Next Article in Journal
The Phononic Properties and Optimization of 2D Multi-Ligament Honeycombs
Previous Article in Journal
Sustainable Hybrid Lightweight Aggregate Concrete Using Recycled Expanded Polystyrene
Previous Article in Special Issue
High-Performance Nanoscale Metallic Multilayer Composites: Techniques, Mechanical Properties and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microwave-Assisted Synthesis of SnO2@ZnIn2S4 Composites for Highly Efficient Photocatalytic Hydrogen Evolution

1
Department of Materials Science and Engineering, Feng Chia University, Taichung 40724, Taiwan
2
Department of Materials Science and Engineering, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan
*
Author to whom correspondence should be addressed.
Materials 2024, 17(10), 2367; https://doi.org/10.3390/ma17102367
Submission received: 17 April 2024 / Revised: 10 May 2024 / Accepted: 13 May 2024 / Published: 15 May 2024
(This article belongs to the Special Issue Nanocomposite Based Materials for Various Applications)

Abstract

:
This research successfully synthesized SnO2@ZnIn2S4 composites for photocatalytic tap water splitting using a rapid two-step microwave-assisted synthesis method. This study investigated the impact of incorporating a fixed quantity of SnO2 nanoparticles and combining them with various materials to form composites, aiming to enhance photocatalytic hydrogen production. Additionally, different weights of SnO2 nanoparticles were added to the ZnIn2S4 reaction precursor to prepare SnO2@ZnIn2S4 composites for photocatalytic hydrogen production. Notably, the photocatalytic efficiency of SnO2@ZnIn2S4 composites is substantially higher than that of pure SnO2 nanoparticles and ZnIn2S4 nanosheets: 17.9-fold and 6.3-fold, respectively. The enhancement is credited to the successful use of visible light and the facilitation of electron transfer across the heterojunction, leading to the efficient dissociation of electron–hole pairs. Additionally, evaluations of recyclability demonstrated the remarkable longevity of SnO2@ZnIn2S4 composites, maintaining high levels of photocatalytic hydrogen production over eight cycles without significant efficiency loss, indicating their impressive durability. This investigation presents a promising strategy for crafting and producing environmentally sustainable SnO2@ZnIn2S4 composites with prospective implementations in photocatalytic hydrogen generation.

1. Introduction

Recently, excessive environmental pollution and severe energy shortages have adversely affected human life and health, prompting research into renewable energy sources [1]. Hydrogen, characterized by its zero-carbon emission properties, is cost-effective, sustainable, and environmentally friendly, making it poised to become a crucial energy source shortly [2]. As an efficient and green method, photocatalysis has been extensively studied for energy conversion, utilizing inexhaustible solar energy to address the issues [3]. Through photocatalytic reactions, solar energy is converted into chemical energy using photocatalysts to decompose water and produce hydrogen [4]. Various semiconductor metal oxide photocatalysts have recently emerged, showcasing elevated photocatalytic efficiency, non-hazardous properties, economic feasibility, and robust chemical resilience [5,6]. Among them, semiconductor metal oxide materials like tin dioxide (SnO2) have attracted considerable attention because of their outstanding physical and chemical characteristics [7,8,9]. While predominantly employed for the photocatalytic degradation of organic pollutants, their use in the photocatalytic water splitting process for hydrogen generation remains somewhat limited [10].
Tin dioxide (SnO2) is classified as an n-type semiconductor material, possessing a wide band gap of approximately 3.6 eV at room temperature [8,11,12]. It is currently regarded as one of the most promising metal oxides due to its non-toxicity, reliable stability, and excellent optoelectronic properties [13]. SnO2 has been widely utilized in gas sensors, electrode materials, lithium-ion batteries, solar cells, and photocatalytic degradation of organic and inorganic pollutants [14,15]. It possesses good electron mobility and a large specific surface area, facilitating its use in photocatalysis and electron transfer [16]. However, its relatively wide bandgap exhibits optimal photocatalytic activity only in ultraviolet, with reduced performance under visible light [17]. Therefore, incorporating other materials or doping other metals is necessary to enhance SnO2’s absorption of visible light and consequently improve its visible light photocatalytic activity [18,19,20].
Zinc indium sulfide (ZnIn2S4) is a ternary metal sulfide known for its layered structure, exhibiting three different crystal phases: cubic, hexagonal, and rhombohedral [21,22,23,24]. This photocatalyst is considered a conventional material for visible light absorption and is categorized as a direct band gap semiconductor with adjustable band gaps ranging from 2.06 to 2.85 eV [23,25,26]. ZnIn2S4 exhibits minimal toxicity, straightforward synthesis, and exceptional chemical durability, rendering it extensively investigated in areas such as photocatalysis, thermoelectricity, and electrochemical energy storage [27,28]. However, the material’s ability to catalyze through light exposure is limited by quick recombination rates of electron–hole pairs, low mobility of charge carriers, and structural flaws [29,30]. To address these limitations, elemental doping and heterojunction construction increase catalytic active sites and enhance photocatalytic reaction rates [31,32,33]. In previous studies, the combination of SnO2 and ZnIn2S4 has not been explored for the photocatalytic decomposition of tap water to produce hydrogen. This study addresses the respective limitations of SnO2 and ZnIn2S4 by combining them to form SnO2@ZnIn2S4 composites for enhancing their photocatalytic tap water splitting. The composites showed significantly higher efficiency than pure SnO2 nanoparticles or ZnIn2S4 nanosheets, credited to visible light utilization and improved electron transfer across the heterojunction, leading to efficient separation of electron–hole pairs. Additionally, the composites demonstrated remarkable recyclability, maintaining high levels of photocatalytic hydrogen production over eight cycles without significant efficiency loss, suggesting their impressive durability and potential for environmentally sustainable applications in photocatalytic hydrogen generation.

