Next Article in Journal
Size Uncertainty in Individual Nanoparticles Measured by Single Particle Inductively Coupled Plasma Mass Spectrometry
Previous Article in Journal
Efficient Dual-Function Catalyst: Palladium–Copper Nanoparticles Immobilized on Co-Cr LDH for Seamless Aerobic Oxidation of Benzyl Alcohol and Nitrobenzene Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rapid Microwave-Assisted Synthesis of ZnIn2S4 Nanosheets for Highly Efficient Photocatalytic Hydrogen Production

1
Department of Materials Science and Engineering, Feng Chia University, Taichung 407102, Taiwan
2
Pritzker School of Molecular Engineering, University of Chicago, Chicago, IL 60637, USA
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(13), 1957; https://doi.org/10.3390/nano13131957
Submission received: 3 June 2023 / Revised: 25 June 2023 / Accepted: 27 June 2023 / Published: 27 June 2023

Abstract

:
In this study, a facile and rapid microwave-assisted synthesis method was used to synthesize In2S3 nanosheets, ZnS nanosheets, and ZnIn2S4 nanosheets with sulfur vacancies. The two-dimensional semiconductor photocatalysts of ZnIn2S4 nanosheets were characterized by XRD, FESEM, BET, TEM, XPS, UV–vis diffuse reflectance, and PL spectroscopy. The ZnIn2S4 with sulfur vacancies exhibited an evident energy bandgap value of 2.82 eV, as determined by UV–visible diffuse reflectance spectroscopy, and its energy band diagram was obtained through the combination of XPS and energy bandgap values. ZnIn2S4 nanosheets exhibited about 33.3 and 16.6 times higher photocatalytic hydrogen production than In2S3 nanosheets and ZnS nanosheets, respectively, under visible-light irradiation. Various factors, including materials, sacrificial reagents, and pH values, were used to evaluate the influence of ZnIn2S4 nanosheets on photocatalytic hydrogen production. In addition, the ZnIn2S4 nanosheets revealed the highest photocatalytic hydrogen production from seawater, which was about 209.4 and 106.7 times higher than that of In2S3 nanosheets and ZnS nanosheets, respectively. The presence of sulfur vacancies in ZnIn2S4 nanosheets offers promising opportunities for developing highly efficient and stable photocatalysts for photocatalytic hydrogen production from seawater under visible-light irradiation.

1. Introduction

Microwave-assisted synthesis is a popular technique for the rapid and facile preparation of organic/inorganic nanostructured materials in various media [1,2]. In addition, it offers control for internal and volumetric heating of materials, making it an environmentally friendly method for fabricating new material structures, components, and devices [3]. Microwave-assisted synthesis has been widely used to synthesize nanostructured materials, including metals, nanoporous materials, colloidal nanocrystals, polymer nanocomposites, and inorganic or semiconducting nanomaterials [2,4,5,6,7,8,9]. Microwave-assisted synthesis of nanostructures offers several advantages, including contactless heat transfer to the reactant sample, short reaction time, high selectivity and yield, energy efficiency, uniform and selective distribution of energy, control over size and temperature, improved safety and reproducibility, and excellent control over experimental parameters, making it an inexpensive, quick, clean, and versatile technique for the preparation of nanostructures [10,11,12,13,14]. In addition, microwave-assisted synthesis has distinct characteristics that render it viable for achieving extensive-scale industrial manufacturing [15,16].
Zinc indium sulfide (ZnIn2S4) is a highly efficient two-dimensional semiconductor photocatalyst from the AB2X4 family of ternary compounds, widely used in solar energy conversion and environmental purification [17,18,19,20]. ZnIn2S4 typically forms two-dimensional nanosheets due to its layered crystal structure, but these nanosheets quickly aggregate and assemble into three-dimensional nanostructures to minimize the surface energy for most reported ZnIn2S4 [19,21]. ZnIn2S4 has two crystal structures: hexagonal and cubic, with hexagonal ZnIn2S4 being the most researched in photocatalysis [22,23,24]. Recently, ZnIn2S4 has gained attention in photocatalytic hydrogen production, carbon dioxide conversion, and pollutant removal, owing to its excellent light absorption and strong redox capabilities [25,26,27,28]. However, the low separation degree of photogenerated charge carriers and the low electron-transport efficiency of ZnIn2S4 limit its application in photocatalysis [29]. In order to improve photocatalytic efficiency, introducing sulfur or zinc vacancies has proven effective [30,31]. So far, most of the research on ZnIn2S4 has focused on combining different materials to improve its photocatalytic hydrogen production performance, but there are few reports on preparing ZnIn2S4 with sulfur defects via the microwave-assisted synthesis method [32].
Previous research has reported significant instances of photocatalytic hydrogen production, predominantly employing deionized or pure water [33,34]. Nevertheless, the preparation of deionized or pure water necessitates the purification of fresh water, which results in energy wastage and contributes to the depletion of freshwater resources [35,36]. Furthermore, the increasing occurrence of abnormal climate changes, such as heavy rainfall or drought, has led to a scarcity of freshwater resources [37,38]. Consequently, developing a photocatalyst capable of effectively decomposing seawater to generate hydrogen would address this challenge and further enhance the situation [39]. Therefore, this study aimed to produce hydrogen through the photocatalytic decomposition of seawater using ZnIn2S4 nanosheets with sulfur defects, potentially reducing freshwater consumption and addressing the scarcity of freshwater resources.
This study successfully synthesized In2S3 nanosheets, ZnS nanosheets, and ZnIn2S4 nanosheets with sulfur vacancies for photocatalytic hydrogen production by microwave-assisted synthesis. The photocatalytic activities of ZnIn2S4 nanosheets with sulfur vacancies were approximately 33.3 times higher than those of In2S3 nanosheets and about 16.6 times higher than those of ZnS nanosheets under visible-light irradiation. Notably, utilizing ZnIn2S4 nanosheets with sulfur vacancies as a photocatalyst demonstrated their exceptional performance in efficiently splitting seawater under visible-light irradiation.

2. Materials and Methods

2.1. Chemicals

Anhydrous zinc chloride (ZnCl2, 98%), anhydrous indium(III) chloride (InCl3, 98%), thioacetamide (TAA, C2H5NS, 98%), sodium sulfide nonahydrate (Na2S·9H2O, 98%), sodium sulfite (Na2SO3, 98%), methanoic acid (CH2O2, 97%), folic acid dihydrate (C19H19N7O6∙2H2O, 97%), methanol (CH3OH, 99%), and sodium chloride (NaCl, 99%) were purchased from Alfa Aesar (Haverhill, MA, USA) without further purification. Deionized water (>18 MΩ·cm) was used throughout the experimental processes.

