Next Article in Journal
Production of Hydrogen from Lignocellulosic Biomass: A Review of Technologies
Previous Article in Journal
Effect of Decoration of C@TiO2 Core-Shell Composites with Nano-Ag Particles on Photocatalytic Activity in 4-Nitrophenol Degradation
Previous Article in Special Issue
Photocatalytic Degradation of Diclofenac by Nitrogen-Doped Carbon Quantum Dot-Graphitic Carbon Nitride (CNQD)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in the Use of SnO2 Quantum Dots: From Synthesis to Photocatalytic Applications

1
School of Mechanical Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
2
Physics Division, Department of Basic Sciences and Humanities, GMR Institute of Technology, GMR Nagar, Rajam 532127, India
3
Department of Automotive Engineering, Hanyang University, Seoul 04763, Republic of Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(4), 765; https://doi.org/10.3390/catal13040765
Submission received: 11 March 2023 / Revised: 14 April 2023 / Accepted: 16 April 2023 / Published: 17 April 2023
(This article belongs to the Special Issue Nanomaterials for Photocatalysis II)

Abstract

:
This review article provides current developments in SnO2 quantum dots (QDs) as effective catalysts over the last five years. SnO2 QDs are exceptional prospects for catalytic applications because of their high surface area, compact size, and tunable optical features. SnO2 QDs have recently made strides in their production and functionalization, which has enabled successful use of them as photocatalytic catalysts. The basic concepts of SnO2 QDs, including their electrical and optical characteristics, are described in this review paper, along with the most current findings on their production and functionalization. Additionally, it covers the fundamental mechanisms that support SnO2 QDs’ catalytic activity and emphasizes the difficulties involved in using them as catalysts. Lastly, it offers a forecast for the direction of research in this quickly evolving topic. Overall, our analysis demonstrates SnO2 QDs’ potential as a successful and cutting-edge catalytic system in recent years.

1. Introduction

Water pollution is a major issue for modern society, caused by the discharge of untreated waste into water sources [1]. This contamination can cause health problems, ecological damage, and economic losses. Quantum dots (QDs) are semiconductor crystals with sizes ranging from 1 to 10 nanometers in diameter. Owing to their small size and distinctive physical and chemical characteristics, QDs are appealing for a variety of applications [2]. Several materials, including semiconductor, metal, and carbon QDs, can be used to create QDs. Because of their remarkable optical and electrical capabilities, semiconductor QDs have been the subject of most research [3]. For example, II–VI and III–V compounds with high quantum yield, narrow emission spectra, and high charge carrier mobility, such as CdSe, CdTe, InP, and GaAs, have received a great deal of attention [4]. Due to their distinctive qualities and possible uses, metal oxide QDs, such as SnO2, TiO2, and ZnO, have also received a lot of attention [5]. The huge surface area of metal oxide QDs, which enables effective catalytic reactions, is one of their key advantages. SnO2 QDs, in particular, have a number of distinctive qualities that set them apart from other QD kinds. First off, SnO2 is an environmentally beneficial substitute for other harmful or expensive materials because it is a substance that is readily available on Earth and is not toxic [6]. Second, SnO2 QDs are promising candidates for a variety of applications, including photocatalysis, sensing, and energy conversion, thanks to their superior optical and electrical characteristics, which include a high surface area, strong stability, and a broad bandgap [7].
Furthermore, SnO2 QDs have demonstrated strong catalytic activity in the oxidation of organic contaminants and the water-splitting process that produces hydrogen [8]. This is a result of their special electrical characteristics, which enable effective charge separation and transfer. These characteristics include high electron mobility and an appropriate conduction band location [9]. SnO2 QDs’ tiny size and the large surface area also make more active sites for catalytic reactions possible, improving the efficiency of such reactions. Many techniques for regulating SnO2 QDs’ size, shape, and surface properties have been developed as a result of in-depth research into their synthesis and functionalization [10]. This has created new possibilities for modifying SnO2 QDs’ catalytic performance for certain applications. More study is required to completely comprehend their underlying mechanisms and improve their catalytic efficiency, given the intriguing features and possible uses of SnO2 QDs. In order to accomplish this, new synthesis and functionalization techniques must be created, and the relationship between a substance’s structure, characteristics, and performance must be better understood. These investigations have demonstrated that SnO2 QDs may degrade a variety of pollutants. Furthermore, they have demonstrated effectiveness in disinfecting water sources by inactivating pathogenic bacteria and viruses. Additionally, SnO2 QDs can be easily modified and functionalized to enhance their performance in water treatment [11]. As the size of the SnO2 QDs decreases, the bandgap energy increases, which leads to enhanced light absorption and higher photocatalytic activity. This unique characteristic makes SnO2 QDs more effective photocatalysts than bulk SnO2 [6]. SnO2 QDs exhibit enhanced reactivity compared to bulk SnO2 due to their high surface energy and the presence of surface defects. These characteristics make it easier for reactive oxygen species (ROS) to develop, including hydroxyl radicals (•OH), which are in charge of degrading contaminants [12]. The bandgap energy of SnO2 QDs can be adjusted by controlling their size and surface chemistry. This property enables the optimization of the photocatalytic activity of SnO2 QDs for specific applications, such as the degradation of particular pollutants or the production of hydrogen through water splitting [13]. SnO2 QDs display excellent stability under different environmental conditions, making them suitable for long-term water treatment applications [14].
This article will focus on the optical and electronic characteristics of SnO2 QDs that are relevant to photocatalysis. The optical properties of SnO2 QDs are size-dependent and result from the quantum confinement phenomenon [15]. The energy levels within QDs are discrete, and the bandgap energy increases as the particle size decreases. Consequently, SnO2 QDs have a tunable bandgap that can be customized by altering their size. These QDs have an optical absorption edge that reaches into the visible range, allowing them to be utilized for visible light photocatalysis [16]. Furthermore, SnO2 QDs exhibit strong light scattering and luminescence, which can enhance photocatalytic efficiency. With respect to electronic properties, SnO2 QDs have a high surface area to volume ratio, resulting in an increased number of surface defects and trap states [8]. The photocatalytic effectiveness of the QDs may decrease as a result of these defects and trap states acting as recombination foci for photogenerated electron–hole pairs. The electrical characteristics of SnO2 QDs can be changed by adding surface changes or doping to overcome this restriction [10]. For instance, SnO2 QDs that have been modified with other materials, such as TiO2, have demonstrated improved electron extraction and transport properties as well as suppressed charge recombination, resulting in better photocatalytic performance [17]. Additionally, doping SnO2 QDs with nitrogen or carbon can introduce new energy levels that can act as electron traps, reducing the recombination rate and increasing the photocatalytic activity [18]. SnO2 QDs can be synthesized utilizing a variety of methods, including the hydrothermal method, solvothermal method, and sol–gel method [19]. They have demonstrated great potential as a solution to water pollution, and this review aims to provide an in-depth analysis of recent advances in SnO2 quantum dot synthesis, modification, and application in water treatment. To encourage the wider deployment of SnO2 QDs for environmental applications, this review will emphasize major discoveries, problems, and prospects for future research in this area.
The size-dependent optical characteristics of SnO2 QDs are ascribed to the quantum confinement effect [20]. The energy levels in the QDs are discrete, and the bandgap energy increases with decreasing particle size. As a result, SnO2 QDs have a tunable bandgap that can be tailored by controlling their size. SnO2 QDs have been shown to possess a remarkable light-harvesting capacity, thereby enhancing photocatalytic efficiency. Additionally, the high surface area to volume ratio of SnO2 QDs has been found to contribute to improved photocatalytic performance. This can be improved by modifying their electronic properties through surface modifications or doping. Strategies, such as doping, composite construction, and adjustment of co-catalyst size, can be employed to improve the photocatalytic performance of SnO2 QDs [21]. Ultimately, the information in this review will give readers an understanding of the potential of SnO2 QDs as a water pollution remedy and the methods for improving their photocatalytic performance for environmental applications. In recent years, there has been a lot of interest in the usage of nanocomposites of SnO2 QDs with other semiconductors to enhance photocatalytic performance. Over the past 5 years, multiple research articles have focused on the importance of making these nanocomposites to create heterostructure formations that improve photocatalytic activity. In order to create nanocomposites and heterostructure composites with improved photocatalytic activity for water treatment and other applications, these investigations looked into several ways to combine SnO2 with other semiconductors [22,23,24]. The higher visible light absorption and effective separation of photogenerated electron–hole pairs at the nanocomposite interface are two reasons for the enhanced photocatalytic activity of these materials. The authors also reported a synergistic effect between the SnO2 QDs and other selected components in the nanocomposite, which improved the visible light absorption and the separation of photogenerated electron–hole pairs in the heterojunctions. Overall, the studies on nanocomposites of SnO2 QDs with other semiconductors have demonstrated promising results for enhancing photocatalytic activity. These findings could have important implications for developing more efficient and sustainable technologies for water treatment and other applications.

2. Synthesis Protocols

SnO2 QDs are prepared through various methods, such as solution synthesis, hydrothermal synthesis, and chemical reduction synthesis, among others [25,26,27,28,29]. While each approach has benefits and drawbacks, the choice of synthesis process is dictated by the desired QD qualities and the particular application. Alternative techniques for creating SnO2 QDs include sonochemical synthesis, microwave-assisted synthesis, solvothermal synthesis, and template-assisted synthesis, among others [30]. Understanding the SnO2 QD synthesis methods is crucial since the synthesis process can affect the QDs’ characteristics and performance, which can affect their potential usage in a variety of applications, such as photocatalysis. SnO2 QDs can have their size, shape, and crystal structure controlled throughout the fabrication process, which may have an effect on how well they function as photocatalysts. Thus, the advancement of SnO2 QD applications in areas including water purification, air purification, and solar energy conversion depends on the development of effective and controllable synthesis methods [31]. To improve the synthesis techniques for SnO2 QDs and to comprehend the underlying mechanisms that control their characteristics and performance, more research is required.

