5,412
Views
29
CrossRef citations to date
0
Altmetric
Miscellany

Highlights on advances in SnO2 quantum dots: insights into synthesis strategies, modifications and applications

, , , , , , & show all
Pages 462-488 | Received 30 Jan 2018, Published online: 25 Jun 2018

ABSTRACT

The applications of SnO2 are benefited from its nanostructure with different sizes and novel morphologies. When the size of nanoparticles reduces to 1–10 nm, the unique physical and chemical properties will make prominent. SnO2 quantum dots (QDs), a type of zero-dimensional ultrasmall SnO2 nanomaterials with a size in 1–10 nm, have displayed unique physical and chemical properties, which are different from those of their larger-sized ones. This review summarizes various synthesis strategies of SnO2 QDs and the methods of their modifications, discusses their applications in lithium-ion batteries, photocatalysis, and gas sensors. These applications profit from the characteristic properties inherent in SnO2 QDs.

GRAPHICAL ABSTRACT

IMPACT STATEMENT

This paper provides a comprehensive understanding for fabricating SnO2 quantum dots and their modifications via various methods for the applications in lithium-ion batteries, photocatalytic degradations, and solid-state gas sensors.

1. Introduction

Tin dioxide (SnO2), a kind of n-type semiconductor with wide-band-gap (Eg = 3.64 eV at 300 K), has very wide applications in the fields of gas sensors [Citation1–10], lithium-ion batteries (LIBs) [Citation11,Citation12], photocatalytic degradations [Citation13,Citation14] and dye-sensitized solar cells (DSSCs) [Citation15–19], photodetectors [Citation20] and heterojunction diode [Citation21]. The wide applications of SnO2 are greatly benefited from the synthesis methods in its nanostructured materials with different sizes and novel morphologies. Nanostructured SnO2 with various morphologies, such as three-dimensional nanospheres [Citation22–24], two-dimensional nanofilms [Citation21,Citation25–28] or nanosheets [Citation29,Citation30], and one-dimensional nanowires [Citation20] or nanorods [Citation29–32] has been successfully fabricated and played an important role in their applications. When the size of the SnO2 nanoparticles reduces to 1–10 nm, the unique physical and chemical properties will make prominent due to quantum size effect [Citation33]. So, SnO2 quantum dots (SnO2 QDs), a type of zero-dimensional ultrasmall SnO2 nanomaterials with size in the 1–10 nm range, have attracted considerable attention owning to their unique physical and chemical properties which are different from those of their larger-sized ones. In recent years, many efforts have been devoted to the investigations on the synthesis of SnO2 QDs. The synthetic methods of SnO2 QDs mainly include hydrothermal and solvothermal synthesis, pulsed laser ablation approach, microwave assistant synthesis, and electron-beam irradiation. Applicable performances of nanomaterials rely on not only their sizes and nanostructures but also their components. In addtion to fabricating SnO2 in the scale of 1–10 nm, enhanced and optimized performances can also be obtained by modifying the SnO2 host materials. Recently, modifications of SnO2 QDs have become one of the most important research fields. The modifications are commonly conducted in two ways: (1) doping SnO2 QDs with other chemical elements; (2) or embedding SnO2 QDs in suitable matrix. The most important applications of SnO2 QDs and their modified ones are in the fields as follows: lithium-ion batteries, photocatalytic degradations, and gas sensors. In this paper, by focusing on the SnO2 QDs, we manage to facilitate the readers’ comprehensive understanding on the synthetic approaches, modification strategies and their potential applications of SnO2 QDs.

2. Synthesis strategies of SnO2 QDs

Various synthesis strategies have been employed to prepare SnO2 QDs. These synthesis strategies include the bottom-up methods (such as hydrothermal synthesis, solvothermal synthesis, microwave-assisted synthesis and electron-beam irradiation synthesis) and top-down method (including pulsed laser ablation decomposition). The following subsections will discuss in details the synthesis strategies commonly employed to fabricate SnO2 QDs.

2.1. Hydrothermal synthesis

Hydrothermal synthesis is a widely used way for the fabrication of nanoparticles, which is normally performed in Teflon-lined stainless steel autoclaves under a controlled temperature and certain pressure. The temperature can be raised above the boiling point of water, reaching the saturated vapor pressure. The interior pressure in autoclave is determined by the reaction temperature and the filling degree of solution in the autoclave. In the past few years, hydrothermal synthesis has been widely used for the fabrication of SnO2 QDs [Citation34–44]. The spherical SnO2 QDs have been obtained by the hydrothermal treatment of a aqueous solution containing certain amount SnCl4·5H2O and aqueous ammonia [Citation34]. For example, 2.8 g SnCl4·5H2O and suitable amount of aqueous ammonia were dissolved in deionized water, followed by ultrasonic vibration, and then transferred to a 120 ml Teflon-lined stainless steel autoclave. The stainless steel autoclave was heated at 200°C for 24 h. The obtained SnO2 QDs were almost spherical shape with diameters in the range of 2.6–7.8 nm. Besides spherical shape SnO2 QDs, cuboid SnO2 QDs have also been fabricated by similar method [Citation35,Citation40]. For example, cuboid SnO2 QDs were prepared by hydrothermal reaction of SnCl4·5H2O and CO(NH2)2 in an aqueous solution [Citation35]. 0.08 g SnCl4·5H2O, 0.8 g CO(NH2)2 and 1.6 mL HCl fume were dissolved in 32 mL deionized water. The above solution was then transferred into a stainless steel autoclave, followed by ultrasonic treatment and heating to 90°C for 15 h. The particles were cuboid shape with the widths of 4.0 nm and the lengths ranging from 6.0 to 13.5 nm.

During these hydrothermal processes, the sizes and morphologies of SnO2 QDs depend on the reaction conditions, including reaction temperature, duration of time and pH of solution. Zhang et al. [Citation36] have investigated the influences of reacting temperature and time on the size of spherical SnO2 QDs. In order to find growth kinetics of SnO2 QDs, the hydrothermal reactions were conducted in different temperatures (140–220°C) and varied times (2–200 h). As shown in Figure , the spherical SnO2 QDs with different particle sizes, varying from 2.2 to 7.0 nm, have been obtained under different reaction conditions.

Figure 1. TEM, HRTEM, SAED images, and the corresponding size distributions of SnO2 quantum dots prepared under different conditions: (a1—a4) 140°C, 2 h; (b1—b4) 200°C, 4 h; and (c1—c4) 200°C, 240 h. Reprinted with permission from Ref. [Citation36]. Copyright 2015 American Chemical Society.

Figure 1. TEM, HRTEM, SAED images, and the corresponding size distributions of SnO2 quantum dots prepared under different conditions: (a1—a4) 140°C, 2 h; (b1—b4) 200°C, 4 h; and (c1—c4) 200°C, 240 h. Reprinted with permission from Ref. [Citation36]. Copyright 2015 American Chemical Society.

Besides reaction temperature and reaction time, the pH value of the solution also has a vital influence on the size and morphology of SnO2 QDs. Several spherical and cuboid SnO2 QDs with different size have been prepared by hydrothermal reaction of SnCl4·5H2O with OH in aqueous solutions with different pH values [Citation37]. To hydrothermally fabricate SnO2 QDs with varied shapes and sizes, different volumes of 25% aqueous tetramethyl ammonium hydroxide [N(CH3)4OH; TMAH] solution were added into aqueous solutions of SnCl4·5H2O, and the pH values of the mixed solution were adjusted from 1.2 to 14. The volume of the mixed solutions was then scaled to 50 mL with pure water. After that, the mixed solutions were transferred to a Teflon-lined stainless steel autoclave, and then heated at 150°C for 24 h. It was found that the sizes and shapes of the resulted SnO2 QDs are dependent on pH values (Figure ). All the SnO2 QDs achieved at pH values between 7.3 and 13.7 exhibit narrow size distribution of 3.2∼3.6 nm. However, the SnO2 QDs prepared at pH < 10.7 were spherical. When pH was changed to 12.6, the shape of SnO2 QDs turned into cubes, while the similar size distribution was kept with those achieved from pH < 10.7. When pH value was increased further, only cubic SnO2 QDs with increasing size were obtained.

Figure 2. TEM images of the nanocrystals grown with TMAH at (a) pH 7.3, (b) pH 9.2, (c) pH 10.7, (d) pH 12.6, (e) pH 13.3, and (f) pH 13.7. Reprinted with permission from Ref. [Citation37]. Copyright 2013 American Chemical Society.

Figure 2. TEM images of the nanocrystals grown with TMAH at (a) pH 7.3, (b) pH 9.2, (c) pH 10.7, (d) pH 12.6, (e) pH 13.3, and (f) pH 13.7. Reprinted with permission from Ref. [Citation37]. Copyright 2013 American Chemical Society.

2.2. Solvothermal synthesis

Solvothermal synthesis is considered to be identical to hydrothermal method except that the solvents used in the solvothermal process are organic solvents, such as methanol, ethanol, oleylamine, and so on. Many articles presented that the SnO2 QDs can be successfully obtained by solvothermal synthesis [Citation45–49]. For instances, Xu et al. [Citation47] have prepared SnO2 QDs through solvothermal treatment of SnCl4·5H2O in a mixed organic solvent. In a typical process, 1.7 mmol SnCl4·5H2O was added into a mixed solvent of 2.5 mL oleylamine, 20 mL oleic acid and 10 mL ethanol to form a mixed solution. Then, the mixed solution was transferred to a stainless steel autoclave, and heated at 180°C for 30 min. In this process, SnO2 QDs with size of 0.5–2.5 nm were obtained. It was found that these SnO2 QDs can be further assembled into SnO2 nanowires by prolonging the reaction time.

Some organics, e.g. N-methylimidazole, oleylamine and oleic acid, can be used to inhibit the growth and aggregation of the nanoparticles because of their outstanding blocking effect. Monodispersed SnO2 QDs can be prepared under solvothermal conditions using this type of organics as reaction solvent. Chen et al. [Citation49] have developed a gram-scale fabrication of SnO2 QDs by using an N-methylimidazole-based solvothermal method. The SnO2 QDs are well-dispersed 4 nm nanoparticles. In the reaction process, N-methylimidazole not only facilitated the growth of the ultrasmall SnO2 QDs by providing an alkaline environment, but also inhibited the growth and aggregation of SnO2 QDs by capping the SnO2 QDs through the hydrogen bonds.

2.3. Microwave-assisted synthesis

As a high-speed synthesis approach, microwave-assisted synthesis has grabbed more attention owning to its unique advantages compared with conventional heating: (1) fast heating rate, (2) uniform heating without thermal gradients, (3) and without superheating of the solvents. Microwave-assisted synthesis has become a useful and rapid approach for the fabrication of nanoparticles with shortening reaction time from hours to minutes.

Recently, microwave-assisted synthetic technique has been applied to prepare SnO2 QDs [Citation17,Citation50–57]. For examples, Liu et al. [Citation57] have synthesized SnO2 QDs with diameters of 3–5 nm by microwave irradiating the solution of SnCl4·5H2O and alkaline. In this work, a dilute NaOH solution was dropped into 20 mL solution of 0.1 M SnCl4·5H2O until the pH value of the mixed solution reached 5. After the mixed solution was sonicated for 30 min, the solution was irradiated at 150°C for 10 min under a microwave irradiation power of 150 W. The synthesized SnO2 QDs were well uniform and dispersive with size ranging from 3 to 5 nm. In these conventional microwave-assisted methods, strong alkaline solutions, such as NaOH, KOH, H2NCH2CH2NH2, are widely used. However, these strong alkaline solutions are severely harmful to environment. Bhattacharjee et al. [Citation50,Citation54] have established a green, environment-friendly microwave-assisted synthesis of SnO2 QDs. In their works, strong alkaline solutions were replaced by biological molecules, namely, serine [Citation50] and sugar cane juice [Citation54]. For example, for the biosynthesis of SnO2 QDs, 0.01 M SnCl2·2H2O was treated with 10% sugar cane solution. The mixture was then placed into a microwave oven and irradiated for thirty 10 s shots. Figure (a,b) show the formation of highly uniformed and well-aligned SnO2 QDs. The size of SnO2 QDs depicted in the Figure (a,b) is 3–4.5 nm, consistent with the crystalline size calculated from their XRD data. The lattice planes depicted in Figure (c) is in good agreement with their XRD results of the SnO2 QDs. Figure (d) demonstrated the presence of Sn and O.

Figure 3. (a) TEM microphotograph of biosynthesized SnO2 QDs, (b) HRTEM image of biosynthesized SnO2 QDs, (c) SAED pattern of biosynthesized SnO2 QDs, (d) EDS spectrum of biosynthesized SnO2 QDs. Reprinted with permission from Ref. [Citation54]. Copyright 2015 the Royal Society of Chemistry.

Figure 3. (a) TEM microphotograph of biosynthesized SnO2 QDs, (b) HRTEM image of biosynthesized SnO2 QDs, (c) SAED pattern of biosynthesized SnO2 QDs, (d) EDS spectrum of biosynthesized SnO2 QDs. Reprinted with permission from Ref. [Citation54]. Copyright 2015 the Royal Society of Chemistry.

The SnO2 QDs with narrow size-distribution and better crystallinity can also be prepared via the reaction of tin precursors with nonaqueous solution. However, extreme operating conditions are often required, which makes it difficult for the large-scale synthesis of SnO2 QDs with an economic way. Recently, ionic liquids, one of nonaqueous media, have been widely applied in nano-synthesis due to their special advantages including high polarizability, low toxicity, recyclability and high thermal stability. Ionic liquids are an excellent nonaqueous medium for the absorption of microwave due to their high polarizability. Xiao et al. [Citation52] have integrated the advantages of ionic liquid and microwave irradiation to control nucleation of SnO2 QDs. In their work, highly crystalline and monodisperse SnO2 QDs were prepared by dissolving Sn(OtBu)4 in dried [BMIM]BF4 (1-butyl-3-methylimidazoliumtetrafluoroborate) under Ar atmosphere, followed by microwave decomposition of Sn(OtBu)4 for 1 min. For comparison, another sample was prepared by oil bath heating the same precursor for 1 h. Figure (a) shows that the SnO2 QDs synthesized by the microwave decomposition of Sn(OtBu)4 in [BMIM] BF4 have an obviously better crystallinity than that of the SnO2 QDs synthesized by oil bath. Figure (b) exhibits the SnO2 QDs synthesized by the microwave decomposition were nearly monodispersed with a narrow size distribution (4.27 ± 0.67 nm). It can be concluded that the microwave could initiate the reaction much faster than the oil bath method and might offer an additional energy for the crystal formation of SnO2 QDs. Compared with the convenient hydrothermal and solvothermal method, the microwave-assisted method is more effective in time and energy saving, facile composition, size control.

