Next Article in Journal
Effects of Bias Voltage and Substrate Temperature on the Mechanical Properties and Oxidation Behavior of CrWSiN Films
Next Article in Special Issue
Modelling the Impact of Graphene Coating of Different Thicknesses on Polyimide Substrate on the Secondary Electron Yield
Previous Article in Journal
Impact of Bulk and Nano Bismuth Oxide on the Attenuation Parameters of Bentonite Barite Composites
Previous Article in Special Issue
Effect of ZrH2 Doping on Electron Emission Performance of Rare Earth Tungsten Electrode
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Properties of Carbon Nanobeads and Mesophase-Pitch-Based Graphite Fibers as Anodes for Rechargeable Lithium-Ion Batteries

1
School of Energy and Power Engineering, North University of China, Taiyuan 030001, China
2
National Key Laboratory of Advanced Composites, AECC Beijing Institute of Aeronautical Materials, Beijing 100095, China
*
Authors to whom correspondence should be addressed.
Coatings 2023, 13(10), 1671; https://doi.org/10.3390/coatings13101671
Submission received: 21 August 2023 / Revised: 18 September 2023 / Accepted: 22 September 2023 / Published: 24 September 2023

Abstract

:
Various strategies have been devised to enhance the specific capacity (over 372 mA h g−1) and the cycle life of pure carbon-based anode materials for lithium-ion batteries (LIBs). Herein, we have prepared carbon nanobeads (CNBs) with inner onion-like carbon layer structures using chemical vapor deposition. Additionally, mesophase-pitch-based graphite fibers (MPGFs) were fabricated via the melt spinning method. MPGFs had a reversible capacity of 328.2 mA h g−1 with a steady Coulombic efficiency after 105 cycles (at 0.05 mA g−1). When cycled at 0.2 A g−1, the CNBs kept a reversible capacity of 586.2 mA h g−1 after 256 cycles with stable Coulombic efficiency. The onion-like carbon layer structures of CNBs were beneficial to the storage of lithium ions. In this work, CNBs were fabricated with inner onion-like carbon layer structures to enhance the electrochemical performance of LIBs.

