ and •O2 were the two main reactive species that contributed to the NPX photocatalytic degradation, the effect of •OH and 1O2 were negligible.
Despite the main reactive species contributed to the photodegradation of NPX have been identified, it is still of significance to study the underling generation and transfer route of photoinduced electron–hole pairs. Based on above experiment results and established energy band structure, the possible transfer route of photogenerated carriers and photocatalytic mechanism were determined and shown in Figure 7. ZnIn2S4 can easily generate electron–hole pairs under the visible light irradiation. the photogenerated electrons on the conduction band (CB) reduce the dissolved oxygen (O2) absorbed on the surface of photocatalyst to superoxide radical (•O2), because of the CB edge potential (−0.58 eV) is more negative than Eϴ (O2/•O2) (−0.33 V vs. NHE) [30]. However, the photogenerated holes on the valence band (VB) cannot oxide the OH/H2O into •OH due to the VB edge potential EVB (+1.3 eV) is less positive than Eϴ (•OH/OH, 1.99 eV vs. NHE; •OH /H2O, 2.27 eV vs. NHE). rGO not only acts as a support, but also serves as a good electron acceptor and transmission medium, which enhanced the effective separation of photogenerated charge carriers. Then, the generated reactive species h+ and •O2 will directly degrade NPX relying on its strong redox ability. The related reactions can be summed up as below:
rGO / ZIS h v   h + + e  
e + O2•O2
•O2 + H2OHO2• + OH
HO2• + HO2H2O2 + O2
H2O2 + e•OH + OH
Reactive species (h+/•O2) + NPX/DCF Photodegradation products

2.4. Reusability and Stability

For the sake of evaluating the reusability and stability of photocatalysts, cycling experiments were carried out using rGO/ZIS-1 as photocatalyst under the same conditions and the results were displayed in Figure 8a. After five repeated runs of photocatalytic test, rGO/ZIS-1 still has a degradation efficiency of 96.1% for NPX, which is slightly lower than that of first run (98.4%). Besides, there is no distinct change in the survey XPS spectra after five consecutive cycles (Figure 8b), illustrating that the rGO/ZIS-1 composite photocatalyst has relatively high reusability and stability in the photocatalytic degradation of NPX and DCF, which is beneficial to practical application.

2.5. Photodegradation Intermediates and Suggested Pathways

To further understand the photodegradation mechanism of NPX during the reaction process, the intermediates of NPX were analyzed by LC-MS/MS. The proposed pathway of degradation was shown in Figure 9. Seven main intermediates with m/z of 262, 216, 216, 200, 188, 170, and 358 were identified in the process of NPX photodegradation. The detection of intermediates was based on the transformation pathway of NPX degradation in the previous research literatures [4,6,31,32,33,34]. The methyl and methoxyl groups on the aromatic ring are the two sites easily attacked by reactive oxygen species [6]. After electron transfer and decarboxylation NPX was transferred to m/z 184, which was further oxidized to produce m/z 200. m/z 216 was the demethoxylation product of NPX after the methoxy group on the ring was attacked by ROS. m/z 170 and m/z 188 could be generated by the oxidation attack on the methyl position of m/z 216. m/z 262 was the hydroxylation product through the electrophilic adduct reaction [4]. In addition, m/z 358 was identified as dimeric photoproducts of m/z 188 [34], which has been detected in other work [35]. These intermediates will be further degraded into small molecules, which are finally oxidized to CO2 and H2O.

3. Materials and Methods

3.1. Materials

All chemicals and reagents were of analytical grade and used without further purification. Zinc acetate (Zn (CH3COO)2⋅2H2O), indium (III) nitrate hydrate (In (NO3)3⋅4.5H2O), and L-cysteine were all obtained from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Acetonitrile (HPLC grade; ≥99.8%) was acquired from ANPEL Tech Co., Ltd. (Shanghai, China). Naproxen (NPX) was purchased from Sigma-Aldrich Chemical Co., Ltd. (St. Louis, MI, USA). Graphene oxide (GO) was sourced from XFNANO Materials Tech Co., Ltd. (Jiangsu, China). Ultrapure water was produced from a Milli-Q device (18.2 MΩ·cm−1) and used in the all experiments.