Figure 4. X-Ray powder diffraction patterns (a) of as-synthesized SnO2 (i) by MW irradiation, (ii) by oil bath heating and TGA trace of the precursor (inset); HR-TEM image (b) of as-synthesized SnO2 QDs and electron diffraction pattern (inset). Reprinted with permission from Ref. [Citation52]. Copyright 2010 The Royal Society of Chemistry.

Figure 4. X-Ray powder diffraction patterns (a) of as-synthesized SnO2 (i) by MW irradiation, (ii) by oil bath heating and TGA trace of the precursor (inset); HR-TEM image (b) of as-synthesized SnO2 QDs and electron diffraction pattern (inset). Reprinted with permission from Ref. [Citation52]. Copyright 2010 The Royal Society of Chemistry.

2.4. Pulsed laser ablation decomposition

The above summarized hydrothermal, solvothermal and microwave-assisted approaches are ‘bottom-up’ method for the fabrication of SnO2 QDs. Pulsed laser ablation of starting material, as a ‘top-down’ strategy, is a versatile method to produce nanoparticles. Using Sn or SnO as starting materials, some research groups have fabricated SnO2 QDs by pulsed laser ablation of Sn or SnO [Citation40, Citation58–62]. Singh et al. [Citation58] have synthesized SnO2 QDs with the average diameter of 2.5 nm by laser ablation of metal Sn in water. In detail, tin pellet was put into a glass vessel containing 20 mL of deionized water. The mixture of Sn and water was then irradiated for one hour by the 1064 nm wavelength Nd:YAG laser beam (35 mJ, 10 ns, 10 Hz). The fabricated SnO2 QDs are ultrafine nanoparticles with a diameter from 1 to 5 nm.

The starting material with a low boiling-point is usually considered to be a better material to generate tiny nanoparticles in the pulsed laser ablation process. SnO is regard as a better precursor for the formation of SnO2 QDs, because of its lower boiling-point (1527°C) compared to that of Sn (2602°C). Some efforts for the fabrication of SnO2 QDs by using SnO as precursor have been made [Citation59,Citation60,Citation62]. The schematic illustration of the formation of the SnO2 QDs is presented in Figure (a). SnO powders absorb the laser power and break down into tiny figments of SnO, the SnO nanoparticles are then transformed into SnO2 QDs via the reaction SnO + H2O → SnO2 + H2.

Figure 5. (a) Schematic illustration of the formation of the SnO2 QDs using SnO as precursors by the pulsed laser ablation process; (b) TEM image, (c) histogram of diameter distribution, and (d) HRTEM of the SnO2 QDs in water; (f) Photos of the mixture of SnO powder and H2O (left), the mixture after laser fragmentation (middle) and the mixture of SnO2 QDs and water after centrifugation (right). Reprinted with permission from Ref. [Citation59]. Copyright 2010 the Royal Society of Chemistry.

Figure 5. (a) Schematic illustration of the formation of the SnO2 QDs using SnO as precursors by the pulsed laser ablation process; (b) TEM image, (c) histogram of diameter distribution, and (d) HRTEM of the SnO2 QDs in water; (f) Photos of the mixture of SnO powder and H2O (left), the mixture after laser fragmentation (middle) and the mixture of SnO2 QDs and water after centrifugation (right). Reprinted with permission from Ref. [Citation59]. Copyright 2010 the Royal Society of Chemistry.

Pan et al. [Citation59] have reported the synthesis of SnO2 QDs by laser ablation method using SnO powder as material. First, 60 mg SnO was added into 20 mL deionized water. Second, the stirring mixture was irradiated by a pulsed laser beam (6 ns, 10 Hz, 1064 nm) for 30 min. Figure (b,c) show that the fabricated SnO2 QDs have uniform diameter of 1.8 ± 0.3 nm. Figure (d) shows the lattice spacing of the SnO2 QDs is 0.212 nm, which is related to the (210) crystal plane and consistent with the rutile SnO2 phase data. Figure (e) described the transformation from SnO to SnO2 QDs. It is found that using SnO powder as precursor can produce SnO2 QDs with better size uniformity and smaller particle size compared to widely employed Sn powder. Pan et al. [Citation60] have also investigated the influences of the laser irradiation time on preparation of SnO2 QDs. With increasing laser irradiation time, the average size of the prepared SnO2 QDs remarkably increases along with a wider size distribution.

2.5. Electron-beam irradiation synthesis

The electron-beam irradiation synthesis has many highly advantageous properties as follows: (i) this method is rapid, simple, and convenient, (ii) the method can be carried out at room temperature without catalysts, and (iii) this strategy is useful for mass production of nanomaterials. Our group has performed systematic investigations on the fabrication of SnO2 QDs by electron-beam irradiation strategy [Citation63,Citation64]. For example, we have prepared a serial of SnO2 QDs by using electron-beam irradiation and evaluated their microstructure evolution. The primal SnO2 QDs samples were prepared by adding dropwise 2 mol/L aqueous ammonia into a 0.2 mol/L SnCl4 solution under stirring until the pH value of the mixed solution reached 7. The as-prepared SnO2 samples were spreaded to a thickness of 2–3 mm, which was irradiated by a GJ-2-II dynamitron electron accelerator with an accelerating voltage of 2 MeV and a current of 8 mA at 700, 980, 1260, and 1400 kGy, respectively. Figure (a) shows that the SnO2 QDs irradiated at 1400 kGy have a better crystallinity compared to that of the unirradiated one. In addition, it was found that the surface area of the irradiated SnO2 QDs is higher than that of unirradiated one as shown in Figure (b). When the irradiated dose increased to 700 kGy, the BET surface area of the SnO2 QDs increased from 105.39 to 170.47 m2/g. However, the BET value has slightly decreased to 165.42 m2/g at 980 kGy radiation dose. With increasing irradiated doses to 1260 and 1400 kGy, the BET values increased. It was reasonable to speculate that the electron-beam irradiation is favorable for the formation of more SnO2 crystal nucleus in SnO2 QDs. This speculation can also be proved by the HRTEM results as shown in Figure (c,d). The research results reveal that the electron-beam irradiation is a potentially powerful technique to achieve SnO2 nucleation and QD growth.

Figure 6. Typical XRD patterns (a) of SnO2 QDs: (i) unirradiated SnO2 QDs, (ii) irradiated SnO2 QDs which was irradiated at 1400 kGy; Specific surface area (b) of unirradiated SnO2 QDs and irradiated SnO2 QDs at 700, 980, 1260, and 1400 kGy, respectively. HRTEM images of (c) unirradiated SnO2 QDs, (d) irradiated SnO2 QDs which was irradiated at 1400 kGy. Scale bar = 5 nm. Reprinted with permission from Ref. [Citation63]. Copyright 2011 American Chemical Society.

Figure 6. Typical XRD patterns (a) of SnO2 QDs: (i) unirradiated SnO2 QDs, (ii) irradiated SnO2 QDs which was irradiated at 1400 kGy; Specific surface area (b) of unirradiated SnO2 QDs and irradiated SnO2 QDs at 700, 980, 1260, and 1400 kGy, respectively. HRTEM images of (c) unirradiated SnO2 QDs, (d) irradiated SnO2 QDs which was irradiated at 1400 kGy. Scale bar = 5 nm. Reprinted with permission from Ref. [Citation63]. Copyright 2011 American Chemical Society.

2.6. Other synthesis strategies

In addition to those above mentioned commonly used methods to prepare SnO2 QDs, there are some other ones, such as sonication [Citation65], solution combustion [Citation66], interfacial reaction [Citation67], liquid phase refluxing [Citation68], grinding [Citation69]. Lee et al. [Citation67] have fabricated a serials of SnO2 QDs with average diameters from 2 to 2.7 nm, energy band gaps between 5.70 and 4.39 eV through the interfacial reaction of Sn4+ and OH- in interface between water and chloroform. Kamble et al. [Citation66] have synthesized the SnO2 QDs with average crystallite size of 7 nm by heating tin nitrate solution with fuel (urea) at 500°C. Kida et al. [Citation68] have prepared monodispersed SnO2 QDs with size of 3.5 nm by refluxing tin (IV) acetylacetonate in dibenzyl ether at 280°C under the aid of oleylamine and oleic acid. Cui et al. [Citation69] have produced SnO2 QDs with particle size of about 1.2 nm at an ultra large scale by grinding the solid mixture of SnCl2·H2O, ammonium persulphate and morpholine in a mortar at room temperature. Although these methods have the uniquely attractive advantages of facility, such as large scale, room temperature, and surfactant-free, the explorations of these approaches have just begun and thus will be a subject of future studies.

SnO2 QDs have been successfully fabricated via a variety of methods mentioned above. Being conducted in different conditions, these synthesis strategies have their own advantages and disadvantages (). These methods usually require harsh conditions, complex procedures and/or special devices, high cost, time-consuming and/or environmental pollution. It is necessary to further explore facile, room temperature, environment-friendly approaches for the fabrication of SnO2 QDs. In addition, these methods are normally low yield. In order to facilitate their practical application, it is of great importance to develop large scale synthesis of SnO2 QDs.

Table 1. Advantages and disadvantages of synthesis strategies for the fabrication of SnO2 QDs.

3. Modified SnO2 QDs

3.1. Doped SnO2 QDs

Many applications of SnO2 QDs are related to their physical properties, such as optical, electrical and magnetic properties. To extend their applications, the enhancement of these properties of SnO2 QDs is very important. Doping the pure SnO2 QDs with extrinsic dopants is a facile and effective way to promote their physical properties. In the doping process, the stannum or oxygen component in the SnO2 QDs can be partially replaced to modify the physical properties of SnO2 QDs. Recently, some efforts have been devoted to the doped SnO2 QDs.

3.1.1. Metal-doped SnO2 QDs

Some metals, including transition and rare earth metals, have been successfully doped into SnO2 QDs [Citation18,Citation70–83]. The fabrications of metal-doped SnO2 QDs are very similar to those of pure SnO2 QDs, which can also be divided into the following types: hydrothermal, solvothermal, sol–gel method and microwave-assisted synthesis strategies, etc.

Some transition metals, such as Ti [Citation83], Mn [Citation77], Pd [Citation70], Cu [Citation84], Cr [Citation85] and Zn [Citation18], have also been successfully doped into SnO2 QDs. Sakthiraj et al. [Citation83] have performed a systematic study of the SnO2 QDs, which were doped with different amount of titanium by sol–gel method, and investigated their room temperature ferromagnetism. Figure  shows that different amount of titanium ion dopants significantly influences the room temperature ferromagnetism (RTFM) of the SnO2 QDs. The undoped and 2% titanium doped SnO2 QDs (i.e. undoped SnO2 and Sn0.98Ti0.02O2) show perfect RTFM, but the 5% and 7% of titanium doped SnO2 QDs (namely Sn0.95Ti0.05O2 and Sn0.93Ti0.07O2) exhibit a weak ferromagnetism with diamagnetic contribution. It is obvious that the RTFM of the doped SnO2 becomes weak with the increase of titanium dopants. The researchers believed that the amount of oxygen vacancies in SnO2 QDs greatly influence the ferromagnetic property of the samples. Li et al. [Citation18] have synthesized undoped and zinc-doped SnO2 QDs by a simple hydrothermal method. The doping with zinc into SnO2 induced a negative shift in the flat-band potential and increased the isoelectric point. As a result, the dye-sensitized solar cells based on the zinc-doped SnO2 QDs showed higher dye loading and longer electron lifetimes compared to that based on pure SnO2 QDs. Sabergharesou et al. [Citation77] have reported the sub-3 nm diameter manganese-doped SnO2 QDs, synthesized from SnCl4·5H2O and MnCl2 precursors. The manganese ions in the SnO2 QDs exhibit two different oxidation states (i.e. Mn2+ and Mn3+); Mn2+ species dominated at low doping levels, however, the fraction of Mn3+ species increased with doping concentrations. The electronic structure and magnetic properties of the samples were also studied in detail. Both the electronic structure and magnetic properties of the SnO2 QDs were primarily determined by the density ratio of the two different Mn oxidation states.

Figure 7. M-H magnetic hysteresis loop of (a) undoped SnO2 nanoparticles, (b) 2% Ti doped sample, (c) 5% Ti doped sample and (d) 7% Ti doped sample. Inset of the figure shows the focused view of 5% and 7% Ti doped samples. Reprinted with permission from Ref. [Citation83]. Copyright 2015 Elsevier.

Figure 7. M-H magnetic hysteresis loop of (a) undoped SnO2 nanoparticles, (b) 2% Ti doped sample, (c) 5% Ti doped sample and (d) 7% Ti doped sample. Inset of the figure shows the focused view of 5% and 7% Ti doped samples. Reprinted with permission from Ref. [Citation83]. Copyright 2015 Elsevier.

Several kind of SnO2 QDs doped with rare earth metals, such as Eu [Citation72], Cs [Citation75] and Gd [Citation80], have also been fabricated by some research groups. Using microwave synthesis method, Patria et al. [Citation72] have prepared the pure and Eu3+ doped SnO2 QDs and studied their optical and electrical properties. Figure  displays the absorption spectra of 1.0 mol% Eu-doped SnO2 QDs annealed at different temperatures. With the increasing temperatures, the quantum size effect results in the red shift in the absorption edge. The doped SnO2 QDs showed higher conductivity than that of pure SnO2 QDs. Using oleic acid as a stabilizing agent, Yang et al. [Citation80] have prepared pure SnO2 QDs and doped the SnO2 QDs with Gd ions via a solvothermal method. The pure SnO2 QDs with quasi-sphere morphology can be converted to dope SnO2 QDs with nanorods shape due to Gd3+ doping. The absorption edge of the SnO2 QDs exhibited an obvious blue shift with increasing Gd3+ dopant concentration. They attributed the blue shift to size effect induced by the reduced average size of the Gd3+ doped SnO2 QDs.

Figure 8. UV-vis spectra of 1.0 mol % Eu3+-doped SnO2 QDs at different temperatures, (a) 100°C, (b) 200°C, (c) 400°C and (d) 800°C. Inset shows (R) vs. photon energy plot of pure and Eu3+-doped SnO2 nanocrystals. Reprinted with permission from Ref. [Citation72]. Copyright 2009 American Chemical Society.

Figure 8. UV-vis spectra of 1.0 mol % Eu3+-doped SnO2 QDs at different temperatures, (a) 100°C, (b) 200°C, (c) 400°C and (d) 800°C. Inset shows (Rhυ) vs. photon energy plot of pure and Eu3+-doped SnO2 nanocrystals. Reprinted with permission from Ref. [Citation72]. Copyright 2009 American Chemical Society.