1. Introduction

Compared to the past, with the increasing development of global society, the demand for energy and quality of life have further improved. Facing environmental pollution and fossil resource consumption, renewable energy has been developed and is utilized in industrial production and daily life [1,2,3,4]. Hence, energy storage devices have become one of the most important key links. Rechargeable lithium-ion batteries (LIBs) can store and release electric energy for electronic products. During 2019, John B. Goodenough, M. Stanley Whittingham, and Akira Yoshino won the Nobel Prize due to their great contributions to LIBs [5]. In LIBs, artificial graphite was prepared as the anode material for commercial application [6,7,8,9]. Many factories produce graphite for LIBs in the market. In previous research, carbon-based materials were prepared and used for LIBs. They can be classified as one-dimensional carbon nanotubes (CNTs) [10,11,12], two-dimensional graphene [13,14,15,16], and graphite [17,18,19,20,21]. Meanwhile, other materials have also been studied as anode materials for LIBs [22,23,24,25]. Silicon possesses a superhigh capacity of 4200 mA h g−1. It has been considered a potential anode for LIBs [26,27,28]. However, it undergoes structure and volume changes with a 400% volume expansion during alloying and conversion reactions. Zhu et al. fabricated a nano-porous silicon network with controllable porosity and thickness. Its first discharge capacity was 2570 mA h g−1, and its corresponding current rate was at C/10. After 20 cycles, the discharge capacity was kept at 1800 mA h g−1. The porous structure effectively reduced the attenuation of energy storage performance in the silicon anode. When tested at a 1 C current rate, its capacity dropped to 1200 mA h g−1 after 200 charge–discharge cycles [29]. Yi cui et al. designed the core–shell structure for silicon-anode materials to obtain high capacity and good Coulombic efficiency [30]. They showed a way to encapsulate silicon microparticles via conformally synthesized cages of multilayered graphene. The prepared graphene cages permitted silicon to expand and shrink during the cycles in order to obtain a high capacity. The graphene cage also formed a stable solid electrolyte interface (SEI) to increase Coulombic efficiency in the early process. At the C/2 current rate, silicon with a graphene cage was kept over 1200 mA h g−1 after 400 cycles. The above method restrained the fracture of silicon microparticles and stabilized SEI during the long cycle life.
Germanium is also a new candidate for carbon-based anode material in LIBs [31,32,33]. It has a theoretical capacity of over 1624 mA h g−1. A two-step method was presented to fabricate the binder-free germanium anode [34]. The two-step realization process contained a plasma-enhanced chemical vapor deposition (CVD) to grow a germanium film. Then, they were treated using hydrofluoric acid electrochemical etching. The germanium anode kept a 95% capacity retention after a long cycle life (at 1 C current rate, 600 cycles, RT). In particular, it could obtain a 1060, 450 mA h g−1 at 10 C and 60 C current rates, respectively. Though the capacity of non-carbon-based anode material is higher than that of ideal graphite, carbon materials have an absolute advantage, considering technical and safety factors in the production process.
Carbon nanotubes have good thermal and electrical conductivity, which is beneficial to enhance the electrical conductivity of the electrode. However, the carbon nanotube needs to combine other assistant materials to obtain a high specific capacity. Using a template synthesis, a three-dimensional architecture composed of one-dimensional graphitized carbon nanotubes and coated with the nitrogen and oxygen elements was prepared [35]. Due to this one-dimensional path, lithium ions can diffuse fast in the intercalation/extraction process. As a freestanding anode for LIBs, it obtained a high capacity of 624 mA h g−1 at a rate of 100 mA g−1. Similarly, boron-doped carbon nanotubes were prepared as anode material for LIBs [36]. The electrode delivered a high capacity of 169 mA h g−1 at 10 A g−1 with excellent long-term stability (5000 cycles). Because of its one-dimensional nanostructure, the carbon nanotube has a lower capacity compared with that of graphene. The ideal graphene’s capacity is 837 mA h g−1 in theory [37]. It has unique thermal and electrical conductivity, which can also be applied in different fields, such as energy storage, composite materials, and the catalysis industry [38,39]. Porous graphene was synthesized via the CVD method [40]. Porous graphene had a pore size of 3–8 nm with a lot of edges. Its initial reversible capacity was 1827 mA h g−1 at a 0.1 C current rate. It delivered a capacity of over 800 mA h g−1 at a 1 C current rate after 100 cycles. Surprisingly, it kept about 200 mA h g−1 at a 20 C current rate after 100 cycles. The porous structure and high specific surface gave rise to a rather high capacity for graphene used as anode materials. Graphene nanobuds were prepared on Cu foil using the CVD method [41]. As an electrode in LIBs, the graphene nanobuds exhibited a high specific capacity of 2813 mA h g−1 at a current density of 4 A g−1 in the first discharge–charge process. It maintained a stable capacity of 1600 mA h g−1 at 8 A g−1 for over 330 cycles. It also had a good Coulombic efficiency of 99%. The anode still obtained a capacity of 870 mA h g−1 when tested at a higher current density of 40 A g−1. This outstanding electrochemical performance resulted from the loose space and hollow spaces of few-layered graphene.
Carbon materials with an onion-like structure were used as the anode material for LIBs via the one-step laser-etched method [42]. It delivered a capacity of 436 mA h g−1 at a current density of 0.05 A g−1. Even at a rather high current rate of 10 A g−1, a capacity of 89 mA h g−1 was obtained. Its concentric structure ensured structural integrity for the electrode, facilitating intrinsic cyclic stability during lithium storage. Graphite could be modified using the pulsed laser annealing method, which introduced a defect into graphite in a controlled way [43]. This also increased the density of lithium-ion trapping sites. The capacity was 430 mA h g−1 after the laser treatment. The trapping sites enhanced lithium-ion absorption during the charge–discharge process.
Carbon materials can obtain a high capacity over the theoretical capacity of graphite through structural design and regulation in view of the above work. Herein, carbon nanobeads (CNBs) were prepared via the CVD method using nickel particles as a catalyst and acetylene as a carbon source. CNBs had an onion-like carbon layer structure on its insides. This effectively enhanced the kinetics of ion diffusion, which enhanced the reversible capacity and rate performance of the CNB anode. As anode materials, CNBs maintained the first discharge capacity and charge capacity of 723.7 and 693.6 mA h g−1 at 50 mA g−1, respectively. After four cycles, the CNBs obtained a high reversible capacity of 1273 mA h g−1. When cycled at a current density of 0.2 A g−1, CNBs kept a reversible capacity of 586.2 mA h g−1 after 256 cycles. The onion-like carbon layer structures of CNBs were beneficial to the storage of lithium ions. This work provides a simple strategy to fabricate CNBs with inner onion-like carbon layer structures to enhance the electrochemical performance of the LIBs.

