Next Article in Journal
Dynamic Mechanism of Dendrite Formation in Zhoukoudian, China
Previous Article in Journal
Exclusive Bus Lane Allocation Considering Multimodal Traffic Equity Based on Bi-Level Programming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Template-Free Synthesis of One-Dimensional SnO2 Nanostructures Using Highly Efficient Hydrothermal Method

1
Faculty of Materials Science and Engineering, Kunming University of Science and Technology, Kunming 650093, China
2
School of Chemistry, Biology and Environment, Yuxi Normal University, Yuxi 653100, China
3
State Key Laboratory of Advanced Technologies for Comprehensive Utilization of Platinum Metals, Kunming Institute of Precious Metals, Kunming 650106, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(4), 2048; https://doi.org/10.3390/app13042048
Submission received: 9 December 2022 / Revised: 21 January 2023 / Accepted: 22 January 2023 / Published: 4 February 2023
(This article belongs to the Section Nanotechnology and Applied Nanosciences)

Abstract

:
One-dimensional (1D) SnO2 nanostructures, as n-type semiconductors with a wide band gap, exhibit unique photoelectric properties that offer potential applications in electronic, photoelectric, gas sensing, and energy generation devices. A detailed study of template-free reaction systems is essential to regulate and efficiently synthesise 1DSnO2 nanostructures. This study employed the hydrothermal method to prepare 1DSnO2 nanostructures, with SnCl4·5H2O as the tin source. The impact of various experimental conditions on SnO2 morphology is analysed. Here, 1DSnO2 nanostructures were characterised by analytical methods such as X-ray powder diffraction, transmission electron microscopy, scanning electron microscopy, and field emission double-beam electron microscopy. The results confirmed the formation of 1DSnO2 nanostructures with a mixed morphology of nanorods and nanowires. The nanorods exhibited a length of 422.87 ± 110.15 nm, a width of 81.4 ± 16.75 nm, and an aspect ratio of 5:1, whereas the nanowires displayed a length of 200 ± 45.24 nm, a width of 15 ± 5.62 nm, and an aspect ratio of 13:1. With the addition of 50 mg of polyvinylpyrrolidone and seed crystal, the acquisition time of the 1DSnO2 nanostructures decreased from 48 to 12 h. The 1DSnO2 nanostructures were efficiently obtained without a template, laying the foundation for large-scale production and application.