2.2. Synthesis of In2S3, ZnIn2S4, and ZnS Nanosheets

A facile microwave-assisted synthesis method can be used to synthesize In2S3 nanosheets, ZnIn2S4 nanosheets, and ZnS nanosheets. For In2S3 nanosheets, 0.0045 g of InCl3 and 0.0075 g of TAA were dissolved in the 20 mL solvent (deionized water: ethanol = 3:1) and put into a Teflon-lined digestion vessel. For ZnIn2S4 nanosheets, 0.0068 g of ZnCl2, 0.0045 g of InCl3, and 0.0075 gTAA were dissolved in the 20 mL solvent (deionized water: ethanol = 3:1) and put into the Teflon-lined digestion vessel. For ZnS nanosheets, 0.0136 g of ZnCl2, and 0.0075 g TAA were dissolved in the 20 mL solvent (deionized water: ethanol = 3:1) and put into the Teflon-lined digestion vessel. A safety shield was placed around the vessel after sealing the reaction solution inside the vessel, using the vessel cover as an overpressure release valve. The vessel was then heated at 140 °C for 1 h using a microwave synthesizer (ETHOS EASY, Milestone, Sorisole, Italy). Following the cooling stage, the as-prepared samples were thoroughly cleaned with deionized water, purified through centrifugation, and dried at 70 °C for 2 h.

2.3. Characterization

The crystal structure of the as-prepared photocatalysts was investigated by X-ray diffractometry (XRD) using a Bruker D2 phaser system (Billerica, MA, USA) in the 2θ range of 25–65° with Cu Kα (λ = 1.5418 Å). In addition, field-emission scanning electron microscopy (FESEM) was used to investigate the surface morphology of the as-prepared photocatalysts using a Hitachi S-4800 microscope (Tokyo, Japan) operating at a 15 kV accelerating voltage. The specific surface areas of the as-prepared photocatalysts were measured using the nitrogen adsorption technique (ASAP 2020, Micromeritics, GA, USA). Field-emission transmission electron microscopy (FETEM) at an accelerating voltage of 200 kV was employed to characterize the microstructures and composition of the ZnIn2S4 nanosheets using a JEOL 2100F microscope (Tokyo, Japan). The surface chemical composition and sulfur vacancies of the ZnIn2S4 nanosheets were detected by using X-ray photoelectron spectroscopy (XPS, ULVAC-PHI PHI 5000 Versaprobe II system, Chigasaki, Japan) with an Al K source. The as-prepared photocatalysts’ UV–vis diffuse reflectance spectra were recorded using a UV–vis spectrophotometer (PerkinElmer Lambda 650 S, Waltham, MA, USA) and analyzed using a Protrustech MRI532S instrument (Tainan, Taiwan). A three-electrode system was utilized with the assistance of the Zennium electrochemical workstation (Zahhner, Kronach, Germany) for measuring electrochemical impedance spectroscopy (EIS).

2.4. Photocatalytic Hydrogen Production Experiment

In order to investigate photocatalytic hydrogen production, an experiment was conducted with the PCX50B Discover photocatalytic reaction system (Perfect Light Technology, Beijing, China). For the experiment, 25 mg of the as-prepared photocatalysts was mixed with 50 mL of different water bases, including deionized water, reverse-osmosis water, tap water, and seawater. Different sacrificial agents, such as sodium sulfide, sodium sulfate, methanol, and methanoic acid, were added to the water bases at 0.1 M concentrations. The mixtures were then placed in 60 mL quartz tubes, and Ar gas was introduced for 30 min to eliminate air. The tubes were sealed with rubber stoppers and irradiated for 3 h using a 5 W blue LED (λ = 420 nm) as a visible-light source. The hydrogen produced was quantified using a gas chromatograph equipped with a thermal conductivity detector (TCD).