3.2. Synthesis of Photocatalysts

Different mass ratios of reduced graphene oxide/Znln2S4 (donated as rGO/ZIS) composites were synthesized via a hydrothermal procedure with some modification [8]. In a typical route, 0.01568 g Graphene oxide (1 wt%) and 0.2195 g (1 mmol) Zn(CH3COO)2⋅2H2O, 0.76384 g (2 mmol) In(NO3)3⋅4.5H2O and 0.4846 g (4 mmol) L-cysteine were firstly added into 30 mL of ultrapure water, then the mixture was magnetically stirred for 30 min and transferred into a 50 mL Teflon-lined stainless steel autoclave. Afterward, the autoclave was sealed and maintained at 180 °C for 12 h in an oven. Finally, the obtained 1 wt% rGO/ZIS was collected, washed with ethanol and deionized water for several times, and dried in an oven at 60 °C for 24 h. On this basis, different mass ratios of rGO/ZIS with loadings of 0.5 wt%,1 wt%, 2.5 wt%, 5 wt%, and 8wt % were prepared and denoted as rGO/ZIS-0.5, rGO/ZIS-1, rGO/ZIS-2.5, rGO/ZIS-5, and rGO/ZIS-8, respectively. The pure Znln2S4 sample was prepared under the same experimental conditions without the addition of GO.

3.3. Characterization of Photocatalysts

The details of characterizations were provided in Text S1.

3.4. Photocatalytic Degradation Experiments

The photocatalytic performance was examined by the degradation of NPX under visible-light irradiation (λ > 420 nm). A 300W xenon lamp (Perfect Light, PLS-SXE300,) with a 420 nm cutoff filter was used as the light source, and the intensity of the light was 115 mW/cm2. In a typical experiment, 20 mg as-prepared photocatalyst were dispersed into the quartz reactors which contains 100 mL NPX aqueous solution (10 mg·L−1, without pH adjustment). Then the quartz reactor is placed on a magnetic stirrer and the temperature was maintained at 25 °C using a recirculated cooling water system. Prior to irradiation, the solution was stirred for 30 min in dark to research the adsorption/desorption equilibrium. 3 mL aliquots were collected at the given time intervals and filtrated through 0.22 µm PES syringe filters to remove the photocatalyst. Finally, the filtrates were tested by HPLC.

3.5. Analytical Methods

The concentrations of NPX were measured by a reversed-phase high-performance liquid chromatography system (Aglient1200, CA, USA). Equipped with a UV detector and a Syncronis column (4.6 × 250 mm, 5 μm particle size, Thermo Fisher Scientific Inc., Waltham, MA, USA). Acetonitrile and water were used as the mobile phase (65:35, v/v) at a flow rate of 1 mL·min−1. The column temperature was maintained at 30 °C and the detection wavelength was set at 240 nm. The degradation intermediates were detected by LC-MS/MS (Thermo Fisher Scientific TSQ Quantum, Waltham, MA, USA) during the photocatalytic degradation process and more details of analysis can be found in Text S2.

4. Conclusions

We have successfully prepared nanocomposite photocatalyst reduced graphene oxide/ZnIn2S4 (rGO/ZIS) via a simple one-step hydrothermal method. rGO/ZIS shown superior photocatalytic degradation activity in contrast to the pure ZnIn2S4 under visible light. The results of photoelectric characterization demonstrated that the higher charge carrier mobility and expansion of absorption range of light are responsible for the improvement of photocatalytic performance. Furthermore, reactive species trapping experiments and EPR signals proved that superoxide radical (•O2) and hole (h+) are the main reactive species in the photocatalytic degradation reaction. Finally, seven degradation byproducts were identified and the possible degradation route was proposed based on the LC-MS/MS results. We hope that this work can provide some insights for the design and application of ternary chalcogenide semiconductor composites in the field of photocatalytic environmental remediation.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/6/710/s1, Figure S1: High-resolution XPS spectra of C 1s in GO, Figure S2: Nitrogen adsorption-desorption isotherms (a) and corresponding pore size distribution(b), Figure S3: The Zeta potential of 1% rGO/ZIS in different solution pH at 25 °C, Figure S4: (a) photocatalytic degradation of DCF with different photocatalysts and (b) plots of −ln(C/C0) vs irradiation time. Reaction conditions: [DCF] = 10 mg·L−1, [photocatalyst] = 0.2 g·L−1, natural pH, λ > 420 nm, Table S1: BET experimental results of GO, ZnIn2S4 and rGO/ZnIn2S4.