1. Introduction

Metal oxides are ionic compounds composed of positive metal and negative oxygen ions. The completely filled s-shell of the metal oxides provides it with adequate thermal and chemical stability, whereas the d-shells may not be fully filled, which renders them with diverse unique properties. Owing to the quantification of state density, the electrical, optical, and chemical properties of high-aspect-ratio nanostructures and various material-dependent work functions can be modulated [1,2].
Nanostructure-based one-dimensional (1D) metal oxides have gained significant attention in recent years because of their unique electronic, optical, and thermal properties [3,4]. Moreover, because 1DSnO2 nanostructures are n-type semiconductors with a wide band gap (e.g., 3.6 eV at 300 K), they exhibit unique photoelectric properties that offer potential applications of SnO2 in electronic, photoelectric, gas sensing, and energy generation devices. The hydrothermal method is a bottom–up synthetic method that has become the most widely adopted method for preparing nanomaterials because of its facilely applicable procedure, the high crystallinity of the product, convenient control of morphology, and high sintering activity. Xi and Ye [5] adopted the hydrothermal method without exploiting any template or surfactant. They used SnCl4∙5H2O as the tin source and added CO(NH2)2 and HCl to ultrapure water, which was insulated at 90 °C for 24 h in a stainless-steel autoclave lined by polytetrafluoroethylene (PTFE). Ultrathin SnO2 nanorods (SnO2(NR)) with a diameter 2.0 ± 0.5 nm, a length 12 ± 3 nm, and a growth orientation in the [001] direction were obtained. Wang et al. [6] adopted the hydrothermal method to enable hydrothermal growth in a sealed PTFE-lined stainless-steel autoclave for 24 h at 200 °C. They obtained a rutile-structured SnO2(NR) of thickness 100 nm, length 1 mm, and a growth orientation in the [001] direction. In addition, Chen et al. [7] implemented the same experimental conditions and obtained rutile-structured SnO2(NR) with 4–15 nm diameter, 100–200 nm length, and (110)-directed growth orientation. Cheng et al. [8] implemented the hydrothermal method at 150 °C to induce hydrothermal growth for 24 h in a sealed PTFE-lined stainless-steel autoclave. They obtained SnO2(NR) with 2.5–5 nm diameter, 15–20 nm length, and [001]-directed growth orientation. Thereafter, Chen et al. [9] adopted the sodium dodecyl sulfate-assisted hydrothermal method to prepare nano-polyhedrons that were self-assembled from ultrathin nanowires in a sealed PTFE-lined stainless-steel autoclave at 180 °C after hydrothermal growth for 24 h. Qin et al. [10] applied the hydrothermal method at 285 °C to prompt hydrothermal growth for 24 h in a sealed PTFE-lined stainless-steel autoclave; they obtained rutile-structured SnO2(NR) with 80 ± 5 nm diameter, 2.5 ± 0.1 μm length, and [001]-directed growth orientation. Wang et al. [11] applied the hydrothermal method at 200 °C for 24 h in a sealed PTFE-lined stainless-steel autoclave to obtain a rutile-structured SnO2(NR) with 15 nm diameter and 50 nm length. Jeong et al. [12] placed a precursor solution with four pieces of SnO2 seed (SnO2 nanoparticles deposited using a spin-coating technique) layer-coated F-doped tin oxide in an autoclave to conduct hydrothermal synthesis at 200 °C for 18 h. They obtained rutile-structured SnO2(NR) with a diameter 85 nm and growth orientation in the [001] direction. Parveen et al. [13] implemented the hydrothermal method in an autoclave at 180 °C for 12 h, after which the SnO2 precipitate was calcined at 350 °C for 2 h. They obtained a rutile-structured SnO2(NR) with a diameter 250 nm and length 1 μm. Yadav et al. [14] applied the hydrothermal method at 180 °C for 24 h to obtain a rutile-structured SnO2(NR) with a diameter of 32.83 nm.
However, these methods require long processing times, which is not ideal for controlling 1DSnO2 structures and morphology. Thus, further investigations on such template-free reaction systems are vital for regulating the nanostructures of 1DSnO2. Although certain scholars [15,16,17] adopted template methods to prepare 1DSnO2, the high cost of the templates is not conducive for commercial production. Therefore, a highly efficient template-free hydrothermal synthesis of 1DSnO2 is reported here. The morphology and composition evolution of nanomaterials were investigated by adjusting the experimental parameters such as reaction period, temperature, and type of surfactant (polyvinylpyrrolidone; PVP) and seed crystals.

2. Materials and Methods

The materials used in the study are listed in Table 1.
In the experimental process, 1.05 g of stannic chloride pentahydrate and 1.3 g of sodium hydroxide were added into a sealed PTFE-lined stainless-steel autoclave along with a certain volume of absolute alcohol and deionised water (total volume: 80 mL). The solution reacted at a specified temperature for a reaction period. Here, SnO2 with varying morphologies was obtained after seal cooling, vacuum filtration, ethanol rinsing five times, and drying at 60 °C for 24 h. Subsequently, we investigated the structure and morphology evolution of the synthesised 1DSnO2 nanomaterials. For parameter optimisation in the reaction, the material preparation was conducted at 150, 160, 170, 180, 190, and 200 °C with adjustments in the volume ratios of the absolute alcohol and deionised water along with modifying the reaction period from 12–48 h.
The synthesised material was characterised by X-ray powder diffraction (XRD; Rigaku multi-bending X-ray diffractometer), transmission electron microscopy (TEM; JEOL 2100F; accelerating voltage: 200 kV), scanning electron microscopy (SEM; Phenom XL), and field emission double-beam electron microscopy (FESEM; FEI-Versa3D).

