Next Article in Journal
The Effect of Lath Martensite Microstructures on the Strength of Medium-Carbon Low-Alloy Steel
Previous Article in Journal
Halogen Bonding Provides Heterooctameric Supramolecular Aggregation of Diaryliodonium Thiocyanate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Environmentally Friendly and Cost-Effective Synthesis of Carbonaceous Particles for Preparing Hollow SnO2 Nanospheres and their Bifunctional Li-Storage and Gas-Sensing Properties

1
Key Laboratory of Functional Molecular Solids, Ministry of Education, College of Chemistry and Materials Science, Anhui Normal University, Wuhu, Anhui 241000, China
2
Institute of Intelligent Machines, Chinese Academy of Sciences, Hefei 230031, China
3
Department of Chemistry, University of Science and Technology of China, Hefei 230026, China
*
Authors to whom correspondence should be addressed.
Crystals 2020, 10(3), 231; https://doi.org/10.3390/cryst10030231
Submission received: 9 March 2020 / Revised: 20 March 2020 / Accepted: 21 March 2020 / Published: 23 March 2020
(This article belongs to the Section Crystalline Materials)

Abstract

:
The templated preparation of hollow nanomaterials has received broad attention. However, many templates are expansive, environmentally-harmful, along with involving a complicated preparation process. Herein, we present a cost-effective, environmentally friendly and simple approach for making carbonaceous particles which have been demonstrated as efficient templates for preparing hollow nanospheres. Natural biomass, such as wheat or corn, is used as the source only, and thus other chemicals are not needed. The carbonaceous particles possess abundant hydroxyl and carboxyl groups, enabling them to efficiently adsorb metal ions in solution. The prepared SnO2 hollow spheres were used in a lithium-ion (Li-ion) battery anode, and as the sensing layer of a gas sensor, respectively. After charge–discharge for 200 times at a rate of 1 C, the anodes exhibit a stable capacity of 500 mAh g−1, and a Coulombic efficiency as high as 99%. In addition, the gas sensor based on the SnO2 hollow spheres shows a high sensing performance towards ethanol gas. It is expected that the presented natural biomass-derived particles and their green preparation method will find more applications for broad research fields, including energy-storage and sensors.

1. Introduction

Since the discovery of fullerenes and carbon nanotubes, carbonaceous materials have attracted considerable attention, owing to their fascinating properties and applications [1,2,3]. Among them, particles as an important member of the carbonaceous family have received more and more interest because of their wide applications in many fields, such as lubricants and catalyst supports [4,5,6]. In addition, carbonaceous particles can be applied as templates for fabricating a large variety of hollow and porous spheres, including metal oxides (e.g., ZnO, TiO2, SnO2, CeO2, etc.) which could greatly enhance their performance in electronic devices [7,8,9,10,11,12]. Recently, the preparation of carbonaceous particles, and further exploring their potential applications, have attracted extensive attention [13,14]. Several methods have been investigated, e.g., chemical vapor deposition, a hydrothermal approach, the carbonization of polymers and pyrolysis [15,16]. Regarding to the environment issues, “green” approaches have been of great interest. Several efforts focus on using environmentally friendly reagents and reducing the preparation steps. For instance, Zhang et al. reported a multiscale recombined porous Si/C composites, prepared via a simple Ag-assisted chemical etching process [17]. In their study, chitosan was chosen as the carbon source, which was nontoxic and biodegradable.
Xiao and co-workers showed a convenient and green strategy for the synthesis of highly luminescent and water-soluble carbon dots, by carbonizing precursors, such as Bovine serum albumin nanoparticles, in water solution [18]. Interestingly, Hao et al. developed a one-pot green synthesis of luminescent carbon dots and carbon-coated metal (Au, Pt, Pd) particles by using the vitamin B2 salt as the carbon source and reducing agent [19]. In addition, Ren et al. in situ fabricated Ru, Fe-containing zinc–trimesic acid metal organic framework fibers, then they were able to get the RuFe@Fe2O3 particles confined into mesoporous carbon nanofibers by using one-step pyrolysis [20]. The thermal treatment and chemicals used for reactions were simplified significantly. Similar green approaches were also widely studied [21]. In spite of those achievements, high temperature, expensive instruments and chemicals, and/or toxic reagents are commonly required. Taking the commonly used polymethyl methacrylate (PMMA) particle template as an example, it is usually prepared through a polymerization route [22]. Usually, methyl methacrylate is used as the polymer monomer, while an initiator (such as potassium peroxodisulfate) and an emulgator (sodium dodecyl benzene sulfonate) are also added into an emulsion system. Under controlled polymerization temperature and stirring speed, PMMA particles are obtained. As seen, the procedure is high-cost, and multiple, complicated steps with poisonous chemicals are needed. Currently, it requires a preparation route of template particles, which is cost-effective, simple and environmentally friendly.
Herein, we present a natural biomass-derived approach that is cost-effective, environmentally friendly and simple for the preparation of carbonaceous particles. The particles were further used as templates to prepare hollow metal oxide spheres. In our investigation, natural wheat and corn were used as our carbon source, respectively, as shown in Figure 1. Compared to the conventional PMMA templates and their preparation approach indicated above, the presented route is completely “green”, in which other chemicals are unnecessary during the hydrothermal synthesis. The carbon sources are low-cost; and the route is simple without complicated procedures and instruments. The carbonaceous particles were used as sacrificing templates for the fabrication of porous and hollow SnO2, which is a representative transition metal oxide. The SnO2 spheres-based Li-ion battery anodes exhibit a stable capacity of 500 mAh g−1 when cycling at a rate of 1 C for 200 times, along with a stable Coulombic efficiency of about 99%. In addition, the sensor based on the SnO2 possesses a good gas-sensing performance towards ethanol gas.