2. Materials and Methods

2.1. Materials

CNBs were prepared via the CVD method at 873 K. Acetylene was used as a carbon source, and nickel particles were used as catalysts. Mesophase pitch-based graphite fibers (MPGFs) were fabricated through the a melt-spinning method. The pitch’s softening point was 553 K. Initially, the pitch was spun into raw fibers. These as-spun fibers were treated at 553 K under an air atmosphere. Subsequently, the fibers were heated at 1273 K under a nitrogen atmosphere. At last, the resulting carbon fibers were ultimately heated at 3073 K, and the atmosphere was argon.

2.2. Method Characterizations and Electrochemical Evaluation

The samples were analyzed via the X-ray diffraction system (XRD; Bruker AXS D8, Karlsruhe, Germany), field-emission scanning electron microscopy (SEM; JSM-7001F, JEOL, Tokyo, Japan), and field-emission transmission electron microscopy (TEM; JEM-2100F, JEOL Tokyo, Japan) for crystalline structure examination morphology observation and microstructure analysis, respectively, along with molecular structure information studies.
Working electrodes were made of active material, acetylene black, and polyvinylidene fluoride (PVDF). Their weight ratio was 8:1:1. CR Li foil was used as the counter electrode and reference electrode in the cells. The separator was a microporous polypropylene film (Celgard 2400). The used electrolyte was composed of 1 M LiPF6 in ethylene carbonate–dimethyl carbonate (1:1 by volume) and 5 vol. % vinylene carbonates. The cells were tested using a battery test system (LAND CT 2001A model, Wuhan Jinnuo Electronics Ltd., Wuhan, China) at room temperature. The electrochemical workstation was applied to study the electrochemical properties of the samples (CHI660E, Shanghai Chenhua Instrument Ltd., Shanghai, China). The specific surface area adsorption instrument was fabricated by Beijing Jingwei Gao Bo Science and Technology Co., Ltd., Beijing, China (BK112T).

