Next Article in Journal
Single-Sensor Vibration-Scanning Method for Assessing the Mechanical Properties of 3D Printed Elements
Next Article in Special Issue
High Capacity Nanocomposite Layers Based on Nanoparticles of Carbon Materials and Ruthenium Dioxide for Potassium Sensitive Electrode
Previous Article in Journal
Experimental Investigation on the Effects of Photocatalysis in Ultraviolet-Induced Nanoparticle Colloid Jet Machining
Previous Article in Special Issue
Electrochemical Performance Enhancement of Micro-Sized Porous Si by Integrating with Nano-Sn and Carbonaceous Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Electrochemical Performance Promoted by Tin in Silica Anode Materials for Stable and High-Capacity Lithium-Ion Batteries

School of Science, Jiangsu University of Science and Technology, 666 Changhui Road, Zhenjiang 212100, China
*
Author to whom correspondence should be addressed.
Materials 2021, 14(5), 1071; https://doi.org/10.3390/ma14051071
Submission received: 18 January 2021 / Revised: 9 February 2021 / Accepted: 9 February 2021 / Published: 25 February 2021
(This article belongs to the Special Issue Electrode Materials: Fabrication, Properties, and Applications)

Abstract

:
Although the silicon oxide (SiO2) as an anode material shows potential and promise for lithium-ion batteries (LIBs), owing to its high capacity, low cost, abundance, and safety, severe capacity decay and sluggish charge transfer during the discharge–charge process has caused a serious challenge for available applications. Herein, a novel 3D porous silicon oxide@Pourous Carbon@Tin (SiO2@Pc@Sn) composite anode material was firstly designed and synthesized by freeze-drying and thermal-melting self-assembly, in which SiO2 microparticles were encapsulated in the porous carbon as well as Sn nanoballs being uniformly dispersed in the SiO2@Pc-like sesame seeds, effectively constructing a robust and conductive 3D porous Jujube cake-like architecture that is beneficial for fast ion transfer and high structural stability. Such a SiO2@Pc@Sn micro-nano hierarchical structure as a LIBs anode exhibits a large reversible specific capacity ~520 mAh·g−1, initial coulombic efficiency (ICE) ~52%, outstanding rate capability, and excellent cycling stability over 100 cycles. Furthermore, the phase evolution and underlying electrochemical mechanism during the charge–discharge process were further uncovered by cyclic voltammetry (CV) investigation.

Graphical Abstract

1. Introduction

Lithium-ion batteries (LIBs) have been regarded as one of the critical energy storage technologies that can be widely used in portable electronics and grid-scale energy storage due to their high energy density and cycle longevity to make a fossil fuel-free environment possible [1,2,3,4,5]. With the advent of electric vehicles (EV) in recent years, the traditional commercialized LIBs are obviously insufficient to meet the requirement owning to their limited capacities. Therefore, it is highly desired for LIBs with higher energy and power densities as well as lower cost to be developed [6,7,8].
According to the working principle of LIBs, the electrode materials play a critical role in the further improvement of the battery performance [9,10,11,12,13,14,15,16]. High-capacity and low-cost materials have triggered vast interest in the past few years [15,16,17,18,19,20,21], which can bring great promise for next-generation LIBs with a higher price–performance ratio. Silica (SiO2) has recently captured great attention as a promising candidate anode materials for LIBs because of its suitable working potential (~0.25 V vs. Li/Li+), proper theoretical specific capacity (~1960 mAh·g−1), lesser volume variation (~100%), and expanded cycling lifespan compared to silicon and other alloys [22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39]. In addition, SiO2 is one of the most abundant materials on earth, and its environmentally friendly and low-cost nature further turns it into an alternative electrode material [25,26,27,28,29,30]. However, the development of SiO2-based anode materials so far has been impeded due to its poor electrical conductivity and sluggish charge transfer kinetics. To overcome these limitations, extensive research efforts have been dedicated to the development of SiO2-based composite materials and structures, such as carbon-coated SiO2 particles [32,35,36], SiO2/Cu polyacrylonitrile-C composite [33], Sn(SnO2)–SiO2/graphene nanocomposites [37], Bi2S3@SiO2 core-shell microwires [38], Ni/SiO2 hierarchical hollow spheres [39], and so on [34,40]. Even though significant progress has been achieved, the commercialization of SiO2-based anodes is still restricted by the low electrochemical activity. On the other hand, metallic tin (Sn) has a good electrical conductivity (8.7 × 106 S·m−1) and low melting point (~232 °C). When used as an electrode material, it shows a high theoretical capacity of 994 mAh·g−1 with the formation of Li4.4Sn and fitting working potential (~0.5 V). However, the huge volume changes (~260%) it suffers during Li alloying/dealloying can always lead to rapid fading of capacity and subsequently poor cyclability [41,42,43,44,45,46,47].
Taking advantages of both SiO2 and Sn, herein, a feasible tactics was developed to construct porous silicon oxide@Pourous Carbon@Tin (SiO2@Pc@Sn) composites with tunable SiO2 to Sn molar ratios to synergistically storage Li in both porous SiO2 and Sn. The SiO2@Pc@Sn composite was fabricated using a simple and scalable freezing-drying and low-temperature thermal-melting combined method. The obtained composites possessed several advantageous features: Firstly, the porous structure in the composites largely shortened the transport path for Li ions and provided the buffering space for volume change during the charging/discharging process; secondly, porous C (Pc) and SiO2 provided a rigid skeleton with long cycle stability; thirdly, the presence of Sn and Pc could improve the electrical conductivity of the SiO2-based electrode. The synergetic effect of porous SiO2, Pc, and Sn nano-ball empowered the fabricated SiO2@Pc@Sn composite electrodes to be competent to show good electrochemical performance, including a stable and long cycling life, low electrochemical impedance, and enhanced specific capacity, which demonstrated a fascinating potential as a promising anode for the next-generation LIBs.