2. Experimental

2.1. Reagents and Solutions

All chemicals used in the study were of analytical grade, and were used without further purification. The wheat and corn were purchased from the China Oil and Foodstuffs Corporation (COFCO Corp.). Tin(IV) chloride (SnCl4·5H2O) and chemicals used in sensors and batteries were bought from the Sinopharm Chemical Reagent Company.

2.2. Preparation of the Carbonaceous Particles

Firstly, wheat or corn was mechanically milled into powders (ca. 500 mesh). Then, 0.1 g of the wheat powders (or corn powders) were dispersed in 40 mL of distilled water by ultrasonication. The solution was transferred into a Teflon-sealed autoclave and maintained at 180 °C for 6 h. After the autoclave was cooled down naturally to room temperature, the products were collected by using centrifugation, and washed several times with distilled water. At last, the products were dried in an oven at 60 °C for further use.

2.3. Preparation of the Hollow SnO2 Spheres

In order to prepare the porous hollow metal oxides by using the synthesized particles, 0.1 g of the carbonaceous particles were dispersed in 10 mL of ethanol by ultrasonication. Then, 1 mmol of SnCl4·5H2O was dissolved into the mixture obtained above. After being left statically at room temperature for 24 h, the precipitate was collected by centrifugation. The samples were put into an oven for drying at 50 °C. At last, the dried powders were annealed in a furnace at 500 °C for 1 h to remove the carbonaceous particles.

2.4. Characterizations

The samples were characterized on a Sirion−200 field emission scanning electron microscopy (FESEM, FEI, USA), a H-800 transmission electron microscopy (TEM, Hitachi, Japan) and an ESCALab MK II X-ray photoelectron spectrometer (XPS, Thermo Scientific, USA). The SEM was operated at an accelerating voltage of 5 kV. All of the samples were dispersed on Si substrates. For the TEM characterizations, the samples were dispersed in ethanol, then dropped onto Cu grids for further observation. The accelerating voltage of TEM was 200 kV. During XPS measurements, the binding energies in the XPS spectrum were calibrated by using that of C 1s (284.6 eV). The Fourier transform infrared (FTIR) spectrum was recorded on a IR-440 spectrometer (Shimadzu, Japan) from 4000 to 500 cm−1. The samples (powders) were mixed with KBr, then ground and pressed into a small piece for measurement.

2.5. Energy-Storage Performance Study

The electrochemical properties of the prepared SnO2 hollow spheres were measured by using a CR2032 coin-cell system. The electrode prepared by mixing SnO2 (80 wt%), carbon black (10 wt%) and polyvinylidene fluoride binder (10 wt%) dissolved in N-methyl pyrrolidinone to form a slurry, which was then coated onto a Cu foil and dried in a vacuum at 110 °C. The substrates were tailored into circular pieces with a 14 mm-diameter. The SnO2 anodes were set as working electrodes. A microporous polypropylene film was employed in the coin cell as the separator. Li foil was used as the counter electrode. Cells were assembled in a Super 1220/750 glovebox (Mikrouna, Germany) filled with Ar gas (O2 and H2O < 0.01 ppm). 1 M of LiPF6 in a mixture of EC/DEC (1:1 by volume) was used as the electrolyte. The tests were conducted through the galvanostatic discharge–charge method on a CT-4008 battery tester (Shenzhen Neware Technology Co., Ltd, China).