2. Materials and Methods

2.1. Chemicals

All chemical reagents were procured from commercial suppliers and employed without additional purification processes. Tin(IV) Chloride Pentahydrate (SnCl4, 98%, Alfa Aesar, Ward Hill, MA, USA), sodium hydroxide (NaOH, 96%, Showa, Osaka, Japan), citric acid anhydrous (99%, Showa, Osaka, Japan), indium (III) chloride anhydrous (InCl3, 98%, Alfa Aesar, Ward Hill, MA, USA), zinc chloride anhydrous (ZnCl2, 98%, Alfa Aesar, Ward Hill, MA, USA), thioacetamide (TAA, 98%, Alfa Aesar, Ward Hill, MA, USA), ethanol (99.8%, Sigma-Aldrich, Darmstadt, Germany), methanol (99.8%, Sigma-Aldrich, Darmstadt, Germany), formic acid (96%, Alfa Aesar, Ward Hill, MA, USA), sodium sulfide nonahydrate (Na2S, 98%, Alfa Aesar, Ward Hill, MA, USA), sodium sulfite anhydrous (Na2SO3, 98%, Alfa Aesar, Ward Hill, MA, USA) were utilized in these trials. Deionized water with a higher resistivity greater than 18.2 MΩ prepared all reaction solutions.

2.2. Fabrication of SnO2 Nanoparticles

An adapted protocol from the existing literature was employed to synthesize SnO2 nanoparticles utilizing a microwave-assisted synthesis approach [16]. A total of 0.45 g of Tin(IV) chloride, 0.1 g of citric acid, and 0.4 g of sodium hydroxide were dissolved in 50 mL of deionized water and subjected to ultrasonic vibration for approximately 20 min to ensure homogeneous mixing. Subsequently, the solution was transferred to a 100 mL Teflon reaction vessel and subjected to microwave-assisted hydrothermal heating at 180 °C for 60 min. Upon cooling the reaction to ambient temperature, the solution underwent multiple washes with deionized water and ethanol, followed by centrifugation for 3 min, and ultimately dried in a 70 °C oven to produce the white powder.

2.3. Fabrication of SnO2@ZnIn2S4 Composites

The SnO2@ZnIn2S4 composites were synthesized using a microwave-assisted synthesis method. Different weights of SnO2 nanoparticles were mixed uniformly with zinc chloride (1.25 mM), indium chloride (2.5 mM), and TAA (5 mM) in a ratio of 1:3 with ethanol and de-ionized water, treated with ultrasonication, transferred into 100 mL Teflon reaction bottle, and underwent a microwave-assisted hydrothermal method to heat at 180 °C for 60 min. Once the reaction concluded and cooled to ambient temperature, the solution underwent rinsing with deionized water and ethanol, followed by centrifugation for 3 min, and subsequently dried in a 70 °C oven to yield the yellow powder of the composite material.

2.4. Characterization

The overall crystal structure of the SnO2 nanoparticles and SnO2@ZnIn2S4 composites was analyzed using X-ray diffraction (XRD) at a facility in Billerica, MA, USA, utilizing a Bruker D2 phaser system with Cu Kα radiation (λ = 1.5418 Å). The surface characteristics of the SnO2 nanoparticles and SnO2@ZnIn2S4 composites were examined through field-emission scanning electron microscopy (FESEM) in Tokyo, Japan, using a Hitachi S-4800 microscope operating at a 15 kV accelerating voltage. Field-emission transmission electron microscopy (FETEM) at 200 kV was employed to study the microstructures and composition of the SnO2@ZnIn2S4 composites with a JEOL 2100F microscope in Tokyo, Japan. The surface chemical composition of the SnO2 nanoparticles and SnO2@ZnIn2S4 composites was investigated using X-ray photoelectron spectroscopy (XPS) at a facility in Chigasaki, Japan, equipped with an Al K source. The newly synthesized photocatalysts’ diffused reflectance spectra were analyzed with a UV–visible spectrophotometer (PerkinElmer Lambda 650 S, Waltham, MA, USA). Furthermore, a spectrofluorophotometer obtained room-temperature photoluminescence spectra (Shimadzu, RF-5301PC, Kyoto, Japan). The qualitative analysis of tap water compositions can be conducted using an Inductively Coupled Plasma Optima Optical Emission Spectrometer (ICP-OES, PerkinElmer, OPTIMA 2000DV, Waltham, MA, USA). The components present in tap water include magnesium (Mg), nickel (Ni), barium (Ba), calcium (Ca), cesium (Cs), copper (Cu), iron (Fe), potassium (K), lithium (Li), manganese (Mn), sodium (Na), strontium (Sr), iridium (Ir), boron (B), zinc (Zn), silicon (Si), and tungsten (W).

2.5. Photocatalytic Hydrogen Production Experiment

A multi-port reaction system produced hydrogen through photocatalysis with specially made photocatalysts. This setup involved stirring with a magnetic stirrer and a 5 W blue LED light source with a peak wavelength of 420 nm powered via PCX50 B Discover with Perfect Light Technology from Beijing, China. In accordance with the standard procedure, 25 mg of the photocatalysts were introduced into a solution comprising 0.1 M of different sacrificial agents (including Na2S, Na2SO3, CH2O2, and CH3OH) and 50 mL of deionized water. When using 0.1 M Na2S as the sacrificial reagent in deionized or tap water, the pH values were 13.0 and 12.9, respectively. Before the commencement of the experiment, a degassing process was conducted for a duration of 10 min to eliminate air from the system. Subsequently, the generated hydrogen was measured utilizing gas chromatography (GC, Shimadzu GC-2014, Kyoto, Japan) equipped with a thermal conductivity detector (TCD).