3. Results and Discussion

The macroscopic morphology of CNBs is illustrated in Figure 1a. To facilitate a clear observation of their morphology, high-magnification images are presented in Figure 1b–d. A three-dimensional morphology was formed by the random stacking of numerous CNBs, which exhibited an average diameter of approximately 100 nm. Moreover, as shown in Figure 1c,d, CNBs could arrange themselves into quasi-one-dimensional chain structures. In Figure 1e, a selected area scan is displayed to provide further detail. These CNBs were synthesized using the CVD method and possessed a high purity of carbon content with a minimal oxygen content, as demonstrated in Figure 1f. Typically, graphite anode materials for LIBs underwent treatment at temperatures exceeding 2500 K for purification purposes. For convenient comparison and control purposes, mesophase-pitch-based graphite fibers (MPGFs) were prepared, as depicted in Figure 1g. There were a lot of graphite folds in the inner fibers, as depicted in Figure 1h, which is the typical structure of MPGFs. The diameter of MPGFs was approximately 11 μm. Notably, the inner structure of these fibers exhibited abundant graphite folds, as shown in Figure 1h. This represents a typical characteristic feature of MPGFs.
The microstructure of the material primarily determines its physical and electrochemical properties. The TEM images of CNBs are shown in Figure 2a,b, revealing quasi-one-dimensional bead chain structures. The electron diffraction pattern exhibited diffraction rings, as depicted in Figure 2c, indicating a disordered structure resulting from low-temperature preparation processes. In Figure 2d, the carbon layer was distinctly visible and arranged circularly, resembling an onion-like structure. Notably, Figure 2e demonstrates evident graphite stripes within MPGFs that were uniformly distributed using a characteristic commonly observed in internationally recognized MPGFs on the market. Furthermore, highly graphitic structures are indicated by highlighted diffraction spots obtained from the entire area, as illustrated in Figure 2f, owing to the presence of graphite grain ribbons within MPGFs, which contributed to their excellent electrical conductivity.
The structural characteristic of the samples was measured using XRD, as shown in Figure 3. Theoretical interplanar spacing is 0.3354 nm. The interplanar spacing of the CNBs was 0.351 nm. Hence, CNBs had a slightly turbostratic structure. In the case of MPGFs, it was 0.339 nm. MPGFs had a good crystal structure due to graphitization at high temperatures. The above data analysis was consistent with that of TEM.
MPGF were tested at a current density of 50 mA g−1, as depicted in Figure 4a,b. The MPGFs electrode first obtained a charge capacity of 351.25 mA h g−1 (Figure 4a). With the increase in the graphitization degree, the structure of graphite material was more stable, and its conductivity was much better. During the whole process, the highest charge capacity was 357.96 mA h g−1 at the 70th cycle. It was near the theoretical capacity of ideal graphite. However, it maintained a capacity of 328.2 mA h g−1 after 105 cycles.
Rate capability is shown in Figure 4. The current density increased from 50 mA g−1 to 800 mA g−1 and then switched back to the initial current density. MPGFs delivered the reversible capacity of 350.4, 314.2, 244.5, and 173.5 mA h g−1 after 5, 10, 15, and 20 cycles. At last, MPGF electrodes were tested at 50 mA g−1 and kept a reversible capacity of 340.5 mA h g−1. The above data show that its rate capability was good, as shown in Figure 4b.
In the case of CNBs, they were cycled at a current density of 50 mA g−1. The CNBs’ initial discharge and charge capacity were 723.7 and 693.6 mA h g−1, respectively (Figure 4c). After four cycles, the CNBs obtained a high reversible capacity of 1273 mA h g−1. Then, they were cycled at 0.2 A g−1 to examine the cycle performance. At this stage, the initial discharge capacity was 606.2 mA h g−1, and the first charge capacity was 599.7 mA h g−1. In the 256 subsequent cycles, CNBs kept a capacity of 586.2 mA h g−1, which was near the initial charge capacity. During the whole process at the second stage, the whole Coulombic efficiency was very stable with no huge ups and downs. The capacity was more than 572 mA h g−1 during the whole process. It mainly came from the disordered microstructure of CNBs. Though the CNBs had a lower graphitization degree, they obtained wider carbon layer spacing for the transport of lithium ions. MPGFs had two-dimensional carbon layer spacing to store the lithium ions. As CNBs, they had a three-dimensional spherical structure with an inner onion-like carbon layer microstructure. An onion-like carbon layer structure in CNBs was beneficial to the storage of lithium ions. This is also one of the reasons why CNBs can have a high lithium-ion storage performance.