2.6. Gas Sensor Fabrication and Gas-Sensing Property Measurements

At first, the gas sensors using the obtained SnO2 as the sensing layer were constructed. The SnO2 spheres were dispersed in ethanol by using ultrasonication for 10 min. Then, the slurry was coated onto the surface of an Al2O3 tube in which there was a Ni–Cr wire used as a heater. The sensors were dried at 50 °C under vacuum. Before each gas-sensing measurement, the sensors were aged at working conditions for two days, so that a stable and reliable performance could be obtained. During the gas-sensing tests, ethanol gas at concentrations of 10 ppm, 20 ppm, 30 ppm, 40 ppm, 150 ppm, 200 ppm and 250 ppm were employed as targets, respectively. All gas-sensing measurements were conducted at the same working temperature of gas sensors (∼200 °C) and humidity of 60% RH. The tests were conducted by using a computer-controlled gas detecting system in which a Keithley 6487 picoameter/voltage sourcemeter was used as both the current recorder and power source.

3. Results and Discussion

The FESEM images of the samples prepared by using wheat and corn are presented in Figure 2a,c, respectively. The wheat- and corn-derived samples are similar in morphology, which is spherical with a narrow size distribution from 650 to 800 nm. This is confirmed by the TEM photographs, as shown in Figure 2b,d. The surface state and functional groups of particles were investigated by using XPS and FTIR. Taking the wheat-derived carbonaceous particle as an example, in the XPS spectrum (Figure 3a), the peaks at 284.6 eV, 285.9 eV, 287.1 eV and 288.5 eV are assigned to the carbon atoms bonded to graphitized carbon, C–OH groups, C–O–C and C=O groups, and O=C–O groups, respectively [23,24,25]. The surface of the carbonaceous particles is hydrophilic, because of the existence of the C–OH and O=C–O groups, and thus the products are unnecessary to be further functionalized before their use as templates to adsorb metal ions. Moreover, aromatization under the hydrothermal condition is demonstrated. It is further supported by the FTIR spectrum (Figure 3b), in which the carbonyl and hydroxyl groups were detected. In Figure 3b, the peak at 3413 cm−1 is assigned to the O–H; while the ones at 1308 cm−1 and 1271 cm−1 correspond to the C–H group. Moreover, the peaks at 1700 cm−1 and 1608 cm−1 are indexed to the C=O and C=C [26,27], respectively, which are in good agreement with the XPS results shown above.
On the basis of the study on XPS and FTIR spectra, the formation mechanism of carbonaceous particles derived from natural biomasses is explained as follows: Firstly, both wheat and corn are consisting of starch dominantly together, with a few saccharine components. Among various components in wheat, this starch and maltose are overwhelming, with percentages of ca. 75% and 8%, respectively. Under hydrothermal conditions, starch and saccharine are released into the solution. The dehydration and aromatization would occur, which are similar to the glucose [28,29], resulting in the formation of carbonaceous particles. In order to confirm this, pure starch and maltose were also employed as raw materials. The hydrothermal conditions are the same as those for wheat and corn. Figure 2e,f show the FESEM images of samples prepared from pure starch and maltose, respectively. They are close to the ones synthesized from wheat, supporting the suggested mechanism that starch and saccharine components would contribute to the formation of the carbonaceous particles.
Figure 4 shows the morphology and structure of the SnO2 hollow spheres prepared by using the wheat-derived carbonaceous particle templates. It should be indicated that those SnO2 spheres are presented as a demonstrating example. Under the same preparation procedures, similar SnO2 spheres were also obtained by using the corn-derived carbonaceous particles. In Figure 4a, the samples exhibit a porous spherical morphology with a diameter of about 300 nm. Compared to the carbonaceous templates, the size of the hollow spheres is small. It is ascribed to the shrink of the carbonaceous particles during annealing, which is similar to some other reports [30,31]. The shell of the hollow spheres is assembled by nanoparticles, as seen from the high-magnification FESEM image (Figure 4b). The particles are about 20 nm-large without obvious morphology. In the TEM image (Figure 4c), the samples display a hollow structure with a shell thickness of ca. 20 nm, which is close to the size of the particles. It indicates that the shell of the spheres is assembled by monolayered nanoparticles. In order to confirm if there is some carbon residual within the SnO2 spheres, we have used the pure carbonaceous particles (without adsorbing Sn(IV) ions) for calcination under the same annealing condition as the SnO2 spheres. After the thermal treatment, no residual in the ceramic boat was observed. Therefore, it is considered that there is no residual carbon in the SnO2 spheres after calcination.