3. Results and Discussion

The crystalline phases of the prepared photocatalysts were analyzed using X-ray diffraction (XRD). Figure 1 shows the XRD patterns of In2S3 nanosheets, ZnIn2S4 nanosheets, and ZnS nanosheets grown via rapid microwave-assisted synthesis at 140 °C for 1 h. For In2S3 nanosheets (Figure 1a), the eight peaks at 27.7°, 28.9°, 33.6°, 43.9°, 48.0°, 51.9°, 56.3°, and 59.8° correspond to the (311), (222), (400), (511), (440), (610), (533), and (444) planes of typical cubic In2S3 (JCPDS card No. 32-0456), respectively. For ZnIn2S4 nanosheets (Figure 1b), the five peaks at 27.7°, 28.8°, 47.3°, 52.2°, and 56.4°correspond to the (102), (103), (110), (1012), and (203) planes of typical hexagonal ZnIn2S4 (JCPDS card No. 72-0773), respectively. For ZnS nanosheets (Figure 1c), the four peaks at 27.7°, 28.8°, 47.3°, and 56.4°correspond to the (100), (002), (110), and (112) planes of typical hexagonal ZnS (JCPDS card No. 80-0007), respectively. The absence of diffraction peaks corresponding to impurities further confirms the phase purity of the prepared photocatalysts.
The surface morphologies of the synthesized In2S3 nanosheets, ZnIn2S4 nanosheets, and ZnS nanosheets were characterized using FESEM, as depicted in Figure 2. In Figure 2a, small and interconnected In2S3 nanosheets can be observed, and the thickness of the nanosheets is about 15–40 nm. ZnIn2S4 nanosheets are larger and interconnected, with a thickness of about 15–60 nm, as shown in Figure 2b. Figure 2c reveals that the ZnS nanosheets are stacked on one another and form a spherical structure. We utilized a BET (Brunauer–Emmett–Teller) analyzer to examine the specific surface areas of In2S3 nanosheets, ZnIn2S4 nanosheets, and ZnS nanosheets. The nitrogen adsorption–desorption isotherms revealed that the specific surface areas of these photocatalysts were as follows: 99.8 (In2S3 nanosheets), 77.3 (ZnIn2S4 nanosheets), and 35.8 (ZnS nanosheets) m2g−1. The specific surface areas of the In2S3 nanosheets and ZnIn2S4 nanosheets were significantly larger than that of the ZnS nanosheets, which is consistent with the FESEM characterization results.
Figure 3a shows the FETEM image of ZnIn2S4 nanosheets, which is consistent with the FESEM image of the nanosheet structure made of lamellar stacks. Additionally, the polycrystalline nature of the ZnIn2S4 nanosheets is substantiated by the presence of characteristic polycrystalline diffraction rings observed in the selected-area electron diffraction (SAED) patterns, as shown in Figure 3b. The prominent diffraction ring patterns correspond to the crystal planes (102), (103), (110), (1012), and (203), which are indicative of the typical hexagonal structure of ZnIn2S4 (JCPDS card No. 72-0773). These findings are consistent with the results obtained from the XRD analysis, confirming the polycrystalline properties of the ZnIn2S4 nanosheets. Figure 3c reveals that the HRTEM image of the lattice fringes of 0.331 nm and 0.322 nm are assigned to the (101) and (102) planes of hexagonal ZnIn2S4, respectively. Moreover, the mapping images obtained from the energy-dispersive spectrometry (EDS) analysis (Figure 3d) reveal a uniform and interconnected distribution of elemental Zn, In, and S. These results provide conclusive evidence of the successful synthesis of ZnIn2S4 nanosheets.
X-ray photoelectron spectroscopy (XPS) was utilized to determine the elemental composition and chemical configurations of the ZnIn2S4 nanosheets. Figure 4a displays the complete XPS spectrum, which confirms the presence of elemental Zn, In, S, C, and O in the ZnIn2S4 nanosheets. The appearance of C 1s and O 1s signals can be explained by the presence of pump oil in the XPS equipment’s vacuum system and the adsorption of oxygen atoms, respectively. Figure 4b shows that the high-resolution XPS spectrum of Zn 2p reveals two peaks at 1021.2 eV and 1044.2 eV, which are assigned to Zn 2p3/2 and Zn 2p1/2, respectively, indicating the presence of Zn2+ in ZnIn2S4. Figure 4c shows that the high-resolution XPS spectrum of In 3d reveals two peaks at 444.2 eV and 451.8 eV, which are assigned to In 3d5/2 and In 3d3/2, respectively, indicating the presence of In3+ in ZnIn2S4. Compared with pure ZnIn2S4 (S 2p1/2 = 161.4 eV and S 2p3/2 = 162.7 eV), the XPS peaks of S 2p1/2 and 2p3/2 in ZnIn2S4 nanosheets with sulfur defects shift to lower binding energies of 160.9 and 161.9 eV, respectively [25,31,40]. This peak shift can primarily be attributed to the presence of sulfur vacancies, indicating the influence of these vacancies on the electronic structure of the materials [41]. This result demonstrates that rapid microwave-assisted synthesis can facilitate the formation of ZnIn2S4 nanosheets with sulfur defects during the same reaction process.
In order to assess the photocatalytic characteristics of the catalyst, a photocatalytic hydrogen evolution experiment was conducted in deionized water. The hydrogen evolution rate (HER) activities of the as-prepared photocatalysts (In2S3 nanosheets, ZnIn2S4 nanosheets, and ZnS nanosheets) and commercial photocatalysts (ZnO nanopowder and TiO2 nanopowder) were evaluated under visible-light irradiation (5 W blue LED, λ = 420 nm) using 0.1 M Na2S as a sacrificial reagent in deionized water with pH = 12, as shown in Figure 5a. The ZnIn2S4 nanosheets exhibited an HER value of 24.95 μmolh−1g−1, which is about 33.3 and 16.6 times higher than 0.75 (In2S3 nanosheets) and 1.5 (ZnS nanosheets) μmolh−1g−1, respectively. On the other hand, the pure ZnIn2S4 nanosheets synthesized by a hydrothermal method revealed significantly lower HER values of 0.065 μmolh−1g−1 under the same photocatalytic reaction conditions [25]. These outcomes indicate that utilizing the rapid microwave-assisted synthesis approach facilitates the growth of ZnIn2S4 nanosheets with sulfur vacancies, leading to a notable improvement in their photocatalytic hydrogen evolution.
In photocatalytic hydrogen production, sacrificial reagents are often employed to enhance the efficiency of oxidation reactions in aqueous environments, compensating for the inefficiency of pure water oxidation [42]. Figure 5b illustrates the HER values of ZnIn2S4 nanosheets under visible-light irradiation using four different sacrificial reagents: sodium sulfide (Na2S), sodium sulfite (Na2SO3), folic acid (FA), and methanol (CH3OH). The HER values of the ZnIn2S4 nanosheets were 24.95 (Na2S), 21.6 (Na2SO3), 0 (FA), and 0 μmolh−1g−1 (CH3OH). In photocatalytic hydrogen generation, the sacrificial agent serves the dual role of providing electrons for proton reduction and scavenging holes to impede the recombination of electron–hole pairs, thereby improving the process’s efficiency [43,44]. Consequently, using sodium sulfide or sodium sulfate as a sacrificial reagent can enhance the reduction/oxidation reaction, reduce the photocorrosion of metal sulfide materials, and enhance their photocatalytic hydrogen production capability [42,45].
In general, concentrated sacrificial reagents are preferable, as they promote better diffusion of the reacting species toward the surface of the photocatalysts [46,47]. However, it is essential to note that the highest hydrogen evolution rate cannot be achieved with diluted or highly concentrated sacrificial reagents [48]. The impact of Na2S concentrations on the photocatalytic efficiency of ZnIn2S4 nanosheets is shown in Figure 5c. The average HER values of the ZnIn2S4 nanosheets were recorded as follows: 0 (without Na2S), 8.1 (0.01 M Na2S), 21.4 (0.025 M Na2S), 29.55 (0.05 M Na2S), and 24.95 μmolh−1g−1 (0.1 M Na2S). The ZnIn2S4 nanosheets with appropriate Na2S concentrations can exhibit the highest HER under visible-light irradiation.
The optimal pH level for photocatalytic hydrogen production is primarily determined by the characteristics of the sacrificial agent and its affinity for adsorption onto the surface of the photocatalyst [49]. The impact of pH values on the photocatalytic efficiency of ZnIn2S4 nanosheets is illustrated in Figure 5d. The average HER values of ZnIn2S4 nanosheets were recorded as follows: 134.5 (pH = 1), 143.9 (pH = 3), 54.4 (pH = 6), 31.25 (pH = 9), 29.55 (pH = 12), and 16.3 μmolh−1g−1 (pH = 12.9, without adjustment). An increase in the pH value from 1 to 3 significantly enhanced the photocatalytic hydrogen production efficiency. This enhancement can be attributed to the increased dissociation of HS and S2− species with the increasing pH value [47]. However, low pH values may result in photocorrosion of the catalyst and a reduction in hydrogen production efficiency [35]. Hence, the ZnIn2S4 nanosheets exhibited the highest photocatalytic hydrogen production efficiency at pH = 3.