3. Results and Discussion

Experimental principles:
SnCl4 + 6NaOH = Na2Sn(OH)6 + 4NaCl
Na2Sn(OH)6 = SnO2 + 2NaOH + 2H2O
SnCl4∙5H2O dissolves rapidly when added to an aqueous solution of NaOH, which forms Na2Sn(OH)6 because of the quick hydrolysis reaction of Sn4+. When NaOH is excessive, Na2Sn(OH) exists as colloids in the aqueous solution of NaOH. Upon exposure to absolute alcohol, the colloidal particles of Na2Sn(OH)6 agglomerate to form pre-condensation nuclei in the hydrothermal process, which evolve into condensation nuclei for the growth of SnO2 crystals at high temperature and pressure.

3.1. Effects of Varying Reaction Times in Absolute Alcohol on SnO2 Morphology

When absolute alcohol was used as the solvent and PVP was not added, the reaction temperature was 200 °C, and the reaction periods were 8, 12, and 48 h.
As observed in Figure 1, the morphology evolution of the products indicated that small particles with smooth edges aggregate into large particle clusters after reacting for 8 h. Furthermore, irregular flake morphology appears in the vicinity of aggregated particle clusters after reacting for 12 h, whereas the product morphology is dominated by lamelliform structures without any 1D structures after reacting for 48 h. Because of the absence of templates, surfactant, and aqueous solution under this condition, the chemical potential in the solution was insufficient to trigger the nucleation and growth of 1D structures.

3.2. Effects of Varying Volume Ratios between Absolute Alcohol and Deionised Water on SnO2 Morphology

The effects of varying the volume ratios (V) between absolute alcohol and deionised water on SnO2 morphology were investigated at a reaction temperature of 200 °C for 48 h without adding PVP.
As observed in Figure 2, with a decreased content of absolute alcohol and an increased content of deionised water, the SnO2 morphology evolves from a lamelliform into a microsphere structure via self-assembled 1DSnO2. Eventually, these microspheres aggregate into a quasi-microsphere structure. Note that 1D structures appear in final products only if the value of V approaches one. Based on the aggregation degree of 1D structures self-assembled into quasi-microspheres, the edges of the quasi-microspheres in Figure 2e are more prominent at V = 1, which is distinct from the quasi-microsphere aggregation indicated in Figure 2d,f. The formed microstructure is relatively clear in Figure 2e for V = 1:1.

3.3. Effects of Reaction Time on SnO2 Morphology

The effects of the reaction time on SnO2 morphology were investigated by adding 1DSnO2, which was acquired by the same method as the seed crystal in Figure 2e: 200 °C, 50 mg PVP, and V = 1:1.
The PEG [18] and CTAB [19] were used as surfactants, the hydrothermal time was too long for 24h. However, the addition of PVP can inhibit the precipitation of solute on the particles, thereby inhibiting the growth of crystals and yielding smaller particles. So, we chose to use PVP as the surfactant. When the amount of PVP added was too low, the stability performance of small particles was weak [20]. However, the addition of too much PVP [21] (100 mg) had little effect on the morphology of 1DSnO2. Therefore, we chose to add 50mg of PVP. As the reaction time decreased, the product morphology displayed microsphere structures caused by 1DSnO2 self-assembly. The results indicated that the mutual adhesion between the particles absorbed on the surfaces of tin-oxide nanocrystals can be prevented by the effects of the surfactant PVP. Moreover, the presence of 1DSnO2 seed crystal favours the acceleration of 1DSnO2 nucleation and growth. Therefore, the required reaction time reduced from 48 h in Figure 3e to 12 h in Figure 3f.