2.1. Microwave-Assisted Synthesis

In order to create SnO2 QDs, microwave-assisted synthesis is a practical and efficient method. This technique has the advantage of producing uniformly sized, high-quality QDs in a short period of time. Inducing the creation of nuclei and accelerating particle growth using microwave irradiation can lead to the formation of well-defined, extremely uniform SnO2 QDs [32]. This method offers a highly precise, scalable, and energy-efficient method for producing SnO2 QDs. Furthermore, microwave-assisted synthesis is a low-temperature technique that uses less energy and produces fewer undesirable byproducts. This process is simple and affordable, making it simple to scale up for the large-scale manufacturing of SnO2 QDs. For creating SnO2 QDs, microwave-assisted synthesis has a number of benefits [33]. Microwave irradiation, which accelerates particle growth, can make it easier to produce SnO2 QDs with precise size and shape. This method is also very adaptable and may be used to make vast amounts of SnO2 QDs. This process is appealing for the mass production of SnO2 QDs since it is simple, affordable, and energy-efficient. This method was used by researchers Xu et al. to produce densely packed, extremely crystalline SnO2 QDs [26]. Smaller than the SnO2 Bohr radius, they were able to produce homogenous particles with an average radius of 1–1.5 nm. This raises the possibility of a substantial quantum-size effect, in which a dot’s wave function extends into that of its neighbors, resulting in an electrically active QD solid. According to these results, the distinctive properties of SnO2 QDs and their potential applications in several fields are now well characterized [26,32,33].
Geng et al. also employed a microwave-assisted method to synthesize SnO2 QDs [32]. Their technique involved adding tin (II) chloride dehydrate to a mixed solvent of ethanol and water, which was then subjected to microwave irradiation at 180 °C for 30 min. The resulting SnO2 QDs had an average size of 4–6 nm and a consistent spherical shape. This method allowed for a shorter reaction time and higher yield compared to conventional hydrothermal methods. Additionally, the smaller particle size of the SnO2 QDs obtained through microwave-assisted synthesis is advantageous for higher surface area and improved electrochemical performance. Parthibavarman et al. employed a similar strategy but different precursor materials to create SnO2 QDs [33]. They were successful in producing particles with a narrow size range and a mean diameter of 2.8 nm. When exposed to visible light, these particles demonstrated remarkable photocatalytic activity in degrading rhodamine B. High-quality SnO2 QDs with features that could make them helpful in a variety of applications, from electronics to photocatalysis, are being produced successfully by using microwave-assisted synthesis.

2.2. Hydrothermal Synthesis

Hydrothermal synthesis is a promising technique for synthesizing SnO2 QDs due to its several advantages. It provides precise control over particle size and morphology by adjusting reaction conditions such as temperature, pressure, and reaction time. This control over the particle size and shape is crucial as it directly affects the properties and performance of the QDs. Additionally, hydrothermal synthesis leads to high-quality SnO2 QDs with a high degree of crystallinity, purity, and uniformity due to the high-temperature and high-pressure conditions that promote the growth of single crystals and reduce the formation of impurities. Moreover, hydrothermal synthesis provides a somewhat easy and affordable way to produce SnO2 QDs on a large scale. Recent research studies have also shown the effectiveness of hydrothermal synthesis in the preparation of SnO2 QDs. For instance, Praveen et al. utilized hydrothermal synthesis to prepare SnO2 QDs for use in photodetectors [25], while Yu et al. synthesized SnO2 QDs by varying the concentration of C2H3ClO2 to enhance their photocatalytic activity (Figure 1) [34]. Overall, the production of high-quality and tightly controlled SnO2 QDs using a straightforward and scalable procedure is made possible by hydrothermal synthesis.

2.3. Wet Chemical Synthesis

Solution-based synthesis is a highly popular method for creating colloidal SnO2 QDs due to its ability to produce uniform particles with controlled size and shape [27]. The most often used method involves injecting a heated organic solvent containing a capping agent with a precursor solution of SnCl4 and a reducing agent. The capping agent stabilizes the nanoparticles and prevents agglomeration, while the organic solvent serves as a reaction medium. Since the reaction is commonly conducted at high temperatures, SnO2 QDs with a restricted size distribution can form quickly and expand rapidly. The quality of the colloidal QDs can be improved by adding further purification procedures such as centrifugation and solvent washing. With the help of this technique, colloidal SnO2 QDs with customized properties can be made in a variety of applications, such as sensing, solar cells, and catalysis.
Liu et al. employed a straightforward solution processing approach to convert SnCl2·2H2O into a SnO2 QD colloidal solution, with anhydrous ethanol and deionized water serving as the solvent [35]. Then, for 48 h, the suspension was constantly stirred to create a SnO2 colloidal precursor solution. The use of this solution as an electron transport layer led to increased electron extraction, reduced charge recombination, and dramatically improved total solar cell efficiency. By dispersing SnCl2·2H2O and thiourea in DI water for one to two days, Gao et al. were able to create a colloidal solution of SnO2 QDs, which was then coupled with rGO for battery applications (Figure 2) [36]. Jin et al. also utilized a similar approach to create SnO2 QDs and functionalize MoO3 nanobelts for high-selectivity ethylene sensing [22]. Meanwhile, Chen et al. utilized SnCl4·5H2O, NaOH, H2O, and ethanol to make very stable SnO2 QDs, which were made in a Teflon-lined, stainless steel autoclave and heated at 150 °C for 6 h [37]. They employed tetramethylammonium hydroxide in the synthesis to obtain highly stable QDs, which were used for high-performance QLEDs with a SnO2-based ETL. Luo et al. used a different method to prepare SnO2 QDs for NO2 gas sensing [38]. They mixed SnCl4·5H2O, oleylamine, oleic acid, and deionized water in a three-neck flask and subjected it to vacuum conditions at 30 °C for an hour. After the mixture became clear and transparent, the temperature of the flask was raised to 80 °C under an N2 gas flow for six hours. The resulting mixture was then put into autoclaves and kept there for three hours at 180 °C. The studies mentioned above have demonstrated the potential of colloidal SnO2 QDs for various applications, including photocatalysis, electrochemical energy storage, and sensing. However, it is crucial to carefully consider the synthesis parameters, as the choice of synthesis method and reaction conditions can impact the size, morphology, and properties of the resulting QDs. In conclusion, a solution-based synthesis is a promising approach for creating colloidal SnO2 QDs with well-controlled sizes and properties.

2.4. Chemical Reduction Approach

Another extensively used technique for creating SnO2 QDs is chemical reduction. In this process, a reducing agent and a stabilizing agent are utilized to reduce the precursor metal ions into the QDs. This method allows for the use of both organic and inorganic reducing agents, such as sodium borohydride and hydrazine. The size and shape of the resultant QDs might vary greatly depending on the reducing agent used. The reduction reaction produces SnO2 QDs with a limited size range when it occurs at low temperatures, usually between 60 and 90 °C. The stabilizing agent, typically a surfactant or polymer, helps to prevent the particles from agglomerating and stabilizes the surface of the QDs. Post-synthesis purification steps, such as centrifugation and washing with solvents, can further enhance the quality of the colloidal QDs. The chemical reduction method offers a versatile and straightforward approach for preparing SnO2 QDs with tailored properties, making them promising candidates for various applications, including sensing, energy storage, and catalysis (Figure 3). Dutta et al. utilized a straightforward chemical reduction method to produce SnO2 QDs with SnCl4·5H2O and hydrazine hydrate as precursors [28]. The reaction was conducted at 100 °C for 18 h, yielding QDs ranging from 2 to 3 nm. Likewise, Babu et al. employed a similar method to synthesize in situ Ag/SnO2 and Au/SnO2 QDs for photocatalytic degradation of water pollutants, such as MO, MB, and RhB dyes [39,40].

2.5. Other Methods

Wahab et al. synthesized colloidal SnO2 QDs via a reflux method [42]. Firstly, tin chloride was added to methanol and continuously stirred until fully dissolved. Using NH3 solution, the pH was brought down to 9.5 before the mixture was transferred to a refluxing pot with two necks and cooked at 65 °C for three hours. The generated SnO2 QDs were examined for possible biological uses on cancerous myoblast (C2C12) cells. SnO2 QDs were synthesized using a biosynthesis approach by Fatimah et al., in which Clitoria ternatea flower extract and SnCl2·2H2O were utilized as the tin precursor [43]. The size range of the QDs that the authors were able to obtain was 4–6 nm. The prepared SnO2 QDs showed efficient degradation of the rhodamine B pollutant. Kumar et al. utilized the solution combustion method to synthesize SnO2 QDs [29]. They used urea (CO(NH2)2) and tin chloride pentahydrate (SnCl4·5H2O) as precursors in DI water for their study. Urea was used as a fuel to start the reaction, and the solution was heated in a furnace to a temperature range of 350 to 550 °C. The SnO2 QDs obtained through this method had a crystallite size of less than 4 nm. Similarly, Babu et al. utilized the solution combustion synthesis method to prepare various transition metal-doped SnO2 QDs for photocatalytic applications [40]. Table 1 describes the synthesis techniques of SnO2 QDs.