Author Contributions

Data curation, K.F.; Funding acquisition, H.D.; Investigation, C.D. and J.S.; Methodology, Y.P. and J.S.; Project administration, J.S. and H.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Major Science and Technology Program for Water Pollution Control and Treatment (grant number. 2017ZX07201003 and 2017ZX07501001), China.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liang, R.; Luo, S.; Jing, F.; Shen, L.; Qin, N.; Wu, L. A simple strategy for fabrication of Pd@MIL-100(Fe) nanocomposite as a visible-light-driven photocatalyst for the treatment of pharmaceuticals and personal care products (PPCPs). Appl. Catal. B 2015, 176–177, 240–248. [Google Scholar] [CrossRef]
  2. Kumar, A.; Khan, M.; Fang, L.; Lo, I.M.C. Visible-light-driven N-TiO2@SiO2@Fe3O4 magnetic nanophotocatalysts: Synthesis, characterization, and photocatalytic degradation of PPCPs. J. Hazard Mater. 2019, 370, 108–116. [Google Scholar] [CrossRef] [PubMed]
  3. Ray, S.K.; Dhakal, D.; Lee, S.W. Rapid degradation of naproxen by AgBr-α-NiMoO4 composite photocatalyst in visible light: Mechanism and pathways. Chem. Eng. J. 2018, 347, 836–848. [Google Scholar] [CrossRef]
  4. Wang, F.; Wang, Y.; Feng, Y.; Zeng, Y.; Xie, Z.; Zhang, Q.; Su, Y.; Chen, P.; Liu, Y.; Yao, K.; et al. Novel ternary photocatalyst of single atom-dispersed silver and carbon quantum dots co-loaded with ultrathin g-C3N4 for broad spectrum photocatalytic degradation of naproxen. Appl. Catal. B 2018, 221, 510–520. [Google Scholar] [CrossRef]
  5. Begum, S.; Ahmaruzzaman, M. Biogenic synthesis of SnO2/activated carbon nanocomposite and its application as photocatalyst in the degradation of naproxen. Appl. Surf. Sci. 2018, 449, 780–789. [Google Scholar] [CrossRef]
  6. Chi, H.; He, X.; Zhang, J.; Wang, D.; Zhai, X.; Ma, J. Hydroxylamine enhanced degradation of naproxen in Cu2+ activated peroxymonosulfate system at acidic condition: Efficiency, mechanisms and pathway. Chem. Eng. J. 2019, 361, 764–772. [Google Scholar] [CrossRef]
  7. Chi, H.; Wang, Z.; He, X.; Zhang, J.; Wang, D.; Ma, J. Activation of peroxymonosulfate system by copper-based catalyst for degradation of naproxen: Mechanisms and pathways. Chemosphere 2019, 228, 54–64. [Google Scholar] [CrossRef]
  8. Guo, F.; Cai, Y.; Guan, W.; Huang, H.; Liu, Y. Graphite carbon nitride/ZnIn2S4 heterojunction photocatalyst with enhanced photocatalytic performance for degradation of tetracycline under visible light irradiation. J. Phys. Chem. Solids 2017, 110, 370–378. [Google Scholar] [CrossRef]
  9. Wang, S.; Guan, B.Y.; Wang, X.; Lou, X.W.D. Formation of Hierarchical Co9S8@ZnIn2S4 Heterostructured Cages as an Efficient Photocatalyst for Hydrogen Evolution. J. Am. Chem. Soc. 2018, 140, 15145–15148. [Google Scholar] [CrossRef]
  10. Gou, X.L.; Cheng, F.Y.; Shi, Y.H.; Zhang, L.; Peng, S.J.; Chen, J.; Shen, P.W. Shape-controlled synthesis of ternary chalcogenide ZnIn2S4 and CuIn(S,Se)2 nano-/microstructures via facile solution route. J. Am. Chem. Soc. 2006, 128, 7222–7229. [Google Scholar] [CrossRef]