Some other metals, such as Sb [Citation71], Mg [Citation78], and Sn [Citation81] have also been doped into SnO2 QDs. For example, Fan et al. [Citation81] have fabricated a series of Sn2+ self-doped SnO2-x QDs (with different molar ratios (r) of Sn/SnCl4, r = 1:16, 1:8, 1:4, and 1:2, respectively) via a facile one-pot temperature synproportionation reaction, and also studied the self-doping effects on the particle size and band gap modification. The particle size of Sn2+ self-doped SnO2-x QDs decreased a little with the increase in the molar ratios (r) of Sn/SnCl4. The size of Sn2+ self-doped SnO2-x QDs (r = 1:4) was about 5 nm. With the increase of tin powder proportion, the absorption edges for the different Sn2+ self-doped SnO2-x QDs were red-shifted relative to stoichiometric SnO2 QDs. This self-doping effect on these absorption edges was attributed to the incorporation of Sn2+ into the SnO2-x lattice and accompanying oxygen vacancies, which can lead to significant narrowing of the band gap.

3.1.2. Nonmetal-doped SnO2 QDs

It is found that two nonmetal elements, namely nitrogen [Citation79], fluorine [Citation73,Citation74,Citation76], have also been doped into SnO2 QDs. For instance, Zhou et al. [Citation79] have fabricated nitrogen-doped SnO2 QDs with sizes of 2.2, 5.4 nm by continuously stirring the solution of tin powers, nitric acid and deionized water for 40 h, followed by annealing in O2 atmosphere at 500, 700°C, respectively. Compared with bare SnO2 QDs, the PL peaks of nitrogen-doped SnO2 QDs are blueshifted with the increasing of annealing temperature from 500 to 700°C. Infrared and Raman spectra showed the existence of local disorder and oxygen vacancies in the samples. Spectral analysis and theoretical calculation suggested that the PL band of the nitrogen-doped SnO2 QDs originates from the mutual effects of nitrogen dopants and oxygen vacancies. Wu et al. [Citation73] have prepared fluorine-doped SnO2 QDs by a sol–gel procedure accompanied by a hydrothermal process using NH4F as fluorine dopant. The precursor of SnO2 QDs was obtained by refluxing the precipitation of Sn(OH)4 in oxalic acid at 100°C for 4 h. Subsequently, the resulted solution was added with NH4F, and hydrothermally treated at 180°C for 72 h. The electrical resistivity properties of the fluorine-doped SnO2 QDs are dependent on the NH4F/Sn molar ratio. The sheet resistances of fluorine-doped SnO2 QDs decline with improving the NH4F/Sn molar ratio from 0 to 2. However, the sheet resistances of fluorine-doped SnO2 QDs increase with further improving NH4F/Sn molar ratio from 2 to 5. When NH4F/Sn molar ratio is 2, the sheet resistances of fluorine-doped SnO2 QDs is 110 Ω. Nagarajan et al. [Citation74,Citation76] have successfully fabricated heavily fluorine-doped (up to 21 mol%) SnO2 QDs using air-stable KSnF3 as single-source precursor and investigated their optical and pohotocatalytic properties. A blue shift of the exciton absorption, estimated from the Kubelka–Munk function, was observed from 3.52 to 3.87 eV in the fluorine-doped SnO2 QDs. The fluorine-doped SnO2 QDs displayed a broad green emission arising from the singly ionized oxygen vacancies created by high dopant concentration. One year later, Nagarajan et al. [Citation76] have also prepared heavily F-doped SnO2 QDs by a similar method and studied their thermoluminescence property. As shown in Figure , an intense broad thermoluminescence is detected in the fluorine-doped SnO2 QDs in the temperature range of 350–550 K, signifying the presence of trapped states, but no glow peak is observed for the pure SnO2 QDs. These results indicate that it is easier to substitute the Sn4+ in SnO2 with extrinsic metals, and it is more difficult to replace the O2- with other anions which may be due to their differences in ionic radii and charge states.

Figure 9. Thermoluminescence glow curve of (a) F-doped SnO2 QDs and (b) pure SnO2 QDs. Reprinted with permission from Ref. [Citation76]. Copyright 2012 Elsevier.

Figure 9. Thermoluminescence glow curve of (a) F-doped SnO2 QDs and (b) pure SnO2 QDs. Reprinted with permission from Ref. [Citation76]. Copyright 2012 Elsevier.

3.2. SnO2 QDs embedded nanocomposites

Nanocomposites can achieve advantageous optical, electronic, magnetic and mechanic properties due to their potential to combine desirable properties of different components. In recent years, many efforts on SnO2 QDs have been focused on the preparation of SnO2 QDs embedded nanocomposites, including SnO2 QDs embedded reduced graphene oxide (SnO2 QDs/RGO), SnO2 QDs embedded carbon nanotubes (SnO2 QDs/CNTs), SnO2 QDs embedded amorphous carbon (SnO2 QDs/AC), and SnO2 QDs embedded metal oxide, and so on.

3.2.1. SnO2 QDs/RGO

3.2.1.1. Strategy for embedding SnO2 QDs into graphene

Graphene is a single atomic layer of sp2 hybridized carbon atoms arranged in a honeycomb lattice. It has attracted widespread attention because of their large specific surface area, outstanding mechanical flexibility, and excellent optical, electronic and electrochemical properties. Embedding SnO2 QDs into graphene matrix is a facile, efficient way to enhance their physical and chemical properties. An increasing amount of effort is being dedicated to embedding SnO2 QDs into graphene by using various methods, such as hydrothermal method [Citation86–96], refluxing method [Citation97,Citation98], microwave-assisted approaches [Citation99], ultrasonic method [Citation100].

At present, the hydrothermal method is one of the most commonly used approaches to assembly SnO2 QDs onto graphene. Mishra et al. [Citation91] have anchored the SnO2 QDs with a size less than 6 nm on the surface of RGO by a surfactant assisted hydrothermal method. In this process, firstly, RGO (1.0 mg) was added to a transparent colorless aqueous solution containing SnCl4·5H2O (10 mmol). Subsequently, hexamethyldisilazane (1 mL) was added to the above solution with mild stirring. Next, the pH of the solution was adjusted to 8.5 by adding a NaOH aqueous solution. The resulted solution was transferred to a teflon-lined stainless steel autoclave (100 mL) and treated at 120°C for 20 h. The structural characterizations confirmed that the SnO2 QDs with a size less than 6 nm have been anchored on the surface of RGO. Using SnCl4·5H2O or SnCl2·2H2O as precursor, Li et al. [Citation92] have respectively prepared two type of SnO2 QDs/RGO aerogels by hydrothermally treating the solution containing graphene oxide and precursor at 180°C for 16 h, and then freeze-dried the obtained samples under −50°C. The SnO2 QDs/RGO aerogel prepared from SnCl4·5H2O showed a 3D column-like freestanding structure. However, the SnO2 QDs/RGO aerogel prepared from SnCl2·2H2O exhibited a fragile structure. The aerogel prepared from SnCl4·5H2O has plentiful nanopores and thus exhibited extremely large surface area (441.9 m2/g), which could be beneficial for its applications. Li et al. [Citation94] have also successfully planted the well dispersed SnO2 QDs with size of 3–5 nm into nitrogen-doped graphene (containing 18 at% N atoms) via hydrothermally treating the homogeneous mixture of graphene oxide, SnCl2·2H2O, dicyandiamide and sulfourea. In the hydrothermally treating process, graphene oxide was modified by dicyandiamide to produce N-doped graphene, which was then planted with the SnO2 QDs derived from SnCl2·2H2O. Here, the sulfourea was applied to decrease the size of the SnO2 QDs, simultaneously enhance their uniformity, and also facilitate the combination of the SnO2 QDs with nitrogen-doped graphene.

Compared with the convenient hydrothermal method, the microwave-assisted method for the fabrication of SnO2 QDs planted-RGO is more effective in energy and time saving. Zhou et al. [Citation99] have densely anchored the SnO2 QDs with an average size of about 3 nm on N-doped graphene by radiating a solution containing 60 mg graphite oxide, 1 g urea, 0.3 g SnCl4·5H2O and 10 mL ethylene glycol by microwave at 180°C for 5 min. In this process, Sn4+ ions were firstly anchored into graphene oxide because of electrostatic attraction. Subsequently, the urea acted as an N source for doping graphene and the synthesis of Sn(OH)4 was facilitated by ammonia which was released from the decomposition of urea. Then, ethylene glycol reduced graphite oxide to RGO.

The above mentioned approach for embedding SnO2 QDs into graphene is usually time-consuming, often carried out in high pressure and temperature and suffered from poor manipulation. Therefore, it is meaningful to develop a novel facile route to obtain SnO2 QDs embedded-RGO with favored microstructures and performances. By using a facile one-step ultrasonication, our group has assembled SnO2 QDs into graphene nanosheets at room temperature [Citation100]. The mechanism of planting SnO2 QDs into graphene nanosheets is illustrated in Figure . In a typical synthesis, 100 mL GO solution was dispersed in 50 mL distilled water to form a suspension. Then, 10 mL aqueous solution of 0.45 g SnCl2·2H2O was added into the above suspension and mixed with the suspension. The resulting mixture was treated by ultrasonic wave for 60 min at ambient temperature. We have discovered that the SnO2 QDs were uniformly dispersed on both sides of the graphene. The size of SnO2 QDs ranged from 4 to 6 nm and their average size was about 4.8 ± 0.2 nm. In this ultrasonic method, the loading of SnO2 QDs was an effective approach to prevent graphene nanosheets from self-restacking during the reduction. On the other hand, the graphene nanosheets distributed between SnO2 QDs have also prevented the SnO2 QDs from agglomerating.

Figure 10. Schematic formation mechanism of SnO2 QDs/GNS composites: (a) GO, (b) electrostatic interaction between oxide functional groups of GO and Sn2+, (c) graphene decorated with SnO2 QDs (filled circles) after the calcine treatment, and (d) TEM image of SQDs/GNS composites. Reprinted with permission from Ref. [Citation100]. Copyright 2013 American Chemical Society.

Figure 10. Schematic formation mechanism of SnO2 QDs/GNS composites: (a) GO, (b) electrostatic interaction between oxide functional groups of GO and Sn2+, (c) graphene decorated with SnO2 QDs (filled circles) after the calcine treatment, and (d) TEM image of SQDs/GNS composites. Reprinted with permission from Ref. [Citation100]. Copyright 2013 American Chemical Society.

3.2.1.2. Microstructural evolution of SnO2 QDs embedded into graphene

The size, morphology and microstructure evolution of SnO2 QDs on graphene greatly affect their physical and chemical properties as well as their applications in many fields. For instance, the decrease in the size of the particle will lead to increase in its band gap and surface-volume ratio. This effect will finally result in the improvement of the sensitivity of gas sensors. In order to study the microstructural evolution and enhance the performance of SnO2 QDs anchored on graphene, our group has intentionally applied an electron beam to modify the surface structure of graphene and studied the microstructural evolution of SnO2 QDs which was loaded graphene [Citation64]. In this work, firstly, four identical suspensions of graphene nanosheets were respectively irradiated by an electron accelerator with four different absorbed doses (70, 140, 210, and 280 kGy) at room temperature. Then, certain amount of cetyltrimethylammonium bromide, SnCl4·5H2O and NaOH were added to the irradiated suspensions under stirring. The resulted mixtures were continuously stirred for 30 min. Subsequently, the black suspensions were heated in a Teflon vessel at 160°C for 20 h. Electron beam can create lattice defects or damage in the graphene nanosheets. Since lattice defects and grain boundaries are favorable nucleation sites, these defects and damage are beneficial for the formation of more and smaller SnO2 QDs. From the Figure (b,e,h,k,n), it can be seen that a large amount of SnO2 QDs are loaded on the surface of graphene nanosheets. As shown in Figure (k), the SnO2 QDs loaded on graphene nanosheets irradiated at 210 kGy exhibits much smaller size distribution and much better homogeneous distribution. These can be attributed to the influences of the radiation dose on the grain boundaries, defects, and oxygen-containing groups in the graphene nanosheets (Figure (d,g,j)). Figure (n) indicated that the SnO2 QDs planted on graphene nanosheets irradiated at 280 kGy have large agglomeration. The reason is that the high radiation dose causes deterioration or collapse of the graphene nanosheets as shown in Figure (m), resulting in agglomeration of SnO2 QDs.

Figure 11. HRTEM images of (a) an unirradiated GNS and an irradiated GNS under different irradiation doses: (d) 70, (g) 140, (j) 210, and (m) 280 kGy; TEM images of (b) an unirradiated SGNS and an irradiated SGNS under different irradiation doses: (e) 70, (h) 140, (k) 210, and (n) 280 kGy; HRTEM image of (c) an unirradiated SGNS and an irradiated SGNS under different irradiation doses: (f) 70, (i) 140, (l) 210, and (o) 280 kGy. Reprinted with permission from Ref. [Citation64]. Copyright 2015 American Chemical Society.

Figure 11. HRTEM images of (a) an unirradiated GNS and an irradiated GNS under different irradiation doses: (d) 70, (g) 140, (j) 210, and (m) 280 kGy; TEM images of (b) an unirradiated SGNS and an irradiated SGNS under different irradiation doses: (e) 70, (h) 140, (k) 210, and (n) 280 kGy; HRTEM image of (c) an unirradiated SGNS and an irradiated SGNS under different irradiation doses: (f) 70, (i) 140, (l) 210, and (o) 280 kGy. Reprinted with permission from Ref. [Citation64]. Copyright 2015 American Chemical Society.

3.2.2. SnO2 QDs/CNTs

Owing to the high aspect ratio and excellent electrical conductivity, the carbon nanotubes (CNTs) are attractive. Some efforts have been devoted to embed SnO2 QDs into CNTs to form SnO2 QDs/CNTs nanocomposites [Citation101–106]. The methods of embedding SnO2 QDs into CNTs can be divided into two classes that is in-situ and ex-situ assembly. Jin et al. [Citation102] have prepared SnO2 QDs/MWCNTs by a in-situ assembly method. Firstly, MWCNTs were treated with H2SO4/HNO3 (3:1 ratio). Then, the treated MWCNTs were mixed with SnCl4·5H2O, N2H4 and deionized water to form a homogeneous solution, followed by heating at 150°C for 24 h. The SnO2 QDs (less than 3 nm) were uniformly planted onto the MWCNT, providing a large BET surface area (240 m2 g−1). Compared with the pure SnO2 QDs, there is a smaller decline in the redox peaks of the SnO2 QDs/MWCNT until the tenth cycle, implying its improved cyclability in lithium storage. Song et al. [Citation103] have also embedded SnO2 QDs into MWCNTs using a similar method. In the process, the ethanol suspension of SnCl4, activated MWCNTs and deionized water was hydrothermally treated at 100°C for 6 h. Eventually, monodisperse SnO2 QDs (∼3 nm) are firmly embedded into the MWCNTs. This hybrid displayed excellent cycling performance with high reversible capacity about 700 mAh g−1 after 150 cycles at 0.1 A g−1, surpassing both bare SnO2 QDs and MWCNTs. The ex-situ assembly is another way to anchor SnO2 QDs onto MWCNTs. Lu et al. [Citation105] have developed a facile ex-situ method of loading SnO2 QDs on carbon nanotubes. In this method, colloidal SnO2 QDs with average size of ∼3.5 nm have been successfully fabricated using thiourea as stabilizing and accelerating agent at room temperature. For the ex-situ loading of SnO2 QDs on CNTs, the colloidal SnO2 QDs solution was added into the suspension of acid-treated CNTs under magnetic stirring. During this process, the SnO2 QDs were ex-situ loaded on CNTs, forming SnO2 QDs/CNTs. The SnO2 QDs/CNTs displayed excellent lithium storage properties, with a discharge capacity of 845 mAh/g at 100 mA/g after 90 cycles. Liu et al. [Citation104] have also loaded SnO2 QDs on the surface of MWCNT by mixing chloroform suspension of MWCNT and toluene solution of SnO2 QDs under magnetic stirring at room temperature. They investigated their gas sensing performance for H2S. Compared to the pure SnO2 QDs, the sensor based on the SnO2 QDs/MWCNT displayed a higher response upon H2S.