3. Results and Discussion

Characterization of SnO2@ZnIn2S4 Composites

Figure 1 illustrates the reaction process of SnO2@ZnIn2S4 composites using a two-step microwave-assisted hydrothermal method. Firstly, SnO2 nanoparticles were prepared by microwave-assisted hydrothermal treatment of SnO2 precursor (SnCl4, NaOH, and citric acid) at 180 °C for 60 min. Subsequently, varying amounts of SnO2 nanoparticles were dispersed in the precursor solution of ZnIn2S4 (ZnCl2, InCl3, and TAA), followed by another round of microwave-assisted hydrothermal treatment at 180 °C for 60 min to obtain the SnO2@ZnIn2S4 composites.
The morphology of the SnO2 nanoparticles and SnO2@ZnIn2S4 composites synthesized using the two-step microwave-assisted hydrothermal method was examined using an FESEM, as shown in Figure 2. The SnO2 nanostructures are depicted, comprising numerous nanoparticles with a diameter of approximately 20 nm, uniformly dispersed and aggregated, as shown in Figure 2a. Figure 2b displays the FESEM image of the SnO2@ZnIn2S4 composites following the incorporation of ZnIn2S4. Notably, a multitude of nanoparticles is observed covering the nanosheet structure. Furthermore, flower-like structures with a similar laminar stacking arrangement can be observed in the low-magnification FESEM image (Figure 2c). Further analysis of the distribution of individual elements through FESEM-EDS mapping is illustrated in Figure 2d. It can be observed from the figures that the SnO2@ZnIn2S4 composites contain Sn, O, Zn, In, and S, and they are evenly distributed. Therefore, it can be demonstrated that SnO2 and ZnIn2S4 successfully combine to form SnO2@ZnIn2S4 composites.
To further analyze the overall changes in crystal structure, X-ray diffraction spectroscopy (XRD) was utilized. The XRD analysis shows the crystalline structure of SnO2 nanoparticles and SnO2@ZnIn2S4 composites, as shown in Figure 3. Figure 3a presents the XRD pattern of SnO2 nanoparticles, with diffraction peak angles at 26.6°, 33.9°, 38.0°, 39.0°, 51.8°, 54.7°, 58.1°, 61.8°, 64.9°, 65.8°, 71.4°, and 78.6°, corresponding to the crystallographic planes (110), (101), (200), (111), (211), (220), (002), (310), (112), (301), (202), and (321), respectively, as compared to PDF# 01-075-2893. This crystal phase structure indicates a pure tetragonal structure of SnO2. Figure 3b shows the XRD diffraction pattern of the SnO2@ZnIn2S4 composites. Apart from the SnO2 diffraction peaks, the diffraction peaks are observed at 28.0°, 30.3°, 45.6°, 47.6°, and 56.2°, corresponding to the crystallographic planes (011), (012), (105), (111), and (114), respectively, as compared to PDF# 04-009-4783. This crystal phase structure indicates a hexagonal structure of ZnIn2S4. Additionally, no other impurity phases were detected in the XRD pattern. This result serves to confirm the successful generation of SnO2@ZnIn2S4 composites.
XPS was utilized to examine the chemical compositions of SnO2 nanoparticles and SnO2@ZnIn2S4 composites [34]. The XPS survey spectra of the prepared samples, as shown in Figure 4a, distinctly confirm the binding energies associated with C 1s, Sn 3P, Sn 3d, Sn 4d, and O 1s in both SnO2 nanoparticles and SnO2@ZnIn2S4 composites. Therefore, Zn, In, and S binding energy characteristics are used to identify SnO2 nanoparticles and SnO2@ZnIn2S4 composites. The binding energy for Sn is observed in the SnO2 nanoparticles and SnO2@ZnIn2S4 composites at 486.7 eV and 495.1 eV, respectively, as depicted in Figure 4b. These peaks are identified as the Sn 3d5/2 and Sn 3d3/2 peaks of SnO2 [35]. The O 1s binding energy of the SnO2 nanoparticles and SnO2@ZnIn2S4 composites is analyzed in Figure 4c. It is separated into two peaks with binding energies at 530.6 eV and 531.8 eV, representing lattice oxygen (O lattice, OL) and oxygen-deficient region (O defect, OD) in SnO2, respectively [36,37]. This outcome shows many oxygen defects in the SnO2 nanoparticles and SnO2@ZnIn2S4 composites. The Zn 2p spectrum displays two distinct peaks at 1021.3 eV and 1044.3 eV, corresponding to the Zn 2p3/2 and Zn 2p1/2 states, respectively, as illustrated in Figure 4d [32]. The In 3d spectrum (Figure 4e) reveals peaks at 445.0 eV and 452.5 eV, which are associated with the In 3d5/2 and In 3d3/2 states of ZnIn2S4. These findings align with existing literature [38]. The S 2p spectrum (Figure 4f) exhibits two distinct peaks located at 161.6 eV and 162.8 eV, which have been identified as corresponding to the S 2p3/2 and S 2p1/2 states, respectively [39]. These results collectively confirm the presence of Zn2+, In3+, and S2− chemical states, with no indications of impurity peaks in the spectra.
TEM analysis was performed to examine the morphology and crystal structure of the SnO2@ZnIn2S4 composites more efficiently. Figure 5a displays the TEM image, revealing the complete integration of SnO2 nanoparticles with ZnIn2S4 nanosheets. Figure 5b exhibits the SAED pattern of SnO2@ZnIn2S4 composites, showing a polycrystalline ring. The crystal structures consist of tetragonal SnO2 (PDF# 01-075-2893) and hexagonal ZnIn2S4 (PDF# 04-009-4783). Moreover, upon examination using high-resolution transmission electron microscopy (HRTEM) imaging (Figure 5c), lattice spacings of 0.264 nm indicative of the (011) plane of SnO2 (PDF# 01-075-2893) and 0.322 nm corresponding to the (101) plane of ZnIn2S4 (PDF# 04-009-4783) were identified. This observation serves to validate the crystal structures of SnO2 and ZnIn2S4. TEM-EDS mapping image (Figure 5d) was performed for qualitative and semi-quantitative compositional analysis of SnO2@ZnIn2S4 composites, revealing an even distribution of the individual elements Sn, O, Zn, In, and S within the composites. This outcome confirms the effective production of composite materials consisting of SnO2 and ZnIn2S4.
The photocatalytic efficiency of the fabricated SnO2 nanoparticles was assessed by quantifying the hydrogen evolution rate (HER) in 50 mL deionized water under blue LED light irradiation. In aqueous settings, sacrificial agents are commonly employed to boost the effectiveness of oxidation reactions in the photocatalytic water splitting process for hydrogen production [40]. Figure 6a illustrates the impact of four sacrificial agents (Na2S, Na2SO3, CH2O2, and CH3OH) on the photocatalytic performance of SnO2 nanoparticles without pH adjustment. All sacrificial agents were utilized at a uniform concentration of 0.1 M. The average hydrogen evolution rate (HER) with SnO2 nanoparticles follows the order: Na2S (22.8 μmol·h−1·g−1·L−1), Na2SO3 (13.1 μmol·h−1·g−1·L−1), CH2O2 (11.3 μmol·h−1·g−1·L−1), and CH3OH (9.8 μmol·h−1·g−1·L−1). A possible rationale behind this phenomenon could be the adsorption of sulfide ions (S2−, originating from dissociated Na2S) onto the photocatalyst. These ions could interact with photogenerated positive charges, thereby impeding the recombination of electron–hole pairs [41]. Therefore, Na2S was used as a sacrificial reagent in the subsequent reactions.
To further boost the effectiveness of photocatalytic water splitting for hydrogen production using SnO2 nanoparticles, this investigation also integrated 0.02 g of SnO2 nanoparticles into different precursor materials before the reaction. Subsequently, the efficacy of photocatalytic water splitting for hydrogen generation using SnO2 nanoparticles in conjunction with various materials was examined under blue LED light exposure, as depicted in Figure 6b. The optimal efficiency for photocatalytic water splitting for hydrogen production (408.9 μmol·h−1·g−1·L−1) is achieved when SnO2 nanoparticles react with ZnIn2S4 precursor materials to form SnO2@ZnIn2S4 composites. This result indicates a 17.9-fold increase in efficiency compared to SnO2 nanoparticles alone, demonstrating that the composite with ZnIn2S4 effectively enhances the photocatalytic efficiency of the material.