2. Experimental Section

2.1. Preparation of SiO2@Pc Composites Material

Diatomite (325 mesh, Sinopharm Chemical Reagent Co., Ltd. Shanghai, China) was ground for 10 h by a high-energy ball mill, then the sample was dispersed in the glucose aqueous solution by ultrasonic for 15 mi. After that, the freeze-drying process for 60 h was carried out, in which the mass ratio of SiO2 to glucose was 1:1 (w/w). Then, the freeze-drying samples were transferred to a tube furnace and carbonized for 3 h at 500 °C in an Ar/H2 gas environment to obtain SiO2@Pc composites.

2.2. Preparation of SiO2@Pc @Sn Composites Material

The previously obtained SiO2@Pc from the above step was weighed at ratio of 1:1 (w/w) with Sn powders (325 mesh, Sinopharm Chemical ReagentCo., Ltd. Shanghai, China) and mixed fully. Then the mixture was transferred to a tubular furnace (OTF-1200X), and heated at a rate of 5 °C/min to 300 °C, keeping for 1 h in an Ar/H2 protect gas. After that, the sample of SiO2@Pc@Sn was obtained via rapid cooling.

2.3. Battery Assembly and Electrochemical Measurements

The Celgard 2320 (Shenzhen, China) film was used as a membrane and lithium foil as a pair electrode to conduct electrochemical experiments on the CR2032 (Shenzhen, China) coin battery. The experimental electrolyte was configured of 1.0 M LiPF6 dissolved in ethylene carbonate (EC) and diethyl carbonate (DEC) by volume 1:1. The working electrode was composed of 70 wt.% active materials, 15 wt.% polyvinylidene fluoride (PVDF) binder, and 15 wt.% Super P. After fully mixing and grinding, the slurry was spread on the copper foil evenly, and then dried in an oven at 50 °C for 12 h. The battery was assembled in a glove box filled with Ar gas and the oxygen and water content below 1.0 PPM. After assemblage, the batteries were set aside for 8 h at room temperature. The electrochemical performance was tested by electrochemical impedance spectroscopy (EIS) and cyclic voltammetry (CV) on a DH7001 electrochemical workstation, and the scanning rate of CV was set in the range of 0.1–0.5 mV s−1 with an applied potential 2–0 V, and the frequency range for EIS measurement was set in 1.0 MHz–0.1 Hz. All batteries’ simulation cycling and charge/discharge were conducted on a land battery cabinet (LAND CT2001A, Wuhan, China). In the batteries’ evaluation process, the cut-off voltage was 0.005 V vs Li/Li+ for discharge and 1.5 V for charge. All specific capacity was calculated based on the proportion of the active material in the whole electrode.

2.4. Characterization

The morphology and structure of SiO2@Pc@Sn were obtained by a field-emission scanning electron microscope (FESEM, JEOL JMS-7001-F, JEOL, Tokyo, Japan). The element mapping was measured by the EDS instrument equipped in the FESEM. The phase composition of the material was obtained by X’Pert PRO diffractometer (XRD, Shimadzu, Japan: XRD-6000, Cu–K radiation, 0.15406 nm, λ = 1.5406 Å), the measurement angle was between 10–80°, and the scanning rate was 10°/min. Raman spectroscopy was used to characterize the form of carbon, and the excited wavelength of the laser was 532 nm (Raman, InVia and Ntegra Spectra, Renishaw & NT-MDT, London, UK). The thermogravimetry (TG) analysis was performed by the vertical zero friction dilatometer L75VS Linseis (Selb, Germany) from 25 to 800 °C in air to calculate the carbon weight percent in the composite.