3.2.3. SnO2 QDs/AC

SnO2 is a good anode material for lithium ion batteries (LIBs) due to its relatively high theoretical reversible capacity. However, the poor cycling stability of SnO2 is caused by its volume swell during lithium releasing and storing. An effective approach to avoid the volume swell is embedding SnO2 QDs into amorphous carbon (AC) to form SnO2 QDs/AC nanocomposites [Citation107–110]. The SnO2 QDs can be easily embedded in amorphous carbon by a one-step hydrothermal route. Song et al. [Citation109] have encapsulated SnO2 QDs with a size of 2 nm into mesotunnels of mesoporous carbon by this method. In a typical process, the ethanol suspension containing certain amounts of SnCl4, mesoporous carbon and deionized water was heated to 100°C and maintained for 6 h. The ultrafine SnO2 QDs of 2 nm have been evenly encapsulated in the mesotunnels of mesoporous carbon and anchored on the surface of mesoporous carbon. These SnO2 QDs SnO2 QDs in mesoporous carbon showed excellent cycling performance with high reversible capacity retention above 95% for 200 cycles, which was superior to that of bare SnO2 QDs. As shown in , these SnO2 QDs in mesoporous carbon also showed much better cycling and rate performances than that of larger-sized SnO2 nanoparticle encapsulated in mesoporous carbon. This improved cycling performance can be ascribed to two factors: redox capacitance and interfacial capacitance. A two-step method has also been developed to embed SnO2 QDs in amorphous carbon. Jin et al. [Citation106] have embedded SnO2 QDs into amorphous carbon (SnO2 QDs/AC) by a solvothermal method accompanied by calcination process in a high temperature. Firstly, the precursor of SnO2 QDs was obtained by heating aqueous solution of SnCl4, glucose and urea at 180°C for 12 h. Then, SnO2 QDs of 3–5 nm were embedded in amorphous carbon through calcinating the resulted precursor in argon atmosphere. This hybrid composite exhibits excellent cycle stability and high rate capability with the capacities of 364 mAh g−1 at the current density of 5 A g−1after 300 cycles. Wang et al. [Citation107] have coated SnO2 QDs of 5 nm with amorphous carbon by refluxing Sn4+ and Bovine Serum Albumin (BSA), followed by a carbonization process. In the refluxing process, the BSA promotes the formation of SnO2 QDs and simultaneously restricts the size of SnO2 to quantum dot scale. In the following carbonization process, the BSA acts as a carbon source to form amorphous carbon. The carbon-coated SnO2 QDs exhibited a markedly improved cycle performance in lithium storage with charge capacity of 473.1 mAh g−1 at the current rates of 250 mA g−1. Hu et al. [Citation108] have planted SnO2 QDs of 5 nm into amorphous carbon matrix by boiling Sn powders and crystal sugar in deionized water, followed by carbonizing the precursor in Ar atmosphere. Planting SnO2 QDs in amorphous carbon matrix can well accommodate the volume change of Li+ insertion/extraction in SnO2 anode materials. The SnO2 QDs planted carbon matrix exhibited stable reversible capacity of 400 mAh/g after 100th cycle at a high current rate of 3.3 A g−1.

Table 2. Comparisons of electrochemical properties of SnO2 QDs encapsulated in mesoporous carbon with larger-sized SnO2 nanoparticles encapsulated in mesoporous carbon (‘C’ is ‘capacity’; ‘—’ is null).

3.2.4. Other SnO2 QDs based hybrid

In order to improve its electrochemical or photocatalytic property, SnO2 QDs have also successfully been bonded with some other matrix, such as TiO2 [Citation67,Citation114,Citation115], ZnO [Citation116], MgO [Citation117], Li4Ti5O12 [Citation118], SnS2 [Citation119], amorphous silica [Citation120,Citation121], conducting polymer [Citation122,Citation123], C3N4 [Citation124] and polyvinylpyrrolidone (PVP) [Citation125]. For example, using Zn(NO3)2·6H2O and SnCl4·5H2O as precursor, our group has successfully prepared ZnO/SnO2 heterojunctions by sol–gel method and supercritical fluid drying processes [Citation116]. These ZnO/SnO2 heterojunctions were adorned uniformly with SnO2 QDs and ZnO quantum dots (ZnO QDs), whose size varied from 3 to 7 nm. The optical properties of the ZnO/SnO2 heterojunctions were affected by calcination temperature. When the calcination temperature was increased from 500 to 700°C, the photoluminescence (PL) intensity of the heterojunctions increased. This increase in PL intensity can be attributed to the decrease in the number of surface dangling bonds of ZnO and SnO2 QDs. Ma et al. [Citation119] have successfully loaded SnO2 QDs with sizes of 4 nm on the surface of SnS2 nanosheets by hydrothermal treating a aqueous suspension, which was composed of SnS2 nanosheets, SnCl4. ascorbic acid (Ac) and Na2CO3. In this method, the Ac acted as a stabilizer to enhance the stability of Sn4+ ions in solution by forming Sn4+-Ac complex. The Sn4+-Ac can gradually separate to release Sn4+ under hydrothermal process. As a precipitator, the Na2CO3 can react with the Sn4+ to form SnO2 QDs. These SnO2 QDs loaded on SnS2 nanosheet through intermolecular forces or electrostatic gravitation. The SnO2 QDs/SnS2 hybrids demonstrated obviously higher photocatalytic activities than both the bare SnO2 QDs and SnS2 nanosheets for the reduction of Cr (VI) under visible light. Lee et al. [Citation67] have decorated TiO2 nanoparticles with SnO2 QDs through stirring the mixture of ethanol suspension of SnO2 QDs and ethanol suspension of P25 TiO2 at room temperature. The SnO2 QDs decorated TiO2 nanoparticles showed much enhanced photocatalytic activity for the degradation of Rhodamine B. The reaction rate constant was significantly enhanced from 0.025 to 0.055 min−1 by decorating SnO2 QDs on TiO2 nanoparticles. Du et al. [Citation122] have loaded SnO2 QDs with size less than 10 nm onto polypyrrole (PPy) nanowires by a electrodeposition. The PPy nanowires can accommodate the volumetric change of SnO2 QDs and increase the interface of electrode/electrolyte. So, the SnO2 QDs/PPy nanowires delivered enhanced Li+ storage performance with a reversible capacity of 690 mAh/g after 80 cycles at a rate of 0.3 C.

4. Applications of SnO2 QDs

4.1. Applications in lithium-ion batteries

Compared with commercially used graphite with an theoretical lithium storage of 372 mAh/g, SnO2 is a promising anode material for next-generation lithium-ion batteries (LIBs) due to its low cost, safety, natural abundance and high theoretical lithium storage capacity (about 782 mAh/g). However, the lithiation and delithiation processes cause large volume expansion and severe particle aggregation, and inevitably bring on the pulverization, loss of interparticle contact and blocking of the electrical contact pathways, which consequently result in a rapid capacity fading and poor cycling stability [Citation90,Citation122,Citation126,Citation127]. To eliminate these problems, two strategies have been explored. One strategy is to minimize the particle size of SnO2 and optimize their dispersity [Citation44,Citation49], which can suppress the huge volume expansion during Li+ insertion and extraction processes. Chen et al. [Citation49] have reported that the electrode constructed from ultrasmall pure SnO2 QDs exhibited an improved cyclic capacity and a better rate capacity. The discharge capacity was about 718 mAh g−1 after 60 cycles at 0.1 C. The electrode also showed excellent discharge ability at a high current density. Even after 200 cycles, the discharge capacities at 1 and 5 C were respectively maintained at 454 and 376 mAh g−1. These excellent performances of the ultrasmall SnO2 QDs electrode for lithium-ion batteries can be attributed to the reduction of the SnO2 particle size and optimization of their dispersivity, which can shorten the distance for Li+ diffusion and enlarge electrode–electrolyte contact area for a high Li flux across interface.

Another strategy is to combine ultrasmall SnO2 with other nanomaterials, which can buffer the volume expansion or enhance the electronic conductivity of the electrodes. The currently used nanomaterials in this aspect include various carbonaceous materials, such as graphene nanosheets [Citation66,Citation69–73,Citation128], carbon nanotubes [Citation75–77], amorphous carbon [Citation107,Citation109,Citation129]; several kind of metal oxide, such as TiO2 [Citation114,Citation115], and Li4Ti5O12 [Citation86]; conducting high polymer [Citation122] and amorphous silica [Citation120]. Song et al. [Citation86] have loaded the SnO2 QDs with particle size below 5 nm onto graphene oxide (GO) to form a SnO2 QDs/GO hybrid as shown in Figure (a,b). As shown in Figure (c), the charge–discharge curve for the SnO2 QDs/GO electrode exhibited an obviously high initial discharging capacity (about 2000 mA h g−1). This composite electrode also displayed a high reversible capacity of 800 mAh g−1 at the rate of 100 mA/g as shown in Figure (d), maintaining about 90% of its initial capacity after 200 cycles. More importantly, the composite electrode based on SnO2 QDs/GO hybrid also delivers superior rate performance. Figure (d) also shows that the composite electrode exhibited a superior rate performance at the high rates of 1 A g−1 and 2 g−1, retaining capacities of 600 mAh g−1 and 400 mAh g−1, respectively. As exhibited in Figure (e), increasing rates from 100 mA g−1, 500 mA g−1 to 5 A g−1, 10 A g−1 and then returning to 100 mA g−1 were performed for galvanostatic cycling. Even under these harsh conditions, the electrode fabricated from SnO2 QDs/GO hybrid can still recover to its initial capacity. The long term cycling testing at 5 A g−1 (Figure (f)) exhibited a stable cycling performance for 1000 cycles with a capacity similar to that of the 2nd cycle. Compared with the electrode fabricated from the graphene loaded with different morphologies of SnO2, such as SnO2 nanosheets [Citation30,Citation130], SnO2 nanorods [Citation131,Citation132], SnO2 hollow nanospheres [Citation133], SnO2 mesoporous spheres [Citation134] or SnO2 nanotubes [Citation135], the electrode fabricated from SnO2 QDs/GO has displayed much improved cycling and rate performances for lithium-ion storage. The reasons for these excellent electrochemical performances are as follows: (1) the extremely small size of SnO2 QDs makes the process of redox reaction between SnO2 and Sn partially reversible; (2) the large specific surface area and good distribution of the SnO2 QDs on graphene guarantee the abundant active sites for Li+ insertion and extraction, and the efficient utilization of the active sites in the electrode materials; (3) the intimate contact of the conductive graphene with the SnO2 QDs provides transport pathway for the alloying-dealloying of Li+; (4) the graphene sheets withstand the volume expansion of the electrode materials and prevent the aggregation of SnO2 QDs. Ren et al. [Citation101] have applied multiwalled carbon nanotubes uniformly loaded with 3–5 nm SnO2 QDs as the electrode materials. The electrode exhibited a superior cycling stability with a capacity of 800 mAh g−1 after 300 cycles. It can be explained as follows: (1) the MWCNTs, as a conductive nanomaterial, facilitate transportation of Li+ and electrons during charge and discharge; (2) the MWCNTs serve as volume buffers in the electrode. Wang et al. [Citation107] have reported that the 5 nm SnO2 QDs with a 16 wt% carbon coating exhibited a high discharge capacity of 502 mAh g−1 at current rate of 100 mA g−1 after 100 cycles. They ascribed high discharge capacity to the carbon coating on the surface of SnO2 QDs. The carbon coating can resist the volume expansion/contraction during Li-Sn alloying-dealloying.

Figure 12. (a) SEM and (b) TEM image of the SnO2 QDs–GO hybrid; (c) Typical charge–discharge curve for the composite electrode tested at 100 mA g−1 in a voltage window of 0.005–2.5 V; (d) Cycling performance for the composite electrode galvanostatically tested at different rates; (e)Rate performance for the composite electrode discharging at various rates from 100 mA g−1 to 10 A g−1; (f) Long term cycling for the composite electrode dischargingat 5 A g−1 for 1000 cycles. Reprinted with permission from Ref. [Citation86]. Copyright 2013 the Royal Society of Chemistry.

Figure 12. (a) SEM and (b) TEM image of the SnO2 QDs–GO hybrid; (c) Typical charge–discharge curve for the composite electrode tested at 100 mA g−1 in a voltage window of 0.005–2.5 V; (d) Cycling performance for the composite electrode galvanostatically tested at different rates; (e)Rate performance for the composite electrode discharging at various rates from 100 mA g−1 to 10 A g−1; (f) Long term cycling for the composite electrode dischargingat 5 A g−1 for 1000 cycles. Reprinted with permission from Ref. [Citation86]. Copyright 2013 the Royal Society of Chemistry.

It is well known that the TiO2 has a volume variation of less than 4% with a very low capacity (170 mAh/g) during lithium intercalation. However, SnO2 has a relatively higher capacity (784 mAh/g) with a volume change of more than 250%. Du et al. [Citation115] have prepared a three-dimensional SnO2/TiO2 anodes for lithium-ion battery by planting SnO2 QDs (diameters < 5 nm) into TiO2 nanotube arrays. The three-dimensional anodes displayed an excellent capacity retention of 70.8% after 100 cycles in the voltage range of 0.05–2.5 V. This may contribute by the synergistic effect between the SnO2 QDs and TiO2 nanotube arrays. The SnO2 QDs can improve the capacity of the electrode, while the TiO2 nanotube array can sustain the volume change and maintain the structural integrity of the electrode.

4.2. Application in photocatalysis

SnO2 is considered to be one of the most efficient and nontoxic photocatalysts. Their photocatalytic mechanisms are summarized as following: (1) upon irradiation by a light with photons energy larger than the band gap of SnO2, the electrons from the SnO2 are excited and jump from the valence band (VB) to the conduction band (CB), producing electron–hole pairs; (2) the charge carriers produced by photons move from interior to the surface of photocatalyst; (3) the charge carriers promote the formation of hydroxyl radicals (·OH) and superoxide radical anions (); (4) the organic pollutants are adsorbed on the surface of photocatalyst, then are decomposed by the powerful hydroxyl radicals (·OH) and superoxide radical anions (). The photocatalytic performance of SnO2 is significantly influenced by (1) the light absorption properties of catalytic agent, (2) reduction and oxidation rates on the surface by the electron and hole, and (3) the electron–hole recombination rate [Citation136]. These influence factors depend on particle size, crystallinity, specific surface area, and so on.

It is known that the smaller SnO2 particles have the higher specific surface area. The higher the specific surface area has, the more the active sites locate in the surface. The active sites can improve the photocatalytic activity. The catalytic activity of SnO2 QDs is higher than that of the one with larger particles size. In recent years, many groups have dedicated themselves to the photocatalytic performances of pure SnO2 QDs [Citation50,Citation53,Citation54,Citation57,Citation137–140]. Liu et al. [Citation57] have studied the photocatalytic activity of SnO2 QDs with diameters of 3–5 nm on the degradation of methylene blue under visible light irradiation. The SnO2 QDs exhibited a good photocatalytic activity with a degradation rate of 90% on methylene blue after 240 min. However, at the same condition, the commercial SnO2 photocatalyst showed very low photocatalytic activity with degradation rate of 3% after 240 min. Jia et al. [Citation138] have fabricated the SnO2 QDs with a size of 6.7 nm and investigated their photocatalytic activity on degradation of Rhodamine B (RhB) under UV light irradiation. The SnO2 QDs showed high photocatalytic activities, photodegrading 100% after exposure to UV light for 80 min. The photocatalytic activity of the SnO2 QDs was superior to that of SnO2 nanorods [Citation31], SnO2 nanoflowers [Citation141,Citation142], SnO2 nanospheres [Citation45] and SnO2 films [Citation28]. This enhanced photocatalytic activity of SnO2 QDs may be ascribed to: (1) a easier adsorption of methylene blue on the surface, (2) a higher visible light absorption intensity, and (3) a lower rate of electron–hole pair recombination. Shajira et al. [Citation139] have investigated the photocatalytic performances of the SnO2 with three different particle sizes (2.5, 5 and 12 nm) on the decomposition of methyl orange. Their investigation indicated that the smaller particle shows the better photocatalytic activity. For example, after 6 h of sunlight irradiation, the SnO2 with 2.5 nm particle size decomposed the aqueous vanillin solution (0.01 mM) to 93% of its initial concentration. In contrast, the SnO2 with 5 and 12 nm particle sizes only decomposed the solution to 51% and 35% of its initial concentration, respectively. Two factors may donate the better photocatalytic activity of the SnO2 with smaller particle size. The decrease of crystallite size leads to the increase in the specific surface area, which facilitates the absorption of methyl orange. On the other hand, the decrease of crystallite size results in the increase in population of defects, which accelerates the reaction between electron–hole pairs and methyl orange.

The photocatalytic activity of the SnO2 QDs is significantly hindered by the high recombination rate of electron–hole pairs. It is worth noting how to manipulate the chemical composition and surface chemistry of SnO2 QDs. Metal-ion dopants in the SnO2 QDs significantly influence the recombination rate of electron–hole pairs and interfacial electron transfer rates. Fan et al. [Citation81] have discussed that the optical response of Sn2+ self-doped SnO2 QDs and their catalytic activity on photodegradation of methyl orange. The UV-vis diffuse reflectance spectra illustrated in Figure (a) reveal that with the increase in the added tin powder, the absorption edges for the each samples of Sn2+ self-doped SnO2-x QDs are red-shifted compared to bare SnO2 QDs. This improved visible-light absorbance of the Sn2+ self-doped SnO2-x QDs is beneficial for their photocatalytic applications in visible light range. Figure (b) exhibits that the calculated band gaps for SnO2-x (1:2), SnO2-x (1:4), SnO2-x (1:8) and SnO2-x (1:16) are 2.6, 2.65, 2.8 and 2.9 eV, respectively, which are all obviously narrowed relative to that of the bare SnO2 QDs (3.5 eV). These narrowed band gaps are crucial for the applications of the doped SnO2-x QDs samples in the photodecolorization of organic dyes in visible light illumination, since the bare SnO2 QDs sample can only respond to ultraviolet light. Compared with the bare SnO2 QDs, the Sn2+ doped SnO2-x QDs (1:4) showed an excellent photocatalytic activity on methyl orange (MO) in visible light illumination. As shown in Figure (e), the Sn2+ doped SnO2-x QDs (1:4) decompose 99.5% of methyl MO after irradiated by visible light for 15 min. However, the bare SnO2 QDs show no photocatalytic activity on the degradation of MO at the same experimental conditions. This superior photocatalytic activity in the Sn2+ doped SnO2-x QDs sample can be attributed to the self-doping of Sn2+ into the lattice of SnO2 QDs and accompaniment of oxygen vacancies, which can narrow the band gap and separate the electron–hole pairs in SnO2 QDs. As is exhibited in the photocurrent response experiments (Figure (f)), the photocurrent of the Sn2+ doped SnO2-xQDs (1:4) is more stable and higher than that of the bare SnO2 QDs. This confirms the decrease of the band gap due to the Sn2+ doping and oxygen vacancies. Kumar et al. [Citation74] has studied the photocatalytic properties of heavily F-doped SnO2 QDs with an average diameter of 5–7 nm. The F-doped SnO2 QDs displayed much better photocatalytic properties on the degradation of Rhodamine B compared to pure SnO2 QDs. In the presence of F-doped SnO2 QDs, the Rhodamine B aqueous solution turns colorless within 20 min after irradiated by UV light. They ascribed this enhancement to the large pore diameter and very high concentration of oxygen vacancies in SnO2 produced by F doping.

Figure 13. UV-vis diffuse reflectance spectra (a) and band gap evaluation (b) of the SnO2QDs and Sn2+ self-doped SnO2−x QDs with different molar ratios of Sn/SnCl4; TEM (c) and HRTEM (d) images of Sn2+ self-doped SnO2−x QDs (1:4), the inset in (d) is the corresponding SAED pattern; UV-vis spectra (e) of MO in aqueous Sn2+ self-doped SnO2−x QDs (1:4) dispersions as a function of irradiation time with visible light irradiation (λ ≥ 400 nm); Current density response vs time (f) for the Sn2+ self-doped SnO2−x (1:4) QDs and SnO2 QDs under chopped irradiation at a bias potential of 0.5 V vs Ag/AgCl (λ ≥ 400 nm). Reprinted with permission from Ref. [Citation81]. Copyright 2013 American Chemical Society.

Figure 13. UV-vis diffuse reflectance spectra (a) and band gap evaluation (b) of the SnO2QDs and Sn2+ self-doped SnO2−x QDs with different molar ratios of Sn/SnCl4; TEM (c) and HRTEM (d) images of Sn2+ self-doped SnO2−x QDs (1:4), the inset in (d) is the corresponding SAED pattern; UV-vis spectra (e) of MO in aqueous Sn2+ self-doped SnO2−x QDs (1:4) dispersions as a function of irradiation time with visible light irradiation (λ ≥ 400 nm); Current density response vs time (f) for the Sn2+ self-doped SnO2−x (1:4) QDs and SnO2 QDs under chopped irradiation at a bias potential of 0.5 V vs Ag/AgCl (λ ≥ 400 nm). Reprinted with permission from Ref. [Citation81]. Copyright 2013 American Chemical Society.

Planting SnO2 QDs onto other nanomaterials, e.g. TiO2 [Citation67,Citation143], silica [Citation121] and poly (ethylene glycol methyl ether) [Citation123], are another efficient way to improve the photocatalytic properties on the degradation of organic pollutant. For example, Lee et al. [Citation67] have compared the photocatalytic performances of undecorated P25 TiO2 and the P25 TiO2 decorated with SnO2 QDs on the degradation toward Rhodamine B. Compared with undecorated P25 TiO2, the P25 TiO2 decorated with SnO2 QDs with size of about 5.3 nm showed an improved photocatalytic performance on the degradation of toward Rhodamine B under UV light illumination. The apparent reaction rate constant of the P25 TiO2 planted with SnO2 QDs (0.055 min−1) is much higher than that of unplanted P25 TiO2 (0.025 min−1). Through planting of the SnO2 QDs, the photon-induced charges (namely, electrons and holes) separation in the P25 TiO2 is much boosted, which is beneficial for the enhancing in the photocatalytic performances. Babu et al. [Citation124] have reported the improved photocatalytic activities of SnO2 QDs/g-C3N4 for degradation of MO under illuminated by sunlight. The SnO2 QDs below 3 nm are well dispersed on g-C3N4 layers. At the same conditions, the photodegradation efficiency of SnO2 QDs/g-C3N4 for the degradation of MO is about 94%, which is obviously larger than that of bare SnO2 QDs (21%). Figure  displays the mechanism of SnO2 QDs/g-C3N4 with improved photocatalytic activities under illumination sunlight. The CB level of g-C3N4 is negative than that of SnO2 QDs, therefore, the electrons generated by photons move from CB of g-C3N4 to that of SnO2 QDs. The VB level of SnO2 QDs is positive than that of g-C3N4, therefore, the holes generated by photons still stay at the VB of g-C3N4. So, the electrons and holes generated by photons are efficiently separated and promote the formation of hydroxyl radicals (·OH) and superoxide radical anions (). These and ·OH are powerful enough to oxidize the toxic MO molecules adsorbed on the surface of SnO2 QDs/g-C3N4 to nontoxic H2O and CO2.

Figure 14. The mechanism of SnO2 QDs/g-C3N4 with improved photocatalytic activities. Reprinted with permission from Ref. [Citation124]. Copyright 2018 Elsevier.

Figure 14. The mechanism of SnO2 QDs/g-C3N4 with improved photocatalytic activities. Reprinted with permission from Ref. [Citation124]. Copyright 2018 Elsevier.

4.3. Application in gas sensors

It is important that the working mechanism of the semiconductor oxide gas sensors. Upon exposure to target gases, the resistance of semiconductor oxide may significantly change. This resistance change reflects the gas sensing performance of semiconductor oxides. SnO2, as an n-type semiconductor, possesses an electron-depletion layer on the surface due to the chemisorbed oxygen, which lead to the structure with resistive shell and semiconducting core. When the oxygen species with negative charges are adsorbed on the surface of SnO2, the resistive shell is oxidized by these oxygen species. The conductivity of SnO2 increases with the increase of electrons induced by the oxidation reactions. On the contrary, the conductivity of SnO2 will decrease when it is exposed in reducing gas. In order to boost gas sensing performances, it is of importance to fabricate SnO2 with large surface areas or large porosity, which can ensure its easy access to target gases.

Some researchers have discovered that the SnO2 QDs have gas sensoring performance [Citation44,Citation144,Citation145]. Ma et al. [Citation44] have reported the ethanol sensing behavior by using the sensor based on the SnO2 QDs with a small size of 2.5–4.1 nm. It was found that the sensor response is enhanced obviously with the increase of ethanol concentration from 5 to 100 ppm. The sensor based on 2.5–4.1 nm SnO2 QDs also showed good repeatability at 100 ppm ethanol concentration. In addition, this sensor showed more superior response and recovery speeds performances than that based on the SnO2 with size of 20–110 nm. Two factors significantly contribute to this strongly improved gas sensing activities. Firstly, the SnO2 QDs with smaller size have a higher BET surface area (189.4 m2 g−1), which can supply more active sites for sensing reactions. Secondly, the SnO2 QDs have a small size of 2.5–4.1 nm, which is smaller than the thickness of the electron depletion layer. It has been demonstrated that the gas sensing properties can be enhanced if the size of an active material is smaller than the thickness of the electron depletion layer. He et al. [Citation145] have fabricated the SnO2 QDs with different size by annealing primal SnO2 QDs, and studied their gas sensing properties towards ethanol. It was found that the SnO2 QDs with smaller size exhibited higher sensing activities. Moreover, the SnO2 QDs with a size of 3.7 nm showed fast response (1 s) and recovery (1 s) toward ethanol. These excellent gas sensing properties may be contributed to the fast adsorption of ethanol on the surface of SnO2, which is promoted by the small size of the SnO2 QDs.

It is known that the bare SnO2 QDs have an undesirable tendency to aggregate because of their ultrasmall size. This agglomeration will seriously abate the specific surface area and thus degrade the gas sensing performances. In order to improve the gas sensing property, it is necessary to conquer the above mentioned drawbacks. The dispersion of SnO2 QDs can be significantly enhanced by planting them on a conductive and stable matrix. Carbon materials, especially graphene [Citation87,Citation91,Citation92,Citation146] and carbon nanotubes [Citation92], have been used as the matrix to support SnO2 QDs due to their superior stability and conductivity. Li et al. [Citation92] have planted the SnO2 QDs with an average size of 2–7 nm into three-dimensional mesoporous graphene aerogel (SnO2 QDs/rGO-4), and studied the NO2 gas sensing performance of obtained SnO2 QDs/rGO-4. At 55°C, the optimal operating temperature, the sensor displayed low detection limit (2 ppm), high sensitivity (0.001 ppm−1), an excellent linearity in a wide NO2 gas concentration (from 14 to 110 ppm). These excellent sensing performances for NO2 gas result from the supporting effect of graphene, such as the 3D graphene aerogel, which boost the dispersity of SnO2 QDs, enhance the electron transfer to target gas as well as improve the gas diffusion. The sensor also showed high sensing selectivity for NO2 gas in the presence of other vapors (∼110 ppm) at 55°C, including CO, ethanol, phenylcarbinol, ethylene glycol, toluene, acetone, trihalomethanes (THMS) ammonia, and formaldehyde. In addition, due to the flammable nature of liquefied petroleum gas (LPG), some groups have focused on the LPG gas sensing by using SnO2 QDs/rGO. Mishra et al. [Citation91] have reported the LPG gas sensing behavior of SnO2 QDs/RGO. This sensor showed a high response of ∼92.4% to 500 ppm LPG at temperatures of 250°C, and good selectivity for LPG upon exposure to various volatile gases, including LPG, ammonia, toluene, chloroform, benzene, n-butylacetate, acetone, formic acid, and acetic acid. The response for LPG was 17.8 times higher than that for formic acid. Nemade et al. [Citation146] have investigated the LPG gas sensing behavior of the SnO2 QDs/RGO at low temperatures. They found that the response increased with the wt% of graphene at room temperature. When exposed to towards 50 ppm LPG during 30 days, the 1.6 wt% graphene/SnO2 QDs show a fast response time of 16 s, good recovery time of 16 s, and good stability. Liu et al. [Citation104] have studied the gas sensing property of SnO2 QDs/MWCNTs towards H2S. The sensor based on SnO2 QDs/MWCNT was fabricated by spin-coating method as shown in Figure (a). As is shown in Figure (b), in the SnO2 QDs/MWCNT, ultrasmall SnO2 QDs are efficiently attached and covered on MWCNT. As is exhibited in the response curves (Figure (c)), the sensor based on SnO2 QDs/MWCNT was more sensitive to H2S gas detection than that fabricated from the pristine SnO2 QDs at the same experiment conditions. For example, at the gas concentration of 100 ppm, the response of the SnO2 QDs/MWCNT sensor reached to 181, while that of pristine SnO2 QDs sensor only was 47. The response performances of the SnO2 QDs/MWCNT gas sensors toward H2S, NH3, SO2, and NO2 were compared in Figure (d). The sensors were highly selective toward H2S with response value of 108 at 70°C, while with very small response values of 0.28, 4.3 and 1.0 for NO2, NH3, and SO2, respectively. In addition to the excellent access of H2S molecules to SnO2 QDs surfaces and the electron transport in MWCNTs, they ascribed this higher response toward H2S to the synergetic effect between SnO2 QDs and MWCNTs. The favorable energy band alignment of SnO2 QDs/MWCNT facilitates the electron transport, thereby boosts the H2S sensing performance.