Figure 5 shows the charge–discharge performance of the SnO2 hollow spheres-based anodes. In Figure 5a, the plateaus at the discharge and charge curves are ascribed to the lithiation and delithiation. On the basis of the reaction mechanism of SnO2, an alloying–dealloying mechanism would be dominated from the second cycle [32]. Cycling at a rate of 1 C (1 C equals to the complete charge or discharge of theoretical capacity of 782 mAh g−1 in 1 h), the SnO2 anode exhibits a high first-cycle capacity of about 1200 mAh g−1, which is ascribed to the formation of an unstable solid electrolyte interphase (SEI) [33,34]. In Figure 5b, during the subsequent cycles, the capacity decreases gradually. However, from the 30th cycle, the capacity becomes stable, indicating the formation of a stable SEI. After cycling for 200 times, the capacity remains exceeding 500 mAh g−1, along with a high Coulombic efficiency of about 99%, indicating a good reversibility. This is ascribed to the hollow structure which enables the volume-change to be accommodated [35,36].
Figure 6 displays the gas-sensing responses of the sensor towards ethanol gases. In Figure 6a, once the ethanol gas was injected into the detecting chamber, the current increased rapidly, indicating a short response time, which was commonly defined as the time for the sensor to achieve 90% of the total resistance-change. As seen from the real-time sensing curves, the response times of the sensor are about 8.5 s, 6.5 s, 4.5 s and 5 s for 10 ppm, 20 ppm, 30 ppm and 40 ppm ethanol gases, respectively. The fast response is attributed to the porous and hollow morphology and the nanospheres assembled by small nanoparticles. As for the sensing mechanism, this is explained as follows: In air ambient conditions, the oxygen is adsorbed on SnO2, forming some negative ionic species, like O and O2−. Because of that, the electrons in SnO2 would be captured, resulting in a high resistance. When ethanol gas is injected into the chamber, it will be oxidized by those ionic species on SnO2, releasing the electrons and exhibiting a low resistance. In contrast, when the fresh gas was introduced to replace the target ethanol gas, the reversible reactions happened, resulting in the recovery of the current. For instance, for detecting 20 ppm ethanol, the resistance of the fresh gas sensor was 2.73 × 106 Ω. While the target gas was injected into the chamber, the resistance decreased, and reached to about 0.21 × 106 Ω at the equilibrium stage.
In Figure 6b, the gas response is defined as the ratio of the current in pure air (𝐼air) and in the gas mixture (𝐼gas) at a constant voltage, which also reflects the change of resistances. In our investigation, the gas responses are about 7, 13, 22 and 39, when the concentrations of ethanol gases are 10 ppm, 20 ppm, 30 ppm and 40 ppm, respectively. The sensing response is competitive with some other materials, such as the TiO2/SnO2 nanofibers [37] and micro-/mesoporous SnO2 spheres prepared by a solvothermal method [38]. It is ascribed to the nanosphere with a hollow and porous structure which provides a good contact and numerous sensing sites for gas molecules. In addition, during target gas injecting and releasing, the diffusion kinetics of the hollow spheres is fast, enabling short response and recovery times. Moreover, it is noted that the response curve is not in a linear profile. It is ascribed to this that the gas concentrations in the measurements would be not in the linear region of the sensing performance. Since that, some relatively higher concentration gases were detected, as shown in Figure 6c,d. The responses of the sensor towards 200 ppm and 250 ppm gases become similar, indicating a saturation trend. Furthermore, compared to the sensing responses towards lower concentrations (10–40 ppm), it indicates that the sensor has a linear sensing range towards ethanol from about 30 to 150 ppm. It is expected that further investigations on humidity and stability would also be a benefit for the practical applications of the presented gas sensor.