3. Modification of SnO2 QDs for Photocatalytic Applications

3.1. Doping of SnO2 QDs

Enhancing the photocatalytic efficiency of SnO2 QDs through doping is a crucial method. Doping can inject impurity levels within the band gap, improving the characteristics of light absorption and lowering the rate at which electron–hole pairs recombine. Dopants can be included in SnO2 QDs to modify the band gap energy, which can enhance photocatalytic activity in the visible light spectrum [50]. Additionally, doping can also alter the surface properties of SnO2 QDs, which can enhance the surface reactivity and increase the adsorption capacity for target pollutants [51]. SnO2 QDs are, therefore, a desirable material for environmental cleanup applications due to the large improvements in photocatalytic efficiency and selectivity that can be achieved through doping. The choice of dopant elements for SnO2 QDs in photocatalytic applications depends on the desired properties and specific application requirements. Generally, transition metals such as Fe, Co, Ni, Mn, and Cu, as well as non-metals, such as N and C, have been successfully used as dopants in SnO2 QDs for photocatalytic applications [29,45,52,53]. SnO2 QDs’ electrical and optical characteristics can be changed by these dopants, improving the photocatalytic activity, charge transfer efficiency, and absorption spectrum. For example, nitrogen doping can narrow the bandgap of SnO2 QDs, increasing their visible light absorption, while transition metal doping can enhance their catalytic activity toward specific reactions [54]. However, the doping concentration, kind of dopant, and manufacturing process should be carefully regulated in order to achieve the desired properties and avoid any negative effects on photocatalytic performance. Chen et al. recently created a quick and easy approach for creating Ni-doped SnO2 QDs with improved photocatalytic efficiency for breaking down Rhodamine B [44]. The SnO2 QDs were created using a hydrothermal technique, and the researchers then looked at how Ni doping affected the photocatalytic activity of the SnO2 QDs. After 28 min of solar light irradiation, the results showed that QDs with a Ni/Sn atomic ratio of 1% had the best photocatalytic performance, destroying 91.5% of Rhodamine B. Ni was added to the SnO2 QDs, which improved photodegradation performance by increasing light harvesting and reducing band gaps. SnO2 QDs also offered more active sites for photocatalytic reactions due to their small particle size and high specific surface area. The 1% Ni-doped SnO2 QDs outperformed their pure counterparts in terms of photocatalytic activity.
In their study, Wang et al. looked at how niobium alteration could enhance the visible-light photocatalytic abilities of SnO2 QDs (Figure 4) [45]. One-step hydrothermal synthesis of niobium-modified SnO2 QDs was carried out by the researchers, who then evaluated their photocatalytic capabilities. The changed band structure brought on by the presence of the transition metal was blamed for the increase in photocatalytic activity. The rate constant of Nb-SnO2 QDs in MO deterioration was found to be 2.34 times higher than that of pristine SnO2 QDs after characterizing the crystal structures and optical characteristics of the Nb-SnO2 QDs. According to the study’s findings, the changed band structure caused by the transition metal plays a significant role in increasing the photocatalytic abilities of SnO2 QDs when exposed to visible light. For the breakdown of antibiotic pollutants caused by visible light, Wang et al. presented a study on the modification of tin oxide QDs’ band structure and the improvement of their photocatalytic capabilities through Mo doping [45]. The effective production of extremely potent hydroxyl radicals is made possible by the photogenerated holes’ strong oxidizing properties in the valence band edge over 3 eV. This study explores the impact of Mo inclusion on the microstructural, compositional, electrical, and optical properties of SnO2 QDs and describes a method for producing them. The modified QDs demonstrated antibiotic elimination capabilities via visible-light-driven photocatalysis. Mo-doped SnO2 QDs were easily synthesized by the authors utilizing a green synthesis technique. Mo dopants changed the electrical band structure of SnO2 QDs, lowering the band gap and bringing about photocatalytic functions controlled by visible light. With a degradation efficiency of up to 96.5%, the nanostructured photocatalysts showed effective capabilities in the tetracycline hydrochloride degradation process. In a study, Chu et al. looked at how well SnO2 QDs doped with Bi3+ worked in combination with BiPO4 to clean various organic contaminants [46]. The synthetic composite material successfully broke down a number of organic dyes and medicines. In this article, a broad-spectrum photocatalyst called BiPO4/Bi-SnO2 QD is synthesized in detail, and its enhanced photocatalytic performance in comparison to pure Bi-SnO2 and BiPO4 is investigated. The composite material was created by a simple hydrothermal one-pot process and demonstrated enhanced photocatalytic efficacy when exposed to artificial sunlight. The effects of catalyst dosage, starting concentration, and common anions were all studied in relation to the breakdown of bisphenol A (BPA). The composite material also showed success in degrading a number of organic dyes and medicines.
High photocatalytic activity of pure and Ag-doped SnO2 QDs produced using a one-step microwave-assisted technique was reported by Parthibavarman et al. The optical characteristics of SnO2 were greatly enhanced by Ag doping [33]. Under visible light irradiation, the Ag-doped SnO2 catalyst demonstrated a maximum RhB degrading efficiency of 97.5% with good reusability and stability after seven cycles. SnO2 nanoparticles’ structural, morphological, compositional, and optical characteristics, as well as their photocatalytic performance, were all influenced by the amount of Ag dopant present. With Ag doping, SnO2’s band gap was reduced from 3.54 to 3.09 eV. According to a study by Shao et al., Nb-modified SnO2 QDs produced through aqueous synthesis were used to clean agricultural waste in situ while also photocatalyzing the breakdown of polyethylene [36]. According to the study, Nb-modified SnO2 QDs can be used to remove polyethylene in an innovative and environmentally benign way. After 6 h of exposure to visible light, Nb-SnO2 QDs had a 29% degradation efficiency. Due to the effective photocatalytic degradation occurring in visible light, this low-cost technology offers a fresh approach to the in situ treatment of agricultural waste, and their potential for real-world use is highlighted. Babu et al. studied the improved visible light photocatalytic activity of solution combustion-synthesized Cu-doped SnO2 QDs [40]. In this study, solution combustion synthesis was successfully used to create Cu-doped SnO2 QDs and the photocatalytic degradation of the methyl orange dye under visible light irradiation was examined. The findings demonstrated that the optical band gap energy of the QDs decreased with increasing Cu content. The best photocatalytic performance was demonstrated by the QDs doped with 0.03 mol% Cu, which degraded the methyl orange dye by 99% under visible light in about 180 min. These results point to the potential of solution combustion-produced Cu-doped SnO2 QDs as efficient photocatalysts for the degradation of organic impurities when exposed to visible light.

3.2. SnO2 QDs’ Decorated 1D Nanostructures

Due to their distinct optical and electrical characteristics, one-dimensional core–shell nanorods are a potential material for photocatalytic applications. Efficient charge separation is facilitated by the core–shell structure, which is essential for high photocatalytic performance. The nanorods’ high aspect ratio and anisotropic shape provide a large surface area and directional charge transfer, respectively, which are beneficial for catalytic reactions. Core–shell nanorods can be made from a variety of substances, including metal oxides and semiconductors, and they can be customized for particular applications by varying the size and makeup of the core and shell. The application of QDs as a shell material is one fascinating field of research. For example, SnO2 QDs as the shell material in core–shell nanorods enhance their performance. SnO2 QDs have unique electronic and optical properties that allow efficient charge separation, and their anisotropic shape facilitates directional charge transfer, which enhances photocatalytic activity. Various methods, including hydrothermal synthesis and sol–gel methods, can be used to synthesize core–shell nanorods with SnO2 QDs as the shell material [55]. The potential applications of 1-dimensional core-shell nanorods with SnO2 QDs as the shell material are significant and include environmental remediation and energy production.
The increased photoelectrochemical performance of cadmium sulfide–tin oxide QD core–shell nanorods was examined in a study by Babu et al. [56]. Using hydrothermal synthesis and straightforward chemical synthesis, core–shell nanorods with a cadmium sulfide core and a tin oxide quantum dot shell were created. The researchers found that the core–shell nanorods’ photoelectrochemical performance outperformed that of pure cadmium sulfide nanorods. This increase was due to the tin oxide quantum dot shells’ effective lowering of the core–shell nanorods’ bandgap from 2.37 eV to 2.35 eV, which led to larger photocurrents and lower charge transfer resistance. Overall, the research shows that core–shell nanorods with tin oxide QDs are a viable material for photoelectrochemical applications that are driven by solar light.
Babu et al. looked into optimizing core–shell nanostructures to improve the photocatalytic activity of colloidal SnO2 QD-decorated CdS NRs with visible light (Figure 5a) [40]. It was revealed that the colloidal SnO2 QDs’ synergistic effect helped these nanorods perform better in the dye degradation process. This paper describes a novel two-step method for creating core–shell heterojunction photocatalysts for SnO2 QDs using hydrothermal and ultrasonication procedures. Colloidal SnO2 QDs were used to completely decorate the exterior of the CdS NRs that were created, with the quantity chosen to maximize degrading efficiency. The colloidal SnO2 QDs protected the CdS NRs’ surface and helped explain the NRs’ superior dye-degradation performance (Figure 5b). These findings show that CdS@SnO2 core–shell NRs have the potential to be an effective material for photocatalytic processes involving visible light. The visible-light photocatalytic performance of CdS@SnO2 nanorods for the selective oxidation of benzyl alcohol to benzaldehyde was the focus of Liu et al. studies [51]. The core–shell structure of the nanorods was found to effectively accelerate the separation rate of electron–hole pairs and the electron lifetime of CdS nanorods, increasing visible light absorption and improving photocatalytic activity. Utilizing the use of a solvent-assisted interfacial reaction technique and ultrasonic stirring to anchor SnO2 nanoparticles to the surface of CdS NRs, the scientists created the core–shell structure CdS@SnO2 nanorods. The photocatalytic activity of the nanorods for the selective oxidation of benzyl alcohol to benzaldehyde under visible light irradiation was shown to be greatly improved by the core–shell configuration. This improvement was attributed to the core–shell structure’s improved visible light absorption, increased surface area, and more effective electron and hole separation.
Lee et al. have described the extremely effective interfacial charge transfer-based photocatalytic performance of SnO2-deposited ZnS nanorods [57]. The higher separation rate of photogenerated charges is principally responsible for the increased photocatalytic degradation efficiency of SnO2/ZnS nanocomposites. In this study, SnO2/ZnS nanocomposites with improved photocatalytic activity and photostability are made using a two-step hydrothermal procedure assisted by hydrazine. In comparison to pristine ZnS nanorods and commercial ZnS, the produced SnO2/ZnS nanocomposites exhibit 17-fold greater photostability and a 3-fold increase in photocatalytic activity. The increased separation rates of photogenerated charges, the increased number of active surface sites, and the larger light absorption range are all responsible for the improved photocatalytic performance of SnO2/ZnS nanocomposites. According to these results, SnO2/ZnS nanocomposites have a lot of potential for use as waste-water treatment photocatalysts. Table 2 provides a detailed description of recent research on the photocatalytic abilities of SnO2 QD-based nanocomposites.