3.4. Effects of Reaction Temperature on SnO2 Morphology

The effects of the reaction temperature on SnO2 morphology were investigated with 50 mg of PVP, a reaction time of 12 h, and V = 1:1.
As observed in Figure 4b–f, SnO2 morphology is dominated by lamelliform structures at high reaction temperatures. Meanwhile, microelements of 1DSnO2 nucleate at multiple sites on the flakes and grow via aggregation induced by the reaction conditions, thereby forming the microsphere structures of 1DSnO2 self-assembly. More such microspheres are formed as the reaction temperature increases. At 200 °C, the lamelliform morphology disappears and is replaced by the microspheres of 1DSnO2 self-assembly. Based on the morphology and structures, 1DSnO2 is formed at the reaction temperature of 200 °C.
Only 1DSnO2 were observed in Figure 4a. So, the optimum experimental conditions were at the reaction temperature of 200 °C, reaction time of 12 h, and V = 1:1. With the addition of 50 mg PVP, the product morphology exhibited microsphere structures of 1DSnO2 self-assembly. The composition and microstructure of 1DSnO2 nanocrystals are significantly affected by the reaction temperature. Below 200 °C, the condensation degree among Na2Sn(OH)6 mesophases is low, which may significantly decelerate the growth of 1DSnO2. However, by raising the temperature to 200 °C and reacting for 12 h, 100% 1DSnO2 can be formed. The conversion of Na2Sn(OH)6 into tin oxide is more apparent at a higher temperature, and the grain size of samples increases slightly upon increasing the hydrothermal temperature. This is because the pressure inside the reactor increases with the hydrothermal temperature and provides sufficient energy for grain growth. During the process, small grains are dissolved to increase the solute concentration, thereby facilitating the growth of large grains, which conforms to the mechanism of Ostwald ripening [22]. Ostwald ripening occurs via the dissolution and re-precipitation of matter at regions with small and large radii of curvatures, respectively. Ostwald ripening results in the dissolution of smaller solid grains, the diffusion of the solute throughout the liquid, and the re-precipitation of the solid into large grains. The net result is grain growth.
The XRD, TEM, and SEM results of 1DSnO2, prepared at the condition, are stated as follows. The diffraction peaks of the products appear sharp and free of impurity peaks, indicating superior crystallinity and high purity of the final products. Here, (110), (101), and (211) are the three relatively strong peaks. Compared to the standard card, the results are consistent with those of the tetragonal rutile SnO2 (JCPDS No. 41-1445, S.G.:P42/nm, a = b = 0.4738 nm, c = 0.3187 nm). Contrary to the standard card, the peak intensities of (101), (211), and (002) in the sample are significantly enhanced. According to Wang et al. [6], peak (002) is the characteristic peak of SnO2(NR) array, and its enhancement suggests an improvement in the degree of order in the sample. The crystal structure is demonstrated in Figure 5a. Each Sn coordinate contains six oxygen atoms at the vertices of the twisted octahedron. Among these, four oxygen atoms are in the same plane, and the Sn–O bonds are shorter than those of the other two oxygen atoms.
TEM analysis is conducted on the products, as illustrated in Figure 6.
As verified from the TEM and SEM images, the products displayed a morphology of 100% 1D nanostructures with a length of 422.87 ± 110.15 nm, width of 81.4 ± 16.75 nm, and aspect ratio of 5:1, indicating the prevalence of strictly anisotropic growth during the entire process. The SEM result in Figure 6a indicates that the front end of a single nanorod resembles a pyramid, which is consistent with the FESEM result illustrated in Figure 6c. The inset portrays that the spacing between adjacent lattice planes corresponding to the [110] plane of rutile tin oxides is 0.34 nm, which indicates that the preferential growth direction is [001]. Projecting outward from the centre of the nanolayers, Figure 6b,c demonstrate the quasi-microspheres with a diameter of 2 μm, which are formed by the self-assembly of 1D nanorods. The highly magnified SEM image further indicate that all nanorods resemble a square (Figure 6c). The growth model with thermodynamic stability of square tin-oxide nanowires (SnO2(NW)) in the solution is demonstrated in Figure 6f. The square tin-oxide nanorods grow from the centre of the quasi-microspheres, then elongate and protrude along the c-axis, thereby forming layered nanostructures. Figure 6 and Figure 7 show the different morphological types at different sites of same product. Figure 7 shows the morphologies of SnO2 nanowires. These nanowires are estimated to be about 200 ± 45.24 nm in length and 15 ± 5.62 nm in width, and the aspect ratio of the nanowires is up to 1:13.
The specific surface energy of crystalline SnO2 has been calculated earlier [23,24]. The same tendency is demonstrated in the results, i.e., the crystal surface energy increased in the order of (110) < (100) < (101) < (001) with incremental energy. As the lowest and highest surface energies appear on (110) and (001) surfaces, the growth in the [001] orientation yielded a higher aspect ratio of the particles than that in other orientations. The experimental results of SnO2(NR) obtained in this study are consistent with the calculation results, as single crystal nanorods grow along the [001] orientation of the long axis.
Thus, 1DSnO2 was prepared, adopting the hydrothermal method using SnCl4∙5H2O as the tin source, absolute alcohol and deionised water as the solvent, and PVP as the surfactant. Our hypothesis of the formation process is stated as follows. At the initial stage of the hydrothermal reaction, Na2Sn(OH)6 colloids are formed as condensation nuclei. As the temperature and pressure in the reaction system increases, small particles aggregate into large particles according to the mechanism of Ostwald ripening. The particle morphology evolves into lamelliform structures with temperature, pressure, and prolonged reaction period. As observed in the highly magnified SEM, small protrusions appear on the lamelliform morphology, which grow gradually during the process of hydrothermal crystallisation. The crystals will grow preferentially along the crystal plane with low nucleation energy, i.e., the growth rate of crystals is higher along the direction with lower nucleation energy on the corresponding crystal plane. Because the (110) plane of SnO2 crystals exhibits the lowest nucleation energy, the crystals will grow preferentially along the [001] orientation, thus forming nanorods and nanowires. Because of various interactions between nanocrystals [25] (such as van der Waals force between coated molecules, dipole interaction between nanocrystals, and hydrophilic/hydrophobic interaction), quasi-microsphere structures are formed via the self-assembly of nanorods. The high surface to volume ratio of nanomaterials can provide more positions for absorbing gas molecules. The 1DSnO2 nanostructures is used for preparing gas-sensitive components, which can improve the sensitivity and accuracy of gas detection. They are promising for further applications.