Figure 6c illustrates the performance variation in photocatalytic water splitting to produce hydrogen for pure SnO2 nanoparticles, ZnIn2S4 nanosheets, and SnO2@ZnIn2S4 composites formed by reacting different weights of SnO2 nanoparticles with ZnIn2S4 reaction precursors under blue LED light irradiation. Adding 0.02 g of SnO2 nanoparticles exhibits the optimal photocatalytic water splitting efficiency for hydrogen production (408.9 μmol·h−1·g−1·L−1). This result represents a significant enhancement compared to pure SnO2 nanoparticles (22.8 μmol·h−1·g−1·L−1) and ZnIn2S4 nanosheets (65.1 μmol·h−1·g−1·L−1) by approximately 17.9-fold and 6.3-fold, respectively. This finding verifies that incorporating an appropriate weight of SnO2 nanoparticles assists in preparing SnO2@ZnIn2S4 composites with optimal hydrogen production efficiency.
Figure 7a displays the UV–visible absorption spectra of SnO2 nanoparticles and SnO2@ZnIn2S4 composites. In contrast to SnO2 nanoparticles, the absorption edge of SnO2@ZnIn2S4 composites demonstrates a significant redshift. Moreover, the absorption intensity of SnO2@ZnIn2S4 composites is notably enhanced in the wavelength range of λ > 450 nm, which augments the generation of photogenerated carriers and enhances photocatalytic performance under visible light. Photoluminescence (PL) spectroscopy is a technique that utilizes the fluorescence produced when unbound charge carriers recombine, providing valuable information on the movement, transfer, and separation of electron–hole pairs generated by light in semiconductor materials [6,42]. Figure 7b reveals the measured PL emission spectra of SnO2 nanoparticles and SnO2@ZnIn2S4 composites. SnO2@ZnIn2S4 composites exhibit significantly lower emission intensity than SnO2 nanoparticles, indicating suppression of photogenerated hole-electron recombination in the composites. Photocurrent measurements are conducted to evaluate the efficiency of photogenerated electron–hole pair separation for photocatalytic performance. As depicted in Figure 7c, the photocurrent intensity of SnO2@ZnIn2S4 composites exceeds that of SnO2 nanoparticles, demonstrating the synergistic effect between SnO2 and ZnIn2S4 in suppressing photogenerated electron–hole recombination.
Figure 8 illustrates the potential mechanism underlying the photocatalytic water-splitting process on the SnO2@ZnIn2S4 composites when subjected to blue LED light irradiation. An ion exchange resin is first applied to the indium tin oxide (ITO) glass, creating a coating of SnO2 and ZnIn2S4. Subsequently, the flat band potential is determined using cyclic voltammetry [43]. The determined valence band (VB) and conduction band (CB) positions of SnO2 and ZnIn2S4 align with established findings. In particular, the conduction band positions for SnO2 and ZnIn2S4 are recorded at −0.14 eV and −0.58 eV, respectively, while their valence band positions are noted at 3.35 eV and 1.78 eV, respectively. [15,44,45,46]. In addition, previous XPS O1s (Figure 4c) can verify that SnO2 nanoparticles and SnO2@ZnIn2S4 composites exhibited oxygen defects. Notably, the presence of oxygen defects leads to the emergence of new electronic state bands near the lower boundary of the CB in SnO2. This phenomenon reduces the band gap, enhancing visible light absorption [47]. Upon exposure to blue LED light irradiation, photogenerated electrons within the VB of SnO2 and ZnIn2S4 are excited to the CB. Subsequently, the photogenerated electrons in the CB of ZnIn2S4 can migrate to the CB of SnO2. SnO2 acts as an electron sink, effectively capturing and reducing hydrogen ions to generate hydrogen. Concurrently, the photogenerated holes within the VB of SnO2 can transfer to ZnIn2S4, where they participate in water oxidation to produce oxygen or hydrogen ions. This photocatalytic process facilitates the efficient separation of photogenerated charge carriers, thereby promoting the photocatalytic production of hydrogen. The collaborative influence notably mitigates carrier recombination, consequently amplifying the photocatalytic efficacy of SnO2@ZnIn2S4 and facilitating effective hydrogen generation. This synergy of photocatalysis helps to separate charge carriers, promoting the efficient production of hydrogen efficiently. Additionally, the combination of SnO2@ZnIn2S4 composites dramatically enhances the ability to absorb light, thereby increasing the effectiveness of generating hydrogen through photocatalysis.
Deionized water undergoes purification by eliminating diverse ions and contaminants from tap water [46,48]. Hence, if photocatalysts could directly generate hydrogen via the decomposition of tap water, it could significantly decrease the expenses and duration associated with purifying tap water into deionized water, thereby facilitating the practical utilization of photocatalysts [42]. Additionally, this approach could lead to more sustainable and cost-effective water treatment methods, contributing to environmental conservation efforts [49]. To demonstrate the practicality of utilizing SnO2@ZnIn2S4 composites for photocatalytic tap water splitting, 50 mg of the composites was dispersed in 50 mL of tap water solution containing 0.1 M Na2S without any pH modification. This experimental setup was subjected to blue or white LED light irradiation, as depicted in Figure 9a. The average HER of SnO2@ZnIn2S4 composites is 408.9 μmolh−1g−1L−1 in deionized water and 670.8 μmolh−1g−1L−1 in tap water under blue LED light irradiation. The average HER of SnO2@ZnIn2S4 composites is 160.1 μmolh−1g−1L−1 in deionized water and 395.8 μmolh−1g−1L−1 in tap water under white LED light irradiation. Notably, the average HER of SnO2@ZnIn2S4 composites in tap water is approximately 1.64 (blue LED light) and 2.47 (white LED light) times higher than in deionized water, respectively. It is speculated that the possible reason is that more ions or chlorine ions in tap water can facilitate the transfer of photogenerated electron and hole pairs, thereby enhancing photocatalytic hydrogen production [50]. These findings highlight the exceptional photocatalytic efficacy of SnO2@ZnIn2S4 composites in hydrogen generation, even when dealing with intricate water compositions, without needing pH adjustments. In addition, this result also confirms that blue LED light is more effective than white LED light when employing SnO2@ZnIn2S4 composites for the photocatalytic decomposition of tap water to produce hydrogen. The lower photocatalytic efficiency of white LED light compared to blue LED light may be due to the dispersion of light source intensity [41].
Efficiently recycling catalysts is paramount in photocatalysis [41]. Designing photocatalysts that maintain consistent photoactivity across multiple cycles is crucial for minimizing waste and ensuring sustainable processes [51]. In the context of photocatalysts, the gradual decline in efficiency over time can lead to increased waste generation throughout their life cycle. Hence, creating photocatalysts that can be easily separated and recycled is crucial to mitigate the depletion of valuable resources within the waste stream [32]. This study conducted eight consecutive cycles of photocatalytic tap water splitting to assess the durability of SnO2@ZnIn2S4 composites. The average HER of the SnO2@ZnIn2S4 composites exhibits a sustained high level throughout eight cycles, as depicted in Figure 9b. Additionally, the XRD pattern of the sample after the eight cycles, shown in Figure 9c, does not exhibit any new peaks, indicating the preservation of the composite’s structural integrity. These results prove that the SnO2@ZnIn2S4 composites exhibit better durability and recyclability.