The UV–vis diffuse reflectance spectra of In2S3, ZnIn2S4, and ZnS nanosheets within the range of 300–800 nm are illustrated in Figure 6a. For In2S3 nanosheets, a pronounced absorption can be observed below 600 nm, indicating strong intrinsic interband transition absorption specific to In2S3. For ZnIn2S4 nanosheets, a notable absorption can be observed below 500 nm, corresponding to the intrinsic interband transition absorption of In2S3. Finally, for ZnS nanosheets, a significant absorption can be observed below 400 nm in the UV region, attributed to the intrinsic interband transition absorption of ZnS. These observations indicate that both In2S3 nanosheets and ZnIn2S4 nanosheets exhibit substantial absorption of visible light. The direct optical bandgap, Eg, of the materials was determined by employing the equation (αhν)2 = A(hν − Eg), where hν represents the photon energy in electron volts (eV), α is the absorption coefficient, and A is a material constant [50]. This analysis was conducted based on the data presented in Figure 6b. Consistent with prior findings, the results indicate that the In2S3, ZnIn2S4, and ZnS bandgaps were measured to be 2.05 eV, 2.82 eV, and 3.32 eV, respectively [25,51,52].
Photoluminescence (PL) spectra are a valuable tool for examining photogenerated charge carriers’ trapping, migration, and transfer efficiency in semiconductors [53,54]. By analyzing the PL spectra, it is possible to gain insights into the recombination and annihilation processes of photogenerated electron–hole pairs in semiconductors [55]. Figure 7a displays the PL spectra of In2S3, ZnIn2S4, and ZnS nanosheets. The ZnS nanosheets demonstrate more pronounced emission characteristics than the In2S3 and ZnIn2S4 nanosheets, suggesting a higher degree of recombination of photogenerated charge carriers within the photocatalyst. Electrochemical impedance spectroscopy (EIS) is presented in Figure 7b. The arc radii of the EIS Nyquist curves of the samples were in the following order: ZnS > In2S3 > ZnIn2S4. ZnIn2S4′s arc radius was smaller than those of In2S3 and ZnS. The smaller radius observed in the EIS test indicates lower charge transfer resistance and higher efficiency in separating charge carriers, suggesting a faster electron transfer process [56,57,58]. Consequently, the EIS results confirm that ZnIn2S4 nanosheets with sulfur vacancies facilitate the migration of photogenerated charge carriers, resulting in enhanced photocatalytic activity for hydrogen evolution.
The band structure greatly influences the generation and migration of charge carriers. Therefore, the valence band (VB) XPS was conducted on the obtained ZnIn2S4 nanosheets to investigate the valence band location. Figure 8a illustrates the typical valence band density of states (DOS) characteristics of ZnIn2S4 nanosheets with the valence band position at 0.76 eV. Based on the equation ECB = EVB − Eg and the measured bandgap of ZnIn2S4 (Figure 6b), the conduction band (CB) potential is estimated to occur at −2.06 eV. After carefully examining the findings above, we developed a hypothetical reaction mechanism for photocatalytic hydrogen production using ZnIn2S4 nanosheets. The proposed mechanism is visually depicted in Figure 8b, providing a realistic representation of the process. The presence of sulfur vacancies in ZnIn2S4 leads to the formation of defect levels within the conduction band (CB). In addition, these sulfur vacancies act as electron traps, effectively capturing photogenerated electrons and preventing their direct recombination with photoinduced holes. The trapped electrons in the sulfur vacancies of ZnIn2S4 can subsequently participate in the reduction of 2H+ to produce H2. At the same time, the photogenerated holes in ZnIn2S4 can oxidize sulfide ions (S2−, sodium sulfide dissociated).
Given that approximately 93% of the Earth’s water exists in the form of seawater, utilizing seawater as a source for hydrogen production through water splitting offers a practical solution to conserve freshwater resources for various purposes, such as agriculture, industry, and human consumption [39,59]. Furthermore, the hydrogen produced through the photocatalytic decomposition of seawater can be utilized to generate pure water through a fuel cell generator, adding to its potential benefits. The exceptional versatility of ZnIn2S4 nanosheets is demonstrated by their photocatalytic capability to generate hydrogen using various water sources. As depicted in Figure 9a, the dispersion of ZnIn2S4 nanosheets in 50 mL of different water sources, including deionized water (DI), reverse-osmosis water (RO), tap water (TW), and seawater (SW), was observed under visible-light irradiation with a pH value of 3 and the presence of 0.1 M Na2S as a scavenger. The average HER values of ZnIn2S4 nanosheets were measured as 143.9 (DI), 158.9 (RO), 207.7 (TW), and 992.4 µmolh−1g−1 (SW). With the increasing complexity of the water matrix, there is a gradual enhancement in the efficiency of photocatalytic hydrogen production. In order to investigate the influence of sodium chloride (NaCl) on the photocatalytic effect, this study included the addition of 35 psu NaCl (35 g/L) to deionized water to assess its impact on hydrogen production, as shown in Figure 9b. This result indicates that adding sodium chloride to deionized water did not significantly improve the efficiency of photocatalytic hydrogen production. Therefore, it can be inferred that sodium chloride might not be the primary factor for enhancing photocatalytic hydrogen production. The average HER values of photocatalysts, such as In2S3 nanosheets, ZnIn2S4 nanosheets, ZnS nanosheets, ZnO nanopowder, and TiO2 nanopowder, were assessed under visible-light irradiation using 0.1 M Na2S as a sacrificial reagent in seawater with a pH value of 3, as depicted in Figure 9c. The average HER values of the photocatalysts were measured as 4.74 (In2S3 nanosheets), 992.4 (ZnIn2S4 nanosheets), 9.3 (ZnS nanosheets), 3.7 (ZnO nanopowder), and 0 µmolh−1g−1 (TiO2 nanopowder). The ZnIn2S4 nanosheets exhibited the highest photocatalytic hydrogen production from seawater, which was about 209.4 and 106.7 times higher than that of In2S3 nanosheets and ZnS nanosheets, respectively. These results demonstrate that the microwave-assisted synthesis method promotes the formation of ZnIn2S4 nanosheets with sulfur vacancies, significantly enhancing their photocatalytic activity for hydrogen production from seawater.
The reusability of ZnIn2S4 nanosheets was investigated through an eight-cycle photocatalytic process, wherein the seawater containing 0.1 M Na2S at pH = 3 was renewed and subjected to visible-light irradiation, as shown in Figure 10. As a result, the average HER values of ZnIn2S4 nanosheets were recorded as follows: 992.4, 1277.5, 1445.6, 1694.8, 1782.3, 1688.2, 1676.9, and 1518.0 μmolh−1g−1. The photocatalytic hydrogen production efficiency from seawater exhibited an initial increase followed by a gradual decline as the number of cycles increased. The HER of ZnIn2S4 nanosheets showed a notable increase of 1.8 times after the fifth cycle compared to the first cycle. A noticeable distinction was observed between the XRD spectrum before the cycle (Figure 1b) and after the eighth cycle (Figure 11a). Notably, three additional crystal phases of ZnS (JCPDS card No. 72-0163), Zn2SO4 (JCPDS card No. 86-0802), and NaZn(OH)3 (JCPDS card No. 87-0762) were identified following the cycling process. This phenomenon can be attributed to the interaction between ZnIn2S4 nanosheets and seawater, forming ZnIn2S4/ZnS nanocomposites. ZnS possesses sulfur vacancies at the defect level, which can facilitate the migration of photogenerated charge carriers to enhance photocatalytic performance [60,61]. The morphology of ZnIn2S4 nanosheets exhibited similarity before the cycle (Figure 2b) and after the eighth cycle (Figure 11b), and noticeable changes were obtained in the roughness of the surface following the cycling process. However, a slight decrease in HER was observed in the sixth cycle, possibly due to the loss of photocatalysts during multiple centrifugations. Furthermore, these results confirm the outstanding stability and reusability of ZnIn2S4 nanosheets, highlighting their potential for wide-ranging and diverse applications in various fields.

4. Conclusions

This study employed a convenient and rapid microwave-assisted synthesis method to produce In2S3 nanosheets, ZnS nanosheets, and ZnIn2S4 nanosheets with sulfur vacancies. The ZnIn2S4 nanosheets demonstrated significantly enhanced photocatalytic hydrogen production performance compared to In2S3 nanosheets and ZnS nanosheets, achieving approximately 33.3 times and 16.6 times higher activity under visible-light irradiation, respectively. We further investigated the impacts of various factors, such as materials, sacrificial reagents, and pH values, on the photocatalytic hydrogen production of ZnIn2S4 nanosheets. The ZnIn2S4 nanosheets exhibited the highest photocatalytic hydrogen production efficiency from seawater, surpassing In2S3 nanosheets and ZnS nanosheets by approximately 209.4 times and 106.7 times, respectively. The presence of sulfur vacancies in ZnIn2S4 nanosheets holds great promise for developing highly efficient and stable photocatalysts for hydrogen production from seawater under visible-light irradiation. This research can potentially conserve precious freshwater resources and use light energy to split seawater into hydrogen energy.

Author Contributions

Funding acquisition, resources, methodology, project administration, resources, software, supervision, validation, writing—original draft, writing—review and editing, Y.-C.C. (Yu-Cheng Chang); formal analysis, investigation, data curation, Y.-C.C. (Yung-Chang Chiao); funding acquisition, resources, software, P.-C.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Technology of Taiwan, Taiwan (MOST 109-2221-E-035-041-MY3).

Data Availability Statement

Not applicable.

Acknowledgments

The authors appreciate the Precision Instrument Support Center of Feng Chia University for providing the fabrication and measurement facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gawande, M.B.; Shelke, S.N.; Zboril, R.; Varma, R.S. Microwave-Assisted Chemistry: Synthetic Applications for Rapid Assembly of Nanomaterials and Organics. Acc. Chem. Res. 2014, 47, 1338–1348. [Google Scholar] [CrossRef]
  2. Wang, J.; Wu, W.; Kondo, H.; Fan, T.; Zhou, H. Recent progress in microwave-assisted preparations of 2D materials and catalysis applications. Nanotechnology 2022, 33, 342002. [Google Scholar] [CrossRef]
  3. Chang, T.-H.; Chang, Y.-C.; Lee, C.-I.; Lin, Y.-R.; Ko, F.-H. Optimization Temperature Programming of Microwave-Assisted Synthesis ZnO Nanoneedle Arrays for Optical and Surface-Enhanced Raman Scattering Applications. Nanomaterials 2022, 12, 3989. [Google Scholar] [CrossRef] [PubMed]
  4. Zhu, Y.-J.; Chen, F. Microwave-Assisted Preparation of Inorganic Nanostructures in Liquid Phase. Chem. Rev. 2014, 114, 6462–6555. [Google Scholar] [CrossRef] [PubMed]
  5. Menon, S.G.; Choudhari, K.S.; Shivashankar, S.A.; Chidangil, S.; Kulkarni, S.D. Microwave solution route to ceramic ZnAl2O4 nanoparticles in 10 minutes: Inversion and photophysical changes with thermal history. New J. Chem. 2017, 41, 5420–5428. [Google Scholar] [CrossRef]
  6. Tompsett, G.A.; Conner, W.C.; Yngvesson, K.S. Microwave Synthesis of Nanoporous Materials. ChemPhysChem 2006, 7, 296–319. [Google Scholar] [CrossRef]
  7. Mirzaei, A.; Neri, G. Microwave-assisted synthesis of metal oxide nanostructures for gas sensing application: A review. Sens. Actuators B Chem. 2016, 237, 749–775. [Google Scholar] [CrossRef]
  8. Meng, L.-Y.; Wang, B.; Ma, M.-G.; Lin, K.-L. The progress of microwave-assisted hydrothermal method in the synthesis of functional nanomaterials. Mater. Today Chem. 2016, 1–2, 63–83. [Google Scholar] [CrossRef]
  9. Shen, P.-S.; Tseng, C.-M.; Kuo, T.-C.; Shih, C.-K.; Li, M.-H.; Chen, P. Microwave-assisted synthesis of titanium dioxide nanocrystalline for efficient dye-sensitized and perovskite solar cells. Sol. Energy 2015, 120, 345–356. [Google Scholar] [CrossRef]
  10. Schwenke, A.M.; Hoeppener, S.; Schubert, U.S. Synthesis and Modification of Carbon Nanomaterials utilizing Microwave Heating. Adv. Mater. 2015, 27, 4113–4141. [Google Scholar] [CrossRef] [PubMed]
  11. Architha, N.; Ragupathi, M.; Shobana, C.; Selvankumar, T.; Kumar, P.; Lee, Y.S.; Kalai Selvan, R. Microwave-assisted green synthesis of fluorescent carbon quantum dots from Mexican Mint extract for Fe3+ detection and bio-imaging applications. Environ. Res. 2021, 199, 111263. [Google Scholar] [CrossRef]
  12. Chin, C.D.; Treadwell, L.J.; Wiley, J.B. Microwave Synthetic Routes for Shape-Controlled Catalyst Nanoparticles and Nanocomposites. Molecules 2021, 26, 3647. [Google Scholar] [CrossRef]
  13. Sharifvaghefi, S.; Zheng, Y. Microwave vs conventional heating in hydrogen production via catalytic dry reforming of methane. Resour. Chem. Mater. 2022, 1, 290–307. [Google Scholar] [CrossRef]
  14. Motasemi, F.; Ani, F.N. A review on microwave-assisted production of biodiesel. Renew. Sustain. Energy Rev. 2012, 16, 4719–4733. [Google Scholar] [CrossRef]
  15. Devi, N.; Sahoo, S.; Kumar, R.; Singh, R.K. A review of the microwave-assisted synthesis of carbon nanomaterials, metal oxides/hydroxides and their composites for energy storage applications. Nanoscale 2021, 13, 11679–11711. [Google Scholar] [CrossRef] [PubMed]
  16. Nayak, S.N.; Bhasin, C.P.; Nayak, M.G. A review on microwave-assisted transesterification processes using various catalytic and non-catalytic systems. Renew. Energ. 2019, 143, 1366–1387. [Google Scholar] [CrossRef]
  17. Pan, J.; Guan, Z.; Yang, J.; Li, Q. Facile fabrication of ZnIn2S4/SnS2 3D heterostructure for efficient visible-light photocatalytic reduction of Cr(VI). Chin. J. Catal. 2020, 41, 200–208. [Google Scholar] [CrossRef]
  18. Janani, R.; Preethi, V.R.; Singh, S.; Rani, A.; Chang, C.-T. Hierarchical Ternary Sulfides as Effective Photocatalyst for Hydrogen Generation Through Water Splitting: A Review on the Performance of ZnIn2S4. Catalysts 2021, 11, 277. [Google Scholar] [CrossRef]
  19. Yang, R.; Mei, L.; Fan, Y.; Zhang, Q.; Zhu, R.; Amal, R.; Yin, Z.; Zeng, Z. ZnIn2S4-Based Photocatalysts for Energy and Environmental Applications. Small Methods 2021, 5, 2100887. [Google Scholar] [CrossRef]
  20. Mamiyev, Z.; Balayeva, N.O. Metal Sulfide Photocatalysts for Hydrogen Generation: A Review of Recent Advances. Catalysts 2022, 12, 1316. [Google Scholar] [CrossRef]
  21. Wang, J.; Sun, S.; Zhou, R.; Li, Y.; He, Z.; Ding, H.; Chen, D.; Ao, W. A review: Synthesis, modification and photocatalytic applications of ZnIn2S4. J. Mater. Sci. Technol. 2021, 78, 1–19. [Google Scholar] [CrossRef]
  22. Zhou, H.; Wang, L.; Shi, H.; Zhang, H.; Wang, Y.; Bai, S.; Yang, Y.; Li, Y.; Zhang, T.; Zhang, H. Highly efficient solar-driven photocatalytic hydrogen evolution by a ternary 3D ZnIn2S4–MoS2 microsphere/1D TiO2 nanobelt heterostructure. New J. Chem. 2021, 45, 14167–14176. [Google Scholar] [CrossRef]
  23. Zhang, G.; Wu, H.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. A mini-review on ZnIn2S4-Based photocatalysts for energy and environmental application. Green Energy Environ. 2022, 7, 176–204. [Google Scholar] [CrossRef]
  24. Chen, Y.; Huang, R.; Chen, D.; Wang, Y.; Liu, W.; Li, X.; Li, Z. Exploring the Different Photocatalytic Performance for Dye Degradations over Hexagonal ZnIn2S4 Microspheres and Cubic ZnIn2S4 Nanoparticles. ACS Appl. Mater. Interfaces 2012, 4, 2273–2279. [Google Scholar] [CrossRef]
  25. Chang, Y.-C.; Tasi, C.-L.; Ko, F.-H. Construction of ZnIn2S4/ZnO heterostructures with enhanced photocatalytic decomposition and hydrogen evolution under blue LED irradiation. Int. J. Hydrogen Energy 2021, 46, 10281–10292. [Google Scholar] [CrossRef]
  26. Shen, S.; Zhao, L.; Guan, X.; Guo, L. Improving visible-light photocatalytic activity for hydrogen evolution over ZnIn2S4: A case study of alkaline-earth metal doping. J. Phys. Chem. Solids 2012, 73, 79–83. [Google Scholar] [CrossRef]
  27. Shen, J.; Zai, J.; Yuan, Y.; Qian, X. 3D hierarchical ZnIn2S4: The preparation and photocatalytic properties on water splitting. Int. J. Hydrogen Energy 2012, 37, 16986–16993. [Google Scholar] [CrossRef]
  28. Sabbah, A.; Shown, I.; Qorbani, M.; Fu, F.-Y.; Lin, T.-Y.; Wu, H.-L.; Chung, P.-W.; Wu, C.-I.; Santiago, S.R.M.; Shen, J.-L.; et al. Boosting photocatalytic CO2 reduction in a ZnS/ZnIn2S4 heterostructure through strain-induced direct Z-scheme and a mechanistic study of molecular CO2 interaction thereon. Nano Energy 2022, 93, 106809. [Google Scholar] [CrossRef]
  29. Chen, J.; Li, K.; Cai, X.; Zhao, Y.; Gu, X.; Mao, L. Sulfur vacancy-rich ZnIn2S4 nanosheet arrays for visible-light-driven water splitting. Mater. Sci. Semicond. Process. 2022, 143, 106547. [Google Scholar] [CrossRef]
  30. 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]
  31. 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] [PubMed]
  32. Shen, S.; Chen, J.; Wang, X.; Zhao, L.; Guo, L. Microwave-assisted hydrothermal synthesis of transition-metal doped ZnIn2S4 and its photocatalytic activity for hydrogen evolution under visible light. J. Power Sources 2011, 196, 10112–10119. [Google Scholar] [CrossRef]
  33. Zhang, R.; Gong, K.; Cao, S.; Du, F. Amorphous sulfur-rich CoSx nanodots as highly efficient cocatalyst to promote photocatalytic hydrogen evolution over TiO2. Int. J. Hydrogen Energy 2022, 47, 39875–39885. [Google Scholar] [CrossRef]
  34. Chen, L.; Wang, D.; Xia, Y.; Liang, R.; Huang, R.; Yan, G. Decoration of nicale phosphide nanoparticles on CdS nanorods for enhanced photocatalytic hydrogen evolution. Int. J. Hydrogen Energy 2022, 47, 28486–28494. [Google Scholar] [CrossRef]
  35. Chang, Y.-C.; Zeng, C.-J.; Chen, C.-Y.; Tsay, C.-Y.; Lee, G.-J.; Wu, J.J. NiS/Pt loaded on electrospun TiO2 nanofiber with enhanced visible-light-driven photocatalytic hydrogen production. Mater. Res. Bull. 2023, 157, 112041. [Google Scholar] [CrossRef]
  36. Liu, Y.; Li, D.; Wan, S.; Wang, F.; Dou, W.; Xu, X.; Li, S.; Ma, R.; Qi, L. A long short-term memory-based model for greenhouse climate prediction. Int. J. Intell. Syst. 2022, 37, 135–151. [Google Scholar] [CrossRef]
  37. Gruda, N.; Bisbis, M.; Tanny, J. Influence of climate change on protected cultivation: Impacts and sustainable adaptation strategies—A review. J. Clean. Prod. 2019, 225, 481–495. [Google Scholar] [CrossRef]
  38. Levy, K.; Woster, A.P.; Goldstein, R.S.; Carlton, E.J. Untangling the Impacts of Climate Change on Waterborne Diseases: A Systematic Review of Relationships between Diarrheal Diseases and Temperature, Rainfall, Flooding, and Drought. Environ. Sci. Technol. 2016, 50, 4905–4922. [Google Scholar] [CrossRef] [Green Version]
  39. Guan, X.; Chowdhury, F.A.; Pant, N.; Guo, L.; Vayssieres, L.; Mi, Z. Efficient Unassisted Overall Photocatalytic Seawater Splitting on GaN-Based Nanowire Arrays. J. Phys. Chem. C 2018, 122, 13797–13802. [Google Scholar] [CrossRef]
  40. Chen, H.; Li, M.; Gao, J.; Yang, D.; Li, Z.; Liu, H.; Xie, Y.; Guo, L.; Zhou, W. Synergism of sulfur vacancy and Schottky junction in Ni/ZnIn2S4 nanosheet assembly for efficient charge separation and photocatalytic hydrogen evolution. Appl. Surf. Sci. 2023, 628, 157385. [Google Scholar] [CrossRef]
  41. Zhang, M.; Tan, P.; Yang, L.; Zhai, H.; Liu, H.; Chen, J.; Ren, R.; Tan, X.; Pan, J. Sulfur vacancy and p-n junction synergistically boosting interfacial charge transfer and separation in ZnIn2S4/NiWO4 heterostructure for enhanced photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2023, 634, 817–826. [Google Scholar] [CrossRef] [PubMed]
  42. Chang, Y.-C.; Chiao, Y.-C.; Fun, Y.-X. Cu2O/CuS/ZnS Nanocomposite Boosts Blue LED-Light-Driven Photocatalytic Hydrogen Evolution. Catalysts 2022, 12, 1035. [Google Scholar] [CrossRef]
  43. Corredor, J.; Rivero, M.J.; Rangel, C.M.; Gloaguen, F.; Ortiz, I. Comprehensive review and future perspectives on the photocatalytic hydrogen production. J. Chem. Technol. Biotechnol. 2019, 94, 3049–3063. [Google Scholar] [CrossRef] [Green Version]
  44. Kumaravel, V.; Imam, M.D.; Badreldin, A.; Chava, R.K.; Do, J.Y.; Kang, M.; Abdel-Wahab, A. Photocatalytic Hydrogen Production: Role of Sacrificial Reagents on the Activity of Oxide, Carbon, and Sulfide Catalysts. Catalysts 2019, 9, 276. [Google Scholar] [CrossRef] [Green Version]
  45. Lin, Y.-R.; Chang, Y.-C.; Chiao, Y.-C.; Ko, F.-H. Au@CdS Nanocomposites as a Visible-Light Photocatalyst for Hydrogen Generation from Tap Water. Catalysts 2022, 13, 33. [Google Scholar] [CrossRef]
  46. Markovskaya, D.V.; Cherepanova, S.V.; Saraev, A.A.; Gerasimov, E.Y.; Kozlova, E.A. Photocatalytic hydrogen evolution from aqueous solutions of Na2S/Na2SO3 under visible light irradiation on CuS/Cd0.3Zn0.7S and NizCd0.3Zn0.7S1+z. Chem. Eng. J. 2015, 262, 146–155. [Google Scholar] [CrossRef]
  47. Bao, N.; Shen, L.; Takata, T.; Domen, K. Self-Templated Synthesis of Nanoporous CdS Nanostructures for Highly Efficient Photocatalytic Hydrogen Production under Visible Light. Chem. Mater. 2008, 20, 110–117. [Google Scholar] [CrossRef]
  48. Gomathisankar, P.; Hachisuka, K.; Katsumata, H.; Suzuki, T.; Funasaka, K.; Kaneco, S. Photocatalytic hydrogen production with CuS/ZnO from aqueous Na2S+Na2SO3 solution. Int. J. Hydrogen Energy 2013, 38, 8625–8630. [Google Scholar] [CrossRef]
  49. Mills, A.; Le Hunte, S. An overview of semiconductor photocatalysis. J. Photochem. Photobiol. A Chem. 1997, 108, 1–35. [Google Scholar] [CrossRef]
  50. Safizade, B.; Masoudpanah, S.M.; Hasheminiasari, M.; Ghasemi, A. Photocatalytic activity of BiFeO3/ZnFe2O4 nanocomposites under visible light irradiation. RSC Adv. 2018, 8, 6988–6995. [Google Scholar] [CrossRef] [Green Version]
  51. Chang, Y.-C.; Syu, S.-Y.; Lu, M.-Y. Fabrication of In(OH)3–In2S3–Cu2O nanofiber for highly efficient photocatalytic hydrogen evolution under blue light LED excitation. Int. J. Hydrogen Energy 2023, 48, 9318–9332. [Google Scholar] [CrossRef]
  52. Khan, M.S.; Shi, L.; Zou, B.; Ullah, H. Effect of Vanadium doping on optoelectronic and magnetic properties of wurtzite ZnS crystal. Optik 2020, 204, 164162. [Google Scholar] [CrossRef]
  53. Hezam, A.; Wang, J.; Drmosh, Q.A.; Karthik, P.; Abdullah Bajiri, M.; Namratha, K.; Zare, M.; Lakshmeesha, T.R.; Shivanna, S.; Cheng, C.; et al. Rational construction of plasmonic Z-scheme Ag-ZnO-CeO2 heterostructures for highly enhanced solar photocatalytic H2 evolution. Appl. Surf. Sci. 2021, 541, 148457. [Google Scholar] [CrossRef]
  54. Munawar, K.; Mansoor, M.A.; Olmstead, M.M.; Zaharinie, T.; Mohd Zubir, M.N.; Haniffa, M.; Basirun, W.J.; Mazhar, M. Fabrication of Ag-ZnO composite thin films for plasmonic enhanced water splitting. Mater. Chem. Phys. 2020, 255, 123220. [Google Scholar] [CrossRef]
  55. Trang, T.N.Q.; Phan, T.B.; Nam, N.D.; Thu, V.T.H. In Situ Charge Transfer at the Ag@ZnO Photoelectrochemical Interface toward the High Photocatalytic Performance of H2 Evolution and RhB Degradation. ACS Appl. Mater. Interfaces 2020, 12, 12195–12206. [Google Scholar] [CrossRef] [PubMed]
  56. Motora, K.G.; Wu, C.-M.; Chala, T.F.; Chou, M.-H.; Kuo, C.-F.J.; Koinkar, P. Highly efficient photocatalytic activity of Ag3VO4/WO2.72 nanocomposites for the degradation of organic dyes from the ultraviolet to near-infrared regions. Appl. Surf. Sci. 2020, 512, 145618. [Google Scholar] [CrossRef]
  57. Zhang, D.; Liang, S.; Yao, S.; Li, H.; Liu, J.; Geng, Y.; Pu, X. Highly efficient visible/NIR photocatalytic activity and mechanism of Yb3+/Er3+ co-doped Bi4O5I2 up-conversion photocatalyst. Sep. Purif. Technol. 2020, 248, 117040. [Google Scholar] [CrossRef]
  58. Cui, L.; Ding, X.; Wang, Y.; Shi, H.; Huang, L.; Zuo, Y.; Kang, S. Facile preparation of Z-scheme WO3/g-C3N4 composite photocatalyst with enhanced photocatalytic performance under visible light. Appl. Surf. Sci. 2017, 391, 202–210. [Google Scholar] [CrossRef]
  59. Li, Y.; Wang, R.; Li, H.; Wei, X.; Feng, J.; Liu, K.; Dang, Y.; Zhou, A. Efficient and Stable Photoelectrochemical Seawater Splitting with TiO2@g-C3N4 Nanorod Arrays Decorated by Co-Pi. J. Phys. Chem. C 2015, 119, 20283–20292. [Google Scholar] [CrossRef]
  60. Peng, H.; Liu, D.; Zheng, X.; Fu, X. N-Doped Carbon-Coated ZnS with Sulfur-Vacancy Defect for Enhanced Photocatalytic Activity in the Visible Light Region. Nanomaterials 2019, 9, 1657. [Google Scholar] [CrossRef] [Green Version]
  61. Zhou, J.; Zhao, J.; Liu, R. Defect engineering of zeolite imidazole framework derived ZnS nanosheets towards enhanced visible light driven photocatalytic hydrogen production. Appl. Catal. B Environ. 2020, 278, 119265. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns of (a) In2S3 nanosheets, (b) ZnIn2S4 nanosheets, and (c) ZnS nanosheets.
Figure 1. The XRD patterns of (a) In2S3 nanosheets, (b) ZnIn2S4 nanosheets, and (c) ZnS nanosheets.
Nanomaterials 13 01957 g001
Figure 2. The FESEM images of (a) In2S3 nanosheets, (b) ZnIn2S4 nanosheets, and (c) ZnS nanosheets.
Figure 2. The FESEM images of (a) In2S3 nanosheets, (b) ZnIn2S4 nanosheets, and (c) ZnS nanosheets.
Nanomaterials 13 01957 g002
Figure 3. (a) FETEM, (b) SAED, (c) HRTEM, and (d) EDS mapping images of ZnIn2S4 nanosheets.
Figure 3. (a) FETEM, (b) SAED, (c) HRTEM, and (d) EDS mapping images of ZnIn2S4 nanosheets.
Nanomaterials 13 01957 g003
Figure 4. XPS spectra of the ZnIn2S4 nanosheets: (a) survey spectrum, (b) Zn 2p, (c) In 3d, and (d) S 2p.
Figure 4. XPS spectra of the ZnIn2S4 nanosheets: (a) survey spectrum, (b) Zn 2p, (c) In 3d, and (d) S 2p.
Nanomaterials 13 01957 g004
Figure 5. (a) The average HER of different photocatalysts. The average HER of ZnIn2S4 nanosheets under the different photocatalytic reaction conditions: (b) sacrificial reagents, (c) Na2S concentrations, and (d) pH values.
Figure 5. (a) The average HER of different photocatalysts. The average HER of ZnIn2S4 nanosheets under the different photocatalytic reaction conditions: (b) sacrificial reagents, (c) Na2S concentrations, and (d) pH values.
Nanomaterials 13 01957 g005
Figure 6. (a) UV–vis diffuse reflectance spectra and (b) Tauc plot of In2S3, ZnIn2S4, and ZnS nanosheets.
Figure 6. (a) UV–vis diffuse reflectance spectra and (b) Tauc plot of In2S3, ZnIn2S4, and ZnS nanosheets.
Nanomaterials 13 01957 g006
Figure 7. (a) PL and (b) EIS spectra of In2S3, ZnIn2S4, and ZnS nanosheets.
Figure 7. (a) PL and (b) EIS spectra of In2S3, ZnIn2S4, and ZnS nanosheets.
Nanomaterials 13 01957 g007
Figure 8. (a) VB XPS spectrum of the ZnIn2S4 nanosheets. (b) The schematic diagram depicts the charge separation process and photocatalytic reaction in the ZnIn2S4 nanosheets under visible-light irradiation.
Figure 8. (a) VB XPS spectrum of the ZnIn2S4 nanosheets. (b) The schematic diagram depicts the charge separation process and photocatalytic reaction in the ZnIn2S4 nanosheets under visible-light irradiation.
Nanomaterials 13 01957 g008
Figure 9. (a) The average HER of ZnIn2S4 nanosheets for the different water sources. (b) The average HER of ZnIn2S4 nanosheets for DI, DI-NaCl, and seawater. (c) The average HER of different photocatalysts for seawater.
Figure 9. (a) The average HER of ZnIn2S4 nanosheets for the different water sources. (b) The average HER of ZnIn2S4 nanosheets for DI, DI-NaCl, and seawater. (c) The average HER of different photocatalysts for seawater.
Nanomaterials 13 01957 g009
Figure 10. Reusability test of ZnIn2S4 nanosheets for eight cycles under visible-light irradiation.
Figure 10. Reusability test of ZnIn2S4 nanosheets for eight cycles under visible-light irradiation.
Nanomaterials 13 01957 g010
Figure 11. (a) XRD pattern and (b) FESEM image of ZnIn2S4 nanosheets after the eighth cycle.
Figure 11. (a) XRD pattern and (b) FESEM image of ZnIn2S4 nanosheets after the eighth cycle.
Nanomaterials 13 01957 g011
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

Chang, Y.-C.; Chiao, Y.-C.; Hsu, P.-C. Rapid Microwave-Assisted Synthesis of ZnIn2S4 Nanosheets for Highly Efficient Photocatalytic Hydrogen Production. Nanomaterials 2023, 13, 1957. https://doi.org/10.3390/nano13131957

AMA Style

Chang Y-C, Chiao Y-C, Hsu P-C. Rapid Microwave-Assisted Synthesis of ZnIn2S4 Nanosheets for Highly Efficient Photocatalytic Hydrogen Production. Nanomaterials. 2023; 13(13):1957. https://doi.org/10.3390/nano13131957

Chicago/Turabian Style

Chang, Yu-Cheng, Yung-Chang Chiao, and Po-Chun Hsu. 2023. "Rapid Microwave-Assisted Synthesis of ZnIn2S4 Nanosheets for Highly Efficient Photocatalytic Hydrogen Production" Nanomaterials 13, no. 13: 1957. https://doi.org/10.3390/nano13131957

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