3. Results and Discussion

The preparation flow chart of the SiO2@Pc@Sn composite is shown in Figure 1. As depicted in the schematic diagram, firstly, the SiO2@Pc composite with a porous structure was prepared by the freeze-drying method, and secondly, the SiO2@Pc@Sn composite was obtained via the low-temperature thermal melting and self-assembly process.
Figure 2a shows the comparison of the XRD pattern of SiO2, SiO2@Pc, SiO2@Pc@Sn, and PDF card of standard XRD, correspondingly. The characteristic peaks at 21.6° and 35.6° belonged to SiO2 [26], and the peak value of SiO2@Pc was consistent with that of SiO2, indicating that SiO2 did not change significantly after Pc coating. In the SiO2@Pc@Sn composite, the characteristic peaks for Sn were centered at 30.6°, 32.1°, 43.9°, 44.9°, and 55.3°. The characteristic peaks that belonged to SiO2 and Sn in the SiO2@Pc@Sn composite were matched well with the standard PDF cards. The synthesized Pc was characterized by the Raman spectrum as indicated in Figure 2. It can be seen that the peaks around 1357 and 1591 cm−1 corresponded to the disordered D peak and graphitized G peak for the obtained Pc. The D peak was generally the crystallization defect of carbon atoms and the G peak represented the in-plane vibration of sp2 hybridization of carbon atoms [48,49]. The existence of the G peak and D peak indicates that the microstructure of Pc in the SiO2@Pc@Sn composite was graphitized carbon. In addition, the main peak of SiO2 at 480–490 cm−1 was not present in the current Raman spectrum 500–3000 cm−1 [27], while the peak around 1080 cm−1 under D peak of carbon was also invisible due to the encapsulation of SiO2 in a carbon shell [26]. From the TG analysis result in the Raman spectrum, the C weight percent in the SiO2@Pc composite was ~27.3%.
The morphology and elements distribution of the obtained SiO2@Pc@Sn composite was measured by scanning electron microscopy (SEM). As shown in Figure 3, the pristine SiO2 was in the shape of a sunflower (Figure 3a), and its average size was between 20–40 μm, with many nano-pores on the surface (Figure 3b). The size of the pores was in the range of 50–600 nm (Figure 3b). From the images as shown in Figure 3c, after rapid cooling, the Sn nano-balls were formed and dispersed uniformly in the SiO2@Pc composite, which filled into the pores in Pc or embedded among the SiO2@Pc blocks. As shown in Figure 3f, the statistical distribution of size for the Sn balls was mainly centered around 100 nm. The element mapping for Si, Sn, and C in the SiO2@Pc@Sn composite is shown in Figure 3g–i. From the result, it is found that three elements are distributed in all the areas detected in the SiO2@Pc@Sn.
The electrochemical performance is displayed in Figure 4. The charge/discharge curves of different samples at the same current density of 100 mA·g−1 are compared in Figure 4a. It was found that the first discharge capacity reached to 1228 for SiO2@Pc@Sn, 990 for SiO2@Pc, 672 for bare SiO2, and 352 mA·h·g−1 for the synthesized Pc. The initial coulomb efficiency (ICE) was 52%, 37.7%, 29.9%, and 27.4%, respectively. The improved specific capacity and ICE of SiO2@Pc@Sn were attributed to the fact that Pc and Sn can improve the electrical conductivity of the composite and enhance the electrochemical activity of SiO2. The poor conductivity of SiO2 was the cause of the low initial coulombic efficiency, and most of Li ions combined with SiO2 to produce irreversible Li4SiO4 and Li2O at the first charge and discharge [50,51], while the presence of Pc and Sn improved the whole electrical conductivity of SiO2@Pc@Sn, which is helpful for the electrons to arrive at the surface of SiO2, and as a result, facilitated the Li ions transfer in the composite. Meanwhile, the existence of Pc further prevented the by-products brought by the direct reaction between electrolyte and SiO2 and Sn, thus improving the ICE of the composite [52]. The cycling performance at 100 mA·g−1 is compared in Figure 4b. It is evident that SiO2@Pc@Sn shows the highest specific capacity and best capacity retention through 100 cycles. While for bare SiO2, the capacity underwent continuous increasing during the initial 100 cycles that changed from the initial 200 to 400 mA·h·g−1 after 100 cycles. Though the capacity of SiO2@Pc could not reach SiO2@Pc@Sn, it was still better than bare SiO2 and Pc. Moreover, the rate capability for different samples was listed in Figure 4c. It is clear that the SiO2@Pc@Sn