4. Conclusions

With absolute alcohol, without the addition of PVP, and after reaction for 8–48 h at a temperature of 200 °C, lamelliform SnO2 was observed in the products, but 1DSnO2 was not detected. Upon varying V from 5:3 to 9:7, the product morphology transformed from lamelliform to quasi-microsphere structures, a process facilitated by the 1DSnO2 self-assembly. Notably, at V = 3:5, the agglomeration of quasi-microspheres became severe. In the absence of templates, the time required to form 1DSnO2 decreased from 48 to 12 h with the addition of PVP and 1DSnO2 seed crystals, thereby enhancing the synthesis efficiency. The products are SnO2(NR) and SnO2(NW), with tetragonal rutile structures. The spacing between the adjacent lattice planes was 0.34 nm. This corresponds to the (110) plane of the rutile tin oxide, indicating that [001] is the preferential growth direction. Nanorods of length 422.87 ± 110.15 nm, width 81.4 ± 16.75 nm, and an aspect ratio of 5:1 were observed, in addition to nanowires of length ~200 ± 45.24 nm, width 15 ± 5.62 nm, and an aspect ratio of 13:1. Owing to the aggregation initiated by various interactions between nanocrystals, quasi-microsphere structures were formed by the self-assembly of 1DSnO2. The obtained 1DSnO2 nanostructures can be used to prepare gas-sensitive components.