3.3. SnO2 QDs’ Integrated 2D Nanostructures

Nanocomposite heterostructures composed of various nanomaterials have demonstrated improved photocatalytic activity compared to their individual components. In particular, zero-dimensional (0D) and two-dimensional (2D) nanocomposite heterostructures exhibit great potential in photocatalysis. QDs, as a type of 0D nanocomposite, possess a high surface area-to-volume ratio, enabling efficient charge separation and transmission [64]. On the other hand, graphene-based materials, as a type of 2D nanocomposite, exhibit high electrical conductivity, excellent mechanical properties, and large surface area, leading to improved light absorption and charge transport [65]. Combining these two types of nanocomposites can achieve improved photocatalytic properties, including high efficiency, increased light absorption range, improved charge transfer, and better chemical stability. Thus, 0D and 2D nanocomposite heterostructures hold promise for various photocatalytic applications, such as water purification, air pollution control, and renewable energy conversion [66]. Due to their distinct structural and optoelectronic characteristics, two-dimensional nanocomposite heterostructures and zero-dimensional QDs are particularly appealing materials for photocatalytic applications. QDs have tunable bandgap properties and discrete energy levels, enabling efficient light absorption and exciton generation, leading to enhanced photocatalytic activity. Additionally, QDs possess a high surface area-to-volume ratio, providing a larger number of active sites for catalytic reactions. On the other hand, 2D nanocomposite heterostructures show improved photocatalytic performance as a result of their distinctive band structures and effective charge separation characteristics [67]. Increased photocatalytic efficiency results from the parting of photogenerated charge carriers being made easier by the heterojunction contact between the materials. Furthermore, 2D nanocomposite heterostructures possess large surface areas and exposed edges, providing more active sites for catalytic reactions [68]. Due to their distinct characteristics, SnO2 QDs and 2D nanocomposite heterostructures are particularly interesting materials for photocatalytic applications [69]. SnO2 QDs possess a high surface area-to-volume ratio, providing more active sites for chemical reactions to occur. Their small size allows for efficient light absorption and effective electron–hole pair separation. Meanwhile, SnO2/graphene and SnO2/MoS2 2D nanocomposite heterostructures exhibit enhanced photocatalytic activity due to the synergistic effect between different materials, resulting in efficient charge transfer and improved electron–hole separation [70]. Additionally, the large surface area of 2D heterostructures allows for more active sites for catalytic reactions to take place. Combining 0D SnO2 QDs and 2D nanosheets in heterostructures can enhance photocatalytic activity by promoting charge separation, improving light absorption, and increasing active surface area. The unique morphology of heterostructures allows for efficient electron transport and reduced recombination of charge carriers. Therefore, creating heterostructures of 0D SnO2 QDs and 2D nanosheets holds great promise for enhancing the effectiveness and selectivity of photocatalytic reactions, which might result in the creation of more effective and long-lasting environmental remediation and energy conversion processes. These initiatives are essential for tackling international problems with sustainability and renewable energy (Figure 6a–c) [71].
Babu et al. have shown that the combination of stacked g-C3N4 nanolayers with SnO2 QDs results in increased sunlight-driven photocatalytic activity [41]. In this study, a composite made of SnO2 QDs-g-C3N4 nanolayers (NLs) for reducing water pollution is synthesized and evaluated. The combination exhibits enhanced photocatalytic activity in sunshine, degrading methyl orange (MO) by 94% in about 180 min. Strong contacts between the SnO2 QDs and g-C3N4 NLs were seen, and the average crystallite size of the SnO2 QDs was less than 3 nm. The synergistic interaction of stacked g-C3N4 and SnO2 QDs is responsible for the observed increase in photocatalytic activity. According to Sreekanth et al., SnO2 QDs may be easily decorated in one pot on the surface of iron phosphate nanosheets for improved catalytic decolorization of methylene blue dye when NaBH4 is present [72].
The degradation of methylene blue showed increased catalytic activity in the SnO2 QDs @ FePO4 nanosheets nanocomposite. Bail et al. developed a novel NH3 gas sensor utilizing SnO2 QDs and SnS2 nanosheets, as discussed in their research [73]. The authors proposed the rational design of heterostructure as an effective method to enhance the room-temperature NH3-sensing performance of the sensor. The SnO2 QDs/SnS2 nanocomposites were synthesized through a two-step hydrothermal method and exhibited remarkable room-temperature NH3-sensing performance with a low detection limit of 0.1 ppm. The nanocomposites also showed good stability and reproducibility. In the study conducted by Bail et al., a new NH3 gas sensor based on SnO2 QDs and SnS2 nanosheets was introduced. The authors proposed the rational design of heterostructure as an effective means to increase the room-temperature NH3-sensing performance of the sensor. Through a two-step hydrothermal process, SnO2 QDs/SnS2 nanocomposites were synthesized, which exhibited outstanding room-temperature NH3-sensing performance with a low detection limit of 0.1 ppm. Furthermore, the nanocomposites demonstrated stability and reproducibility. This research report outlines the synthesis of a nanocomposite called FP-Sn, consisting of SnO2 QDs and FePO4 nanosheets (FPNSs), for the degradation of methylene blue (MB). The preparation of FP-Sn was accomplished using a simple method without any additional instruments. The catalytic efficiency of FP-Sn was improved for the degradation of MB, resulting in the successful decolorization of approximately 92% of the MB in only 6 min in the presence of NaBH4. A unique technique for creating an integrated photocatalytic adsorbent (IPCA) was created by Mohanta et al., employing SnO2 QDs contained in carbon nanoflakes [62]. Due to the material’s synergistic effect, the SnO2-CNF nanocomposite showed improved effectiveness in the removal of bisphenol A (BPA) from water. SnO2-CNF has a 250 mg/g Langmuir adsorption capacity, which is 1.7 and 5.3 times larger than those of bare CNF and bare SnO2, respectively. With a removal effectiveness of about 98%, it was discovered that the coupled adsorption–photodegradation process was synergistically superior. Table 3 provides a detailed description of recent research on the photocatalytic abilities of SnO2 QD-based binary nanocomposites.

3.4. SnO2 QDs in Ternary Nanocomposites

SnO2 QDs have many benefits when used in ternary nanocomposites. SnO2 QDs can be coupled with other materials, such as metal oxides or carbon-based materials, to provide synergistic effects that can improve the photocatalytic activity and stability of the composite material [80]. Ternary nanocomposites are three-material structures. SnO2 QDs have a high surface-to-volume ratio and distinctive electrical characteristics that may facilitate electron transit and separation and increase photocatalytic activity. SnO2 QDs can also be used in visible-light-driven photocatalysis by changing the bandgap due to their small size. Moreover, the addition of SnO2 QDs to ternary nanocomposites can improve their stability and durability by preventing the aggregation of other materials and safeguarding them from degradation [81].
SnO2 QDs are good candidates for the development of highly effective and sustainable technologies for environmental remediation and energy conversion because they provide several advantages in ternary nanocomposites. In a study on a novel ternary nanocomposite for photocatalytic and photoelectrochemical (PEC) water-splitting applications, Babu et al. used g-C3N4, Au-SnO2 QDs [82]. The three-step procedure of sonication, stirring, and annealing was used to create the ternary nanocomposite of g-CN/Au-SnO2 QDs. The amount of Au-SnO2 QDs loaded was tuned to increase these qualities. The inclusion of Au-SQDs into g-CN improved the light absorption and bandgap. In comparison to pure g-CN, the CNAS-20 photoelectrode demonstrated greater PEC performance with a substantially larger photocurrent. Moreover, under visible light irradiation, the ternary nanocomposite showed increased photocatalytic activity in the destruction of RhB. A higher level of photocatalytic activity resulted from the addition of SnO2 QDs to the ternary nanocomposite, which facilitated electron transit and separation. A decreased charge-transfer resistance was also seen in the CNAS-20 photoelectrode, indicating improved PEC water-splitting efficiency. In order to improve solar-driven photocatalytic performance in hydrogen production and dye degradation, Ganesh et al. constructed a ternary composite comprising SnO2 QDs, AgVO3 nanoribbons, and g-C3N4 nanosheets (0D/1D/2D) structures (Figure 7) [24]. The ternary photocatalytic system exhibited a notable improvement in photocatalytic performance for hydrogen production compared to the individual photocatalysts of g-C3N4, AgVO3, and SnO2. The ternary photocatalytic system for H2 synthesis and RhB degradation under light irradiation was successfully synthesized for the investigation. During visible light irradiation, the ternary nanocomposite demonstrated enhanced photocatalytic performance for H2 generation and a significant improvement in RhB degradation. Compared to the AgVO3/g-C3N4 photocatalyst, the ternary nanocomposite showed improved photocatalytic activity for H2 generation, which was 4.5 times higher. Furthermore, the ternary nanocomposite exhibited effective photodegradation of RhB under light irradiation in only 50 min. The removal of carbamazepine (CBZ) from the water was studied by Duan et al. using composites of Ag-SnO2 QDs and silver phosphate (AgSn/AgP) [83]. With CBZ, the composites showed greater photodegradation efficiency. Superoxide, hydroxyl, and hole radicals all play a part in the mechanism underlying this improved performance. To prove that these reactive species cooperate in the breakdown of CBZ, the authors used radical trapping experiments and ESR measurements. Using liquid chromatography–mass spectrometry (LC-MS), the study discovered eight degradation intermediates. Toxicity assessments revealed that most photodegradation intermediates were less harmful than the parent drug CBZ. In order to create effective and ecologically friendly photocatalysts for eliminating organic contaminants from water, the in situ development of Ag-SnO2 QDs on silver phosphate offers a potential strategy. Table 4 provides a detailed description of recent research on the photocatalytic abilities of SnO2 QD-based ternary nanocomposites.

4. Conclusions and Perspectives

In conclusion, this review article systematically presents and reviews recent advances in SnO2 quantum dot-based photocatalytic systems. Various synthesis protocols, including microwave-assisted synthesis, hydrothermal synthesis, wet chemical synthesis, chemical reduction approach, and other methods, have been discussed to synthesize SnO2 QDs. Additionally, the modification of SnO2 QDs through doping, decorating with 1D nanostructures, and integrating with 2D nanostructures and ternary nanocomposites have been thoroughly reviewed to enhance their photocatalytic performance. Based on the analysis conducted in this review article, it is evident that each synthesis method for SnO2 QDs has its own set of merits and drawbacks. This review highlights that selecting the appropriate synthesis method is crucial in determining the type of composite that can be produced with other materials, especially when doping or creating nanocomposites with other semiconductors. Additionally, it is essential to consider the morphology of other materials, such as zero-dimensional, one-dimensional, two-dimensional, or three-dimensional materials, to integrate them with SnO2 QDs effectively. The colloidal synthesis method of SnO2 QDs is most suitable for providing coverage and dispersion throughout the material when preparing one-dimensional or two-dimensional materials for nanocomposites. Conversely, when making nanocomposites with metals, such as Au, Ag, or Pd, it is preferable to use the chemical reduction method. Sol–gel, hydrothermal, and solution combustion syntheses are commonly used methods for doping purposes. While it is not possible to definitively state which synthesis method is the “best,” this review provides valuable insights to enable authors to make informed decisions about which method to use for their specific application. It is recommended that authors consider these factors when fabricating nanocomposites to obtain a better heterojunction interface for fast electron–hole charge transfer.
To enhance the photocatalytic performance of SnO2 QDs, researchers need to focus on selecting a candidate with appropriate band edge positions to create Z-scheme and S-scheme photocatalysts. Despite limited attention in this area, integrating SnO2 QDs with other materials can further improve their photocatalytic performance by forming Z-scheme and S-scheme photocatalysts. Z-scheme photocatalysts enable efficient charge separation by establishing an electron–hole transfer pathway between two different semiconductors, while S-scheme photocatalysts enhance light absorption and utilization efficiency by utilizing a cascade of energy levels between two semiconductors. Therefore, the development of SnO2 QD-based Z-scheme and S-scheme photocatalysts is crucial for achieving better photocatalytic performance and promoting their practical applications in environmental remediation and energy conversion.
Ternary nanostructures incorporating SnO2 QDs received less attention in previous studies. Therefore, researchers are advised to direct their focus toward this area, as it can aid in capturing more solar light. SnO2 QDs possess a small size and can disperse easily in any 1D, 2D, 3D, and hierarchical nanostructures. This characteristic provides researchers with a chance to enhance the photocatalytic efficiency of SnO2 QDs through the proposed approach. In addition, researchers are encouraged to explore the potential of using SnO2 QDs as a protective shell material for other core materials, allowing for the maximum utilization of nanocomposites. Furthermore, researchers should focus on using SnO2 QDs as the 0D material in combination with 1D–2D–3D nanocomposites to achieve enhanced performance. Thus, researchers should invest in further studies to examine the efficacy of SnO2 QDs in the aforementioned areas. Overall, the proposed approach will enable the development of new and improved materials that can be utilized in various applications, including energy and environmental sciences. The development of robust and durable SnO2 QD-based photocatalysts can facilitate their practical application in various fields and contribute to the development of sustainable and green technologies.