exhibited capacities of 650, 610, 580, and 520 mA·h·g−1 at 100, 200, 500, and 1000 mA·g−1, respectively, whereas the bare SiO2 and SiO2@Pc exhibited a lower capacity and faster capacity decay. Obviously, the rate capability of SiO2@Pc@Sn was better than that of the others, especially at high current density due to the fact that Pc and Sn had better conductivity than SiO2, which provided higher mobility for Li ion diffusion through the whole electrode. Without the addition of other assistance, such as an electrolyte additive (for instance, FEC) and so on, the good rate capability and stable cycling performance of SiO2@Pc@Sn was believed to be originated from the unique structure. Firstly, the built-in void in Pc and SiO2 shorted the Li ions transfer distance in the electrode; secondly, Sn and Pc were conductive for electrons and ions, and the face-to-face contact between Pc and SiO2 as well as Sn aroused more efficient channels for fast transfer of electrons and Li ions [15,17]. The CV test could detect electrode surface reaction process, electrochemical activity, and reversibility of the active material. Figure 4d is the CV curve of SiO2@Pc@Sn. As shown, the cathode peak of 0.98 V in the first cathode scan was caused by electrolyte decomposition and the generation of the SEI layer [50], while the reductive peaks at 0.63 and 0.32 V were attributed to the phase LixSiOy and LixSi formed while SiO2 was combined with Li+ in the discharge process [30,40]. On the contrary, the oxidation peaks at 0.64, 0.74, and 0.82 V were caused by the dealloying process of LixSn and Li2Si2O5 [53].
Furthermore, the electrochemical impedance spectra (EIS) were compared and analyzed in Figure 5. From the Nyquist plots diagram of different samples, as shown in Figure 5a, it was found that all impedance spectra consisted of a semicircle in the high frequency region and an inclined line in the low frequency region, which corresponded to the Li+ migration resistance and interface contact resistance in the active materials, respectively [54,55]. The impedance resistance was 205 for bare SiO2, 129 for the SiO2@Pc, and 77 Ω for SiO2@Pc@Sn, indicating that the migration impedance of Li ions was minimal in the active material of SiO2@Pc@Sn. In addition, EIS was often used in the qualitative determination of Li ions’ diffusion coefficient in LIB materials. Figure 5b is the plots of correlation curve of Zre (real part of impedance) and w−1/2 (w is the frequency) within the frequency range of 1–0.1 Hz for the electrode composed of different materials. According to the relation, Zre = G−k·w−1/2, where, k is the slope of the correlation curve between Zre and w−1/2, from which the diffusion coefficient of lithium ions in different electrode materials can be qualitatively deduced [20]. In order to ensure the accuracy of the experiment, each set of data tested 5–10 batteries for analysis. From the fitting results (Figure 5b), the curve slopes k of SiO2, SiO2@Pc, and SiO2@Pc@Sn were 0.33, 0.20 and 0.07, respectively. The result showed that the diffusion coefficient of Li+ was the largest in the SiO2@Pc @Sn composite according to the relation formula [21], DLi+ = A/k, where, A is constant related to the Li ions content and electrode area, etc.
Meanwhile, the CV measurement with different scanning rates (mV·s−1) is shown in Figure 6. The CV curves of SiO2 (Figure 6a), SiO2@Pc (Figure 6b), and SiO2@Pc@Sn (Figure 6c) electrodes under scanning rates of 0.1, 0.2, 0.3, 0.4, and 0.5 mV·s−1 were measured. The obtained peak current Imax and the quadratic root of scanning rate v1/2 were fitted linearly, from which the diffusion strength of Li+ in different electrode materials could be qualitatively determined. According to the Randle–Sevcik equation: Imax = A1/2·DoLi [56], where A is the constant related to charge and surface area, v is the scanning rate of CV, and DoLi is the diffusion coefficient of Li+ in oxide [54]. According to the test results shown in Figure 6d, the slopes were 0.49, 0.44, and 0.12 for SiO2@Pc@Sn, SiO2@Pc, and bare SiO2, respectively, indicating that the diffusion coefficient of Li+ ions was the highest in SiO2@Pc@Sn compared with the two others [57]. The result was also consistent with the fitting results of the correlation curve between Zre and v−1/2 of EIS in the frequency range of 1–0.1 Hz (Figure 5b), which further demonstrated that Sn could improve the electrochemical performance of SiO2-based anode materials for LIBs.