  11. Tian, Q.; Wu, W.; Liu, J.; Wu, Z.; Yao, W.; Ding, J.; Jiang, C. Dimensional heterostructures of 1D CdS/2D ZnIn2S4 composited with 2D graphene: Designed synthesis and superior photocatalytic performance. Dalton Trans. 2017, 46, 2770–2777. [Google Scholar] [CrossRef]
  12. Qiu, P.; Yao, J.; Chen, H.; Jiang, F.; Xie, X. Enhanced visible-light photocatalytic decomposition of 2,4-dichlorophenoxyacetic acid over ZnIn2S4/g-C3N4 photocatalyst. J. Hazard Mater. 2016, 317, 158–168. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, S.; Wang, L.; Liu, C.; Luo, J.; Crittenden, J.; Liu, X.; Cai, T.; Yuan, J.; Pei, Y.; Liu, Y. Photocatalytic wastewater purification with simultaneous hydrogen production using MoS2 QD-decorated hierarchical assembly of ZnIn2S4 on reduced graphene oxide photocatalyst. Water Res. 2017, 121, 11–19. [Google Scholar] [CrossRef] [PubMed]
  14. Zhu, T.; Ye, X.; Zhang, Q.; Hui, Z.; Wang, X.; Chen, S. Efficient utilization of photogenerated electrons and holes for photocatalytic redox reactions using visible light-driven Au/ZnIn2S4 hybrid. J. Hazard Mater. 2019, 367, 277–285. [Google Scholar] [CrossRef] [PubMed]
  15. Yan, A.; Shi, X.; Huang, F.; Fujitsuka, M.; Majima, T. Efficient photocatalytic H2 evolution using NiS/ZnIn2S4 heterostructures with enhanced charge separation and interfacial charge transfer. Appl. Catal. B 2019, 250, 163–170. [Google Scholar] [CrossRef]
  16. Lim, W.Y.; Hong, M.; Ho, G.W. In situ photo-assisted deposition and photocatalysis of ZnIn2S4/transition metal chalcogenides for enhanced degradation and hydrogen evolution under visible light. Dalton Trans. 2016, 45, 552–560. [Google Scholar] [CrossRef]
  17. Jiao, X.; Chen, Z.; Li, X.; Sun, Y.; Gao, S.; Yan, W.; Wang, C.; Zhang, Q.; Lin, Y.; Luo, Y.; et al. Defect-Mediated Electron-Hole Separation in One-Unit-Cell ZnIn2S4 Layers for Boosted Solar-Driven CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 7586–7594. [Google Scholar] [CrossRef]
  18. Ye, L.; Fu, J.; Xu, Z.; Yuan, R.; Li, Z. Facile one-pot solvothermal method to synthesize sheet-on-sheet reduced graphene oxide (RGO)/ZnIn2S4 nanocomposites with superior photocatalytic performance. ACS Appl. Mater. Interfaces 2014, 6, 3483–3490. [Google Scholar] [CrossRef]
  19. Che, W.; Luo, Y.; Deng, F.; Zhang, A.; Zhao, L.; Luo, X.; Ruan, Q. Facile solvothermal fabrication of cubic-like reduced graphene oxide/AgIn5S8 nanocomposites with anti-photocorrosion and high visible-light photocatalytic performance for highly-efficient treatment of nitrophenols and real pharmaceutical wastewater. Appl. Catal. A Gen. 2018, 565, 170–180. [Google Scholar] [CrossRef]
  20. Wang, J.; Tsuzuki, T.; Tang, B.; Hou, X.; Sun, L.; Wang, X. Reduced graphene oxide/ZnO composite: Reusable adsorbent for pollutant management. ACS Appl. Mater. Interfaces 2012, 4, 3084–3090. [Google Scholar] [CrossRef]