Author Contributions

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

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (No. 2021R1C1C1011089).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no conflict of interest regarding the publication of this article, financial and/or otherwise.

References

  1. Sun, C.Y.; Yang, J.K.; Xu, M.; Cui, Y.; Ren, W.W.; Zhang, J.X.; Zhao, H.L.; Liang, B. Recent intensification strategies of SnO2-based photocatalysts: A review. Chem. Eng. J. 2022, 427, 131564. [Google Scholar] [CrossRef]
  2. Jouyandeh, M.; Mousavi Khadem, S.S.; Habibzadeh, S.; Esmaeili, A.; Abida, O.; Vatanpour, V.; Rabiee, N.; Bagherzadeh, M.; Iravani, S.; Reza Saeb, M.; et al. Quantum dots for photocatalysis: Synthesis and environmental applications. Green Chem. 2021, 23, 4931–4954. [Google Scholar] [CrossRef]
  3. Zhou, J.T.; Lyu, M.; Zhu, J.; Li, G.N.; Li, Y.T.; Jin, S.Z.; Song, J.L.; Niu, H.H.; Xu, J.Z.; Zhou, R. SnO2 Quantum Dot-Modified Mesoporous TiO2 Electron Transport Layer for Efficient and Stable Perovskite Solar Cells. ACS Appl. Energy Mater. 2022, 5, 3052–3063. [Google Scholar] [CrossRef]
  4. Kaur, R.; Rana, A.; Singh, R.K.; Chhabra, V.A.; Kim, K.-H.; Deep, A. Efficient photocatalytic and photovoltaic applications with nanocomposites between CdTe QDs and an NTU-9 MOF. RSC Adv. 2017, 7, 29015–29024. [Google Scholar] [CrossRef]
  5. Sun, P.; Xing, Z.; Li, Z.; Zhou, W. Recent advances in quantum dots photocatalysts. Chem. Eng. J. 2023, 458, 141399. [Google Scholar] [CrossRef]
  6. Kumar, N.; Dagupati, R.; Lim, J.; Choi, J. Bright red luminescence from Eu3+-activated noncytotoxic SnO2 quantum dots for latent fingerprint detection. Ceram. Int. 2022, 48, 17738–17748. [Google Scholar] [CrossRef]
  7. Wang, Y.L.; Fan, L.Q.; Sun, S.J.; Chen, J.J.; Wu, Z.X.; Zhu, T.T.; Huang, Y.F.; Wu, J.H. Ti3C2Tx MXene supported SnO2 quantum dots with oxygen vacancies as anode for Li-ion capacitors. Chem. Eng. J. 2022, 428, 131993. [Google Scholar] [CrossRef]
  8. Chen, D.Y.; Huang, S.S.; Huang, R.T.; Zhang, Q.; Le, T.T.; Cheng, E.B.; Hu, Z.J.; Chen, Z.W. Highlights on advances in SnO2 quantum dots: Insights into synthesis strategies, modifications and applications. Mater. Res. Lett. 2018, 6, 462–488. [Google Scholar] [CrossRef]
  9. Wu, T.T.; Zhen, C.; Wu, J.B.; Jia, C.X.; Haider, M.; Wang, L.Z.; Liu, G.; Cheng, H.M. Chlorine capped SnO2 quantum-dots modified TiO2 electron selective layer to enhance the performance of planar perovskite solar cells. Sci. Bull. 2019, 64, 547–552. [Google Scholar] [CrossRef]
  10. Xu, Y.; Matios, E.; Luo, J.M.; Li, T.; Lu, X.; Jiang, S.H.; Yue, Q.; Li, W.Y.; Kang, Y.J. SnO2 Quantum Dots Enabled Site-Directed Sodium Deposition for Stable Sodium Metal Batteries. Nano Lett. 2021, 21, 816–822. [Google Scholar] [CrossRef]
  11. Liu, J.Q.; Zhang, Q.R.; Tian, X.Y.; Hong, Y.; Nie, Y.C.; Su, N.N.; Jin, G.H.; Zhai, Z.X.; Fu, C. Highly efficient photocatalytic degradation of oil pollutants by oxygen deficient SnO2 quantum dots for water remediation. Chem. Eng. J. 2021, 404, 127146. [Google Scholar] [CrossRef]
  12. Sahu, B.K.; Kaur, G.; Das, A. Revealing Interplay of Defects in SnO2 Quantum Dots for Blue Luminescence and Selective Trace Ammonia Detection at Room Temperature. ACS Appl. Mater. Interfaces 2020, 12, 49227–49236. [Google Scholar] [CrossRef]
  13. Han, X.; Zhao, J.J.; An, L.; Li, Z.H.; Yu, X. One-step synthesis of oxygen vacancy-rich SnO2 quantum dots with ultrahigh visible-light photocatalytic activity. Mater. Res. Bull. 2019, 118, 110486. [Google Scholar] [CrossRef]
  14. Das, J.; Dhar, S.S. Synthesis of SnO2 quantum dots mediated by Camellia sinensis shoots for degradation of thiamethoxam. Toxicol. Environ. Chem. 2020, 102, 186–196. [Google Scholar] [CrossRef]
  15. Bhawna; Kumar, S.; Sharma, R.; Gupta, A.; Tyagi, A.; Singh, P.; Kumar, A.; Kumar, V. Recent insights into SnO2-based engineered nanoparticles for sustainable H-2 generation and remediation of pesticides. New J. Chem. 2022, 46, 4014–4048. [Google Scholar] [CrossRef]
  16. Wu, Z.D.; Jing, H.Y.; Zhao, Y.; Lu, K.R.; Liu, B.Y.; Yu, J.; Xia, X.F.; Lei, W.; Hao, Q.L. Grain boundary and interface interaction Co-regulation promotes SnO2 quantum dots for efficient CO2 reduction. Chem. Eng. J. 2023, 451, 138477. [Google Scholar] [CrossRef]
  17. Sun, B.; Chen, Y.Z.; Tao, L.; Zhao, H.B.; Zhou, G.D.; Xia, Y.D.; Wang, H.Y.; Zhao, Y. Nanorod Array of SnO2 Quantum Dot Interspersed Multiphase TiO2 Heterojunctions with Highly Photocatalytic Water Splitting and Self-Rechargeable Battery-Like Applications. ACS Appl. Mater. Interfaces 2019, 11, 2071–2081. [Google Scholar] [CrossRef]
  18. Mallikarjuna, K.; Rafiqul Bari, G.A.K.M.; Vattikuti, S.V.P.; Kim, H. Synthesis of carbon-doped SnO2 nanostructures for visible-light-driven photocatalytic hydrogen production from water splitting. Int. J. Hydrogen Energy 2020, 45, 32789–32796. [Google Scholar] [CrossRef]
  19. Sheikh, S.P.; Reddy, V.R.; Sen, P. Size-dependent nonlinear optical absorption in SnO2 quantum dots. J. Opt. Soc. Am. B-Opt. Phys. 2020, 37, 2788–2795. [Google Scholar] [CrossRef]
  20. Liu, J.Q.; Nie, Y.C.; Xue, W.T.; Wu, L.T.; Jin, H.; Jin, G.H.; Zhai, Z.X.; Fu, C. Size effects on structural and optical properties of tin oxide quantum dots with enhanced quantum confinement. J. Mater. Res. Technol.-JMRT 2020, 9, 8020–8028. [Google Scholar] [CrossRef]
  21. Mishra, S.R.; Ahmaruzzaman, M. Tin oxide based nanostructured materials: Synthesis and potential applications. Nanoscale 2022, 14, 1566–1605. [Google Scholar] [CrossRef]
  22. Jin, W.; Wang, H.; Liu, Y.L.; Yang, S.; Zhou, J.; Chen, W. SnO2 Quantum Dots-Functionalized MoO3 Nanobelts for High-Selectivity Ethylene Sensing. ACS Appl. Nano Mater. 2022, 5, 10485–10494. [Google Scholar] [CrossRef]
  23. Lee, J.H.; Mirzaei, A.; Kim, J.H.; Kim, J.Y.; Nasriddinov, A.F.; Rumyantseva, M.N.; Kim, H.W.; Kim, S.S. Gas-sensing behaviors of TiO2-layer-modified SnO2 quantum dots in self-heating mode and effects of the TiO2 layer. Sens. Actuator B-Chem. 2020, 310, 127870. [Google Scholar] [CrossRef]
  24. Koyyada, G.; Kumar, N.S.; Al-Ghurabi, E.H.; Asif, M.; Mallikarjuna, K. Enhanced solar-driven photocatalytic performance of a ternary composite of SnO2 quantum dots//AgVO3 nanoribbons//g-C3N4 nanosheets (0D/1D/2D) structures for hydrogen production and dye degradation. Environ. Sci. Pollut. Res. 2021, 28, 31585–31595. [Google Scholar] [CrossRef] [PubMed]
  25. Praveen, S.; Veeralingam, S.; Badhulika, S. A Flexible Self-Powered UV Photodetector and Optical UV Filter Based on beta-Bi2O3/SnO2 Quantum Dots Schottky Heterojunction. Adv. Mater. Interfaces 2021, 8, 2100373. [Google Scholar] [CrossRef]
  26. Xu, Z.J.; Jiang, Y.; Li, Z.X.; Chen, C.; Kong, X.Y.; Chen, Y.W.; Zhou, G.F.; Liu, J.M.; Kempa, K.; Gao, J.W. Rapid Microwave-Assisted Synthesis of SnO2 Quantum Dots for Efficient Planar Perovskite Solar Cells. ACS Appl. Energy Mater. 2021, 4, 1887–1893. [Google Scholar] [CrossRef]
  27. Liu, H.R.; Chen, Z.L.; Wang, H.B.; Ye, F.H.; Ma, J.J.; Zheng, X.L.; Gui, P.B.; Xiong, L.B.; Wen, J.; Fang, G.J. A facile room temperature solution synthesis of SnO2 quantum dots for perovskite solar cells. J. Mater. Chem. A 2019, 7, 10636–10643. [Google Scholar] [CrossRef]