Author Contributions

Conceptualization, J.Z. and H.L.; methodology, M.X.; validation, Y.C.; writing—original draft preparation, J.Z.; writing—review and editing, Y.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Technology Talent and Platform Project of Yunnan Province (2018HB110), Technology Talent and Platform Project of Yunnan Province (202105AE160027), and Yunan Provincial General Project of Applied Basic Research (2016FB091, 2017FB145).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Devan, R.S.; Patil, R.A.; Lin, J.H.; Ma, Y.R. One-dimensional metal-oxide nanostructures: Recent developments in synthesis, characterization, and applications. Adv. Funct. Mater. 2012, 22, 3326–3370. [Google Scholar] [CrossRef]
  2. Mathur, S.; Barth, S. One-dimensional semiconductor nanostructures: Growth, characterization and device applications. Z. Phys. Chem. 2008, 222, 307–317. [Google Scholar] [CrossRef]
  3. Huang, M.H.; Mao, S.; Feick, H.; Yan, H.; Wu, Y.; Kind, H.; Weber, E.; Russo, R.; Yang, P. Room-temperature ultraviolet nanowire nanolasers. Science 2001, 292, 1897–1899. [Google Scholar] [CrossRef]
  4. Zhao, Q.X.; Willander, M.; Morjan, R.E.; Hu, Q.H.; Campbell, E.E.B. Optical recombination of ZnO nanowires grown on sapphire and Si substrates. Appl. Phys. Lett. 2003, 83, 165–167. [Google Scholar] [CrossRef]
  5. Xi, G.; Ye, J. Ultrathin SnO2 nanorods: Template-and surfactant-free solution phase synthesis, growth mechanism, optical, gas-sensing, and surface adsorption properties. Inorg. Chem. 2010, 49, 2302–2309. [Google Scholar] [CrossRef]
  6. Wang, Y.L.; Guo, M.; Zhang, M.; Wang, X.D. Hydrothermal preparation and photoelectrochemical performance of size-controlled SnO2 nanorod arrays. CrystEngComm 2010, 12, 4024–4027. [Google Scholar] [CrossRef]
  7. Chen, Y.J.; Xue, X.Y.; Wang, Y.G.; Wang, T.H. Synthesis and ethanol sensing characteristics of single crystalline SnO2 nanorods. Appl. Phys. Lett. 2005, 87, 233503. [Google Scholar] [CrossRef]
  8. Cheng, B.; Russell, J.M.; Shi, W.; Zhang, L.; Samulski, E.T. Large-scale, solution-phase growth of single-crystalline SnO2 nanorods. J. Am. Chem. Soc. 2004, 126, 5972–5973. [Google Scholar] [CrossRef]
  9. Chen, D.; Xu, J.; Xie, Z.; Shen, G. Nanowires assembled SnO2 nanopolyhedrons with enhanced gas sensing properties. ACS Appl. Mater. Interfaces 2011, 3, 2112–2117. [Google Scholar] [CrossRef]
  10. Qin, L.; Xu, J.; Dong, X.; Pan, Q.; Cheng, Z.; Xiang, Q.; Li, F. The template-free synthesis of square-shaped SnO2 nanowires: The temperature effect and acetone gas sensors. Nanotechnology 2008, 19, 185705. [Google Scholar] [CrossRef]
  11. Wang, X.; Wang, S.; Tian, J.; Cui, H.; Wang, X. Synthesis of 1D SnO2 nanorods/2D NiO porous nanosheets pn heterostructures for enhanced ethanol gas sensing performance. Vacuum 2022, 205, 111399. [Google Scholar] [CrossRef]
  12. Jeong, Y.J.; Hong, S.Y.; Cho, I.S. All-solution-processed BiVO4/SnO2 nanorods-axial-heterostructure with improved charge collection properties for solar water-splitting. Ceram. Int. 2022, 48, 33101–33107. [Google Scholar] [CrossRef]
  13. Parveen, A.; Surumbarkuzhali, N.; Shkir, M.; Massoud, E.E.S.; Manjunath, V.; Ahn, C.H.; Park, S.H. Design of SnO2 nanorods/polypyrrole nanocomposite photocatalysts for photocatalytic activity towards various organic pollutants under the visible light irradiation. Inorg. Chem. Commun. 2022, 142, 109685. [Google Scholar] [CrossRef]