4. Summary

The SiO2@Pc@Sn composite anode material was prepared by the freeze-drying and low-temperature thermal melting method, which exhibited improved electrochemical performance and faster Li+ transfer kinetic. The synergetic effect of porous carbon, SiO2, and Sn endows the as-fabricated SiO2@Pc@Sn composites to be competent to show good electrochemical performance. When used as an anode in LIBs, the SiO2@Pc@Sn composite could deliver a large reversible capacity of 650 at 100 mA·g−1, a remarkable rate capability of 500 retained at 1000 mA·g−1, and a long-term cycling durability with ~87% capacity retention over 100 cycles. EIS and CV measurements demonstrated that, with the participation of Sn phase and Pc, the diffusion and migration kinetics of Li ions in SiO2@Pc@Sn composites was significantly improved. The understanding of the synergistic effect of Li storage between SiO2 and Sn in this work will not only provide insight towards exploring new SiO2-based anode materials, but also shed light on the design of other low-cost and environmentally friendly electrode materials for the next-generation LIBs.

Author Contributions

Conceptualization, X.D.; methodology, X.D.; software, D.L.; validation, X.D., D.L. and H.Z.; formal analysis, X.D. and D.L.; investigation, X.D., D.L.; resources, X.D.; data curation, D.L., H.Z.; writing—original draft preparation, X.D.; writing—review and editing, X.D.; visualization, X.D., D.L. and H.Z.; supervision, X.D.; project administration, X.D.; funding acquisition, X.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No.11874282, 11604245, 11981240429), the Six Talent Peaks Project in Jiangsu Province (No. 2019-XNY-074), the Vice President Project of Industry-University-Research Cooperation in Science and Technology of Jiangsu Province (No. BY2020675).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, J.-N.; Li, Q.; Ouyang, C.; Yu, X.; Ge, M.; Huang, X.; Hu, E.; Ma, C.; Li, S.; Xiao, R.; et al. Trace doping of multiple elements enables stable battery cycling of LiCoO2 at 4.6 V. Nat. Energy 2019, 4, 594–603. [Google Scholar] [CrossRef] [Green Version]
  2. Yang, C.; Ji, X.; Fan, X.; Gao, T.; Suo, L.; Wang, F.; Sun, W.; Chen, J.; Chen, L.; Han, F.; et al. Flexible Aqueous Li-Ion Battery with High Energy and Power Densities. Adv. Mater. 2017, 29. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, H.; Wu, H.-H.; Ge, M.; Li, L.; Yuan, Y.; Yao, Q.; Chen, J.; Xia, L.; Zheng, J.; Zhao, Y.; et al. Simultaneously Dual Modification of Ni-Rich Layered Oxide Cathode for High-Energy Lithium-Ion Batteries. Adv. Funct. Mater. 2019, 29, 1808–1825. [Google Scholar] [CrossRef]
  4. Li, L.; Fang, C.; Wei, W.; Zhang, L.; Ye, Z.; He, G.; Huang, Y. Nano-ordered structure regulation in delithiated Si an-ode triggered by homogeneous and stable Li-ion diffusion at the interface. Nano Energy 2020, 72, 104651. [Google Scholar] [CrossRef]
  5. Han, M.; Yu, J. Subnanoscopically and homogeneously dispersed SiOx/C composite spheres for high-performance lithium ion battery anodes. J. Power Sources 2019, 414, 435–443. [Google Scholar] [CrossRef]
  6. Doughty, D.H.; Crafts, C.C. FreedomCAR Electrical Energy Storage System Abuse Test Manual for Electric and Hybrid Electric Vehicle Applications; Sandia Report SAND2005-3123; Sandia National Laboratories: Albuquerque, NM, USA; Livermore, CS, USA, 2006. [Google Scholar]
  7. Commission NDAR, QC/T-743-2006. Automotive standard of the People’s Republic of China: Lithium Ion Bateries for Electric Vehicles. 2006. Available online: https://max.book118.com/html/2017/0726/124432928.shtm (accessed on 13 February 2021).
  8. Perner, A.; Vetter, J. Lithium-ion batteries for hybrid electric vehicles. In Advances in Battery Technologies for Electric Vehicles; Elsevier BV: Amsterdam, The Netherlands, 2015; pp. 173–190. [Google Scholar]
  9. Cabana, J.; Monconduit, L.; Larcher, D.; Palacín, M.R. Beyond Intercalation-Based Li-Ion Batteries: The State of the Art and Challenges of Electrode Materials Reacting Through Conversion Reactions. Adv. Mater. 2010, 22, E170–E192. [Google Scholar] [CrossRef]
  10. Whittingham, M.S. ChemInform Abstract: Ultimate Limits to Intercalation Reactions for Lithium Batteries. ChemInform 2015, 114, 11414–11443. [Google Scholar] [CrossRef]
  11. Zhang, Q.; Uchaker, E.; Candelaria, S.L.; Cao, G. ChemInform Abstract: Nanomaterials for Energy Conversion and Storage. Cheminform 2013, 44. [Google Scholar] [CrossRef]
  12. Masquelier, C.; Croguennec, L. Polyanionic (Phosphates, Silicates, Sulfates) Frameworks as Electrode Materials for Rechargeable Li (or Na) Batteries. Chem. Rev. 2013, 113, 6552–6591. [Google Scholar] [CrossRef] [PubMed]
  13. Song, H.-K.; Lee, K.T.; Kim, M.G.; Nazar, L.F.; Cho, J. Recent Progress in Nanostructured Cathode Materials for Lithium Secondary Batteries. Adv. Funct. Mater. 2010, 20, 3818–3834. [Google Scholar] [CrossRef]
  14. Jia, H.; Gao, P.; Yang, J.; Wang, J.; Nuli, Y.; Yang, Z. Novel Three-Dimensional Mesoporous Silicon for High Power Lithium-Ion Battery Anode Material. Adv. Energy Mater. 2011, 1, 1036–1039. [Google Scholar] [CrossRef]
  15. Ding, X.; Liu, X.; Huang, Y.; Zhang, X.; Zhao, Q.; Xiang, X.; Li, G.; He, P.; Wen, Z.; Li, J.; et al. Enhanced electrochemical performance promoted by monolayer graphene and void space in silicon composite anode materials. Nano Energy 2016, 27, 647–657. [Google Scholar] [CrossRef]
  16. Tao, W.; Wang, P.; You, Y.; Park, K.; Wang, C.-Y.; Li, Y.-K.; Cao, F.-F.; Xin, S. Strategies for improving the storage per-formance of silicon-based anodes in lithium-ion batteries. Nano Res. 2019, 12, 1739–1749. [Google Scholar] [CrossRef]
  17. Ding, X.; Wang, Y. Bilayer-graphene-coated Si nanoparticles as advanced anodes for high-rate lithium-ion bat-teries. Electrochim. Acta 2020, 329, 134975. [Google Scholar] [CrossRef]
  18. Luo, J.; Tao, X.; Zhang, J.; Xia, Y.; Huang, H.; Zhang, L.; Gan, Y.; Liang, C.; Zhang, W. Sn4+ Ion Decorated Highly Conductive Ti3C2 MXene: Promising Lithium-Ion Anodes with Enhanced Volumetric Capacity and Cyclic Performance. ACS Nano 2016, 10, 2491–2499. [Google Scholar] [CrossRef]
  19. Ding, X.; Wang, H.; Liu, X.; Gao, Z.; Huang, Y.; Lv, D.; He, P.; Huang, Y. Advanced anodes composed of graphene encapsulated nano-silicon in a carbon nanotube network. Rsc Adv. 2017, 7, 15694–15701. [Google Scholar] [CrossRef] [Green Version]
  20. Li, W.; Feng, X.; Chen, Y. High performance lithium battery anode materials by coating SiO2 nanowire arrays with PEO. New J. Chem. 2019, 43, 14609–14615. [Google Scholar] [CrossRef]
  21. Yi, X.; Yu, W.J.; Tsiamtsouri, M.A.; Zhang, F.; He, W.; Dai, Q.; Hu, S.; Tong, H.; Zheng, J.; Zhang, B. Highly conductive C-Si@G nanocomposite as a high-performance anode material for Li-ion batteries. Electrochim. Acta 2018, 295, 295. [Google Scholar] [CrossRef]
  22. Liu, Y.; Jing, S.X.; Luo, S.H.; Li, S.H. Silicon oxides: A promising family of anode materials for lithium-ion batteries. Nat. Prod. Rep. 2019, 36, 626–665. [Google Scholar] [CrossRef] [PubMed]
  23. Tang, C.; Liu, Y.; Xu, C.; Zhu, J.; Wei, X.; Zhou, L.; He, L.; Yang, W.; Mai, L. Ultrafine Nickel-Nanoparticle-Enabled SiO2 Hierarchical Hollow Spheres for High-Performance Lithium Storage. Adv. Funct. Mater. 2018, 28, 1704561. [Google Scholar] [CrossRef]
  24. 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, 4605. [Google Scholar]
  25. Chang, W.-S.; Park, C.-M.; Kim, J.-H.; Kim, Y.-U.; Jeong, G.; Sohn, H.-J. Quartz (SiO2): A new energy storage anode mate-rial for Li-ion batteries. Energy Environ. Sci. 2012, 5, 6895. [Google Scholar] [CrossRef]
  26. Meng, J.; Cao, Y.; Suo, Y.; Liu, Y.; Zhang, J.; Zheng, X. Facile Fabrication of 3D SiO2@Graphene Aerogel Composites as Anode Material for Lithium Ion Batteries. Electrochim. Acta 2015, 176, 1001–1009. [Google Scholar] [CrossRef]
  27. Ichikawa, S.; Suda, J.; Sato, T.; Suzuki, Y. Lattice dynamics and temperature dependence of the first-order Raman spectra for? SiO2 crystals. J. Raman Spectrosc. 2003, 34, 135–141. [Google Scholar] [CrossRef]
  28. Brijesh, K.; Dhanush, P.C.; Vinayraj, S.; Nagaraja, H.S. Monoclinic Wolframite ZnWO4/SiO2 nanocomposite as an anode material for lithium ion battery. Mater. Lett. 2020, 275, 128108. [Google Scholar]
  29. Di, F.W.; Wang, N.; Li, L.X.; Geng, X.; Yang, H.M.; Zhou, W.M.; Sun, C.G.; An, B.G. Coral-like porous composite material of silicon and carbon synthesized by using diatomite as self-template and precursor with a good performance as an-ode of lithium-ions battery. J. Alloy. Compd. 2021, 854. [Google Scholar] [CrossRef]
  30. Shen, D.; Huang, C.; Gan, L.; Liu, J.; Gong, Z.; Long, M. Rational Design of Si@SiO2/C Composites Using Sustainable Cellulose as a Carbon Resource for Anodes in Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 7946–7954. [Google Scholar] [CrossRef]
  31. Dong, X.; Zheng, X.; Deng, Y.; Wang, L.; Ju, Z. SiO2/N-doped graphene aerogel composite anode for lithium-ion batteries. J. Mater. Sci. 2020, 55, 13023–13035. [Google Scholar] [CrossRef]
  32. Yao, Y.; Zhang, J.; Xue, L.; Huang, T.; Yu, A. Carbon-coated SiO2 nanoparticles as anode material for lithium ion bat-teries. J. Power Sources 2011, 196, 10240–10243. [Google Scholar] [CrossRef]