4. Conclusions

In summary, we present a cost-effective, environmentally friendly, and simple approach for preparing carbonaceous particles by using natural biomasses including wheat and corn. The particles can be easily used as templates for constructing hollow nanomaterials for energy-storage and gas-sensing applications. The abundant hydroxyl and carboxyl groups on the carbonaceous particles enable them to be suitable templates for the preparation of hollow nanostructures. The SnO2 hollow spheres were prepared by using the carbonaceous particles as a demonstrating example. The SnO2 anode-based Li-ion batteries exhibit a stable capacity of 500 mAh g−1, and a Coulombic efficiency as high as 99% after charge–discharge 200 times at a rate of 1 C. In addition, the SnO2-based gas sensor shows high response and short responding time towards ethanol. We believe that the presented wheat or corn-derived method for preparing carbonaceous particles could be potentially extended to many other natural biomasses. In addition, it is expected that the prepared carbonaceous particles are able to find more applications for a broad set of research fields.

Author Contributions

Investigation, Y.D., and J.L.; resources, T.H., and J.L.; writing—original draft preparation, Y.D., P.Z., T.H., and J.L.; supervision, J.L.; project administration, T.H., and J.L.; funding acquisition, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (grant number: 51672176), Science and Technology Major Project of Anhui Province (grant number: 18030901093), Key Research and Development Program of Wuhu (grant number: 2019YF07), Natural Science Research Project for Universities in Anhui Province (grant numbers: KJ2018ZD034 and KJ2019A0502), and Anhui Provincial Program for Innovation and Entrepreneurship of Returnees from Overseas (grant number: 2019LCX005). The APC was funded by Science and Technology Major Project of Anhui Province (grant number: 18030901093).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, Z.F.; Tian, J.R.; Yin, Z.F.; Cui, C.J.; Qian, W. Carbon nanotube- and graphene-based nanomaterials and applications in high-voltage supercapacitor: A review. Carbon 2019, 141, 467–480. [Google Scholar] [CrossRef]
  2. He, Y.; Zhao, L.D.; Wang, X.J.; Liu, L.Y.; Liu, H. Microstructured hybrid nanocomposite flexible piezoresistive sensor and its sensitivity analysis by mechanical finite-element simulation. Nanotechnology 2020, 31, 185502. [Google Scholar] [CrossRef] [PubMed]
  3. Kim, T.; Cho, M.; Yu, K.J. Flexible and stretchable bio-tntegrated electronics based on carbon nanotube and graphene. Materials 2018, 11, 1163. [Google Scholar] [CrossRef] [Green Version]
  4. He, X.Z. Techno-economic feasibility analysis on carbon membranes for hydrogen purification. Sep. Purif. Techbnol. 2017, 186, 117–124. [Google Scholar] [CrossRef]
  5. Yan, D.Y.; Hu, H.; Gao, N.Y.; Ye, J.S.; Ou, H.S. Fabrication of carbon nanotube functionalized MIL-101(Fe) for enhanced visible-light photocatalysis of ciprofloxacin in aqueous solution. Appl. Surf. Sci. 2019, 498, 143836. [Google Scholar] [CrossRef]
  6. Zhang, Y.X.; Cai, T.; Shang, W.J.; Liu, D.; Guo, Q.; Liu, S.G. Facile synthesis of photoluminescent inorganic-organic hybrid carbon dots codoped with B and N: Towards an efficient lubrication additive. Dalton Trans. 2017, 46, 12306–12312. [Google Scholar] [CrossRef]
  7. Nowak, A.P. Composites of tin oxide and different carbonaceous materials as negative electrodes in lithium-ion batteries. J. Solid State Electrochem. 2018, 22, 2297–2304. [Google Scholar] [CrossRef]
  8. Xiao, Q.; Wang, K.; Wang, X.X.; Huang, S.P.; Cai, N.N.; Li, N. Solvent-free template synthesis of SnO2/C hybrid hollow spheres for superior lithium-sulfur batteries. Mater. Chem. Phys. 2020, 239, 122070. [Google Scholar] [CrossRef]