During the rate of measurement, as depicted in Figure 4d, the charge capacity of CNBs was 751.9 mA h g−1 under a current density of 50 mA g−1 after 5 cycles. At the current density of 100, 200, 400, and 800 mA g−1, CNBs obtained a reversible capacity of 688.3, 632.7, 596.3, and 536.2 mA h g−1, respectively. Then, the current density was restored to the previous current density. The corresponding reversible capacity was 617.1, 661.5, 702.4, and 736.5 mA h g−1, respectively. This should be the result of the microstructure of the CNBs, which enhanced the structural integrity and the capacity of the electrode.
In previous research work, carbon nanomaterials were performed in a different capacity as anode material for LIBs. Several anode materials with their performances are listed in Table 1. Porous graphene obtained a capacity of 142 mA h g−1 at 2 A g−1 after 1200 cycles [44]. The porous structure gave rise to a high capacity for graphene. Hard carbon/graphene composites could obtain 623 mA h g−1 at a current density of 0.1 A g−1 after 500 cycles [45]. Graphene with a nano-mesh structure maintained a capacity of 1221 mA h g−1 at a current density of 0.005 A g−1 after 100 cycles [46]. Graphene sheets maintained a capacity of 730 mA h g−1 at a current density of 0.0744 A g−1 after 200 cycles [47]. Carbon nanofibers kept a capacity of 323 mA h g−1 at a current density of 1 A g−1 after 1000 cycles [48]. Their three-dimensional structure and good electrical conductivity were beneficial to the capacity. In this work, CNBs obtained a capacity of 596.2 mA h g−1 at a current density of 0.2 A g−1 after 256 cycles. The CNBs were prepared at low temperatures compared with graphite and had not been graphitized. Hence, CNBs had no ideal graphite structure, graphene sheets, or a graphene nano-mesh structure. They could not have a rather high capacity of over 1200 mA h g−1 (graphene nanomeshes anode). But, the CNBs had a disordered carbon layer structure. This helps to improve the performance of lithium storage.
Carbon atoms usually rearrange during the graphitization process, which usually eliminates some hole structures or defects. Ideal graphite without a pore structure has a theoretical capacity of 372 mAh g−1. But carbon materials usually have an inner-pore structure and defective structure, which can provide a much more specific capacity over 372 mAh g−1 [49,50,51]. CNBs were prepared at a low temperature compared with graphite. Hence, there was a certain pore structure in the inner parts of CNBs. The pore volume–pore distribution diagram is shown in Figure 5. The mean pore size of adsorption was 8.59 nm. This helped to improve the performance of lithium storage.
A typical cyclic voltammetry (CV) measurement of the samples is shown in Figure 6. Figure 6a shows the CV curves of MPGFs. Its sweep rate of the potential was 1 mV/s. In the first cathodic half-cycle, the discharge current became large at the starting potential of 1.0 V, which resulted in the formation of the SEI film. The first anodic peak appeared at 0.68 V. The anodic peak shifted to 0.63 V, and the cathodic peak was located at 0.12 V in the following process. In addition, the anodic current peaks of the following scan presented a slightly enhanced intensity compared with the first scan. In the case of the CNB electrode, the discharge current increased gradually from 1.41 V, which could be attributed to the formation of an SEI film in the first cycle. This resulted in an initial irreversible capacity, as shown in Figure 6c. The corresponding sweep rate of the potential was 10 mV/s. In the first anodic scan, two obvious anodic peaks appeared at 0.88 and 1.26 V. During the following scan, the corresponding peaks overlapped quite well. Judging from the three curves of CNB electrodes, they overlapped quite well, suggesting their good cycling performance.
The electrolyte resistance is represented by Re in the equivalent circuit shown in Figure 6e,f, while Rct denotes the charge transfer resistance, and Rf represents the resistance of the surface film and contact. Additionally, Q is the double-layer constant phase angle element, and Zw is the Warburg impedance in relation to the diffusion of lithium ions into the bulk of the composite electrode. C is the capacitance. The equivalent circuit of MPGFs is as shown in Figure 6e; the Re was 12 Ω, the Rf was 110 Ω, and the Rct was 119 Ω, respectively. The equivalent circuit of CNBs is depicted in Figure 6f. The corresponding Re was 21 Ω, and the Rf was 122 Ω. The CNBs had a lower graphitization degree. The lower heat treatment temperature resulted in relatively weak conductivity compared with that of MPGFs. Hence, the Rct of CNBs was higher than that of MPGFs. This was because CNBs were prepared at a low temperature, which resulted in a lower electrical conductivity compared with that of the commercial graphite anode. CNBs also had a larger specific surface area than that of the MPGFs. Hence, CNBs had a larger impedance than that of the graphite anode.