  33. Li, M.; Li, J.; Li, K.; Zhao, Y.; Zhang, Y.; Gosselink, D.; Chen, P. SiO2/Cu/polyacrylonitrile-C composite as anode mate-rial in lithium ion batteries. J. Power Sources 2013, 240, 659–666. [Google Scholar] [CrossRef]
  34. Li, W.; Wang, F.; Ma, M.; Zhou, J.; Liu, Y.; Chen, Y. Preparation of SiO2 nanowire arrays as anode material with en-hanced lithium storage performance. RSC Adv. 2018, 8, 33652–33658. [Google Scholar] [CrossRef] [Green Version]
  35. Liang, Y.; Cai, L.; Chen, L.; Lin, X.; Fu, R.; Zhang, M.Q.; Wu, D. Silica nanonetwork confined in nitrogen-doped ordered mesoporous carbon framework for high-performance lithium-ion battery anodes. Nanoscale 2015, 7, 3971–3975. [Google Scholar] [CrossRef]
  36. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Namsar, O.; Autthawong, T.; Laokawee, V.; Boonprachai, R.; Haruta, M.; Kurata, H.; Yu, A.; Chairuangsri, T.; Sarakonsri, T. Improved electrochemical performance of anode materials for high energy density lithium-ion batteries through Sn(SnO2)–SiO2/graphene-based nanocomposites prepared by a facile and low-cost approach. Sustain. Energy Fuels 2020, 4, 4625–4636. [Google Scholar] [CrossRef]
  38. Tang, C.; Li, N.; Sheng, J.; Zhou, L.; He, L.; Zhu, J.; Li, F.; Liu, Y.; Mai, L. Facile Synthesis of Bi2S3@SiO2 Core-Shell Microwires as High-Performance Anode Materials for Lithium-Ion Batteries. J. Electrochem. Soc. 2017, 164, 6110–6115. [Google Scholar] [CrossRef]
  39. Tu, J.; Yuan, Y.; Zhan, P.; Jiao, H.; Wang, X.; Zhu, H.; Jiao, S. Straightforward Approach toward SiO2 Nanospheres and Their Superior Lithium Storage Performance. J. Phys. Chem. C 2014, 118, 7357–7362. [Google Scholar] [CrossRef]
  40. Liu, D.D.; Kong, Z.; Liu, X.H.; Fu, A.P.; Wang, Y.Q.; Guo, Y.G.; Guo, P.Z.; Li, H.L.; Zhao, X.S. Spray-Drying induced assembly of skeleton-structured SnO2/Graphene composite spheres as superior anode materials for high-performance lithium-ion batteries. ACS Appl. Mater. Inter. 2018, 10, 2515–2525. [Google Scholar] [CrossRef]
  41. Xin, F.; Whittingham, M.S. Challenges and Development of Tin-Based Anode with High Volumetric Capacity for Li-Ion Batteries. Electrochem. Energy Rev. 2020, 3, 643–655. [Google Scholar] [CrossRef]
  42. Wang, K.; He, X.; Ren, J.; Jiang, C.; Wan, C. Preparation of Sn/C microsphere composite anode for lithium-ion batteries via carbonthermal reduction. Electrochem. Solid St. 2006, 9, 320–323. [Google Scholar]
  43. Derrien, G.; Hassoun, J.; Panero, S.; Scrosati, B. Nanostructured Sn–C Composite as an Advanced Anode Material in High-Performance Lithium-Ion Batteries. Adv. Mater. 2007, 19, 2336–2340. [Google Scholar] [CrossRef]
  44. Zhao, X.; Wang, W.; Hou, Z.; Yu, Y.; Di, Q.; Wu, X.; Wei, G.; Quan, Z.; Zhang, J. Monodisperse tin nanoparticles and hollow tin oxide nanospheres as anode materials for high performance lithium ion batteries. Inorg. Chem. Front. 2018, 6, 473–476. [Google Scholar] [CrossRef]
  45. Tolosa, A.; Widmaier, M.; Krüner, B.; Griffin, J.M.; Presser, V. Continuous silicon oxycarbide fiber mats with tin nanoparticles as high capacity anode for lithium-ion batterie. Sustain. Energy Fuels 2018, 2, 215–228. [Google Scholar]
  46. Huang, X.; Cui, S.; Chang, J.; Hallac, P.B.; Fell, C.R.; Luo, Y.; Metz, B.; Jiang, J.; Hurley, P.T.; Chen, J. A Hierarchical Tin/Carbon Composite as an Anode for Lithium-Ion Batteries with a Long Cycle Life. Angew. Chem. Int. Ed. 2014, 54, 1490–1493. [Google Scholar] [CrossRef] [PubMed]
  47. Yan, Y.; Ben, L.; Zhan, Y.; Huang, X. Nano-Sn embedded in expanded graphite as anode for lithium ion batteries with improved low temperature electrochemical performance. Electrochim. Acta 2016, 187, 186–192. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Liu, Z.; Zhang, Y.; Mentbayeva, A.; Wang, X.; Maximov, M.Y.; Liu, B.; Bakenov, Z.; Yin, F. Facile Synthesis of SiO2@C Nanoparticles Anchored on MWNT as High-Performance Anode Materials for Li-ion Batteries. Nanoscale Res. Lett. 2017, 12, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Wu, W.; Liang, Y.; Ma, H.; Peng, Y.; Yang, H. Insights into the conversion behavior of SiO-C hybrid with pre-treated graphite as anodes for Li-ion batteries. Electrochim. Acta 2016, 187, 473–479. [Google Scholar] [CrossRef]
  50. Ma, B.; Luo, J.; Deng, X.; Wu, Z.; Luo, Z.; Wang, X.; Wang, Y. Hollow Silicon–Tin Nanospheres Encapsulated by N-Doped Carbon as Anode Materials for Lithium-Ion Batteries. ACS Appl. Nano Mater. 2018, 1, 6989–6999. [Google Scholar] [CrossRef]