4. Conclusions

In this study, we have successfully produced composite materials consisting of SnO2@ZnIn2S4 for photocatalytic splitting of tap water, employing a two-step synthesis method assisted by microwave technology. Our investigation delved into the impact of incorporating fixed quantities of SnO2 nanoparticles into various materials to form composites, thereby enhancing hydrogen production through photocatalysis. Furthermore, we investigated the effect of different concentrations of SnO2 nanoparticles in the ZnIn2S4 reaction precursor to improve the synthesis of SnO2@ZnIn2S4 composites for photocatalytic hydrogen generation. Notably, the efficiency of photocatalytic hydrogen production in SnO2@ZnIn2S4 composites surpasses that of pure SnO2 nanoparticles or ZnIn2S4 nanosheets. The improved efficiency can be credited to successfully harnessing visible light and promoting photogenerated electrons across the heterojunction, leading to the effective dissociation of electron–hole pairs. Furthermore, reusability tests have validated the exceptional performance of the SnO2@ZnIn2S4 composites. Despite undergoing eight cycles, the composites continue to exhibit elevated levels of photocatalytic hydrogen production without experiencing a notable decline in efficiency. This study introduces a forward-looking methodology for synthesizing and fabricating environmentally sustainable SnO2@ZnIn2S4 composites, which hold promise for applications in the field of photocatalytic hydrogen generation.

Author Contributions

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

Funding

The authors thank the National Science and Technology Council, Taiwan (NSTC 112-2221-E-035-017-MY3) for supporting our work with financial resources.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article.

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 conflicts of interest.