  14. Yadav, V.; Singh, N.; Meena, D. Investigation of structural and optical properties of pure SnO2, ZnO and SnO2/ZnO composite nanorods. Mater. Today Proc. 2022, 62, 3368–3375. [Google Scholar] [CrossRef]
  15. Zhu, W.; Wang, W.; Xu, H.; Shi, J. Fabrication of ordered SnO2 nanotube arrays via a template route. Mater. Chem. Phys. 2006, 99, 127–130. [Google Scholar] [CrossRef]
  16. Meng, X.; Zhang, Y.; Sun, S.; Li, R.; Sun, X. Three growth modes and mechanisms for highly structure-tunable SnO2 nanotube arrays of template-directed atomic layer deposition. J. Mater. Chem. 2011, 21, 12321–12330. [Google Scholar] [CrossRef]
  17. Lai, M.; Martinez, J.A.G.; Grätzel, M.; Riley, D.J. Preparation of tin dioxide nanotubes via electrosynthesis in a template. J. Mater. Chem. 2006, 16, 2843–2845. [Google Scholar] [CrossRef]
  18. Zhou, X.; Fu, W.; Yang, H.; Ma, D.; Cao, J.; Leng, Y.; Guo, J.; Zhang, Y.; Sui, Y.; Zhao, W.; et al. Synthesis and ethanol-sensing properties of flowerlike SnO2 nanorods bundles by poly (ethylene glycol)-assisted hydrothermal process. Mater. Chem. Phys. 2010, 124, 614–618. [Google Scholar] [CrossRef]
  19. Chen, D.; Gao, L. Facile synthesis of single-crystal tin oxide nanorods with tunable dimensions via hydrothermal process. Chem. Phy. Lett. 2004, 398, 201–206. [Google Scholar] [CrossRef]
  20. Shen, X.S. Shape-Controlled Synthesis, Self-Assembly and Sers Properties of Noble Metal Nanoparticles. Ph.D. Thesis, University of Science and Technology of China, Hefei, China, 2010. [Google Scholar]
  21. Lin, Z.J. Microstructure Control and Properties of Ag-SnO2 and Ag-Ni Electrical Contact Materials. Ph.D. Thesis, Northeastern University, Shenyang, China, 2016. [Google Scholar]
  22. Wang, W.S.; Zhen, L.; Xu, C.; Shao, W. Hollow inorganic micro/nanostructures synthesized by solution-phase method. Prog. Chem. 2008, 20, 679–689. [Google Scholar]
  23. Slater, B.; Catlow, C.R.A.; Gay, D.H.; Williams, D.E.; Dusastre, V. Study of surface segregation of antimony on SnO2 surfaces by computer simulation techniques. J. Phys. Chem. B 1999, 103, 10644–10650. [Google Scholar] [CrossRef]
  24. Oviedo, J.; Gillan, M.J. Energetics and structure of stoichiometric SnO2 surfaces studied by first-principles calculations. Surf. Sci. 2000, 463, 93–101. [Google Scholar] [CrossRef]
  25. Peng, Q.; Li, Y.D. Controlled synthesis, assembly of functional nanomaterials and their properties exploration based on structures. Sci. China Chem. 2009, 39, 1028–1052. [Google Scholar]
Figure 1. SnO2 morphology at 200 °C, with 80 mL absolute alcohol, without polyvinylpyrrolidone (PVP), and reaction period of (a) 8, (b) 12, and (c) 48 h.
Figure 1. SnO2 morphology at 200 °C, with 80 mL absolute alcohol, without polyvinylpyrrolidone (PVP), and reaction period of (a) 8, (b) 12, and (c) 48 h.
Applsci 13 02048 g001
Figure 2. SnO2 morphology at 200 °C, reacting for 48 h without PVP: (a) absolute alcohol: 80 mL; (b) ratio of absolute alcohol to denoised water (V) = 3:1; (c) V = 5:3; (d) V = 9:7; (e) V = 1:1; (f) V = 3:5.