  51. Blanco, M.V.; Renman, V.; Vullum-Bruer, F.; Svensson, A.M. Nanostructured diatom earth SiO2 negative electrodes with superior electrochemical performance for lithium ion batteries. RSC Adv. 2020, 10, 33490–33498. [Google Scholar] [CrossRef]
  52. Ma, B.; Lu, B.; Luo, J.; Deng, X.; Wu, Z.; Wang, X. The hollow mesoporous silicon nanobox dually encapsulated by SnO2/C as anode material of lithium ion battery. Electrochim. Acta 2018, 288, 61–70. [Google Scholar] [CrossRef]
  53. Gu, Z.; Xia, X.; Liu, C.; Hu, X.; Chen, Y.; Wang, Z.; Liu, H. Yolk structure of porous C/SiO2 /C composites as. anode for lithium-ion batteries with quickly activated SiO2. J. Alloy. Compd. 2018, 757, 265–272. [Google Scholar] [CrossRef]
  54. Chen, S.; Shen, L.; Aken, P.v.A.; Maier, J.; Yu, Y. Dual-Functionalized Double Carbon Shells Coated Silicon Nanoparticles for High Performance Lithium-Ion Batteries. Adv. Mater. 2017, 29. [Google Scholar] [CrossRef]
  55. He, Y.; Xu, G.; Wang, C.; Xu, L.; Zhang, K. Horsetail-derived Si@N-doped carbon as low-cost and long cycle life anode for Li-ion half/full cells. Electrochim. Acta 2018, 264, 173–182. [Google Scholar] [CrossRef]
  56. Park, C.-K.; Park, S.-B.; Oh, S.-H.; Jang, H.; Cho, W.-I. Li Ion Diffusivity and Improved Electrochemical Performances of the Carbon Coated LiFePO 4. Bull. Korean Chem. Soc. 2011, 32, 836–840. [Google Scholar] [CrossRef] [Green Version]
  57. Wang, H.; Yang, X.; Wu, Q.; Zhang, Q.; Chen, H.; Jing, H.; Wang, J.; Mi, S.-B.; Rogach, A.L.; Niu, C. Encapsulating Silica/Antimony into Porous Electrospun Carbon Nanofibers with Robust Structure Stability for High-Efficiency Lithium Storage. ACS Nano 2018, 12, 3406–3416. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of the synthesize process of silicon oxide@Pourous Carbon@Tin (SiO2@Pc@Sn) composite.
Figure 1. Schematic diagram of the synthesize process of silicon oxide@Pourous Carbon@Tin (SiO2@Pc@Sn) composite.
Materials 14 01071 g001
Figure 2. (a) XRD patterns of different samples; (b) Raman spectrum of SiO2@Pc; (c) thermogravimetry (TG) diagram for the SiO2@Pc.
Figure 2. (a) XRD patterns of different samples; (b) Raman spectrum of SiO2@Pc; (c) thermogravimetry (TG) diagram for the SiO2@Pc.
Materials 14 01071 g002
Figure 3. (a,b) SEM of bare SiO2 with different magnification; (ce) SEM of SiO2@Pc@Sn with different magnification; (f) size distribution for Sn in the SiO2@Pc@Sn composite; (g,i) element mapping of SiO2@Pc@Sn for Si (g); Sn (h); and C (i).
Figure 3. (a,b) SEM of bare SiO2 with different magnification; (ce) SEM of SiO2@Pc@Sn with different magnification; (f) size distribution for Sn in the SiO2@Pc@Sn composite; (g,i) element mapping of SiO2@Pc@Sn for Si (g); Sn (h); and C (i).
Materials 14 01071 g003
Figure 4. (a) Charge/discharge curve of different samples; (b) cycling performance curve of different samples; (c) the rate performance for different samples; (d) cyclic voltammetry (CV) curve of SiO2@Pc@Sn.
Figure 4. (a) Charge/discharge curve of different samples; (b) cycling performance curve of different samples; (c) the rate performance for different samples; (d) cyclic voltammetry (CV) curve of SiO2@Pc@Sn.
Materials 14 01071 g004
Figure 5. (a) Comparison for the Nyquist diagram of different samples; (b) impedance real part Zre Vs w−1/2 in the frequency range 1–0.1 Hz.
Figure 5. (a) Comparison for the Nyquist diagram of different samples; (b) impedance real part Zre Vs w−1/2 in the frequency range 1–0.1 Hz.
Materials 14 01071 g005
Figure 6. (ac) CV curve of SiO2, SiO2@Pc, and SiO2@Pc@Sn with different scanning rates of 0.1–0.5 mV·s−1; (d) relationship between scan rate and peak current.
Figure 6. (ac) CV curve of SiO2, SiO2@Pc, and SiO2@Pc@Sn with different scanning rates of 0.1–0.5 mV·s−1; (d) relationship between scan rate and peak current.
Materials 14 01071 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ding, X.; Liang, D.; Zhao, H. Enhanced Electrochemical Performance Promoted by Tin in Silica Anode Materials for Stable and High-Capacity Lithium-Ion Batteries. Materials 2021, 14, 1071. https://doi.org/10.3390/ma14051071

AMA Style

Ding X, Liang D, Zhao H. Enhanced Electrochemical Performance Promoted by Tin in Silica Anode Materials for Stable and High-Capacity Lithium-Ion Batteries. Materials. 2021; 14(5):1071. https://doi.org/10.3390/ma14051071

Chicago/Turabian Style

Ding, Xuli, Daowei Liang, and Hongda Zhao. 2021. "Enhanced Electrochemical Performance Promoted by Tin in Silica Anode Materials for Stable and High-Capacity Lithium-Ion Batteries" Materials 14, no. 5: 1071. https://doi.org/10.3390/ma14051071

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