References

  1. Kumar, R.; Verma, A.; Shome, A.; Sinha, R.; Sinha, S.; Jha, P.K.; Kumar, R.; Kumar, P.; Shubham; Das, S.; et al. Impacts of Plastic Pollution on Ecosystem Services, Sustainable Development Goals, and Need to Focus on Circular Economy and Policy Interventions. Sustainability 2021, 13, 9963. [Google Scholar] [CrossRef]
  2. Sharma, G.D.; Verma, M.; Taheri, B.; Chopra, R.; Parihar, J.S. Socio-economic aspects of hydrogen energy: An integrative review. Technol. Forecast. Soc. Change 2023, 192, 122574. [Google Scholar] [CrossRef]
  3. Armaroli, N.; Balzani, V. The Future of Energy Supply: Challenges and Opportunities. Angew. Chem. Int. Ed. 2007, 46, 52–66. [Google Scholar] [CrossRef] [PubMed]
  4. Yan, Z.; Yin, K.; Xu, M.; Fang, N.; Yu, W.; Chu, Y.; Shu, S. Photocatalysis for synergistic water remediation and H2 production: A review. Chem. Eng. J. 2023, 472, 145066. [Google Scholar] [CrossRef]
  5. Shi, Y.; Yang, A.-F.; Cao, C.-S.; Zhao, B. Applications of MOFs: Recent advances in photocatalytic hydrogen production from water. Coord. Chem. Rev. 2019, 390, 50–75. [Google Scholar] [CrossRef]
  6. Lin, Y.-R.; Chang, Y.-C.; Ko, F.-H. One-pot microwave-assisted synthesis of In2S3/In2O3 nanosheets as highly active visible light photocatalysts for seawater splitting. Int. J. Hydrogen Energy 2024, 52, 953–963. [Google Scholar] [CrossRef]
  7. Sun, C.; Yang, J.; Xu, M.; Cui, Y.; Ren, W.; Zhang, J.; Zhao, H.; Liang, B. Recent intensification strategies of SnO2-based photocatalysts: A review. Chem. Eng. J. 2022, 427, 131564. [Google Scholar] [CrossRef]
  8. Do Nascimento, J.L.; Chantelle, L.; dos Santos, I.M.; Menezes de Oliveira, A.L.; Alves, M.C. The Influence of Synthesis Methods and Experimental Conditions on the Photocatalytic Properties of SnO2: A Review. Catalysts 2022, 12, 428. [Google Scholar] [CrossRef]
  9. Rajput, R.B.; Jamble, S.N.; Kale, R.B. A review on TiO2/SnO2 heterostructures as a photocatalyst for the degradation of dyes and organic pollutants. J. Environ. Manag. 2022, 307, 114533. [Google Scholar] [CrossRef]
  10. Roy, H.; Rahman, T.U.; Khan, M.A.J.R.; Al-Mamun, M.R.; Islam, S.Z.; Khaleque, M.A.; Hossain, M.I.; Khan, M.Z.H.; Islam, M.S.; Marwani, H.M.; et al. Toxic dye removal, remediation, and mechanism with doped SnO2-based nanocomposite photocatalysts: A critical review. J. Water Process Eng. 2023, 54, 104069. [Google Scholar] [CrossRef]
  11. Kaewsuwan, D.; Wongpinij, T.; Euaruksakul, C.; Chanlek, N.; Triamnak, N.; Lertvanithphol, T.; Horprathum, M.; Kaewkhao, J.; Manyum, P.; Yimnirun, R.; et al. Photoluminescence of tin dioxide (SnO2) nanostructure grown on Si(001) by thermal evaporation technique. Radiat. Phys. Chem. 2023, 206, 110805. [Google Scholar] [CrossRef]
  12. Xu, Y.; Zheng, L.; Yang, C.; Zheng, W.; Liu, X.; Zhang, J. Oxygen Vacancies Enabled Porous SnO2 Thin Films for Highly Sensitive Detection of Triethylamine at Room Temperature. ACS Appl. Mater. Interfaces 2020, 12, 20704–20713. [Google Scholar] [CrossRef] [PubMed]
  13. Dalapati, G.K.; Sharma, H.; Guchhait, A.; Chakrabarty, N.; Bamola, P.; Liu, Q.; Saianand, G.; Sai Krishna, A.M.; Mukhopadhyay, S.; Dey, A.; et al. Tin oxide for optoelectronic, photovoltaic and energy storage devices: A review. J. Mater. Chem. A 2021, 9, 16621–16684. [Google Scholar] [CrossRef]
  14. Shabna, S.; Dhas, S.S.J.; Biju, C.S. Potential progress in SnO2 nanostructures for enhancing photocatalytic degradation of organic pollutants. Catal. Commun. 2023, 177, 106642. [Google Scholar] [CrossRef]
  15. Szkoda, M.; Zarach, Z.; Nadolska, M.; Trykowski, G.; Trzciński, K. SnO2 nanoparticles embedded onto MoS2 nanoflakes—An efficient catalyst for photodegradation of methylene blue and photoreduction of hexavalent chromium. Electrochim. Acta 2022, 414, 140173. [Google Scholar] [CrossRef]
  16. Yu-Cheng, C.; Yu-Wen, L. MoS2@SnO2 core-shell sub-microspheres for high efficient visible-light photodegradation and photocatalytic hydrogen production. Mater. Res. Bull. 2020, 129, 110912. [Google Scholar] [CrossRef]
  17. Ren, W.; Yang, J.; Zhang, J.; Li, W.; Sun, C.; Zhao, H.; Wen, Y.; Sha, O.; Liang, B. Recent progress in SnO2/g-C3N4 heterojunction photocatalysts: Synthesis, modification, and application. J. Alloys Compd. 2022, 906, 164372. [Google Scholar] [CrossRef]
  18. Ahmed, A.; Naseem Siddique, M.; Alam, U.; Ali, T.; Tripathi, P. Improved photocatalytic activity of Sr doped SnO2 nanoparticles: A role of oxygen vacancy. Appl. Surf. Sci. 2019, 463, 976–985. [Google Scholar] [CrossRef]
  19. Vattikuti, S.V.P.; Reddy, P.A.K.; Shim, J.; Byon, C. Visible-Light-Driven Photocatalytic Activity of SnO2–ZnO Quantum Dots Anchored on g-C3N4 Nanosheets for Photocatalytic Pollutant Degradation and H2 Production. ACS Omega 2018, 3, 7587–7602. [Google Scholar] [CrossRef]
  20. Wang, Y.; Cheng, J.; Yu, S.; Alcocer, E.J.; Shahid, M.; Wang, Z.; Pan, W. Synergistic effect of N-decorated and Mn2+ doped ZnO nanofibers with enhanced photocatalytic activity. Sci. Rep. 2016, 6, 32711. [Google Scholar] [CrossRef]
  21. Yan, Y.; Chen, Z.; Cheng, X.; Shi, W. Research Progress of ZnIn2S4-Based Catalysts for Photocatalytic Overall Water Splitting. Catalysts 2023, 13, 967. [Google Scholar] [CrossRef]
  22. Yue, Y.; Zou, J. Boosting interfacial charge separation for ZnIn2S4 homojunction by lattice matching effect to enhance photocatalytic performance. J. Alloys Compd. 2023, 966, 171659. [Google Scholar] [CrossRef]
  23. Chang, Y.-C.; Chiao, Y.-C.; Chang, C.-J. Synthesis of g-C3N4@ZnIn2S4 Heterostructures with Extremely High Photocatalytic Hydrogen Production and Reusability. Catalysts 2023, 13, 1187. [Google Scholar] [CrossRef]
  24. Tu, H.; Wang, H.; Zhang, J.; Ou, Y.; Zhang, Z.; Chen, G.; Wei, C.; Xiang, X.; Xie, Z. A novel hierarchical 0D/3D NH2-MIL-101(Fe)/ZnIn2S4 S-scheme heterojunction photocatalyst for efficient Cr(VI) reduction and photo-Fenton-like removal of 2-nitrophenol. J. Environ. Chem. Eng. 2024, 12, 111695. [Google Scholar] [CrossRef]
  25. Zhang, W.; Zhao, S.; Xing, Y.; Qin, H.; Zheng, Q.; Zhang, P.; Zhang, S.; Xu, X. Sandwich-like P-doped h-BN/ZnIn2S4 nanocomposite with direct Z-scheme heterojunction for efficient photocatalytic H2 and H2O2 evolution. Chem. Eng. J. 2022, 442, 136151. [Google Scholar] [CrossRef]
  26. Pan, Y.; Yuan, X.; Jiang, L.; Yu, H.; Zhang, J.; Wang, H.; Guan, R.; Zeng, G. Recent advances in synthesis, modification and photocatalytic applications of micro/nano-structured zinc indium sulfide. Chem. Eng. J. 2018, 354, 407–431. [Google Scholar] [CrossRef]
  27. Zhang, T.; Wang, T.; Meng, F.; Yang, M.; Kawi, S. Recent advances in ZnIn2S4-based materials towards photocatalytic purification, solar fuel production and organic transformations. J. Mater. Chem. C 2022, 10, 5400–5424. [Google Scholar] [CrossRef]
  28. 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]
  29. Zhu, R.; Tian, F.; Che, S.; Cao, G.; Ouyang, F. The photocatalytic performance of modified ZnIn2S4 with graphene and La for hydrogen generation under visible light. Renew. Energy 2017, 113, 1503–1514. [Google Scholar] [CrossRef]
  30. Zheng, X.; Song, Y.; Liu, Y.; Yang, Y.; Wu, D.; Yang, Y.; Feng, S.; Li, J.; Liu, W.; Shen, Y.; et al. ZnIn2S4-based photocatalysts for photocatalytic hydrogen evolution via water splitting. Coord. Chem. Rev. 2023, 475, 214898. [Google Scholar] [CrossRef]
  31. 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]
  32. 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. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, S.; Wang, K.; Yu, H.; Huang, Y.; Guo, P.; Ye, C.; Wen, H.; Zhang, G.; Luo, D.; Jiang, F.; et al. Carbon fibers derived from spent cigarette filters for supporting ZnIn2S4/g-C3N4 heterojunction toward enhanced photocatalytic hydrogen evolution. Mater. Sci. Eng. B 2023, 288, 116214. [Google Scholar] [CrossRef]
  34. Alvarez-Roca, R.; Gouveia, A.F.; de Foggi, C.C.; Lemos, P.S.; Gracia, L.; da Silva, L.F.; Vergani, C.E.; San-Miguel, M.; Longo, E.; Andrés, J. Selective Synthesis of α-, β-, and γ-Ag2WO4 Polymorphs: Promising Platforms for Photocatalytic and Antibacterial Materials. Inorg. Chem. 2021, 60, 1062–1079. [Google Scholar] [CrossRef]
  35. Ilka, M.; Bera, S.; Kwon, S.-H. Influence of Surface Defects and Size on Photochemical Properties of SnO2 Nanoparticles. Materials 2018, 11, 904. [Google Scholar] [CrossRef] [PubMed]
  36. Yang, Y.; Wang, Y.; Yin, S. Oxygen vacancies confined in SnO2 nanoparticles for desirable electronic structure and enhanced visible light photocatalytic activity. Appl. Surf. Sci. 2017, 420, 399–406. [Google Scholar] [CrossRef]
  37. Sangwichien, C.; Aranovich, G.L.; Donohue, M.D. Density functional theory predictions of adsorption isotherms with hysteresis loops. Colloids Surf. A Physicochem. Eng. Asp. 2002, 206, 313–320. [Google Scholar] [CrossRef]
  38. Liu, B.; Liu, X.; Liu, J.; Feng, C.; Li, Z.; Li, C.; Gong, Y.; Pan, L.; Xu, S.; Sun, C.Q. Efficient charge separation between UiO-66 and ZnIn2S4 flowerlike 3D microspheres for photoelectronchemical properties. Appl. Catal. B 2018, 226, 234–241. [Google Scholar] [CrossRef]
  39. 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]
  40. Burek, B.O.; Timm, J.; Bahnemann, D.W.; Bloh, J.Z. Kinetic effects and oxidation pathways of sacrificial electron donors on the example of the photocatalytic reduction of molecular oxygen to hydrogen peroxide over illuminated titanium dioxide. Catal. Today 2019, 335, 354–364. [Google Scholar] [CrossRef]
  41. 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]
  42. Chang, Y.-C.; Lin, Y.-R. Construction of Ag/Ag2S/CdS Heterostructures through a Facile Two-Step Wet Chemical Process for Efficient Photocatalytic Hydrogen Production. Nanomaterials 2023, 13, 1815. [Google Scholar] [CrossRef] [PubMed]
  43. 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]
  44. Wang, X.; Wang, X.; Yang, H.; Meng, A.; Li, Z.; Yang, L.; Wang, L.; Li, S.; Li, G.; Huang, J. Interfacial engineering improved internal electric field contributing to direct Z-scheme-dominated mechanism over CdSe/SL-ZnIn2S4/MoSe2 heterojunction for efficient photocatalytic hydrogen evolution. Chem. Eng. J. 2022, 431, 134000. [Google Scholar] [CrossRef]
  45. Tu, H.; Wang, H.; Xu, H.; Zhang, Z.; Chen, G.; Wei, C.; Xiang, X.; Xie, Z. Construction of novel NH2-MIL-125(Ti)@ZnIn2S4 Z-scheme heterojunction and its photocatalytic performance regulation. J. Alloys Compd. 2023, 958, 170462. [Google Scholar] [CrossRef]
  46. 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 2023, 13, 33. [Google Scholar] [CrossRef]
  47. Shao, M.; Liu, J.; Ding, W.; Wang, J.; Dong, F.; Zhang, J. Oxygen vacancy engineering of self-doped SnO2−x nanocrystals for ultrasensitive NO2 detection. J. Mater. Chem. C 2020, 8, 487–494. [Google Scholar] [CrossRef]
  48. Marotta, E.; Ceriani, E.; Schiorlin, M.; Ceretta, C.; Paradisi, C. Comparison of the rates of phenol advanced oxidation in deionized and tap water within a dielectric barrier discharge reactor. Water Res. 2012, 46, 6239–6246. [Google Scholar] [CrossRef] [PubMed]
  49. Wang, A.; Liang, H.; Chen, F.; Tian, X.; Yin, S.; Jing, S.; Tsiakaras, P. Facile synthesis of C3N4/NiIn2S4 heterostructure with novel solar steam evaporation efficiency and photocatalytic H2O2 production performance. Appl. Catal. B 2022, 310, 121336. [Google Scholar] [CrossRef]
  50. Hippargi, G.; Mangrulkar, P.; Chilkalwar, A.; Labhsetwar, N.; Rayalu, S. Chloride ion: A promising hole scavenger for photocatalytic hydrogen generation. Int. J. Hydrogen Energy 2018, 43, 6815–6823. [Google Scholar] [CrossRef]
  51. Chang, Y.-C.; Hsu, C.-C.; Wu, S.-H.; Chuang, K.-W.; Chen, Y.-F. Fabrication of Cu-doped ZnO nanoneedles on different substrate via wet chemical approach: Structural characterization and photocatalytic performance. Appl. Surf. Sci. 2018, 447, 213–221. [Google Scholar] [CrossRef]
Figure 1. Illustrates a schematic representation of the reaction to forming the SnO2@ZnIn2S4 composites.
Figure 1. Illustrates a schematic representation of the reaction to forming the SnO2@ZnIn2S4 composites.
Materials 17 02367 g001
Figure 2. The FESEM images of (a) SnO2 nanoparticles and (b,c) SnO2@ZnIn2S4 composites. (d) The FESEM EDS mapping images of SnO2@ZnIn2S4 composites.
Figure 2. The FESEM images of (a) SnO2 nanoparticles and (b,c) SnO2@ZnIn2S4 composites. (d) The FESEM EDS mapping images of SnO2@ZnIn2S4 composites.
Materials 17 02367 g002
Figure 3. The XRD pattern of (a) SnO2 nanoparticles and (b) SnO2@ZnIn2S4 composites.
Figure 3. The XRD pattern of (a) SnO2 nanoparticles and (b) SnO2@ZnIn2S4 composites.
Materials 17 02367 g003
Figure 4. (a) The survey XPS spectra of the SnO2 nanoparticles and SnO2@ZnIn2S4 composites. High-resolution XPS spectra of (b) Sn 3d and (c) O 1s for SnO2 nanoparticles and SnO2@ZnIn2S4 composites, respectively. High-resolution XPS spectra of (d) Zn 2p, (e) In 3d, and (f) S 2p for SnO2@ZnIn2S4 composites, respectively.
Figure 4. (a) The survey XPS spectra of the SnO2 nanoparticles and SnO2@ZnIn2S4 composites. High-resolution XPS spectra of (b) Sn 3d and (c) O 1s for SnO2 nanoparticles and SnO2@ZnIn2S4 composites, respectively. High-resolution XPS spectra of (d) Zn 2p, (e) In 3d, and (f) S 2p for SnO2@ZnIn2S4 composites, respectively.
Materials 17 02367 g004
Figure 5. The (a) FETEM image, (b) SAED pattern, (c) HRTEM image, and (d) FETEM EDS mapping images of SnO2@ZnIn2S4 composites.
Figure 5. The (a) FETEM image, (b) SAED pattern, (c) HRTEM image, and (d) FETEM EDS mapping images of SnO2@ZnIn2S4 composites.
Materials 17 02367 g005
Figure 6. (a) The average HER of SnO2 nanoparticles with different sacrificial reagents. (b) The average HER of SnO2 nanoparticles is decorated with various materials. (c) The average HER of SnO2@ZnIn2S4 composites with varying weights of SnO2 nanoparticles.
Figure 6. (a) The average HER of SnO2 nanoparticles with different sacrificial reagents. (b) The average HER of SnO2 nanoparticles is decorated with various materials. (c) The average HER of SnO2@ZnIn2S4 composites with varying weights of SnO2 nanoparticles.
Materials 17 02367 g006
Figure 7. (a) UV–visible absorption spectra, (b) PL spectra, and (c) photocurrent response of SnO2 nanoparticles and SnO2@ZnIn2S4 composites.
Figure 7. (a) UV–visible absorption spectra, (b) PL spectra, and (c) photocurrent response of SnO2 nanoparticles and SnO2@ZnIn2S4 composites.
Materials 17 02367 g007
Figure 8. The schematic diagram delineates the photocatalytic mechanism of SnO2@ZnIn2S4 composites.
Figure 8. The schematic diagram delineates the photocatalytic mechanism of SnO2@ZnIn2S4 composites.
Materials 17 02367 g008
Figure 9. (a) The average HER of SnO2@ZnIn2S4 composites under deionized water and tap water under blue or white LED light irradiation. (b) Reusability of SnO2@ZnIn2S4 composites for eight cycles. (c) XRD spectrum of SnO2@ZnIn2S4 composites after the eight cycles.
Figure 9. (a) The average HER of SnO2@ZnIn2S4 composites under deionized water and tap water under blue or white LED light irradiation. (b) Reusability of SnO2@ZnIn2S4 composites for eight cycles. (c) XRD spectrum of SnO2@ZnIn2S4 composites after the eight cycles.
Materials 17 02367 g009
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.; Bi, J.-N.; Pan, K.-Y.; Chiao, Y.-C. Microwave-Assisted Synthesis of SnO2@ZnIn2S4 Composites for Highly Efficient Photocatalytic Hydrogen Evolution. Materials 2024, 17, 2367. https://doi.org/10.3390/ma17102367

AMA Style

Chang Y-C, Bi J-N, Pan K-Y, Chiao Y-C. Microwave-Assisted Synthesis of SnO2@ZnIn2S4 Composites for Highly Efficient Photocatalytic Hydrogen Evolution. Materials. 2024; 17(10):2367. https://doi.org/10.3390/ma17102367

Chicago/Turabian Style

Chang, Yu-Cheng, Jia-Ning Bi, Kuan-Yin Pan, and Yung-Chang Chiao. 2024. "Microwave-Assisted Synthesis of SnO2@ZnIn2S4 Composites for Highly Efficient Photocatalytic Hydrogen Evolution" Materials 17, no. 10: 2367. https://doi.org/10.3390/ma17102367

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