4. Conclusions

CNBs were prepared with an inner onion-like carbon layer structure using chemical vapor deposition. MPGFs were fabricated via the melt-spinning method. Both kinds of carbon materials were used as anode materials for LIBs. The MPGFs had a reversible capacity of 328.2 mA h g−1 with a steady Coulombic efficiency after 105 cycles at a current rate of 0.05 mA g−1. CNBs were graphitized. They had disordered carbon layers and a certain pore structure. Their mean pore size of adsorption was 8.59 nm, which helped to improve the performance of lithium storage. The CNBs kept a reversible capacity of 586.2 mA h g−1 after 256 cycles with stable Coulombic efficiency at a current rate of 0.2 A g−1. The onion-like carbon layer structures of CNBs gave rise to the storage of lithium ions. CNBs also had a good rate capability compared with MPCFs. In this work, CNBs with an inner onion-like carbon layer structure were prepared as anode materials to enhance the specific capacity of LIBs.

Author Contributions

L.W.: Conceptualization, validation, writing—original draft preparation, drafting the work, writing—review and editing, funding acquisition. T.L., S.Y. and S.J.: investigation, formal analysis. H.W. and J.Y.: project administration, funding acquisition, conceptualization. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Fundamental Research Program of Shanxi Province (No. 20210302123052, 201901D211270), the Key Research and Development (R&D) Projects of Shanxi Province (No. 202102040201003), and the Graduate Student Education Innovation Projects of Shanxi Province (No. 2020SY352, 2020SY355).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Koroneos, C.; Spachos, T.; Moussiopoulos, N. Exergy analysis of renewable energy sources. Renew. Energy 2003, 28, 295–310. [Google Scholar] [CrossRef]
  2. Mahon, H.; O’Connor, D.; Friedrich, D.; Hughes, B. A review of thermal energy storage technologies for seasonal loops. Energy 2022, 239, 122207. [Google Scholar] [CrossRef]
  3. Khezri, R.; Mahmoudi, A.; Aki, H. Optimal planning of solar photovoltaic and battery storage systems for grid-connected residential sector: Review, challenges and new perspectives. Renew. Sustain. Energy Rev. 2022, 153, 111763. [Google Scholar] [CrossRef]
  4. Kant, K.; Biwole, P.H.; Shamseddine, I.; Tlaiji, G.; Pennec, F.; Fardoun, F. Recent advances in thermophysical properties enhancement of phase change materials for thermal energy storage. Sol. Energy Mater. Sol. Cells 2021, 231, 111309. [Google Scholar] [CrossRef]
  5. Service, R.F. Lithium-ion battery development takes Nobel. Science 2019, 366, 292. [Google Scholar] [CrossRef] [PubMed]
  6. Endo, M.; Kim, C.; Nishimura, K.; Fujino, T.; Miyashita, K. Recent development of carbon materials for Li ion batteries. Carbon 2000, 38, 183–197. [Google Scholar] [CrossRef]
  7. Cheng, H.; Shapter, J.G.; Li, Y.; Gao, G. Recent progress of advanced anode materials of lithium-ion batteries. J. Energy Chem. 2021, 57, 451–468. [Google Scholar] [CrossRef]
  8. Azam, M.A.; Safie, N.E.; Ahmad, A.S.; Yuza, N.A.; Zulkifli, N.S.A. Recent advances of silicon, carbon composites and tin oxide as new anode materials for lithium-ion battery: A comprehensive review. J. Energy Storage 2021, 33, 102096. [Google Scholar] [CrossRef]
  9. Liu, K.; Liu, Y.; Lin, D.; Pei, A.; Cui, Y. Materials for lithium-ion battery safety. Sci. Adv. 2018, 4, eaas9820. [Google Scholar] [CrossRef]