  21. Xie, C.; Lu, X.; Deng, F.; Luo, X.; Gao, J.; Dionysiou, D.D. Unique surface structure of nano-sized CuInS2 anchored on rGO thin film and its superior photocatalytic activity in real wastewater treatment. Chem. Eng. J. 2018, 338, 591–598. [Google Scholar] [CrossRef]
  22. Li, H.; Yu, H.; Chen, S.; Zhao, H.; Zhang, Y.; Quan, X. Fabrication of graphene wrapped ZnIn2S4 microspheres heterojunction with enhanced interfacial contact and its improved photocatalytic performance. Dalton Trans. 2014, 43, 2888–2894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Jeevitha, G.; Abhinayaa, R.; Mangalaraj, D.; Ponpandian, N. Tungsten oxide-graphene oxide (WO3-GO) nanocomposite as an efficient photocatalyst, antibacterial and anticancer agent. J. Phys. Chem. Solids 2018, 116, 137–147. [Google Scholar] [CrossRef]
  24. Wang, H.; Zhang, M.; He, X.; Du, T.; Wang, Y.; Li, Y.; Hao, T. Facile prepared ball-like TiO2 at GO composites for oxytetracycline removal under solar and visible lights. Water Res. 2019, 160, 197–205. [Google Scholar] [CrossRef] [PubMed]
  25. Khazaee, Z.; Mahjoub, A.R.; Cheshme Khavar, A.H.; Srivastava, V.; Sillanpää, M. Synthesis of layered perovskite Ag,F-Bi2MoO6/rGO: A surface plasmon resonance and oxygen vacancy promoted nanocomposite as a visible-light photocatalyst. J. Photochem. Photobiol. A Chem. 2019, 379, 130–143. [Google Scholar] [CrossRef]
  26. Zhang, S.; Liu, X.; Liu, C.; Luo, S.; Wang, L.; Cai, T.; Zeng, Y.; Yuan, J.; Dong, W.; Pei, Y.; et al. MoS2 Quantum Dot Growth Induced by S Vacancies in a ZnIn2S4 Monolayer: Atomic-Level Heterostructure for Photocatalytic Hydrogen Production. ACS Nano 2018, 12, 751–758. [Google Scholar] [CrossRef]
  27. Tien, H.N.; Luan, V.H.; Hoa, L.T.; Khoa, N.T.; Hahn, S.H.; Chung, J.S.; Shin, E.W.; Hur, S.H. One-pot synthesis of a reduced graphene oxide–zinc oxide sphere composite and its use as a visible light photocatalyst. Chem. Eng. J. 2013, 229, 126–133. [Google Scholar] [CrossRef]
  28. Yan, T.; Yan, Q.; Wang, X.; Liu, H.; Li, M.; Lu, S.; Xu, W.; Sun, M. Facile fabrication of heterostructured g-C3N4/Bi2MoO6 microspheres with highly efficient activity under visible light irradiation. Dalton Trans. 2015, 44, 1601–1611. [Google Scholar] [CrossRef]
  29. Miao, X.; Ji, Z.; Wu, J.; Shen, X.; Wang, J.; Kong, L.; Liu, M.; Song, C. g-C3N4/AgBr nanocomposite decorated with carbon dots as a highly efficient visible-light-driven photocatalyst. J Colloid. Interface Sci. 2017, 502, 24–32. [Google Scholar] [CrossRef]
  30. Chen, F.; Li, D.; Luo, B.; Chen, M.; Shi, W. Two-dimensional heterojunction photocatalysts constructed by graphite-like C3N4 and Bi2WO6 nanosheets: Enhanced photocatalytic activities for water purification. J. Alloys Compd. 2017, 694, 193–200. [Google Scholar] [CrossRef]
  31. Dulova, N.; Kattel, E.; Trapido, M. Degradation of naproxen by ferrous ion-activated hydrogen peroxide, persulfate and combined hydrogen peroxide/persulfate processes: The effect of citric acid addition. Chem. Eng. J. 2017, 318, 254–263. [Google Scholar] [CrossRef]