Figure 15. (a) Sensor fabrication from the SnO2 QDs/MWCNT; (b) HRTEM image of the SnO2 QDs/MWCNT; (c) Response curves of the sensors fabricated from SnO2 QDs/MWCNT and pristine SnO2 QDs upon H2S exposure/release cycles at 70°C; (d) Selectivity of the SnO2 QDs /MWCNT gas sensor at 70°C. Reprinted with permission from Ref. [Citation104]. Copyright 2013 American Chemical Society.

Figure 15. (a) Sensor fabrication from the SnO2 QDs/MWCNT; (b) HRTEM image of the SnO2 QDs/MWCNT; (c) Response curves of the sensors fabricated from SnO2 QDs/MWCNT and pristine SnO2 QDs upon H2S exposure/release cycles at 70°C; (d) Selectivity of the SnO2 QDs /MWCNT gas sensor at 70°C. Reprinted with permission from Ref. [Citation104]. Copyright 2013 American Chemical Society.

4.4. Applications in some other domains

Up to now, the applications of bare SnO2 QDs and their modified ones mainly focus on the above discussed lithium-ion batteries, photocatalysis and gas sensors domains. However, researchers have also explored their applications in some other domains, such as fast adsorption of methylene blue (MB) [Citation121], nano light emitting [Citation125], microwave absorbing [Citation147] and resonance imaging [Citation148]. Dutta et al. [Citation121] have loaded the SnO2 QDs with size of 3 nm on mesoporous SiO2 nanoparticles (MSN) and investigated their adsorption performances of MB. The SnO2 QDs/MSN nanocomposite showed fast adsorption performances of MB adsorbing 100% of MB in 5 min at room temperature. The SnO2 QDs/MSN nanocomposite also displayed excellent regenerated performance with no obvious loss in adsorption capacity of MB after four cycles. Nath et al. [Citation125] have embedded the SnO2 QDs with size of 8 nm in polyvinylpyrrolidone (PVP) and explored their light emitting performances as a diodes. At room temperature, the SnO2 QDs displayed obvious electroluminescence (EL) intensity at 580 nm. Moreover, the EL intensities were linear with applied voltage (up to 20 V). It is believed that the EL originated from the oxygen vacancies in SnO2 QDs. Lin et al. [Citation147] have anchored the Ni doped SnO2 QDs with a diameter from 3 to 5 nm on MWCNTs and investigated their microwave absorption properties. It is found that the Ni-doped SnO2 QDs/MWCNTs with 28.2% (molar percentage) Ni exhibited the best microwave absorbing performances. Dutta et al. [Citation148] have embedded the 3 nm SnO2 QDs in 30 nm γ-Fe2O3 nanoparticles (NPs) and studied their magnetic resonance imaging properties at room temperature. The SnO2 QDs/γ-Fe2O3 NPs inherit not only the optical properties of SnO2 QDs but also the superparamagnetic property derived from γ-Fe2O3 NPs. Here, the introduction of superparamagnetic γ-Fe2O3 NPs extend the applications of SnO2 QDs to the magnetic resonance imaging. The SnO2 QDs/γ-Fe2O3 NPs were used to image hela cells. Laser scanning confocal microscope (LSCM) of hela cells shows the cellular uptake of SnO2 QDs/γ-Fe2O3 NPs after incubation for 6 h.

5. Conclusions and outlook

In summary, our efforts putted into the SnO2 QDs have resulted in a rich database for the synthesis, modifications, and applications for SnO2 QDs over the past years. Many specified ways can be employed to achieve SnO2 QDs, in which the reaction conditions, including reaction temperature, duration of time and pH influence on the sizes and morphologies. Doping the SnO2 QDs with extrinsic dopants is an effective way to adjust their physical and chemical properties. Strategies for embedding SnO2 QDs into graphene have intentionally been summarized. Electron-beam irradiation can create lattice defects or damage in the graphene nanosheets, which is beneficial for their potential applications. The SnO2 QDs and their modified ones showed improved performances in lithium-ion storage, photochemical catalysis, and gas sensing.

Although significant review has been made in SnO2 QDs, including their synthesis, modifications, and applications, further efforts are still required in the following aspects: (1) in order to make their high-volume production more viable, the explore in the synthesis methods of SnO2 QDs in atmospheric pressure, room temperature is necessary; (2) doping multi-elements into SnO2 QDs may be an appropriate method to optimize their physical and chemical properties; (3) more efforts should be focused on the improvement in the selectivity of SnO2 QDs based sensors through further exploring the fundamental sensing mechanisms; (4) the application of SnO2 QDs and their modified ones can be extended over a much wider fields, such as sodium-ion batteries, dye-sensitized solar cells, photo detection, photocatalytic hydrogen production and electrochemical sensing of heavy metals and organic pollutants. This review will stimulate further development in the synthesis and modification of SnO2 QDs, as well as further studies on their applications in the environment and energy fields.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

The work described in this article was financially supported by the National Natural Science Foundation of China (21601120 and 11375111), the Science and Technology Commission of Shanghai Municipality (17ZR1410500), the Program for Changjiang Scholars and Innovative Research Team in University (IRT13078), the Key Natural Science Foundation of Anhui Provincial Education Commission (KJ2016A510), the Anhui Provincial Science Foundation for Excellent Youth Talents in University (gxyq2017104) and the Educational Quality and Innovation Project of Anhui Province (2015jyxm398).