  9. Kang, Y.R.; Li, Z.; Xu, K.; He, X.J.; Wei, S.X.; Cao, Y.M. Hollow SnO2 nanospheres with single-shelled structure and the application for supercapacitors. J. Alloy. Compd. 2019, 779, 728–734. [Google Scholar] [CrossRef]
  10. Hu, Q.; Huang, B.Y.; Li, Y.; Zhang, S.M.; Zhang, Y.X.; Hua, X.H.; Liu, G.; Li, B.S.; Zhou, J.Y.; Xie, E.Q.; et al. Methanol gas detection of electrospun CeO2 nanofibers by regulating Ce3+/Ce4+ mole ratio via Pd doping. Sens. Actuators B 2020, 307, 127638. [Google Scholar] [CrossRef]
  11. Liu, H.J.; Zhang, S.; Chen, Y.L.; Zhang, J.T.; Guo, P.; Liu, M.; Lu, X.Q.; Zhang, J.; Wang, Z.J. Rational design of TiO2@nitrogen-doped carbon coaxial nanotubes as anode for advanced lithium ion batteries. Appl. Surf. Sci. 2018, 458, 1018–1025. [Google Scholar] [CrossRef]
  12. Jin, L.F.; Chen, W.G.; Zhang, H.; Xiao, G.W.; Yu, C.T.; Zhou, Q. Characterization of reduced graphene oxide (rGO)-loaded SnO2 nanocomposite and applications in C2H2 gas detection. Appl. Sci. 2017, 7, 19. [Google Scholar] [CrossRef]
  13. Hu, L.J.; Qian, Z.X.; Gao, W.; Wang, X.F.; Tian, Y. Nanoengineering of uniform and monodisperse mesoporous carbon nanospheres mediated by long hydrophilic chains of triblock copolymers. J. Mater. Sci. 2020, 55, 2052–2067. [Google Scholar] [CrossRef]
  14. Zhou, Q.Y.; Chen, W.H.; Jiang, X.; Liu, H.Y.; Ma, S.G.; Wang, B.D. Preparation of a novel nitrogen-containing graphitic mesoporous carbon for the removal of acid red 88. Sci. Rep. 2020, 10, 1353. [Google Scholar] [CrossRef] [PubMed]
  15. Han, B.; Zhang, E.Y.; Cheng, G.; Zhang, L.J.; Wang, D.W.; Wang, X.K. Hydrothermal carbon superstructures enriched with carboxyl groups for highly efficient uranium removal. Chem. Eng. J. 2018, 338, 734–744. [Google Scholar] [CrossRef]
  16. Huang, B.B.; Liu, Y.C.; Xie, Z.L. Biomass derived 2D carbons via a hydrothermal carbonization method as efficient bifunctional ORR/HER electrocatalysts. J. Mater. Chem. A 2017, 5, 23481–23488. [Google Scholar] [CrossRef]
  17. Zhao, T.T.; Zhu, D.L.; Li, W.R.; Li, A.J.; Zhang, J.J. Novel design and synthesis of carbon-coated porous silicon particles as high-performance lithium-ion battery anodes. J. Power Sources 2019, 439, 227027. [Google Scholar] [CrossRef]
  18. Yang, Q.X.; Wei, L.; Zheng, X.F.; Xiao, L.H. Single particle dynamic imaging and Fe3+ sensing with bright carbon dots derived from bovine serum albumin proteins. Sci. Rep. 2015, 5, 17727. [Google Scholar] [CrossRef]
  19. Zhang, B.H.; Yang, Q.; Li, Z.H.; Hao, J.C. Green synthesis of luminescent carbon dots and carbon-coated metal particles: Two birds with one stone. Colloid. Surface A 2015, 485, 34–41. [Google Scholar] [CrossRef]
  20. Yang, Y.; Yang, F.; Sun, C.J.; Zhao, H.R.; Hao, S.J.; Brown, D.E.; Zhang, J.; Ren, Y. Ru–Fe alloy mediated α-Fe2O3 particles on mesoporous carbon nanofibers as electrode materials with superior capacitive performance. RSC Adv. 2017, 7, 6818–6826. [Google Scholar] [CrossRef] [Green Version]
  21. Nekoueian, K.; Amiri, M.; Sillanpaa, M.; Marken, F.; Boukherroub, R.; Szunerits, S. Carbon-based quantum particles: An electroanalytical and biomedical perspective. Chem. Soc. Rev. 2019, 48, 4281–4316. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Z.B.; Zhang, C.L.; Xu, C.Q.; Zhu, Z.H.; Chen, C.N. Hollow polypyrrole nanosphere embedded in nitrogen-doped graphene layers to obtain a three-dimensional nanostructure as electrode material for electrochemical supercapacitor. Ionics 2017, 23, 147–156. [Google Scholar] [CrossRef]