Figure 2. SnO2 morphology at 200 °C, reacting for 48 h without PVP: (a) absolute alcohol: 80 mL; (b) ratio of absolute alcohol to denoised water (V) = 3:1; (c) V = 5:3; (d) V = 9:7; (e) V = 1:1; (f) V = 3:5.
Applsci 13 02048 g002
Figure 3. SnO2 morphology at 200 °C, with 50 mg PVP, V = 1:1, and reacting for (a) 48, (b) 36, (c) 30, (d) 24, (e) 18, and (f) 12 h.
Figure 3. SnO2 morphology at 200 °C, with 50 mg PVP, V = 1:1, and reacting for (a) 48, (b) 36, (c) 30, (d) 24, (e) 18, and (f) 12 h.
Applsci 13 02048 g003
Figure 4. SnO2 morphology with 50 mg PVP, reacting for 12 h, and V = 1:1: (a) 200, (b) 190, (c) 180, (d) 170, (e) 160, and (f) 150 °C.
Figure 4. SnO2 morphology with 50 mg PVP, reacting for 12 h, and V = 1:1: (a) 200, (b) 190, (c) 180, (d) 170, (e) 160, and (f) 150 °C.
Applsci 13 02048 g004
Figure 5. (a) Crystal structure and (b) XRD pattern of (1D)SnO2 prepared via the developed method at 200 °C for 12 h.
Figure 5. (a) Crystal structure and (b) XRD pattern of (1D)SnO2 prepared via the developed method at 200 °C for 12 h.
Applsci 13 02048 g005
Figure 6. TEM and FESEM images of SnO2(NR): (a) low-magnification TEM image of single SnO2(NR), (inset) calibration of Fourier transform mode and interplanar crystal spacing; (b) 1 μm FESEM image; (c) 500 nm FESEM image. Frequency statistics charts of (d) nanorod length and (e) nanorod width. (f) Growth model of synthesised SnO2 nanorods with thermodynamic stability.
Figure 6. TEM and FESEM images of SnO2(NR): (a) low-magnification TEM image of single SnO2(NR), (inset) calibration of Fourier transform mode and interplanar crystal spacing; (b) 1 μm FESEM image; (c) 500 nm FESEM image. Frequency statistics charts of (d) nanorod length and (e) nanorod width. (f) Growth model of synthesised SnO2 nanorods with thermodynamic stability.
Applsci 13 02048 g006
Figure 7. FSEM images of SnO2 (NW): (a) 500 nm, (b) 200 nm.
Figure 7. FSEM images of SnO2 (NW): (a) 500 nm, (b) 200 nm.
Applsci 13 02048 g007
Table 1. Raw materials and reagents.
Table 1. Raw materials and reagents.
NameGradeManufacturer
Stannic chloride pentahydrate (SnCl5H2O)Analytically pureSinopharm Chemical Reagent Co., Ltd. (Shanghai, China)
Sodium hydroxide (NaOH)Analytically pureSinopharm Chemical Reagent Co., Ltd. (Shanghai, China)
Absolute alcohol (C2H5OH)Analytically pureSinopharm Chemical Reagent Co., Ltd. (Shanghai, China)
Deionised water (H2O) Self-manufactured
Polyvinylpyrrolidone (PVP-k30)Analytically pureSinopharm Chemical Reagent Co., Ltd. (Shanghai, China)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, J.; Li, H.; Chen, Y.; Xie, M.; Bi, Y. Template-Free Synthesis of One-Dimensional SnO2 Nanostructures Using Highly Efficient Hydrothermal Method. Appl. Sci. 2023, 13, 2048. https://doi.org/10.3390/app13042048

AMA Style

Zhao J, Li H, Chen Y, Xie M, Bi Y. Template-Free Synthesis of One-Dimensional SnO2 Nanostructures Using Highly Efficient Hydrothermal Method. Applied Sciences. 2023; 13(4):2048. https://doi.org/10.3390/app13042048

Chicago/Turabian Style

Zhao, Jingchen, Hongmei Li, Yongtai Chen, Ming Xie, and Yanan Bi. 2023. "Template-Free Synthesis of One-Dimensional SnO2 Nanostructures Using Highly Efficient Hydrothermal Method" Applied Sciences 13, no. 4: 2048. https://doi.org/10.3390/app13042048

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop