Next Article in Journal
Layered Nano-Sheets: Synthesis and Applications
Next Article in Special Issue
A New Interaction Force Model of Gold Nanorods Derived by Molecular Dynamics Simulation
Previous Article in Journal
Gold Nanoparticle Self-Aggregation on Surface with 1,6-Hexanedithiol Functionalization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of Core-Shell Structured SiO2@Carbon Composite Nanorods for High-Performance Lithium-Ion Batteries

1
College of Materials and Energy, Guangdong Provincial Engineering Technology Research Center for Optical Agriculture, South China Agricultural University, Guangzhou 510642, China
2
Guangdong Laboratory of Lingnan Mordern Agriculture, Guangzhou 510642, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2020, 10(3), 513; https://doi.org/10.3390/nano10030513
Submission received: 4 February 2020 / Revised: 26 February 2020 / Accepted: 5 March 2020 / Published: 12 March 2020
(This article belongs to the Special Issue Molecular Interfaces Based Nanotechnology)

Abstract

:
Recently, SiO2 has attracted wide attention in lithium-ion batteries owing to its high theoretical capacity and low cost. However, the utilization of SiO2 is impeded by the enormous volume expansion and low electric conductivity. Although constructing SiO2/carbon composite can significantly enhance the electrochemical performance, the skillful preparation of the well-defined SiO2/carbon composite is still a remaining challenge. Here, a facile strategy of in situ coating of polydopamine is applied to synthesis of a series of core-shell structured SiO2@carbon composite nanorods with different thicknesses of carbon shells. The carbon shell uniformly coated on the surface of SiO2 nanorods significantly suppresses the volume expansion to some extent, as well as improves the electric conductivity of SiO2. Therefore, the composite nanorods exhibit a remarkable electrochemical performance as the electrode materials of lithium-ion batteries. For instance, a high and stable reversible capacity at a current density of 100 mA g−1 reaches 690 mAh g−1 and a capacity of 344.9 mAh g−1 can be achieved even at the high current density of 1000 mA g−1. In addition, excellent capacity retention reaches 95% over 100 cycles. These SiO2@carbon composite nanorods with decent electrochemical performances hold great potential for applications in lithium-ion batteries.

1. Introduction

Lithium-ion batteries (LIBs) have been widely used in portable devices, vehicle mounted equipment, and electrochemical energy storage owning to their environmental friendliness, high energy density, and long lifespan [1,2,3,4,5]. As the commercial anode material in LIBs, graphite has a low theoretical capacity of 372 mAh g−1, which is difficult to meet the ever-growing demands of high-energy-density [6,7,8]. Therefore, momentous efforts have been made to exploring an ideal alternative anode with high lithium storage capacity. Recently, SiO2 has attracted great attention for LIBs due to its low cost, high theoretical capacity (i.e., 1956 mAh g−1), and ease of fabrication [9,10,11]. Unfortunately, the practical application SiO2 anode is restricted by the fast capacity fading and poor rate performance, which mainly derive from its low electrical conductivity and huge volume expansion during the charging-discharging processes [12,13,14,15].
In order to circumvent these intrinsic limitations, two mainstream strategies have been proposed, i.e., design of nanostructures and compositing with high conductive carbon or other materials. For the former one, various nanostructured SiO2 such as nanofilms [16], nanocubes [17], nanospheres [18,19], nanotubes [20], and nanoparticles [21] have been successfully synthesized. This strategy greatly shortens the lithium ion and electron transport paths and increases the contact area between electrodes and electrolytes, and thus improves the electrochemical performance. However, the rate capability performance is still unsatisfactory because of the poor conductivity of SiO2 [22]. Compositing with high conductive carbon materials has been demonstrated to be a promising strategy to improve the electron transport and facilitate the electrochemical reaction kinetics, and thus enhance the rate performance [23,24]. In addition, the carbonaceous materials with a rational nanostructure can provide valuable void space to buffer a large volume expansion of SiO2, thus increasing the structural integrity and long lifespan of the active material [25,26,27]. Nevertheless, the facile preparation of well-defined SiO2/carbon composite anode is still full of challenge. Most of the general methods have various limitations, including rigid reaction conditions, complex multiple synthetic steps, and time-consuming. Moreover, control of the uniform carbon/SiO2 interface remains challenging.
In this contribution, we report a facile method to prepare SiO2@carbon composite nanorods (SiO2@CCNRs) with a core-shell structure based on the in situ coating of polydopamine (PDA), demonstrating its superior ability when it is used as the anode material in LIBs. The synthetic process is exhibited in Scheme 1. PDA with a strong adhesive capability allows easy and quick construction of polymeric coating on the SiO2 nanorods. In addition, the polymerization of dopamine is executed in an aqueous solution at ambient temperature, which can provide a very mild and environmentally friendly technique. Moreover, by adjusting the amount of dopamine, it is easy to control the thickness of the PDA coating and the derivative carbonaceous shells of the SiO2@CCNRs. The SiO2@CCNRs composite architecture with the synergistic effect between highly active SiO2 core and conductive carbonaceous shell makes a prominent improvement to the excellent electrochemical performance. When evaluated as anode materials for LIBs, the as-synthesized SiO2@CCNRs with core-shell structure show excellent performance, such as a high and stable reversible capacity of 690 mAh g−1 after 100 cycles at 100 mA g−1, and a superior rate capability of 344.9 mAh g−1 even at the high current of 1000 mA g−1.

2. Materials and Methods

2.1. Materials

Triblock copolymer F127 was purchased from Sigma-Aldrich (Shanghai) trading Co., Ltd (Shanghai, China). Cetyltrimethylamine bromide (CTAB), tris-hydrochloride buffer (Tris-HCl buffer), and dopamine were purchased from Aladdin Biochemical Technology Co., Ltd (Shanghai, China). Aqueous ammonia (NH3·H2O) and tetraethyl orthosilicate (TEOS) were purchased from Sinopharm Chemical Reagent Co., Ltd (Shanghai, China). All chemical reagents were used as received without further purification.

2.2. Preparation of SiO2@Carbon Nanorods

2.2.1. Synthesis of SiO2 Nanorods

Typically, 0.123 g of triblock copolymer F127 was added in the mixture of 3.50 mL of deionized water and 12.50 mL of the CTAB bromide solution and stirred at room temperature. Subsequently, 15 mL of NH3·H2O and 0.6 mL tetraethyl orthosilicate were added in the above solution under stirring. After quiescing for 4 h, the product was gathered by centrifuging and washed with distilled water until the filtrates was close to neutral. Finally, the SiO2 nanorods were obtained after drying.

2.2.2. Synthesis of SiO2@Carbon Nanorods

In a typical process, 0.25 g of SiO2 nanorods were dispersed in 330 mL of Tris-HCl buffer solution by ultrasonic dispersion. After that, 0.25 g of dopamine was added to the above solution under stirring. After 24 h of stirring reaction, the outcome was gathered by centrifuging and then washed with distilled water several times and dried at 105 °C. The SiO2@CCNRs-1 was obtained after carbonizing at 900 °C for 6 h in the N2 atmosphere. The preparation processes of SiO2@CCNRs-2 and SiO2@CCNRs-3 were similar to the SiO2@CCNRs-1 except for the ratio of dopamine and SiO2 nanorods was 2 and 3, respectively.

2.3. Material Characterization

The composition of all the samples was determined by powder X-ray diffractometer (XRD, Rigaku Ultima IV, Tokyo, Japan) with Cu Kα radiation. The scanning electron microscope (SEM, FEI Quanta FEG650, Hillsboro, OR, USA) was used for testing the morphology of all the samples. The diameter of all the samples was determined from the particle size distribution by using the Nano Measurer analysis software through 100 nanorods randomly selected in the SEM image. The maximum in the resulting particle size distribution curve was referred to as the diameter of nanorods. The pore structure of all the samples was obtained using an ASAP 3Flex automatic volumetric sorption analyzer (Micromeritics sorption analyzer, Norcross, GA, USA) at 77 K. Moreover, the Brunauer–Emmett–Teller (BET) method was used to determine the specific surface area and total pore volume was calculated at a saturation relative pressure of ca. 0.99. The density functional theory (DFT) method was applied to analyze the pore size distribution.

2.4. Electrochemical Measurements

The electrode is mainly composed of active material (SiO2@CCNRs), conductive agent (Super-P), and binder (PVDF), with a mass ratio of 8:1:1. The slurry is coated on the copper foil, which is completely dried and cut into a circular piece with an area of 1.13 cm2. The average loading mass of each electrode piece is about 3.5 mg cm−2. Then, the CR2032 coin-type cells were assembled in the Ar-filled glove box, with lithium foil as the counter electrode. The electrolyte was 1 M LiPF6 in a 1:1:1 v/v/v mixture of ethylene carbonate (EC), methyl ethyl carbonate (EMC), and dimethyl carbonate (DMC). Galvanostatic charge-discharge tests were carried out in the range of 0.01 to 3.0 V on the Neware battery tester (Neware BTS 7.6×, Shenzhen, China). The specific capacity is calculated by dividing the tested capacity with the mass of active material (i.e., SiO2@CCNRs) in the battery. Cyclic voltammogram (CV) measurements data were recorded on the electrochemical workstation (Chenhua CHI660e, Shanghai, China) at a scan rate of 0.1 mV s−1 between 0.01 and 3 V. Electrochemical impedance spectroscopy (EIS) with the frequency range from 100 kHz to 10 mHz and an amplitude of 5 mV s−1 were carried out on a photoelectrochemical test system (Modulab XM DSSC, Bognor Regis, UK).

3. Results

The fabrication procedure of SiO2@CCNRs is schematically illustrated in Scheme 1. Uniform SiO2 nanorods with an average diameter of 147 nm were first synthesized (Figure 1a,d). Subsequently, dopamine was polymerized in an aqueous solution at room temperature to form PDA with strong adhesion ability, which can easily and quickly in situ construct a polymer coating on SiO2 nanorods. The obtained SiO2@PDA-1 with about 5.5 nm thicknesses of PDA coating reveals the nanorod structure consistent with SiO2 (Figure 1b,e). After carbonization under nitrogen protection, the targeted SiO2@CCNRs were obtained. As shown in Figure 1c, the SiO2@CCNRs-1 still maintain a relatively good nanorods morphology, meaning good structural inheritance after carbonization. The average diameter of the SiO2@CCNRs-1 is 109 nm (Figure 1f), which is about 49 nm smaller than that of SiO2@PDA-1. This could be due to the shrinkage of carbon and SiO2 caused by mass loss after carbonization. By changing the amount of dopamine, the thickness of the carbon shell on the SiO2 nanorods can be simply controlled. For example, by increasing the use of dopamine by 1 and 2 times, the average diameters of obtained SiO2@CCNRs-2 and SiO2@CCNRs-3 are increased by 5 and 10 nm compared to SiO2@CCNRs-1, respectively (Figure S1). This indicates that the thickness of carbon shell increases with the increase of dopamine content.
XRD is conducted to study the microcrystalline structure of SiO2@CCNRs. Figure 1g and Figure S2 show the XRD patterns of the SiO2@CCNRs. The peaks located at 22° indexed to SiO2 (JCPD Card No. 29-0085), and the broad peaks located at 23° is associated with carbon, while the weak peak near 44° can be attributed to amorphous carbon [28,29,30]. To explore the pore characteristic of the SiO2@CCNRs, the N2 adsorption-desorption measurement was executed. As shown in Figure 1h, the N2 adsorption-desorption isotherm shows a typical type IV isotherm, which shows that at a lower relative pressure (P/P0), the adsorption amount of N2 increases rapidly, and a hysteresis loop appears in the range of 0.4 to 1.0, demonstrating the presence of the micropores and mesoporous [31]. It can be seen from the pore size distribution curve (Figure 1i) that the pore size of the micropores of SiO2@CCNRs-1 mainly exists at 0.68 nm, and the mesopores ranges from 8.6 to 29.5 nm. In addition, the total pore volume and BET surface area of SiO2@CCNRs-1 is 0.17 cm3 g−1 and 246 m2 g−1.
The lithium ion storage behavior of SiO2@CCNRs as anode material for LIBs is explored. Figure 2a shows representative CV curves for the SiO2@CCNRs-2 electrode at a scanning rate of 0.1 mV s−1 between 0.01 and 3.0 V. In the first cycle, an obvious reduction peak is observed at ~0.7 and ~1.1 V but disappears in the subsequent two cycles, which was related to formation of the solid-electrolyte-interface (SEI) layer and electrolyte decomposition on the surface of the carbon shell [32,33]. In the second and third cycles, a weak cathodic peak appears at about 0.3 V, corresponding to the formation of Li-Si alloy, while the broad anodic peak at 0.5–1.7 V is assigned to the dealloying processes of Li-Si alloy into Si [34]. Furthermore, the second and third CV curves overlap well, demonstrating the reversible alloying/dealloying reaction [35]. Figure 2b shows that there are two voltage plateaus at ~0.7 and ~1.1 V in the initial discharge curve, which is consistent with the results from the CV curve. The voltage plateaus of the later discharge curves are in the range of 0.01–0.5 V.
The charge/discharge cycling performances of SiO2@CCNRs are displayed in Figure 2c. They show a stable cycling stability up to 100 cycles of charge/discharge at a current density of 100 mA g−1. It can be clearly seen that SiO2@CCNRs-2 with moderate carbon shell thickness delivers a stable reversible specific capacity at about 690 mAh g−1 after 100 cycles, which is much higher than that of SiO2@CCNRs-1 (i.e., 423 mAh g−1) and SiO2@CCNRs-3 (i.e., 498 mAh g−1). The degradation of the cycle stability for SiO2@CCNRs-1 could be attributed to the peeling behavior of the thin carbon layer caused by the volume change of SiO2. The rate performances of the SiO2@CCNRs anode are further tested at different current densities from 100 to 5000 mA g−1 (Figure 2d). At current densities of 100, 250, 500, 750, 1000, 2000, 3000, 4000, 5000 mA g−1, the reversible specific capacities of the SiO2@CCNRs-2 anode are ~730, 573, 464, 362, 315, 267, 213, 187, and 165 mAh g−1, respectively. Furthermore, when the current density is returned to 100 mA g−1, SiO2@CCNRs-2 delivers high reversible specific capacity at about 645 mAh g−1, about 95% of the initial specific capacity could be recovered. Similarly, SiO2@CCNRs-1 and SiO2@CCNRs-3 also show good rate performance, demonstrating the superiority of this type of composite material.
The reaction kinetics and electrochemical activities of electrodes is studied by electrochemical impedance spectroscopy (EIS). Figure 3a and Figure S3 show the Nyquist plots and plots of Z’ against ω−1/2, respectively. The relevant kinetic data are calculated and summarized in Table 1. It can be seen that the charge transfer resistance (Rct) of SiO2@CCNRs-2 is 43.14 Ω, which is much lower than that of SiO2@CCNRs-1 (i.e., 83.64 Ω) and SiO2@CCNRs-3 (i.e., 78.26 Ω), indicating that SiO2@CCNRs-2 has a higher charge transfer efficiency. Figure 3b shows the lithium diffusion ion coefficients of different anodes in LIBs. The lithium ion diffusion coefficient of SiO2@CCNRs-2 is calculated to be 8.09 × 10−19 cm2 s−1, which is higher than that of SiO2@CCNRs-1 (4.49 × 10−19 cm2 s−1) and SiO2@CCNRs-3 (5.18 × 10−19 cm2 s−1). It is illustrated that SiO2@CCNRs-1 has a more rapid lithium ion diffusion rate.
From the above analysis and discussion, the SiO2@CCNRs-2 anode displays more excellent lithium ion diffusion rate than the SiO2@CCNRs-1 and SiO2@CCNRs-3 anodes, which mainly comes down to the moderate carbon shell thickness in SiO2@CCNRs. On the one hand, the thin carbon shell of SiO2@CCNRs-1 could not effectively ameliorate the rapid volume expansion of SiO2 during the cycles, and could not constantly inhibit the pulverization of electrode either, resulting in the increase of resistance and the decrease of lithium ion diffusion rate. On the other hand, the thick carbon shell of SiO2@CCNRs-3 increases the ion diffusion distance, which decreases the mass transfer efficiency. As a comparison, the moderate carbon shell thickness of SiO2@CCNRs-2 not only effectively relieves the volume expansion of SiO2 without cracking, but also maintains good kinetic characteristics, thus exhibiting more excellent electrochemical performance.

4. Conclusions

The core-shell structured SiO2@CCNRs composite was successfully synthesized based on the in situ coating of PDA. The carbon shell on the surface of SiO2 not only increases the conductivity of SiO2, but also reduces the volume change of SiO2 during charging and discharging. Therefore, the core-shell structured SiO2@CCNRs exhibit excellent lithium storage capacity, rate performance, and cycle stability. In view of this, the SiO2@CCNRs are a kind of high-performance LIBs electrode materials with excellent application potential. Moreover, the effect of carbon shell thickness on the electrochemical performance of LIBs is studied as well. It can be predicted that further optimization of electrochemical performance could be realized by tuning of the size of SiO2 nanorod, carbonization temperature, and heating rate. Meanwhile, it is envisioned that this strategy can be extended to those carbon precursors that have a strong adhesive capability to form homogeneous and robust organic/inorganic interfaces with SiO2, which can also provide an opportunity for further boosting their electrochemical performance as well. We hope that this work can provide new ideas for the efficient preparation of high-performance SiO2/carbon composite electrode materials, and help the rapid development of silicon-based electrode materials.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/3/513/s1, Figure S1: SEM images of (a) SiO2@CCNRs-2 and (b) SiO2@CCNRs-3. Particle size distributions of (c) SiO2@CCNRs-2 and (d) SiO2@CCNRs-3; Figure S2: XRD patterns of SiO2@CCNRs-2 and SiO2@CCNRs-3; Figure S3: Plots of ω−1/2 versus Z’ of SiO2@CCNRs composites anodes in LIBs.

Author Contributions

Conceptualization and methodology, Y.L. (Yeru Liang) and Y.L. (Yingliang Liu); performed the experiments, N.P. and H.P.; writing—original draft preparation, H.P., W.Z., and P.Y.; writing—review and editing, W.Z., P.Y., Y.L. (Yeru Liang), H.H., M.Z., and Y.X.; funding acquisition, Y.L. (Yeru Liang) and Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (51972121, 21671069), Tip-top Scientific and Technical Innovative Youth Talents of Guangdong Special Support Program (2017TQ04C419), Guangdong Province Universities and Colleges Pearl River Scholar Funded Scheme (2017), Guangdong Basic and Applied Basic Research Foundation (2019A1515011502), and the Program for Pearl River New Star of Science and Technology in Guangzhou (201710010104).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Na, Z.L.; Huang, G.; Liang, F.; Yin, D.M.; Wang, L.M. A Core-Shell Fe/Fe2O3 Nanowire as a High-Performance Anode Material for Lithium-Ion Batteries. Chem. Eur. J. 2016, 22, 12081–12087. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, K.; Pei, S.E.; He, Z.S.; Huang, L.A.; Zhu, S.S.; Guo, J.F.; Shao, H.B.; Wang, J.M. Synthesis of a novel porous silicon microsphere@carbon core-shell composite via in situ MOF coating for lithium ion battery anodes. Chem. Eng. J. 2019, 356, 272–281. [Google Scholar] [CrossRef]
  3. Meng, J.K.; Cao, Y.; Suo, Y.; Liu, Y.S.; Zhang, J.M.; Zheng, X.C. Facile Fabrication of 3D SiO2@Graphene Aerogel Composites as Anode Material for Lithium Ion Batteries. Electrochim. Acta 2015, 176, 1001–1009. [Google Scholar] [CrossRef]
  4. Andre, D.; Kim, S.J.; Lamp, P.; Lux, S.F.; Maglia, F.; Paschos, O.; Stiaszny, B. Future generations of cathode materials: An automotive industry perspective. J. Mater. Chem. A 2015, 3, 6709–6732. [Google Scholar] [CrossRef]
  5. Liang, Y.R.; Zhao, C.Z.; Yuan, H.; Chen, Y.; Zhang, W.C.; Haung, J.Q.; Yu, D.S.; Titirici, M.M.; Chueh, Y.L.; Yu, H.J.; et al. A review of rechargeable batteries for portable electronic devices. InfoMat 2019, 1, 6–32. [Google Scholar] [CrossRef] [Green Version]
  6. Ali, S.; Jaffer, S.; Maitlo, I.; Shehzad, F.K.; Wang, Q.Y.; Ali, S.; Akram, M.Y.; He, Y.; Nie, J. Photo cured 3D porous silica-carbon (SiO2-C) membrane as anode material for high performance rechargeable Li-ion batteries. J. Alloy. Compd. 2020, 812, 8. [Google Scholar] [CrossRef]
  7. Chang, W.S.; Park, C.M.; Kim, J.H.; Kim, Y.U.; Jeong, G.; Sohn, H.J. Quartz (SiO2): A new energy storage anode material for Li-ion batteries. Energy Environ. Sci. 2012, 5, 6895–6899. [Google Scholar] [CrossRef]
  8. Cao, L.M.; Huang, J.L.; Lin, Z.P.; Yu, X.; Wu, X.X.; Zhang, B.D.; Zhan, Y.F.; Xie, F.Y.; Zhang, W.H.; Chen, J.; et al. Amorphous SiO2/C composite as anode material for lithium-ion batteries. J. Mater. Res. 2018, 33, 1219–1225. [Google Scholar] [CrossRef]
  9. Yang, X.Q.; Ma, H.; Zhang, G.Q.; Li, X.X. Silica/Carbon Composites with Controllable Nanostructure from a Facile One-Step Method for Lithium-Ion Batteries Application. Adv. Mater. Interfaces 2019, 6, 7. [Google Scholar] [CrossRef]
  10. Yuan, Z.N.; Zhao, N.Q.; Shi, C.S.; Liu, E.Z.; He, C.N.; He, F. Synthesis of SiO2/3D porous carbon composite as anode material with enhanced lithium storage performance. Chem. Phys. Lett. 2016, 651, 19–23. [Google Scholar] [CrossRef]
  11. Lener, G.; Garcia-Blanco, A.A.; Furlong, O.; Nazzarro, M.; Sapag, K.; Barraco, D.E.; Leiva, E.P.M. A silica/carbon composite as anode for lithium-ion batteries with a large rate capability: Experiment and theoretical considerations. Electrochim. Acta 2018, 279, 289–300. [Google Scholar] [CrossRef]
  12. Xiao, T.T.; Zhang, W.F.; Xu, T.; Wu, J.X.; Wei, M.D. Hollow SiO2 microspheres coated with nitrogen doped carbon layer as an anode for high performance lithium-ion batteries. Electrochim. Acta 2019, 306, 106–112. [Google Scholar] [CrossRef]
  13. Xu, Q.; Sun, J.K.; Yin, Y.X.; Guo, Y.G. Facile Synthesis of Blocky SiOx/C with Graphite-Like Structure for High-Performance Lithium-Ion Battery Anodes. Adv. Funct. Mater. 2018, 28, 7. [Google Scholar] [CrossRef]
  14. Wang, C.W.; Liu, K.W.; Chen, W.F.; Zhou, J.D.; Lin, H.P.; Hsu, C.H.; Kuo, P.L. Mesoporous SiO2/carbon hollow spheres applied towards a high rate-performance Li-battery anode. Inorg. Chem. Front. 2016, 3, 1398–1405. [Google Scholar] [CrossRef]
  15. Lv, P.P.; Zhao, H.L.; Wang, J.; Liu, X.; Zhang, T.H.; Xia, Q. Facile preparation and electrochemical properties of amorphous SiO2/C composite as anode material for lithium ion batteries. J. Power Sources 2013, 237, 291–294. [Google Scholar] [CrossRef]
  16. Sun, Q.; Zhang, B.; Fu, Z.W. Lithium electrochemistry of SiO2 thin film electrode for lithium-ion batteries. Appl. Surf. Sci. 2008, 254, 3774–3779. [Google Scholar] [CrossRef]
  17. Yan, N.; Wang, F.; Zhong, H.; Li, Y.; Wang, Y.; Hu, L.; Chen, Q.W. Hollow Porous SiO2 Nanocubes Towards High-performance Anodes for Lithium-ion Batteries. Sci. Rep. 2013, 3, 6. [Google Scholar] [CrossRef] [Green Version]
  18. Ma, X.M.; Wei, Z.P.; Han, H.J.; Wang, X.B.; Cui, K.Q.; Yang, L. Tunable construction of multi-shell hollow SiO2 microspheres with hierarchically porous structure as high-performance anodes for lithium ion batteries. Chem. Eng. J. 2017, 323, 252–259. [Google Scholar] [CrossRef]
  19. Tu, J.G.; Yuan, Y.; Zhan, P.; Jiao, H.D.; Wang, X.D.; Zhu, H.M.; Jiao, S.Q. Straightforward Approach toward SiO2 Nanospheres and Their Superior Lithium Storage Performance. J. Phys. Chem. C 2014, 118, 7357–7362. [Google Scholar] [CrossRef]
  20. Favors, Z.; Wang, W.; Bay, H.H.; George, A.; Ozkan, M.; Ozkan, C.S. Stable Cycling of SiO2 Nanotubes as High-Performance Anodes for Lithium-Ion Batteries. Sci. Rep. 2014, 4, 7. [Google Scholar] [CrossRef]
  21. Yao, Y.; Zhang, J.J.; Xue, L.G.; Huang, T.; Yu, A.S. Carbon-coated SiO(2) nanoparticles as anode material for lithium ion batteries. J. Power Sources 2011, 196, 10240–10243. [Google Scholar] [CrossRef]
  22. Guo, B.K.; Shu, J.; Wang, Z.X.; Yang, H.; Shi, L.H.; Liu, Y.N.; Chen, L.Q. Electrochemical reduction of nano-SiO2 in hard carbon as anode material for lithium ion batteries. Electrochem. Commun. 2008, 10, 1876–1878. [Google Scholar] [CrossRef]
  23. Liu, X.L.; Chen, Y.X.; Liu, H.B.; Liu, Z.Q. SiO2@C hollow sphere anodes for lithium-ion batteries. J. Mater. Sci. Technol. 2017, 33, 239–245. [Google Scholar] [CrossRef]
  24. Liang, Y.R.; Zhang, W.C.; Wu, D.C.; Ni, Q.Q.; Zhang, M.Q. Interface Engineering of Carbon-Based Nanocomposites for Advanced Electrochemical Energy Storage. Adv. Mater. Interfaces 2018, 5, 1800430. [Google Scholar] [CrossRef]
  25. Li, H.H.; Zhang, L.L.; Fan, C.Y.; Wang, K.; Wu, X.L.; Sun, H.Z.; Zhang, J.P. A plum-pudding like mesoporous SiO2/flake graphite nanocomposite with superior rate performance for LIB anode materials. Phys. Chem. Chem. Phys. 2015, 17, 22893–22899. [Google Scholar] [CrossRef] [PubMed]
  26. Shi, J.W.; Gao, H.Y.; Hu, G.X.; Zhang, Q. Core-shell structured Si@C nanocomposite for high-performance Li-ion batteries with a highly viscous gel as precursor. J. Power Sources 2019, 438, 9. [Google Scholar] [CrossRef]
  27. Jiao, M.L.; Liu, K.L.; Shi, Z.Q.; Wang, C.Y. SiO2/Carbon Composite Microspheres with Hollow Core-Shell Structure as a High-Stability Electrode for Lithium-Ion Batteries. ChemElectroChem 2017, 4, 542–549. [Google Scholar] [CrossRef]
  28. Zhao, Y.; Liu, Z.J.; Zhang, Y.G.; Mentbayeva, A.; Wang, X.; Maximov, M.Y.; Liu, B.X.; Bakenov, Z.; Yin, F.X. Facile Synthesis of SiO2@C Nanoparticles Anchored on MWNT as High-Performance Anode Materials for Li-ion Batteries. Nanoscale Res. Lett. 2017, 12, 7. [Google Scholar] [CrossRef] [Green Version]
  29. Doi, T.; Tagashira, M.; Iriyama, Y.; Abe, T.; Ogumi, Z. Preparation and electrochemical properties of SiO2-non-graphitizable carbon composites as negative electrode materials for Li-ion batteries. J. Appl. Electrochem. 2012, 42, 69–74. [Google Scholar] [CrossRef]
  30. Li, H.H.; Wu, X.L.; Sun, H.Z.; Wang, K.; Fan, C.Y.; Zhang, L.L.; Yang, F.M.; Zhang, J.P. Dual-Porosity SiO2/C Nanocomposite with Enhanced Lithium Storage Performance. J. Phys. Chem. C 2015, 119, 3495–3501. [Google Scholar] [CrossRef]
  31. Liang, Y.R.; Chen, L.Y.; Zhuang, D.Y.; Liu, H.; Fu, R.W.; Zhang, M.Q.; Wu, D.C.; Matyjaszewski, K. Fabrication and nanostructure control of superhierarchical carbon materials from heterogeneous bottlebrushes. Chem. Sci. 2017, 8, 2101–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Hao, S.; Wang, Z.X.; Chen, L.Q. Amorphous SiO2 in tunnel-structured mesoporous carbon and its anode performance in Li-ion batteries. Mater. Des. 2016, 111, 616–621. [Google Scholar] [CrossRef]
  33. Belgibayeva, A.; Taniguchi, I. Synthesis and characterization of SiO2/C composite nanofibers as free-standing anode materials for Li-ion batteries. Electrochim. Acta 2019, 328, 12. [Google Scholar] [CrossRef]
  34. Cui, J.L.; Cheng, F.P.; Lin, J.; Yang, J.C.; Jiang, K.; Wen, Z.S.; Sun, J.C. High surface area C/SiO2 composites from rice husks as a high-performance anode for lithium ion batteries. Powder Technol. 2017, 311, 1–8. [Google Scholar] [CrossRef]
  35. Ren, Y.R.; Li, M.Q. Facile synthesis of SiOx@C composite nanorods as anodes for lithium ion batteries with excellent electrochemical performance. J. Power Sources 2016, 306, 459–466. [Google Scholar] [CrossRef]
Scheme 1. Schematic illustration of synthesis of the core-shell structured SiO2@CCNRs.
Scheme 1. Schematic illustration of synthesis of the core-shell structured SiO2@CCNRs.
Nanomaterials 10 00513 sch001
Figure 1. SEM images of (a) SiO2, (b) SiO2@PDA-1 complex, and (c) SiO2@CCNRs-1. Particle size distributions of (d) SiO2, (e) SiO2@PDA-1, and (f) SiO2@CCNRs-1. (g) XRD pattern, (h) N2 adsorption-desorption isotherm, and (i) pore size distribution curve of SiO2@CCNRs-1.
Figure 1. SEM images of (a) SiO2, (b) SiO2@PDA-1 complex, and (c) SiO2@CCNRs-1. Particle size distributions of (d) SiO2, (e) SiO2@PDA-1, and (f) SiO2@CCNRs-1. (g) XRD pattern, (h) N2 adsorption-desorption isotherm, and (i) pore size distribution curve of SiO2@CCNRs-1.
Nanomaterials 10 00513 g001
Figure 2. (a) Cyclic voltammogram (CV) curves of SiO2@CCNRs-2 cycled between 0.01 and 3.0 V (vs. Li+/Li) at a scanning rate of 0.1 mV s−1 for the first three cycles. (b) Galvanostatic discharge/charge curves of SiO2@CCNRs-2 electrode at current density of 100 mA g−1 for the first three cycles. (c) Comparison of the cycling performance of the SiO2@CCNRs electrode at a current density of 100 mA g−1 for 100 cycles. (d) Rate performance of the SiO2@CCNRs electrode at different current densities.
Figure 2. (a) Cyclic voltammogram (CV) curves of SiO2@CCNRs-2 cycled between 0.01 and 3.0 V (vs. Li+/Li) at a scanning rate of 0.1 mV s−1 for the first three cycles. (b) Galvanostatic discharge/charge curves of SiO2@CCNRs-2 electrode at current density of 100 mA g−1 for the first three cycles. (c) Comparison of the cycling performance of the SiO2@CCNRs electrode at a current density of 100 mA g−1 for 100 cycles. (d) Rate performance of the SiO2@CCNRs electrode at different current densities.
Nanomaterials 10 00513 g002
Figure 3. (a) Nyquist plots and (b) lithium ion diffusion coefficients of different anodes inLIBs.
Figure 3. (a) Nyquist plots and (b) lithium ion diffusion coefficients of different anodes inLIBs.
Nanomaterials 10 00513 g003
Table 1. Kinetic parameters of SiO2@CCNRs-1, SiO2@CCNRs-2, and SiO2@CCNRs-3 anodes.
Table 1. Kinetic parameters of SiO2@CCNRs-1, SiO2@CCNRs-2, and SiO2@CCNRs-3 anodes.
Key ParametersSiO2@CCNRs-1SiO2@CCNRs-2SiO2@CCNRs-3
Rs (Ω)17.4210.225.69
Rct (Ω)83.6443.1478.26
σ (Ω Hz1/2)191.89142.92178.5
DLi (10−19 cm2 s−1)4.498.095.18

Share and Cite

MDPI and ACS Style

Pang, H.; Zhang, W.; Yu, P.; Pan, N.; Hu, H.; Zheng, M.; Xiao, Y.; Liu, Y.; Liang, Y. Facile Synthesis of Core-Shell Structured SiO2@Carbon Composite Nanorods for High-Performance Lithium-Ion Batteries. Nanomaterials 2020, 10, 513. https://doi.org/10.3390/nano10030513

AMA Style

Pang H, Zhang W, Yu P, Pan N, Hu H, Zheng M, Xiao Y, Liu Y, Liang Y. Facile Synthesis of Core-Shell Structured SiO2@Carbon Composite Nanorods for High-Performance Lithium-Ion Batteries. Nanomaterials. 2020; 10(3):513. https://doi.org/10.3390/nano10030513

Chicago/Turabian Style

Pang, Haibo, Weicai Zhang, Peifeng Yu, Ning Pan, Hang Hu, Mingtao Zheng, Yong Xiao, Yingliang Liu, and Yeru Liang. 2020. "Facile Synthesis of Core-Shell Structured SiO2@Carbon Composite Nanorods for High-Performance Lithium-Ion Batteries" Nanomaterials 10, no. 3: 513. https://doi.org/10.3390/nano10030513

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