  23. Fedoseeva, Y.V.; Bulusheva, L.G.; Koroteev, V.O.; Mevellec, J.Y.; Senkovskiy, B.V.; Flahaut, E.; Okotrub, A.V. Preferred attachment of fluorine near oxygen-containing groups on the surface of double-walled carbon nanotubes. Appl. Surf. Sci. 2020, 504, 144357. [Google Scholar] [CrossRef] [Green Version]
  24. Ge, T.; Min, F.F.; Zhang, M.X. A research into the nonthermal effect of thiophenic sulfur structure in the coking coal under mocrowave radiation. Spectrosc. Spect. Anal. 2018, 38, 3495–3501. [Google Scholar]
  25. Cui, X.Q.; Cao, D.; Djellabi, R.; Qiao, M.; Wang, Y.; Zhao, S.; Mao, R.; Gong, Y.; Zhao, X.; Yang, B. Enhancement of Ni/NiO/graphitized carbon and beta-Cyclodextrin/reduced graphene oxide for the electrochemical detection of norfloxacin in water sample. J. Electroanal. Chem. 2020, 851, 113407. [Google Scholar] [CrossRef]
  26. Ravulapalli, S.; Kunta, R. Enhanced removal of chromium (VI) from wastewater using active carbon derived from Lantana camara plant as adsorbent. Water Sci. Technol. 2018, 78, 1377–1389. [Google Scholar] [CrossRef] [Green Version]
  27. Park, E.J.; Kim, H.J.; Han, S.W.; Jeong, J.H.; Kim, I.H.; Seo, H.O.; Kim, Y.D. Assembly of PDMS/SiO2-PTFE and activated carbon fibre as a liquid water-resistant gas sorbent structure. Chem. Eng. J. 2017, 325, 433–441. [Google Scholar] [CrossRef]
  28. Rao, K.T.V.; Souzanchi, S.; Yuan, Z.S.; Ray, M.B.; Xu, C.B. Simple and green route for preparation of tin phosphate catalysts by solid-state grinding for dehydration of glucose to 5-hydroxymethylfurfural (HMF). RSC Adv. 2017, 7, 48501–48511. [Google Scholar] [CrossRef] [Green Version]
  29. Gromov, N.V.; Medvedeva, T.B.; Taran, O.P.; Bukhtiyarov, A.V.; Aymonier, C.; Prosvirin, I.P.; Parmon, V.N. Hydrothermal solubilization-hydrolysis-dehydration of cellulose to glucose and 5-hydroxymethylfurfural over solid acid carbon catalysts. Top. Catal. 2018, 61, 1912–1927. [Google Scholar] [CrossRef]
  30. Shen, J.Y.; Zhang, L.; Ren, J.; Wang, J.C.; Yao, H.C.; Li, Z.J. Highly enhanced acetone sensing performance of porous C-doped WO3 hollow spheres by carbon spheres as templates. Sens. Actuators B 2017, 239, 597–607. [Google Scholar] [CrossRef]
  31. Jin, C.Q.; Zhu, K.X.; Peterson, G.; Jian, Z.Y.; Xu, G.; Wei, Y.X.; Ge, C.H.; Li, J.H. Morphology dependent photocatalytic properties of ZnO nanostructures prepared by a carbon-sphere template method. J. Nanosc. Nanotechnol. 2018, 18, 5234–5241. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Y.; Jin, Y.H.; Zhao, C.C.; Pan, E.Z.; Jia, M.Q. 1D ultrafine SnO2 nanorods anchored on 3D graphene aerogels with hierarchical porous structures for high-performance lithium/sodium storage. J. Colloid. Interf. Sci. 2018, 532, 352–362. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, F.; Liu, Y.; Zhao, Y.F.; Wang, Y.; Wang, Z.J.; Zhang, W.H.; Ren, F.Z. Facile synthesis of two-dimensional porous MgCo2O4 nanosheets as anode for lithium-ion batteries. Appl. Sci. Basel 2018, 8, 22. [Google Scholar]
  34. Tao, T.; Lu, S.G.; Fan, Y.; Lei, W.W.; Huang, S.M.; Chen, Y. Anode improvement in rechargeable lithium-sulfur batteries. Adv. Mater. 2017, 29, 1700542. [Google Scholar] [CrossRef] [PubMed]
  35. Lee, D.Y.; Ahn, S.H. Investigation of laser cutting width of LiCoO2 coated aluminum for lithium-ion batteries. Appl. Sci. Basel 2017, 7, 914. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, Y.P.; Fu, Q.; Li, C.; Li, H.H.; Tang, H. Nitrogen and phosphorus dual-doped graphene aerogel confined monodisperse iron phosphide nanodots as an ultrafast and long-term cycling anode material for sodium-ion batteries. ACS Sustain. Chem. Eng. 2018, 6, 15083–15091. [Google Scholar] [CrossRef]