References

  • Liu X, Chen N, Han B, et al. Nanoparticle cluster gas sensor: Pt activated SnO2 nanoparticles for NH3 detection with ultrahigh sensitivity. Nanoscale. 2015;7(36):14872–14880. doi: 10.1039/C5NR03585F
  • Suematsu K, Ma N, Yuasa M, et al. Antimony-doped tin dioxide gas sensors exhibiting high stability in the sensitivity to humidity changes. ACS Sens. 2016;1:913–920. doi: 10.1021/acssensors.6b00323
  • Singh K, Malakar R, Narzary R, et al. Hydrogen sensing properties of pure and composites of ZnO and SnO2 particles: understanding sensing mechanism. Sens Lett. 2017;15(9):771–778. doi: 10.1166/sl.2017.3870
  • Chen WG, Peng SY, Tang SR, et al. Palladium doped SnO2 nanocrystals for enhanced methane gas sensing application. Sci Adv Mater. 2017;9:1735–1741.
  • Yin X, Tao L, Wang G, et al. Effects of oxygen and carbon monoxide species on the gas sensing properties of SnO2 nanoparticles. J Nanoelectron Optoelectron. 2017;12(8):748–751. doi: 10.1166/jno.2017.2183
  • Hwang BW, Lee SC, Ahn JH, et al. High sensitivity and recoverable SnO2-based sensor promoted with Fe2O3 and ZnO for Sub-ppm H2S detection. J Nanoelectron Optoelectron. 2017;12(6):617–621. doi: 10.1166/jno.2017.2062
  • Li ZG, Jin L, Liu B, et al. Experiment and simulation study of nano-SnO2 for dissolved fault gases analysis of power transformer. Sci Adv Mater. 2017;9:1888–1894.
  • Velumani M, Meher SR, Balakrishnan L, et al. Zn alloyed SnO2 anomaterials based humidity sensors. Sens Lett. 2017;15(5):424–431. doi: 10.1166/sl.2017.3820
  • Narasimman S, Dinesh U, Balakrishnan L, et al. A comparative study on structural, optical and humidity sensing characteristics of ZnO, SnO2 and ZnO:SnO2 nanocomposite. Sens Lett. 2017;15(5):440–447. doi: 10.1166/sl.2017.3828
  • Kaur S, Kumar A, Rajput JK, et al. SnO2—glycine functionalized carbon nanotubes based electronic nose for detection of explosive materials. Sens Lett. 2016;14(7):733–739. doi: 10.1166/sl.2016.3643
  • Chen Y, Lu J, Wen S, et al. Synthesis of SnO2/MoS2 composites with different component ratios and their applications as lithium ion battery anodes. J Mater Chem A. 2014;2(42):17857–17866. doi: 10.1039/C4TA03770G
  • Sun J, Wang Q, Wang Q, et al. High capacity and cyclability of SnO2–In2O3/graphene nanocomposites as the anode of lithium-ion battery. Sci Adv Mater. 2016;8:1280–1285. doi: 10.1166/sam.2016.2713
  • Fu X, Wang J, Huang D, et al. Trace amount of SnO2-decorated ZnSn(OH)6 as highly efficient photocatalyst for decomposition of gaseous benzene: synthesis, photocatalytic activity, and the unrevealed synergistic effect between ZnSn(OH)6 and SnO2. ACS Catal. 2016;6(2):957–968. doi: 10.1021/acscatal.5b02593
  • Ran L, Zhao D, Gao X, et al. Highly crystalline Ti-doped SnO2 hollow structured photocatalyst with enhanced photocatalytic activity for degradation of organic dyes. CrystEngComm. 2015;17(22):4225–4237. doi: 10.1039/C5CE00184F
  • Ramasamy E, Lee J. Ordered mesoporous Zn-doped SnO2 synthesized by exotemplating for efficient dye-sensitized solar cells. Energy Environ Sci. 2011;4(7):2529–2536. doi: 10.1039/c1ee01123e
  • Tiwana P, Docampo P, Johnston MB, et al. The origin of an efficiency improving ‘light soaking’ effect in SnO2 based solid-state dye-sensitized solar cells. Energy Environ Sci. 2012;5(11):9566–9573. doi: 10.1039/c2ee22320a
  • Asdim, Manseki K, Sugiura T, et al. Microwave synthesis of size-controllable SnO2 nanocrystals for dye-sensitized solar cells. New J Chem. 2014;38(2):598–603. doi: 10.1039/c3nj01278f
  • Li X, Yu Q, Yu C, et al. Zinc-doped SnO2 nanocrystals as photoanode materials for highly efficient dye-sensitized solar cells. J Mater Chem A. 2015;3(15):8076–8082. doi: 10.1039/C5TA01176K
  • Park H, Alhammadi S, Bouras K, et al. Nd-doped SnO2 and ZnO for application in Cu(InGa)Se2 solar cells. Sci Adv Mater. 2017;9:2114–2120.
  • Ate A, Tang Z. High photosensitivity and fast response ultraviolet PD based on large-scale individual tin oxide (SnO2) nanowire. Sens Lett. 2016;14(6):606–610. doi: 10.1166/sl.2016.3660
  • Aksoy S, Caglar Y, Caglar M, et al. Influence of annealing temperature on the structural and optical characteristics of nanostructure SnO2 films and their applications in heterojunction diode. J Nanoelectron Optoelectron. 2016;11(1):115–121. doi: 10.1166/jno.2016.1885
  • Li Z, Zhao Q, Fan W, et al. Porous SnO2 nanospheres as sensitive gas sensors for volatile organic compounds detection. Nanoscale. 2011;3(4):1646–1652. doi: 10.1039/c0nr00728e
  • Sun X, Qiao L, Qiao L, et al. Synthesis of nanosheet-constructed SnO2 spheres with efficient photocatalytic activity and high lithium storage capacity. Ionics. 2017;23(11):3177–3185. doi: 10.1007/s11581-017-2115-9
  • Zhou Q, Liu H, Hong C, et al. Fabrication and enhanced acetylene sensing properties of PdO-decorated SnO2 composites chemical sensor. Sens Lett. 2016;14(11):1144–1149. doi: 10.1166/sl.2016.3734
  • Zhang Q, Zhou Q, Yin X, et al. The effect of PMMA pore-forming on hydrogen sensing properties of porous SnO2 thick film sensor. Sci Adv Mater. 2017;9(8):1350–1355. doi: 10.1166/sam.2017.3111
  • Kim T-K, Jang G-E. Optical and electrical properties of flexible SnO2/Ag/SnO2 multilayer thin films on polyethylene terephthalate substrate. J Nanoelectron Optoelectron. 2017;12(9):881–885. doi: 10.1166/jno.2017.2153
  • Lee J-A, Heo Y-W, Lee J-H, et al. Structural and optical characteristics of Zn-rich In2O3–SnO2–ZnO (ITZO) thin films prepared by radio frequency magnetron sputtering. J Nanoelectron Optoelectron. 2017;12(6):598–601. doi: 10.1166/jno.2017.2054
  • Luo Q, Cai QZ, He J, et al. Characterisation and photocatalytic activity of SnO2 films via plasma electrolytic oxidation. Adv Appl Ceram. 2014;113(4):228–233. doi: 10.1179/1743676114Y.0000000143
  • Zhang L, Wu HB, Wen Lou X. Growth of SnO2 nanosheet arrays on various conductive substrates as integrated electrodes for lithium-ion batteries. Mater Horiz. 2014;1(1):133–138. doi: 10.1039/C3MH00077J
  • Liu H, Huang J, Xiang C, et al. In situ synthesis of SnO2 nanosheet/graphene composite as anode materials for lithium-ion batteries. J Mater Sci: Mater Electron. 2013;24(10):3640–3645.
  • Sathishkumar S, Parthibavarman M, Sharmila V, et al. A facile and one step synthesis of large surface area SnO2 nanorods and its photocatalytic activity. J Mater Sci. 2017;28(11):8192–8196.
  • Hu D, Han B, Deng S, et al. Novel mixed phase SnO2 nanorods assembled with SnO2 nanocrystals for enhancing gas-sensing performance toward isopropanol gas. J Phys Chem C. 2014;118(18):9832–9840. doi: 10.1021/jp501550w
  • Burda C, Chen XB, Narayanan R, et al. Chemistry and properties of nanocrystals of different shapes. Chem Rev. 2005;105:1025–1102. doi: 10.1021/cr030063a
  • Liu LZ, Wu XL, Gao F, et al. Determination of surface oxygen vacancy position in SnO2 nanocrystals by Raman spectroscopy. Solid State Commun. 2011;151(11):811–814. doi: 10.1016/j.ssc.2011.03.029
  • Liu LZ, Wu XL, Xu JQ, et al. Oxygen-vacancy and depth-dependent violet double-peak photoluminescence from ultrathin cuboid SnO2 nanocrystals. Appl Phys Lett. 2012;100(12):121903. doi: 10.1063/1.3696044
  • Zhang Y, Li L, Zheng J, et al. Two-step grain-growth kinetics of sub-7 nm SnO2 nanocrystal under hydrothermal condition. J Phys Chem C. 2015;119(33):19505–19512. doi: 10.1021/acs.jpcc.5b05282
  • Sato K, Yokoyama Y, Valmalette J-C, et al. Hydrothermal growth of tailored SnO2 nanocrystals. Cryst Growth Des. 2013;13(4):1685–1693. doi: 10.1021/cg400013q
  • Wu S, Cao H, Yin S, et al. Amino acid-assisted hydrothermal synthesis and photocatalysis of SnO2 nanocrystals. J Phys Chem C. 2009;113:17893–17898. doi: 10.1021/jp9068762
  • Lv T, Chen Y, Ma J, et al. Hydrothermally processed SnO2 nanocrystals for ultrasensitive NO sensors. RSC Adv. 2014;4(43):22487–22490. doi: 10.1039/c4ra03121k
  • Liu LZ, Xu JQ, Wu XL, et al. Optical identification of oxygen vacancy types in SnO2 nanocrystals. Appl Phys Lett. 2013;102(3):031916. doi: 10.1063/1.4789538
  • Liu LZ, Wu XL, Li TH, et al. Morphology-dependent low-frequency Raman scattering in ultrathin spherical, cubic, and cuboid SnO2 nanocrystals. Appl Phys Lett. 2011;99(25):251902. doi: 10.1063/1.3670337
  • Yuasa M, Kida T, Shimanoe K. Preparation of a stable sol suspension of Pd-loaded SnO2 nanocrystals by a photochemical deposition method for highly sensitive semiconductor gas sensors. ACS Appl Mater Interfaces. 2012;4(8):4231–4236. doi: 10.1021/am300941a
  • Bhanjana G, Dilbaghi N, Kumar R, et al. SnO2 quantum dots as novel platform for electrochemical sensing of cadmium. Electrochim Acta. 2015;169:97–102. doi: 10.1016/j.electacta.2015.04.045
  • Ma J, Zhang J, Wang S, et al. Superior gas-sensing and lithium-storage performance SnO2 nanocrystals synthesized by hydrothermal method. CrystEngComm. 2011;13(20):6077–6081. doi: 10.1039/c1ce05320e
  • Ghosh S, Das K, Chakrabarti K, et al. Effect of oleic acid ligand on photophysical, photoconductive and magnetic properties of monodisperse SnO2 quantum dots. Dalton Trans. 2013 Mar 14;42(10):3434–3446. doi: 10.1039/C2DT31764H
  • Pradhan K, Paul S, Das AR. Synthesis of indeno and acenaphtho cores containing dihydroxy indolone, pyrrole, coumarin and uracil fused heterocyclic motifs under sustainable conditions exploring the catalytic role of the SnO2 quantum dot. RSC Adv. 2015;5(16):12062–12070. doi: 10.1039/C4RA12618A
  • Xu X, Z J, Wang X. SnO2 quantum dots and quantum wires: controllable synthesis, self-assembled 2D architectures, and gas-sensing properties. J Am Chem Soc. 2008;130:12527–12535. doi: 10.1021/ja8040527
  • Paul S, Pradhan K, Ghosh S, et al. Uncapped SnO2 quantum dot catalyzed cascade assembling of four components: a rapid and green approach to the pyrano[2,3-c]pyrazole and spiro-2-oxindole derivatives. Tetrahedron. 2014;70(36):6088–6099. doi: 10.1016/j.tet.2014.02.077
  • Chen Y, Ma J, Li Q, et al. Gram-scale synthesis of ultrasmall SnO2 nanocrystals with an excellent electrochemical performance. Nanoscale. 2013;5(8):3262–3265. doi: 10.1039/c3nr00356f
  • 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. doi: 10.1016/j.matlet.2014.10.121
  • Birkel A, Lee YG, Koll D, et al. Highly efficient and stable dye-sensitized solar cells based on SnO2 nanocrystals prepared by microwave-assisted synthesis. Energy Environ Sci. 2012;5(1):5392–5400. doi: 10.1039/C1EE02115J
  • Xiao L, Shen H, von Hagen R, et al. Microwave assisted fast and facile synthesis of SnO2 quantum dots and their printing applications. Chem Commun. 2010;46(35):6509–6511. doi: 10.1039/c0cc01156h
  • Bhattacharjee A, Ahmaruzzaman M. A novel and green process for the production of tin oxide quantum dots and its application as a photocatalyst for the degradation of dyes from aqueous phase. J Colloid Interface Sci. 2015;448:130–139. doi: 10.1016/j.jcis.2015.01.083
  • Bhattacharjee A, Ahmaruzzaman M. Photocatalytic-degradation and reduction of organic compounds using SnO2 quantum dots (via a green route) under direct sunlight. RSC Adv. 2015;5(81):66122–66133. doi: 10.1039/C5RA07578E
  • Sain S, Kar A, Patra A, et al. Structural interpretation of SnO2 nanocrystals of different morphologies synthesized by microwave irradiation and hydrothermal methods. CrystEngComm. 2014;16(6):1079–1090. doi: 10.1039/C3CE42281J
  • Majles Ara MH, Boroojerdian P, Javadi Z, et al. Synthesis and nonlinear optical characterization of SnO2 quantum dots. Optik. 2012;123(22):2090–2094. doi: 10.1016/j.ijleo.2011.11.005
  • Liu X, Pan L, Chen T, et al. Visible light photocatalytic degradation of methylene blue by SnO2 quantum dots prepared via microwave-assisted method. Catal Sci Technol. 2013;3(7):1805–1809. doi: 10.1039/c3cy00013c
  • Singh MK, Mathpal MC, Agarwal A. Optical properties of SnO2 quantum dots synthesized by laser ablation in liquid. Chem Phys Lett. 2012;536:87–91. doi: 10.1016/j.cplett.2012.03.084
  • Pan SS, Yu SF, Zhang WF, et al. Low threshold amplified spontaneous emission from tin oxide quantum dots: a instantiation of dipole transition silence semiconductors. Nanoscale. 2013;5(23):11561–11567. doi: 10.1039/c3nr03523a
  • Pan S, Lu W, Chu Z, et al. Deep ultraviolet emission from water-soluble SnO2 quantum dots grown via a facile ‘top-down’ strategy. J Mater Sci Technol. 2015;31(6):670–673. doi: 10.1016/j.jmst.2014.09.017
  • Santos NF, Rodrigues J, Holz T, et al. Luminescence studies on SnO2 and SnO2:Eu nanocrystals grown by laser assisted flow deposition. Phys Chem Chem Phys. 2015 May 28;17(20):13512–13519. doi: 10.1039/C4CP06114D
  • Pan SS, Li FD, Liu QW, et al. Strong localization induced anomalous temperature dependence exciton emission above 300 K from SnO2 quantum dots. J Appl Phys. 2015;117(17):173101. doi: 10.1063/1.4919595
  • Wang J, Du J, Chen C, et al. Electron-beam irradiation strategies for growth behavior of tin dioxide nanocrystals. J Phys Chem C. 2011;115(42):20523–20528. doi: 10.1021/jp207673r
  • Huang RT, Zhang Q, Chen ZW, et al. Irradiated graphene loaded with SnO2 quantum dots for energy storage. ACS Nano. 2015;9:11351–11361. doi: 10.1021/acsnano.5b05146
  • Mosadegh Sedghi S, Mortazavi Y, Khodadadi A. Low temperature CO and CH4 dual selective gas sensor using SnO2 quantum dots prepared by sonochemical method. Sens Actuators B Chem. 2010;145(1):7–12. doi: 10.1016/j.snb.2009.11.002
  • Kamble VB, Umarji AM. Defect induced optical bandgap narrowing in undoped SnO2 nanocrystals. AIP Adv. 2013;3(8):082120. doi: 10.1063/1.4819451
  • Lee K-T, Lin C-H, Lu S-Y. SnO2 quantum dots synthesized with a carrier solvent assisted interfacial reaction for band-structure engineering of TiO2 photocatalysts. J Phys Chem C. 2014;118(26):14457–14463. doi: 10.1021/jp5045749
  • Kida T, Doi T, Shimanoe K. Synthesis of monodispersed SnO2 nanocrystals and their remarkably high sensitivity to volatile organic compounds. Chem Mater. 2010;22(8):2662–2667. doi: 10.1021/cm100228d
  • Cui H, Xue J, Ren W, et al. Ultra-large scale synthesis of high electrochemical performance SnO2 quantum dots within 5 min at room temperature following a growth self-termination mechanism. J Alloys Compd. 2015;645:11–16. doi: 10.1016/j.jallcom.2015.05.015
  • Epifani M, Arbio J, Pellicer E, et al. Synthesis and gas-sensing properties of Pd-doped SnO2 nanocrystals. A case study of a general methodology for doping metal oxide nanocrystals. Cryst Growth Des. 2008;8:1774–1778. doi: 10.1021/cg700970d
  • Stroppa DG, Beltra A, Conti TG, et al. Unveiling the chemical and morphological features of Sb-SnO2 nanocrystals by the combined use of high-resolution transmission electron microscopy and ab initio surface energy calculations. J Am Chem Soc. 2009;131:14544–14548. doi: 10.1021/ja905896u
  • Kar A, Patra A. Optical and electrical properties of Eu3+-doped SnO2 nanocrystals. J Phys Chem C. 2009;113:4375–4380. doi: 10.1021/jp810777f
  • Wu S, Yuan S, Shi L, et al. Preparation, characterization and electrical properties of fluorine-doped tin dioxide nanocrystals. J Colloid Interface Sci. 2010;346(1):12–16. doi: 10.1016/j.jcis.2010.02.031
  • Kumar V, Govind A, Nagarajan R. Optical and photocatalytic properties of heavily F(-)-doped SnO2 nanocrystals by a novel single-source precursor approach. Inorg Chem. 2011 Jun 20;50(12):5637–5645. doi: 10.1021/ic2003436