  28. Dutta, D.; Bahadur, D. Influence of confinement regimes on magnetic property of pristine SnO2 quantum dots. J. Mater. Chem. 2012, 22, 24545–24551. [Google Scholar] [CrossRef]
  29. Manikandan, D.; Yadav, A.K.; Jha, S.N.; Bhattacharyya, D.; Boukhvalov, D.W.; Murugan, R. XANES, EXAFS, EPR, and First-Principles Modeling on Electronic Structure and Ferromagnetism in Mn Doped SnO2 Quantum Dots. J. Phys. Chem. C 2019, 123, 3067–3075. [Google Scholar] [CrossRef]
  30. do Nascimento, J.L.A.; Chantelle, L.; dos Santos, I.M.G.; de Oliveira, A.L.M.; Alves, M.C.F. The Influence of Synthesis Methods and Experimental Conditions on the Photocatalytic Properties of SnO2: A Review. Catalysts 2022, 12, 428. [Google Scholar] [CrossRef]
  31. Zhao, X.X.; Yang, Q.; Quan, Z.W. Tin-based nanomaterials: Colloidal synthesis and battery applications. Chem. Commun. 2019, 55, 8683–8694. [Google Scholar] [CrossRef] [PubMed]
  32. Geng, J.G.; Ma, C.T.; Zhang, D.; Ning, X.F. Facile and fast synthesis of SnO2 quantum dots for high performance solid-state asymmetric supercapacitor. J. Alloys Compd. 2020, 825, 153850. [Google Scholar] [CrossRef]
  33. Parthibavarman, M.; Sathishkumar, S.; Jayashree, M.; BoopathiRaja, R. Microwave Assisted Synthesis of Pure and Ag Doped SnO2 Quantum Dots as Novel Platform for High Photocatalytic Activity Performance. J. Clust. Sci. 2019, 30, 351–363. [Google Scholar] [CrossRef]
  34. Yu, B.; Li, Y.C.; Wang, Y.J.; Li, H.; Zhang, R.P. Facile hydrothermal synthesis of SnO2 quantum dots with enhanced photocatalytic degradation activity: Role of surface modification with chloroacetic acid. J. Environ. Chem. Eng. 2021, 9, 105618. [Google Scholar] [CrossRef]
  35. Cheng, Y.; Wang, S.H.; Zhou, L.; Chang, L.M.; Liu, W.Q.; Yin, D.M.; Yi, Z.; Wang, L.M. SnO2 Quantum Dots: Rational Design to Achieve Highly Reversible Conversion Reaction and Stable Capacities for Lithium and Sodium Storage. Small 2020, 16, e2000681. [Google Scholar] [CrossRef]
  36. Gao, L.; Wu, G.S.; Ma, J.; Jiang, T.C.; Chang, B.; Huang, Y.S.; Han, S. SnO2 Quantum Dots@Graphene Framework as a High-Performance Flexible Anode Electrode for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12, 12982–12989. [Google Scholar] [CrossRef]
  37. Chen, M.Y.; Chen, X.T.; Ma, W.C.; Sun, X.J.; Wu, L.J.; Lin, X.F.; Yang, Y.X.; Li, R.; Shen, D.Y.; Chen, Y.; et al. Highly Stable SnO2-Based Quantum-Dot Light-Emitting Diodes with the Conventional Device Structure. ACS Nano 2022, 16, 9631–9639. [Google Scholar] [CrossRef]
  38. Huang, Q.Q.; Liao, J.Y.; Zhang, Q.Y.; Lai, N.; Zhang, B.W.; Wang, C.; Yang, J.; Yang, Y.; Wang, J.; Zhang, G.L.; et al. The synergistic effect of lead-free quantum dots and SnO2 in glass-ceramics for broadband white-emission. J. Mater. Chem. C 2022, 10, 18285–18293. [Google Scholar] [CrossRef]
  39. Babu, B.; Koutavarapu, R.; Harish, V.V.N.; Shim, J.; Yoo, K. Novel in-situ synthesis of Au/SnO2 quantum dots for enhanced visible-light-driven photocatalytic applications. Ceram. Int. 2019, 45, 5743–5750. [Google Scholar] [CrossRef]
  40. Babu, B.; Harish, V.V.N.; Koutavarapu, R.; Shim, J.; Yoo, K. Enhanced visible-light-active photocatalytic performance using CdS nanorods decorated with colloidal SnO2 quantum dots: Optimization of core-shell nanostructure. J. Ind. Eng. Chem. 2019, 76, 476–487. [Google Scholar] [CrossRef]
  41. Babu, B.; Cho, M.; Byon, C.; Shim, J. Sunlight-driven photocatalytic activity of SnO2 QDs-g-C3N4 nanolayers. Mater. Lett. 2018, 212, 327–331. [Google Scholar] [CrossRef]
  42. Wahab, R.; Khan, F.; Al-Khedhairy, A.A. Quantization of SnO2 dots: Apoptosis and intrinsic effect of quantum dots for myoblast cancer cells with caspase 3/7 genes. Ceram. Int. 2020, 46, 6383–6395. [Google Scholar] [CrossRef]
  43. Fatimah, I.; Sahroni, I.; Muraza, O.; Doong, R.A. One-pot biosynthesis of SnO2 quantum dots mediated by Clitoria ternatea flower extract for photocatalytic degradation of rhodamine B. J. Environ. Chem. Eng. 2020, 8, 103879. [Google Scholar] [CrossRef]
  44. Chen, D.Y.; Huang, S.S.; Huang, R.T.; Zhang, Q.; Le, T.T.; Cheng, E.B.; Hu, Z.J.; Chen, Z.W. Convenient fabrication of Ni-doped SnO2 quantum dots with improved photodegradation performance for Rhodamine B. J. Alloys Compd. 2019, 788, 929–935. [Google Scholar] [CrossRef]
  45. Wang, Y.; Su, N.N.; Liu, J.Q.; Lin, Y.H.; Wang, J.K.; Guo, X.; Zhang, Y.H.; Qin, Z.K.; Liu, J.F.; Zhang, C.Y.; et al. Enhanced visible-light photocatalytic properties of SnO2 quantum dots by niobium modification. Results Phys. 2022, 37, 105515. [Google Scholar] [CrossRef]
  46. Chu, L.L.; Duo, F.F.; Zhang, M.L.; Wu, Z.S.; Sun, Y.Y.; Wang, C.; Dong, S.Y.; Sun, J.H. Doping induced enhanced photocatalytic performance of SnO2:Bi3+ quantum dots toward organic pollutants. Colloid Surf. A-Physicochem. Eng. Asp. 2020, 589, 124416. [Google Scholar] [CrossRef]
  47. Zhu, L.F.; Wang, M.Y.; Lam, T.K.; Zhang, C.Y.; Du, H.D.; Li, B.H.; Yao, Y.W. Fast microwave-assisted synthesis of gas-sensing SnO2 quantum dots with high sensitivity. Sens. Actuator B-Chem. 2016, 236, 646–653. [Google Scholar] [CrossRef]
  48. Babu, B.; Kadam, A.N.; Rao, G.T.; Lee, S.W.; Byon, C.; Shim, J. Enhancement of visible-light-driven photoresponse of Mn-doped SnO2 quantum dots obtained by rapid and energy efficient synthesis. J. Lumin. 2018, 195, 283–289. [Google Scholar] [CrossRef]
  49. Babu, B.; Cho, M.Y.; Byon, C.; Shim, J. One pot synthesis of Ag-SnO2 quantum dots for highly enhanced sunlight-driven photocatalytic activity. J. Alloys Compd. 2018, 731, 162–171. [Google Scholar] [CrossRef]
  50. Kaur, H.; Bhatti, H.S.; Singh, K. Dopant incorporation in ultrasmall quantum dots: A case study on the effect of dopant concentration on lattice and properties of SnO2 QDs. J. Mater. Sci.-Mater. Electron. 2019, 30, 2246–2264. [Google Scholar] [CrossRef]
  51. Shao, J.; Deng, K.; Chen, L.; Guo, C.M.; Zhao, C.S.; Cui, J.Y.; Shen, T.A.; Li, K.W.; Liu, J.Q.; Fu, C. Aqueous synthesis of Nb-modified SnO2 quantum dots for efficient photocatalytic degradation of polyethylene for in situ agricultural waste treatment. Green Process. Synth. 2021, 10, 499–506. [Google Scholar] [CrossRef]
  52. Wu, C.P.; Xie, K.X.; He, J.P.; Wang, Q.P.; Ma, J.M.; Yang, S.; Wang, Q.H. SnO2 quantum dots modified N-doped carbon as high-performance anode for lithium ion batteries by enhanced pseudocapacitance. Rare Met. 2021, 40, 48–56. [Google Scholar] [CrossRef]
  53. Fan, D.W.; Liu, X.; Shao, X.R.; Zhang, Y.; Zhang, N.; Wang, X.Y.; Wei, Q.; Ju, H.X. A cardiac troponin I photoelectrochemical immunosensor: Nitrogen-doped carbon quantum dots-bismuth oxyiodide-flower-like SnO2. Microchim. Acta 2020, 187, 332. [Google Scholar] [CrossRef]
  54. Liu, Y.; Wei, S.; Wang, G.; Tong, J.Y.; Li, J.; Pan, D.C. Quantum-Sized SnO2 Nanoparticles with Upshifted Conduction Band: A Promising Electron Transportation Material for Quantum Dot Light-Emitting Diodes. Langmuir 2020, 36, 6605–6609. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, D.; Li, Y.; Zhang, X.L.; Yang, L.; Luo, X. Heterostructured Perylene Diimide (PDI) Supramolecular Nanorods with SnO2 Quantum Dots for Enhanced Visible-Light Photocatalytic Activity and Stability. ChemCatChem 2022, 14, e202200087. [Google Scholar] [CrossRef]
  56. Babu, B.; Koutavarapu, R.; Shim, J.; Yoo, K. SnO2 quantum dots decorated NiFe2O4 nanoplates: 0D/2D heterojunction for enhanced visible-light-driven photocatalysis. Mater. Sci. Semicond. Process 2020, 107, 104834. [Google Scholar] [CrossRef]
  57. Hui, W.; Yang, Y.G.; Xu, Q.; Gu, H.; Feng, S.L.; Su, Z.H.; Zhang, M.R.; Wang, J.O.; Li, X.D.; Fang, J.F.; et al. Red-Carbon-Quantum-Dot-Doped SnO2 Composite with Enhanced Electron Mobility for Efficient and Stable Perovskite Solar Cells. Adv. Mater. 2020, 32, e1906374. [Google Scholar] [CrossRef]
  58. Duan, Z.R.; Deng, L.; Shi, Z.; Zhang, H.J.; Zeng, H.X.; Crittenden, J. In situ growth of Ag-SnO2 quantum dots on silver phosphate for photocatalytic degradation of carbamazepine: Performance, mechanism and intermediates toxicity assessment. J. Colloid Interface Sci. 2019, 534, 270–278. [Google Scholar] [CrossRef]
  59. Bhattacharjee, A.; Ahmaruzzaman, M. Facile synthesis of SnO2 quantum dots and its photocatalytic activity in the degradation of eosin Y dye: A green approach. Mater. Lett. 2015, 139, 418–421. [Google Scholar] [CrossRef]
  60. Reddy, C.V.; Babu, B.; Shim, J. Synthesis of Cr-doped SnO2 quantum dots and its enhanced photocatalytic activity. Mater. Sci. Eng. B-Adv. Funct. Solid-State Mater. 2017, 223, 131–142. [Google Scholar] [CrossRef]
  61. Begum, S.; Ahmaruzzaman, M. Green synthesis of SnO2 quantum dots using Parkia speciosa Hassk pods extract for the evaluation of anti-oxidant and photocatalytic properties. J. Photochem. Photobiol. B-Biol. 2018, 184, 44–53. [Google Scholar] [CrossRef]
  62. Mohanta, D.; Ahmaruzzaman, M. Biogenic synthesis of SnO2 quantum dots encapsulated carbon nanoflakes: An efficient integrated photocatalytic adsorbent for the removal of bisphenol A from aqueous solution. J. Alloys Compd. 2020, 828, 154093. [Google Scholar] [CrossRef]
  63. Babu, B.; Kadam, A.N.; Ravikumar, R.; Byon, C. Enhanced visible light photocatalytic activity of Cu-doped SnO2 quantum dots by solution combustion synthesis. J. Alloys Compd. 2017, 703, 330–336. [Google Scholar] [CrossRef]
  64. Ren, W.W.; Yang, J.K.; Zhang, J.X.; Li, W.; Sun, C.Y.; Zhao, H.L.; Wen, Y.T.; 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]
  65. Wei, Z.P.; Liu, M.N.; Li, H.G.; Sun, S.S.; Yang, L.Y. SnO2 Quantum Dots Decorated Reduced Graphene Oxide Nanosheets Composites for Electrochemical Supercapacitor Applications. Int. J. Electrochem. Sci. 2020, 15, 6257–6268. [Google Scholar] [CrossRef]
  66. Sharma, I.; Kumar, K.N.; Choi, J. Highly sensitive chemiresistive detection of NH3 by formation of WS2 nanosheets and SnO2 quantum dot heterostructures. Sens. Actuator B-Chem. 2023, 375, 132899. [Google Scholar] [CrossRef]
  67. Sahu, B.K.; Juine, R.N.; Sahoo, M.; Kumar, R.; Das, A. Interface of GO with SnO2 quantum dots as an efficient visible-light photocatalyst. Chemosphere 2021, 276, 130142. [Google Scholar] [CrossRef]
  68. Shao, S.F.; Chen, X.; Chen, Y.Y.; Zhang, L.; Kim, H.W.; Kim, S.S. ZnO Nanosheets Modified with Graphene Quantum Dots and SnO2 Quantum Nanoparticles for Room-Temperature H2S Sensing. ACS Appl. Nano Mater. 2020, 3, 5220–5230. [Google Scholar] [CrossRef]
  69. Zou, Y.Z.; Xie, Y.; Yu, S.; Chen, L.; Cui, W.; Dong, F.; Zhou, Y. SnO2 quantum dots anchored on g-C3N4 for enhanced visible-light photocatalytic removal of NO and toxic NO2 inhibition. Appl. Surf. Sci. 2019, 496, 143630. [Google Scholar] [CrossRef]
  70. Luo, J.T.; Li, C.; Yang, Q.T.; Yan, L.X.; Zhang, B.H.; Tao, R.; Rauf, S.; Li, H.L.; Fu, C. Facile Fabrication of MoS2 Nanoflowers/SnO2 Colloidal Quantum Dots Nanocomposite for Enhanced NO2 Sensing at Room Temperature. IEEE Sens. J. 2022, 22, 6295–6302. [Google Scholar] [CrossRef]
  71. Ren, L.; Yao, Y.; Wang, K.Y.; Li, S.T.; Zhu, K.J.; Liu, J. Novel one-step in situ growth of SnO2 quantum dots on reduced graphene oxide and its application for lithium ion batteries. J. Solid State Chem. 2019, 273, 128–131. [Google Scholar] [CrossRef]
  72. Sreekanth, T.V.M.; Ramaraghavulu, R.; Yoo, K.; Kim, J. Facile One-Pot Decoration of SnO2 Quantum Dots on the Surface of the Iron Phosphate Nanosheets for Enhanced Catalytic Decolorization of Methylene Blue Dye in the Presence of NaBH4. J. Clust. Sci. 2022, 33, 835–838. [Google Scholar] [CrossRef]
  73. Bai, J.Z.; Shen, Y.B.; Zhao, S.K.; Chen, Y.S.; Li, G.D.; Han, C.; Wei, D.Z.; Yuan, Z.Y.; Meng, F.L. Flower-like MoS2 hierarchical architectures assembled by 2D nanosheets sensitized with SnO2 quantum dots for high-performance NH3 sensing at room temperature. Sens. Actuator B-Chem. 2022, 353, 131191. [Google Scholar] [CrossRef]
  74. Dutta, D.; Thakur, D.; Bahadur, D. SnO2 quantum dots decorated silica nanoparticles for fast removal of cationic dye (methylene blue) from wastewater. Chem. Eng. J. 2015, 281, 482–490. [Google Scholar] [CrossRef]
  75. Du, F.; Zuo, X.; Yang, Q.; Yang, B.; Li, G.; Ding, Z.; Wu, M.; Ma, Y.; Jin, S.; Zhu, K. Facile assembly of TiO2 nanospheres/SnO2 quantum dots composites with excellent photocatalyst activity for the degradation of methyl orange. Ceram. Int. 2016, 42, 12778–12782. [Google Scholar] [CrossRef]
  76. Tajima, T.; Goto, H.; Nishi, M.; Ohkubo, T.; Nishina, Y.; Miyake, H.; Takaguchi, Y. A facile synthesis of a SnO2/Graphene oxide nano-nano composite and its photoreactivity. Mater. Chem. Phys. 2018, 212, 149–154. [Google Scholar] [CrossRef]
  77. Chen, D.; Huang, S.; Huang, R.; Zhang, Q.; Le, T.-T.; Cheng, E.; Yue, R.; Hu, Z.; Chen, Z. Construction of Ni-doped SnO2-SnS2 heterojunctions with synergistic effect for enhanced photodegradation activity. J. Hazard. Mater. 2019, 368, 204–213. [Google Scholar] [CrossRef] [PubMed]
  78. Van Pham, V.; Mai, D.-Q.; Bui, D.-P.; Van Man, T.; Zhu, B.; Zhang, L.; Sangkaworn, J.; Tantirungrotechai, J.; Reutrakul, V.; Cao, T.M. Emerging 2D/0D g-C3N4/SnO2 S-scheme photocatalyst: New generation architectural structure of heterojunctions toward visible-light-driven NO degradation. Environ. Pollut. 2021, 286, 117510. [Google Scholar] [CrossRef]
  79. Fu, Z.; Qin, B.; Guo, X.; Wang, Y.; Xu, Y.; Qiao, Q.; Wang, Q.; Wang, F.; Gao, S.; Yang, Z. Enhancement of photocatalytic dye degradation and photoconversion capacity of graphene oxide/SnO2 nanocomposites. J. Alloys Compd. 2022, 898, 162796. [Google Scholar] [CrossRef]
  80. Niu, J.J.; Pan, J.Q.; Fu, W.D.; Xiao, G.S.; Fu, Y.Y.; Cao, J.; Wang, J.J.; Zheng, Y.Y.; Shi, L.; Li, C.R. The photovoltaic enhancement of Cu2O/SnO2 quantum dots/ZnO arrays transparent pn junction device via carrier injection and potential transition of SiO2 QDs. J. Mater. Sci.-Mater. Electron. 2022, 33, 7300–7311. [Google Scholar] [CrossRef]
  81. Zhu, J.K.; Li, H.; Cui, X.Q.; Yang, Z.Y.; Chen, B.; Li, Y.H.; Zhang, P.F.; Li, J.R. Efficient photocathodic protection performance of ZnIn2S4 nanosheets/SnO2 quantum dots/TiO2 nanotubes composite for 316 SS under visible light. J. Alloys Compd. 2022, 926, 166901. [Google Scholar] [CrossRef]
  82. Babu, B.; Koutavarapu, R.; Shim, J.; Yoo, K. Enhanced visible-light-driven photoelectrochemical and photocatalytic performance of Au-SnO2 quantum dot-anchored g-C3N4 nanosheets. Sep. Purif. Technol. 2020, 240, 116652. [Google Scholar] [CrossRef]
  83. Chen, Y.W.; Jiang, Y.F.; Chen, B.Y.; Ye, F.L.; Duan, H.Q.; Cui, H.Y. Facile fabrication of N-doped carbon quantum dots modified SnO2 composites for improved visible light photocatalytic activity. Vacuum 2021, 191, 110371. [Google Scholar] [CrossRef]
  84. Mohanta, D.; Ahmaruzzaman, M. A novel Au-SnO2-rGO ternary nanoheterojunction catalyst for UV-LED induced photocatalytic degradation of clothianidin: Identification of reactive intermediates, degradation pathway and in-depth mechanistic insight. J. Hazard. Mater. 2020, 397, 122685. [Google Scholar] [CrossRef] [PubMed]
  85. Song, Z.; Wang, C.; Shu, S.; Liu, J.; Liu, J.; Li, Y.; Huang, L.; Huang, L. Facile synthesis CQDs/SnO2-x/BiOI heterojunction photocatalyst to effectively degrade pollutants and antibacterial under LED light. J. Photochem. Photobiol. B Biol. 2022, 236, 112566. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Diagrammatic representation of the SnO2 QD synthesis process [34], Copyright 2021, Elsevier.
Figure 1. Diagrammatic representation of the SnO2 QD synthesis process [34], Copyright 2021, Elsevier.
Catalysts 13 00765 g001
Figure 2. Colloidal synthesis of SnO2 QDs [35], Copyright 2020, ACS.
Figure 2. Colloidal synthesis of SnO2 QDs [35], Copyright 2020, ACS.
Catalysts 13 00765 g002
Figure 3. Synthesis and photocatalytic application of SnO2 QDs [41], Copyright 2018, Elsevier.
Figure 3. Synthesis and photocatalytic application of SnO2 QDs [41], Copyright 2018, Elsevier.
Catalysts 13 00765 g003
Figure 4. Photocatalytic mechanism of pure SnO2 QDs and Nb-SnO2 QDs [45], Copyright 2022, Elsevier.
Figure 4. Photocatalytic mechanism of pure SnO2 QDs and Nb-SnO2 QDs [45], Copyright 2022, Elsevier.
Catalysts 13 00765 g004
Figure 5. (a) CdS@SnO2 core-shell NRs’ photocatalytic degradation and (b) charge-transfer mechanism [40], Copyright 2019, Elsevier.
Figure 5. (a) CdS@SnO2 core-shell NRs’ photocatalytic degradation and (b) charge-transfer mechanism [40], Copyright 2019, Elsevier.
Catalysts 13 00765 g005
Figure 6. TEM images of SnO2 QDs/rGO (a,b), pure SnO2 QDs (c) [71], Copyright 2019, Elsevier.
Figure 6. TEM images of SnO2 QDs/rGO (a,b), pure SnO2 QDs (c) [71], Copyright 2019, Elsevier.
Catalysts 13 00765 g006
Figure 7. Photocatalytic mechanism of the ternary photocatalyst (SnO2/AgVO3/g-C3N4) [24], Copyright 2019, Springer.
Figure 7. Photocatalytic mechanism of the ternary photocatalyst (SnO2/AgVO3/g-C3N4) [24], Copyright 2019, Springer.
Catalysts 13 00765 g007
Table 1. SnO2 quantum dot synthesis techniques.
Table 1. SnO2 quantum dot synthesis techniques.
Synthesis ApproachSn PrecursorProcessing TemperatureReaction TimeRef
HydrothermalSnCl4·5H2O150 °C6 h[37]
HydrothermalSnCl4·5H2O180 °C2 h[34]
HydrothermalSnCl2140 °C4 h[25]
HydrothermalSnCl4·5H2O160 °C8 h[44]
HydrothermalSnCl2·2H2O160 °C6 h[45]
HydrothermalNa2SnO3·3H2O180 °C24 h[46]
MicrowaveSnCl2·2H2O180 °C30 min[32]
MicrowaveSnCl4·5H2O120 °C15 min[33]
MicrowaveSnCl2·2H2O110 °C6 h[26]
MicrowaveSnCl4·5H2O160 °C20 min[47]
Wet chemicalSnCl4·5H2O100 °C18 h[39]
Wet chemicalSnCl2·2H2ORT6 h[22]
RefluxSnCl2·2H2O65 °C3 h[42]
HummersSnCl2·2H2O180 °C24 h[35]
Solution processingSnCl2·2H2ORT48 h[27]
Solution combustionSnCl4·5H2O450 °C-[48]
Self-assemblySnCl2·2H2ORT24 h[11]
One pot BiosynthesisSnCl2·2H2O60 °C12 h[43]
In situSnCl4·5H2O100 °C18 h[49]
Table 2. Detailed description of recent research on the photocatalytic abilities of SnO2 QD-based nanocomposites.
Table 2. Detailed description of recent research on the photocatalytic abilities of SnO2 QD-based nanocomposites.
PhotocatalystBandgapPollutantDosageLight SourceEfficiencyStabilityRef
Au/SnO2 QDs2.71 eVRhB15 mg/50 mLVisible light98.7% in 200 min-[39]
Oxy-SnO2 QDs4.2 eVOctane-UV–visible91.9% in 48 h90% after 4 cycles[11]
Nb-SnO2 QDs2.95 eVPolyethylene-Visible light29% in 6 h-[51]
Ni-doped SnO2 QDs3.69 eVRhB-Solar light 91.5% in 28 min-[44]
Ag doped SnO2 QDs3.09 eVMB10 mg/10 mLVisible light93.7% in 120 min7 cycles[33]
NiFe2O4/SnO21.754 eVRhB20 mg/50 mLVisible light98% in 105 min93% after 3 cycles[56]
Nb-SnO2 QDs4.74 eVMO30 mg/LVisible light99.7% in 180 min3 cycles[45]
Bio-SnO2 QDs3.66 eVRhB0.25 g/LUV light 65% in 180 min-[43]
SnO2 QDs/Ag3PO4-Carbamazepine-Visible light86.5% in 120 min66.6% in 3 cycles[58]
Bi-doped SnO23.75 eVRhB25 mgSunlight98.28% in 100 min85.79% after 5 cycles[46]
Ag-SnO2 QDs2.54 eVRhB75 mg/100 mLDirect sunlight98% in 180 min4 cycles[49]
Mn-doped SnO2 QDs1.74 eVMO50 mg/100 mLVisible light 92% in 240 min-[48]
SnO2 QDs4.5 eVEosin Y10 mg/200 mLDirect sunlight98% in 60 min-[59]
Cr-doped SnO2 QDs4.35 eVMO100mg/100mLUV light 98.9% in 100 min-[60]
Bio-SnO2 QDs4.17 eVAcid yellow 2320 mgUV light98% in 24 min5 cycles[61]
SnO2-CNF3.0 eVBisphenol A50 mg /100 mLUV light 98% in 60 min84% after 3 cycles[62]
Cu-doped SnO2 QDs2.4 eVMO100 mg/100 mLVisible light 99% in 180 min-[63]
Table 3. Detailed description of recent research on the photocatalytic abilities of SnO2 QD-based binary nanocomposites.
Table 3. Detailed description of recent research on the photocatalytic abilities of SnO2 QD-based binary nanocomposites.
Photocatalyst.SynthesisPollutantDosageLight SourceEfficiencyStabilityRef
CdS/SnO2SonochemicalRhB-Visible light99% in 60 min95% after 3 cycles [40]
GO/SnO2SonochemicalMB0.5 mg/mL 250 ppmWhite light ~94% in 30 min-[67]
SnO2/SiO2Wet ChemicalMB50 mg/80mL of 10 ppm-~100% in 5 min4 cycles[74]
SnO2/g-C3N4SolvothermalMO-Sunlight94% in 180 min91% after 5 cycles[41]
TiO2/SnO2HydrothermalMO100 mg/100 mL of 10 ppmUV-visible light99.5% in 15 min-[75]
SnO2/GOSol–gelRhB-Visible light86% in 360 min-[76]
Ni-SnO2/SnS2Wet chemicalMO10 mg /50 mL of 10 ppmsolar light92.7% in 80 min88% after 3 cycles [77]
g-C3N4/SnO2HydrothermalNO200mg/15 mL of 100 ppmVisible light44.17% in 30 min30% after 5 cycles [78]
SnO2/g-C3N4SonochemicalNO400mg/15 mL of 600 ppmVisible light 32% in 30 min~20% after 3 cycles [69]
SnO2-GOSolvothermalMB30 mg /20 mL of 10 ppmVisible light 89.43% in 90 min-[79]
Table 4. Detailed description of recent research on the photocatalytic abilities of SnO2 QD-based ternary nanocomposites.
Table 4. Detailed description of recent research on the photocatalytic abilities of SnO2 QD-based ternary nanocomposites.
PhotocatalystSynthesisPollutantDosageLight SourceEfficiencyStabilityRef
g-C3N4/Au-SnO2SonochemicalRhB15 mg/50 mLVisible light99.15% in 40 min93.45% after 4 cycles[82]
Au-SnO2-rGOHydrothermalclothianidin- 97% in 120 min93% after 4 cycles [84]
CQDs/SnO2−x/BiOISonochemicalMO-LED light91.8% in 75 min-[85]
SnO2/AgVO3/g-C3N4HydrothermalRhB5 mg/100 mL of 5 ppmVisible light 84% in 50 min~82% after 5 cycles[24]
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

Bathula, B.; Gurugubelli, T.R.; Yoo, J.; Yoo, K. Recent Progress in the Use of SnO2 Quantum Dots: From Synthesis to Photocatalytic Applications. Catalysts 2023, 13, 765. https://doi.org/10.3390/catal13040765

AMA Style

Bathula B, Gurugubelli TR, Yoo J, Yoo K. Recent Progress in the Use of SnO2 Quantum Dots: From Synthesis to Photocatalytic Applications. Catalysts. 2023; 13(4):765. https://doi.org/10.3390/catal13040765

Chicago/Turabian Style

Bathula, Babu, Thirumala Rao Gurugubelli, Jihyung Yoo, and Kisoo Yoo. 2023. "Recent Progress in the Use of SnO2 Quantum Dots: From Synthesis to Photocatalytic Applications" Catalysts 13, no. 4: 765. https://doi.org/10.3390/catal13040765

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