  37. Li, F.; Song, H.M.; Yu, W.S.; Ma, Q.L.; Dong, X.T.; Wang, J.X.; Liu, G.X. Electrospun TiO2//SnO2 Janus nanofibers and its application in ethanol sensing. Mater. Lett. 2020, 262, 127070. [Google Scholar] [CrossRef]
  38. Hermawan, A.; Asakura, Y.; Inada, M.; Yin, S. One-step synthesis of micro-/mesoporous SnO2 spheres by solvothermal method for toluene gas sensor. Ceram. Int. 2019, 45, 15435–15444. [Google Scholar] [CrossRef]
Figure 1. Illustration of the preparation of carbonaceous particles from natural biomass.
Figure 1. Illustration of the preparation of carbonaceous particles from natural biomass.
Crystals 10 00231 g001
Figure 2. Field emission scanning electron microscopy (FESEM) images of the particles prepared using (a) wheat, (c) corn, (e) starch and (f) maltose. Transmission electron microscopy (TEM) photographs of the samples synthesized from (b) wheat and (d) corn.
Figure 2. Field emission scanning electron microscopy (FESEM) images of the particles prepared using (a) wheat, (c) corn, (e) starch and (f) maltose. Transmission electron microscopy (TEM) photographs of the samples synthesized from (b) wheat and (d) corn.
Crystals 10 00231 g002
Figure 3. (a) X-ray photoelectron spectrometer (XPS) spectrum of the C1s region of wheat-derived carbonaceous particles. (b) Fourier transform infrared spectroscopy (FTIR) spectrum of the particles prepared from wheat.
Figure 3. (a) X-ray photoelectron spectrometer (XPS) spectrum of the C1s region of wheat-derived carbonaceous particles. (b) Fourier transform infrared spectroscopy (FTIR) spectrum of the particles prepared from wheat.
Crystals 10 00231 g003
Figure 4. (a) Low- and (b) high-magnification FESEM and (c) TEM images of the SnO2 spheres prepared by using the wheat-derived carbonaceous particles as template.
Figure 4. (a) Low- and (b) high-magnification FESEM and (c) TEM images of the SnO2 spheres prepared by using the wheat-derived carbonaceous particles as template.
Crystals 10 00231 g004
Figure 5. (a) Galvanostatic charge-discharge curves of the SnO2 anode cycling at 1C. (b) Capacity and Coulombic efficiency versus cycle number.
Figure 5. (a) Galvanostatic charge-discharge curves of the SnO2 anode cycling at 1C. (b) Capacity and Coulombic efficiency versus cycle number.
Crystals 10 00231 g005
Figure 6. (a) Real-time gas-sensing curve and (b) the response performance of the sensor towards ethanol at concentrations of 10 ppm, 20 ppm, 30 ppm and 40 ppm. (c) Gas-sensing curves and (d) responses towards 150 ppm, 200 ppm and 250 ppm ethanol gas.
Figure 6. (a) Real-time gas-sensing curve and (b) the response performance of the sensor towards ethanol at concentrations of 10 ppm, 20 ppm, 30 ppm and 40 ppm. (c) Gas-sensing curves and (d) responses towards 150 ppm, 200 ppm and 250 ppm ethanol gas.
Crystals 10 00231 g006

Share and Cite

MDPI and ACS Style

Ding, Y.; Zhou, P.; Han, T.; Liu, J. Environmentally Friendly and Cost-Effective Synthesis of Carbonaceous Particles for Preparing Hollow SnO2 Nanospheres and their Bifunctional Li-Storage and Gas-Sensing Properties. Crystals 2020, 10, 231. https://doi.org/10.3390/cryst10030231

AMA Style

Ding Y, Zhou P, Han T, Liu J. Environmentally Friendly and Cost-Effective Synthesis of Carbonaceous Particles for Preparing Hollow SnO2 Nanospheres and their Bifunctional Li-Storage and Gas-Sensing Properties. Crystals. 2020; 10(3):231. https://doi.org/10.3390/cryst10030231

Chicago/Turabian Style

Ding, Yingyi, Ping Zhou, Tianli Han, and Jinyun Liu. 2020. "Environmentally Friendly and Cost-Effective Synthesis of Carbonaceous Particles for Preparing Hollow SnO2 Nanospheres and their Bifunctional Li-Storage and Gas-Sensing Properties" Crystals 10, no. 3: 231. https://doi.org/10.3390/cryst10030231

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