  32. Arany, E.; Szabo, R.K.; Apati, L.; Alapi, T.; Ilisz, I.; Mazellier, P.; Dombi, A.; Gajda-Schrantz, K. Degradation of naproxen by UV, VUV photolysis and their combination. J. Hazard Mater. 2013, 262, 151–157. [Google Scholar] [CrossRef] [PubMed]
  33. Gao, Y.-q.; Gao, N.-y.; Chu, W.-h.; Yang, Q.-l.; Yin, D.-q. Kinetics and mechanistic investigation into the degradation of naproxen by a UV/chlorine process. RSC Adv. 2017, 7, 33627–33634. [Google Scholar] [CrossRef] [Green Version]
  34. Kanakaraju, D.; Motti, C.A.; Glass, B.D.; Oelgemoller, M. TiO2 photocatalysis of naproxen: Effect of the water matrix, anions and diclofenac on degradation rates. Chemosphere 2015, 139, 579–588. [Google Scholar] [CrossRef] [PubMed]
  35. DellaGreca, M.; Brigante, M.; Isidori, M.; Nardelli, A.; Previtera, L.; Rubino, M.; Temussi, F. Phototransformation and ecotoxicity of the drug Naproxen-Na. Environ. Chem. Lett. 2003, 1, 237–241. [Google Scholar] [CrossRef]
Figure 1. SEM images of ZnIn2S4 (a) and rGO/ZIS-1 (b,c); TEM images of rGO/ZIS-1 with different magnification (d–f); TEM image of rGO/ZIS-1 and corresponding energy-dispersive X-ray spectroscopy (EDS) elemental mapping for S, In, Zn, C, and O (g).
Figure 1. SEM images of ZnIn2S4 (a) and rGO/ZIS-1 (b,c); TEM images of rGO/ZIS-1 with different magnification (d–f); TEM image of rGO/ZIS-1 and corresponding energy-dispersive X-ray spectroscopy (EDS) elemental mapping for S, In, Zn, C, and O (g).
Catalysts 10 00710 g001
Figure 2. (a) XRD patterns and (b) FT-IR spectra of the as-prepared samples.
Figure 2. (a) XRD patterns and (b) FT-IR spectra of the as-prepared samples.
Catalysts 10 00710 g002
Figure 3. XPS spectra of the GO/ZIS-1. (a) Survey of the sample, (b) Zn 2p, (c) In 3d, (d) S 2p, (e) C 1s, (f) O 1s.
Figure 3. XPS spectra of the GO/ZIS-1. (a) Survey of the sample, (b) Zn 2p, (c) In 3d, (d) S 2p, (e) C 1s, (f) O 1s.
Catalysts 10 00710 g003
Figure 4. (a) UV–vis diffuse reflectance spectra and the corresponding (b) αhυ2 versus hυ curves of the GO, ZnIn2S4 and rGO/ZIS-1; (c) PL spectra; and (d) EIS Nynquist plots of GO, ZnIn2S4, and rGO/ZIS-1.
Figure 4. (a) UV–vis diffuse reflectance spectra and the corresponding (b) αhυ2 versus hυ curves of the GO, ZnIn2S4 and rGO/ZIS-1; (c) PL spectra; and (d) EIS Nynquist plots of GO, ZnIn2S4, and rGO/ZIS-1.
Catalysts 10 00710 g004
Figure 5. (a) Photocatalytic degradation of NPX with different photocatalysts and (b) plots of −ln(C/C0) vs irradiation time; Effects of initial concentration of NPX (c), photocatalyst dosage (d), HA (e) and pH (f) on NPX photodegradation by rGO/ZIS-1. Reaction conditions: [NPX] = 10 mg·L−1, [photocatalyst] = 0.2 g·L−1, natural pH for (a), (b), (c), (d), and (e), λ > 420 nm.
Figure 5. (a) Photocatalytic degradation of NPX with different photocatalysts and (b) plots of −ln(C/C0) vs irradiation time; Effects of initial concentration of NPX (c), photocatalyst dosage (d), HA (e) and pH (f) on NPX photodegradation by rGO/ZIS-1. Reaction conditions: [NPX] = 10 mg·L−1, [photocatalyst] = 0.2 g·L−1, natural pH for (a), (b), (c), (d), and (e), λ > 420 nm.