  • Kaviyarasu K, Devarajan PA, Xavier SSJ, et al. One pot synthesis and characterization of cesium doped SnO2 nanocrystals via a hydrothermal process. J Mater Sci Technol. 2012;28(1):15–20. doi: 10.1016/S1005-0302(12)60017-6
  • Kumar V, Nagarajan R. Thermoluminescence in heavily F-doped of SnO2 nanocrystals. Chem Phys Lett. 2012;530:98–101. doi: 10.1016/j.cplett.2012.02.021
  • Sabergharesou T, Wang T, Ju L, et al. Electronic structure and magnetic properties of sub-3 nm diameter Mn-doped SnO2 nanocrystals and nanowires. Appl Phys Lett. 2013;103(1):012401. doi: 10.1063/1.4813011
  • Mazumder N, Sen D, Saha S, et al. Enhanced ultraviolet emission from Mg doped SnO2 nanocrystals at room temperature and its modulation upon H2 annealing. J Phys Chem C. 2013;117(12):6454–6461. doi: 10.1021/jp4000329
  • Zhou GX, Xiong SJ, Wu XL, et al. N-doped SnO2 nanocrystals with green emission dependent upon mutual effects of nitrogen dopant and oxygen vacancy. Acta Mater. 2013;61(19):7342–7347. doi: 10.1016/j.actamat.2013.08.040
  • Yu SX, Yang LW, Li YC, et al. Preferred orientation growth and size tuning of colloidal SnO2 nanocrystals through Gd3+ doping. J Cryst Growth. 2013;367:62–67. doi: 10.1016/j.jcrysgro.2012.11.068
  • Fan CM, Peng Y, Zhu Q, et al. Synproportionation reaction for the fabrication of Sn2+ self-doped SnO2-x nanocrystals with tunable band structure and highly efficient visible light photocatalytic activity. J Phys Chem C. 2013;117(46):24157–24166. doi: 10.1021/jp407296f
  • Nilavazhagan S, Muthukumaran S, Ashokkumar M. Structural, optical and morphological properties of La, Cu co-doped SnO2 nanocrystals by co-precipitation method. Opt Mater. 2014;37:425–432. doi: 10.1016/j.optmat.2014.07.003
  • Sakthiraj K, Balachandrakumar K. Influence of Ti addition on the room temperature ferromagnetism of tin oxide (SnO2) nanocrystal. J Magn Magn Mater. 2015;395:205–212. doi: 10.1016/j.jmmm.2015.07.083
  • Babu B, Kadam AN, Ravikumar R, et al. Enhanced visible light photocatalytic activity of Cu-doped SnO2 quantum dots by solution combustion synthesis. J Alloys Compd. 2017;703:330–336. doi: 10.1016/j.jallcom.2017.01.311
  • Reddy CV, Babu B, Shim J. Synthesis of Cr-doped SnO2 quantum dots and its enhanced photocatalytic activity. Mater Sci Eng, B. 2017;223:131–142. doi: 10.1016/j.mseb.2017.06.007
  • Song H, Li N, Cui H, et al. Enhanced capability and cyclability of SnO2–graphene oxide hybrid anode by firmly anchored SnO2 quantum dots. J Mater Chem A. 2013;1(26):7558–7562. doi: 10.1039/c3ta11442b
  • Chang Y, Yao Y, Wang B, et al. Reduced graphene oxide mediated SnO2 nanocrystals for enhanced gas-sensing properties. J Mater Sci Technol. 2013;29(2):157–160. doi: 10.1016/j.jmst.2012.11.007
  • Cai D, Yang T, Liu B, et al. A nanocomposite of tin dioxide octahedral nanocrystals exposed to high-energy facets anchored onto graphene sheets for high performance lithium-ion batteries. J Mater Chem A. 2014;2(34):13990. doi: 10.1039/C4TA01850H
  • Dutta D, Chandra S, Swain AK, et al. SnO2 quantum dots-reduced graphene oxide composite for enzyme-free ultrasensitive electrochemical detection of urea. Anal Chem. 2014;86(12):5914–5921. doi: 10.1021/ac5007365
  • Zhang Y, Jiang L, Wang C. Facile synthesis of SnO2 nanocrystals anchored onto graphene nanosheets as anode materials for lithium-ion batteries. Phys Chem Chem Phys. 2015;17(31):20061–20065. doi: 10.1039/C5CP03305E
  • Mishra RK, Upadhyay SB, Kushwaha A, et al. SnO2 quantum dots decorated on RGO: a superior sensitive, selective and reproducible performance for a H2 and LPG sensor. Nanoscale. 2015;7(28):11971–11979. doi: 10.1039/C5NR02837J
  • Li L, He S, Liu M, et al. Three-dimensional mesoporous graphene aerogel-supported SnO2 nanocrystals for high-performance NO2 gas sensing at low temperature. Anal Chem. 2015;87(3):1638–1645. doi: 10.1021/ac503234e
  • Zhang J, Chang L, Wang F, et al. Ultrafine SnO2 nanocrystals anchored graphene composites as anode material for lithium-ion batteries. Mater Res Bull. 2015;68:120–125. doi: 10.1016/j.materresbull.2015.03.041
  • Li Z, Wu G, Deng S, et al. Combination of uniform SnO2 nanocrystals with nitrogen doped graphene for high-performance lithium-ion batteries anode. Chem Eng J. 2016;283:1435–1442. doi: 10.1016/j.cej.2015.08.052
  • Wang Y, Jin Y, Zhao C, et al. SnO2 quantum dots/graphene aerogel composite as high-performance anode material for sodium ion batteries. Mater Lett. 2017;191:169–172. doi: 10.1016/j.matlet.2016.12.072
  • Zhang W, Li M, Xiao X, et al. In situ synthesis of ultrasmall SnO2 quantum dots on nitrogen-doped reduced graphene oxide composite as high performance anode material for lithium-ion batteries. J Alloys Compd. 2017;727:1–7. doi: 10.1016/j.jallcom.2017.04.316
  • Luo Y, Fan S, Luo Y, et al. Assembly of SnO2 quantum dots on RGO to form SnO2/N doped RGO as a high-capacity anode material for lithium ion batteries. CrystEngComm. 2015;17(8):1741–1744. doi: 10.1039/C4CE02315C
  • Yang Y, Ji X, Lu F, et al. The mechanistic exploration of porous activated graphene sheets-anchored SnO2 nanocrystals for application in high-performance Li-ion battery anodes. Phys Chem Chem Phys. 2013;15(36):15098–15105. doi: 10.1039/c3cp52808a
  • Zhou W, Wang J, Zhang F, et al. SnO2 nanocrystals anchored on N-doped graphene for high-performance lithium storage. Chem Commun. 2015;51(17):3660–3662. doi: 10.1039/C4CC08650C
  • Chen C, Wang L, Liu Y, et al. Assembling tin dioxide quantum dots to graphene nanosheets by a facile ultrasonic route. Langmuir. 2013;29(12):4111–4118. doi: 10.1021/la304753x
  • Ren J, Yang J, Abouimrane A, et al. SnO2 nanocrystals deposited on multiwalled carbon nanotubes with superior stability as anode material for Li-ion batteries. J Power Sources. 2011;196(20):8701–8705. doi: 10.1016/j.jpowsour.2011.06.036
  • Jin Y-H, Min K-M, Seo S-D, et al. Enhanced Li storage capacity in 3 nm diameter SnO2 nanocrystals firmly anchored on multiwalled carbon nanotubes. J Phys Chem C. 2011;115(44):22062–22067. doi: 10.1021/jp208021w
  • Song H, Li N, Cui H, et al. Monodisperse SnO2 nanocrystals functionalized multiwalled carbon nanotubes for large rate and long lifespan anode materials in lithium ion batteries. Electrochim Acta. 2014;120:46–51. doi: 10.1016/j.electacta.2013.12.052
  • Liu H, Zhang W, Yu H, et al. Solution-processed gas sensors employing SnO2 quantum dot/MWCNT nanocomposites. ACS Appl Mater Interfaces. 2016;8(1):840–846. doi: 10.1021/acsami.5b10188
  • Lu X, Wang H, Wang Z, et al. Room-temperature synthesis of colloidal SnO2 quantum dot solution and ex-situ deposition on carbon nanotubes as anode materials for lithium ion batteries. J Alloys Compd. 2016;680:109–115. doi: 10.1016/j.jallcom.2016.04.128
  • Jin R, Meng Y, Li G. Multiwalled carbon nanotubes@C@SnO2 quantum dots and SnO2 quantum dots@C as high rate anode materials for lithium-ion batteries. Appl Surf Sci. 2017;423:476–483. doi: 10.1016/j.apsusc.2017.06.215
  • Wang L, Wang D, Zhang FX, et al. Protein-inspired synthesis of SnO2nanocrystals with controlled carbon nanocoating as anode materials for lithium-ion battery. RSC Adv. 2013;3(5):1307–1310. doi: 10.1039/C2RA22375A
  • Hu R, Sun W, Zeng M, et al. Dispersing SnO2 nanocrystals in amorphous carbon as a cyclic durable anode material for lithium ion batteries. J Energy Chem. 2014;23(3):338–345. doi: 10.1016/S2095-4956(14)60156-X
  • Song H, Li N, Cui H, et al. Significantly improved high-rate Li-ion batteries anode by encapsulating tin dioxide nanocrystals into mesotunnels. CrystEngComm. 2013;15(42):8537–8543. doi: 10.1039/c3ce41410h
  • Yang J, Xi L, Tang J, et al. There-dimensional porous carbon network encapsulated SnO2 quantum dots as anode materials for high-rate lithium ion batteries. Electrochim Acta. 2016;217:274–282. doi: 10.1016/j.electacta.2016.09.086
  • Yu Z, Zhu S, Li Y, et al. Synthesis of SnO2 nanoparticles inside mesoporous carbon via a sonochemical method for highly reversible lithium batteries. Mater Lett. 2011;65:3072–3075. doi: 10.1016/j.matlet.2011.06.053
  • Wang X, Li Z, Yin L. Nanocomposites of SnO2@ordered mesoporous carbon (OMC) as anode materials for lithium-ion batteries with improved electrochemical performance. CrystEngComm. 2013;15(37):7589–7597. doi: 10.1039/c3ce41256c
  • Fan X, Shao J, Xiao X, et al. In situ synthesis of SnO2 nanoparticles encapsulated in micro/mesoporous carbon foam as a high-performance anode material for lithium ion batteries. J Mater Chem A. 2014;2(43):18367–18374. doi: 10.1039/C4TA04278F
  • Yang ZX, Du GD, Meng Q, et al. Dispersion of SnO2 nanocrystals on TiO2(B) nanowires as anode material for lithium ion battery applications. RSC Adv. 2011;1:1834–1840. doi: 10.1039/c1ra00500f
  • Du G, Guo Z, Zhang P, et al. SnO2 nanocrystals on self-organized TiO2 nanotube array as three-dimensional electrode for lithium ion microbatteries. J Mater Chem. 2010;20(27):5689–5694. doi: 10.1039/c0jm00330a
  • Wang J, Chen Z, Liu Y, et al. Heterojunctions and optical properties of ZnO/SnO2 nanocomposites adorned with quantum dots. Sol Energy Mater Sol Cells. 2014;128:254–259. doi: 10.1016/j.solmat.2014.05.038
  • Sahana MB, Sudakar C, Dixit A, et al. Quantum confinement effects and band gap engineering of SnO2 nanocrystals in a MgO matrix. Acta Mater. 2012;60(3):1072–1078. doi: 10.1016/j.actamat.2011.11.012
  • Han SY, Kim IY, Lee SH, et al. Electrochemically active nanocomposites of Li4Ti5O12 2D nanosheets and SnO2 0D nanocrystals with improved electrode performance. Electrochim Acta. 2012;74:59–64. doi: 10.1016/j.electacta.2012.03.175
  • Ma L, Xu L, Xu X, et al. Fabrication of SnO2/SnS2 hybrids by anchoring ultrafine SnO2 nanocrystals on SnS2 nanosheets and their photocatalytic properties. Ceram Int. 2016;42(4):5068–5074. doi: 10.1016/j.ceramint.2015.12.020
  • Cai J, Li Z, Yao S, et al. Close-packed SnO2 nanocrystals anchored on amorphous silica as a stable anode material for lithium-ion battery. Electrochim Acta. 2012;74:182–188. doi: 10.1016/j.electacta.2012.04.045
  • 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. doi: 10.1016/j.cej.2015.06.110
  • Du Z, Zhang S, Jiang T, et al. Facile synthesis of SnO2 nanocrystals coated conducting polymer nanowires for enhanced lithium storage. J Power Sources. 2012;219:199–203. doi: 10.1016/j.jpowsour.2012.07.052
  • Zhuang S, Xu X, Pang Y, et al. PEGME-bonded SnO2 quantum dots for excellent photocatalytic activity. RSC Adv. 2013;3(43):20422–20428. doi: 10.1039/c3ra42774a
  • Babu B, Cho M, Byon C, et al. Sunlight-driven photocatalytic activity of SnO2 QDs-g-C3N4 nanolayers. Mater Lett. 2018;212:327–331. doi: 10.1016/j.matlet.2017.10.110
  • Nath SS, Ganguly A, Gope G, et al. SnO2 quantum dots for nano light emitting devices. Nanosyst: Phys Chem Math. 2017;8:661–664.
  • Chen Z, Pan D, Li Z, et al. Recent advances in tin dioxide materials: some developments in thin films, nanowires, and nanorods. Chem Rev. 2014;114(15):7442–7486. doi: 10.1021/cr4007335
  • Wang H, Rogach AL. Hierarchical SnO2 nanostructures: recent advances in design, synthesis, and applications. Chem Mater. 2014;26(1):123–133. doi: 10.1021/cm4018248
  • Huang M, Feng M, Li H, et al. Rapid microwave-assisted synthesis of SnO2 quantum dots/reduced graphene oxide composite with its application in lithium-ion battery. Mater Lett. 2017;209:260–263. doi: 10.1016/j.matlet.2017.08.006
  • Mi H, Xu Y, Shi W, et al. Polymer-derived carbon nanofiber network supported SnO2 nanocrystals: a superior lithium secondary battery material. J Mater Chem. 2011;21(48):19302–19309. doi: 10.1039/c1jm12262b
  • Ding S, Luan D, Boey FY, et al. SnO2 nanosheets grown on graphene sheets with enhanced lithium storage properties. Chem Commun. 2011;47(25):7155–7157. doi: 10.1039/c1cc11968k
  • Jiang S, Zhao B, Ran R, et al. A freestanding composite film electrode stacked from hierarchical electrospun SnO2 nanorods and graphene sheets for reversible lithium storage. RSC Adv. 2014;4(18):9367–9371. doi: 10.1039/C3RA47840H
  • Xu C, Sun J, Gao L. Direct growth of monodisperse SnO2 nanorods on graphene as high capacity anode materials for lithium ion batteries. J Mater Chem. 2012;22(3):975–979. doi: 10.1039/C1JM14099J
  • Zhou X, Yin YX, Wan LJ, et al. A robust composite of SnO2 hollow nanospheres enwrapped by graphene as a high-capacity anode material for lithium-ion batteries. J Mater Chem. 2012;22(34):17456–17459. doi: 10.1039/c2jm32984k
  • Shahid M, Yesibolati N, Reuter MC, et al. Layer-by-layer assembled graphene-coated mesoporous SnO2 spheres as anodes for advanced Li-ion batteries. J Power Sources. 2014;263:239–245. doi: 10.1016/j.jpowsour.2014.03.146
  • Wu P, Xu X, Zhu Q, et al. Self-assembled graphene-wrapped SnO2 nanotubes nanohybrid as a high-performance anode material for lithium-ion batteries. J Alloys Compd. 2015;626:234–238. doi: 10.1016/j.jallcom.2014.12.037
  • Chen XB, Mao SS. Titanium dioxide nanomaterials: synthesis, properties, modifications, and applications. Chem Rev. 2007;107:2891–2959. doi: 10.1021/cr0500535
  • Wan X, Ma R, Tie S, et al. Effects of calcination temperatures and additives on the photodegradation of methylene blue by tin dioxide nanocrystals. Mater Sci Semicond Process. 2014;27:748–757. doi: 10.1016/j.mssp.2014.07.048
  • Jia ZQ, Sun HJ, Wang Y, et al. Facile synthesis of tin oxide nanocrystals and their photocatalytic activity. Trans Nonferrous Met Soc China. 2014;24(6):1813–1818. doi: 10.1016/S1003-6326(14)63258-1
  • Shajira PS, Prabhu VG, Bushiri MJ. Sunlight assisted photodegradation by tin oxide quantum dots. J Phys Chem Solids. 2015;87:244–252. doi: 10.1016/j.jpcs.2015.09.003
  • Begum S, Devi TB, Ahmaruzzaman M. Surfactant mediated facile fabrication of SnO2 quantum dots and their degradation behavior of humic acid. Mater Lett. 2016;185:123–126. doi: 10.1016/j.matlet.2016.07.028
  • Dai S, Yao Z. Synthesis of flower-like SnO2 single crystals and its enhanced photocatalytic activity. Appl Surf Sci. 2012;258(15):5703–5706. doi: 10.1016/j.apsusc.2012.02.065
  • Wang J, Fan HQ, Yu HW. Synthesis of hierarchical SnO2 microflowers assembled by nanosheets and their enhanced photocatalytic properties. Mater Trans. 2015;56:1911–1914. doi: 10.2320/matertrans.M2015287
  • Du F, Zuo X, Yang Q, et al. Facile assembly of TiO2 nanospheres/SnO2 quantum dots composites with excellent photocatalyst activity for the degradation of methyl orange. Ceram Int. 2016;42(11):12778–12782. doi: 10.1016/j.ceramint.2016.05.036
  • Zhu L, Wang M, Kwan Lam T, et al. Fast microwave-assisted synthesis of gas-sensing SnO2 quantum dots with high sensitivity. Sens Actuators B Chem. 2016;236:646–653. doi: 10.1016/j.snb.2016.04.173
  • He Y, Tang P, Li J, et al. Ultrafast response and recovery ethanol sensor based on SnO2 quantum dots. Mater Lett. 2016;165:50–54. doi: 10.1016/j.matlet.2015.11.092
  • Nemade KR, Waghuley SA. In situ synthesis of graphene/SnO2 quantum dots composites for chemiresistive gas sensing. Mater Sci Semicond Process. 2014;24:126–131. doi: 10.1016/j.mssp.2014.02.047
  • Lin L, Xing H, Shu R, et al. Preparation and microwave absorption properties of multi-walled carbon nanotubes decorated with Ni-doped SnO2 nanocrystals. RSC Adv. 2015;5(115):94539–94550. doi: 10.1039/C5RA17303E
  • Dutta D, Gupta J, Thakur D, et al. Magnetically engineered SnO2 quantum dots as a bimodal agent for optical and magnetic resonance imaging. Mater Res Express. 2017;4(12):125005. doi: 10.1088/2053-1591/aa9ac3