Catalysts 10 00710 g005
Figure 6. (a) NPX photodegradation over rGO/ZIS-1 and (b) corresponding degradation efficiency in the presence of 1.0 mM various scavengers; (c) EPR spectra of DMPO-•O2 and (d) DMPO−•OH adducts over different photocatalysts with irradiation for 5 min. Reaction conditions: [NPX] = 10 mg·L−1, [photocatalyst] = 0.2 g·L−1, [DMPO] = 100 mM, natural pH, λ > 420 nm.
Figure 6. (a) NPX photodegradation over rGO/ZIS-1 and (b) corresponding degradation efficiency in the presence of 1.0 mM various scavengers; (c) EPR spectra of DMPO-•O2 and (d) DMPO−•OH adducts over different photocatalysts with irradiation for 5 min. Reaction conditions: [NPX] = 10 mg·L−1, [photocatalyst] = 0.2 g·L−1, [DMPO] = 100 mM, natural pH, λ > 420 nm.
Catalysts 10 00710 g006
Figure 7. Schematic illustration for photogenerated carriers transfers and photocatalytic mechanism of rGO/ZIS composites under visible-light irradiation (λ > 420 nm).
Figure 7. Schematic illustration for photogenerated carriers transfers and photocatalytic mechanism of rGO/ZIS composites under visible-light irradiation (λ > 420 nm).
Catalysts 10 00710 g007
Figure 8. (a) Cycling tests for the photodegradation of NPX by rGO/ZIS-1. Reaction conditions: [NPX] = 10 mg·L−1, [photocatalyst] = 0.2 g·L−1, natural pH, λ > 420 nm. (b) Survey XPS spectra of rGO/ZIS-1 before and after the five cycle reactions.
Figure 8. (a) Cycling tests for the photodegradation of NPX by rGO/ZIS-1. Reaction conditions: [NPX] = 10 mg·L−1, [photocatalyst] = 0.2 g·L−1, natural pH, λ > 420 nm. (b) Survey XPS spectra of rGO/ZIS-1 before and after the five cycle reactions.
Catalysts 10 00710 g008
Figure 9. Proposed photodegradation pathways for NPX by rGO/ZIS-1. Reaction conditions: [NPX] = 10 mg·L−1, [photocatalyst] = 0.2 g·L−1, natural pH, λ > 420 nm.
Figure 9. Proposed photodegradation pathways for NPX by rGO/ZIS-1. Reaction conditions: [NPX] = 10 mg·L−1, [photocatalyst] = 0.2 g·L−1, natural pH, λ > 420 nm.
Catalysts 10 00710 g009

Share and Cite

MDPI and ACS Style

Fu, K.; Pan, Y.; Ding, C.; Shi, J.; Deng, H. Reduced Graphene Oxide/ZnIn2S4 Nanocomposite Photocatalyst with Enhanced Photocatalytic Performance for the Degradation of Naproxen under Visible Light Irradiation. Catalysts 2020, 10, 710. https://doi.org/10.3390/catal10060710

AMA Style

Fu K, Pan Y, Ding C, Shi J, Deng H. Reduced Graphene Oxide/ZnIn2S4 Nanocomposite Photocatalyst with Enhanced Photocatalytic Performance for the Degradation of Naproxen under Visible Light Irradiation. Catalysts. 2020; 10(6):710. https://doi.org/10.3390/catal10060710

Chicago/Turabian Style

Fu, Kun, Yishuai Pan, Chao Ding, Jun Shi, and Huiping Deng. 2020. "Reduced Graphene Oxide/ZnIn2S4 Nanocomposite Photocatalyst with Enhanced Photocatalytic Performance for the Degradation of Naproxen under Visible Light Irradiation" Catalysts 10, no. 6: 710. https://doi.org/10.3390/catal10060710

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