  10. Liu, X.-M.; Huang, Z.d.; Oh, S.w.; Zhang, B.; Ma, P.-C.; Yuen, M.M.F.; Kim, J.-K. Carbon nanotube (CNT)-based composites as electrode material for rechargeable Li-ion batteries: A review. Compos. Sci. Technol. 2012, 72, 121–144. [Google Scholar] [CrossRef]
  11. de las Casas, C.; Li, W. A review of application of carbon nanotubes for lithium ion battery anode material. J. Power Sources 2012, 208, 74–85. [Google Scholar] [CrossRef]
  12. Zhang, T.; Han, S.; Guo, W.; Hou, F.; Liu, J.; Yan, X.; Chen, S.; Liang, J. Continuous carbon nanotube composite fibers for flexible aqueous lithium-ion batteries. Sustain. Mater. Technol. 2019, 20, e00096. [Google Scholar] [CrossRef]
  13. Chen, K.; Wang, Q.; Niu, Z.; Chen, J. Graphene-based materials for flexible energy storage devices. J. Energy Chem. 2018, 27, 12–24. [Google Scholar] [CrossRef]
  14. Dong, Y.; Wu, Z.-S.; Ren, W.; Cheng, H.-M.; Bao, X. Graphene: A promising 2D material for electrochemical energy storage. Sci. Bull. 2017, 62, 724–740. [Google Scholar] [CrossRef] [PubMed]
  15. Cao, L.; Wang, C.; Huang, Y. Structure optimization of graphene aerogel-based composites and applications in batteries and supercapacitors. Chem. Eng. J. 2023, 454, 140094. [Google Scholar] [CrossRef]
  16. Wen, Y.; Liu, H.; Jiang, X. Preparation of graphene by exfoliation and its application in lithium-ion batteries. J. Alloys Compd. 2023, 961, 170885. [Google Scholar] [CrossRef]
  17. Zhang, H.; Yang, Y.; Ren, D.; Wang, L.; He, X. Graphite as anode materials: Fundamental mechanism, recent progress and advances. Energy Storage Mater. 2021, 36, 147–170. [Google Scholar] [CrossRef]
  18. Natarajan, S.; Divya, M.L.; Aravindan, V. Should we recycle the graphite from spent lithium-ion batteries? The untold story of graphite with the importance of recycling. J. Energy Chem. 2022, 71, 351–369. [Google Scholar] [CrossRef]
  19. Niu, B.; Xiao, J.; Xu, Z. Advances and challenges in anode graphite recycling from spent lithium-ion batteries. J. Hazard. Mater. 2022, 439, 129678. [Google Scholar] [CrossRef]
  20. Zhu, X.; Chen, Y.; Xiao, J.; Xu, F.; Su, F.; Yao, Z.; Zhang, Z.; Tang, L.; Zhong, Q. The strategy for comprehensive recovery and utilization of the graphite anode materials from the end-of-life lithium-ion batteries: Urgent status and policies. J. Energy Storage 2023, 68, 107798. [Google Scholar] [CrossRef]
  21. Yao, N.; Liu, F.; Zou, Y.; Wang, H.; Zhang, M.; Tang, X.; Wang, Z.; Bai, M.; Liu, T.; Zhao, W.; et al. Resuscitation of spent graphite anodes towards layer-stacked, mechanical-flexible, fast-charging electrodes. Energy Storage Mater. 2023, 55, 417–425. [Google Scholar] [CrossRef]
  22. Sun, L.; Liu, Y.; Shao, R.; Wu, J.; Jiang, R.; Jin, Z. Recent progress and future perspective on practical silicon anode-based lithium ion batteries. Energy Storage Mater. 2022, 46, 482–502. [Google Scholar] [CrossRef]
  23. Li, Y.; Arnold, S.; Husmann, S.; Presser, V. Recycling and second life of MXene electrodes for lithium-ion batteries and sodium-ion batteries. J. Energy Storage 2023, 60, 106625. [Google Scholar] [CrossRef]
  24. Wang, Q.; Wu, X.; You, H.; Min, H.; Xu, X.; Hao, J.; Liu, X.; Yang, H. Template-directed Prussian blue nanocubes supported on Ni foam as the binder-free anode of lithium-ion batteries. Appl. Surf. Sci. 2022, 571, 151194. [Google Scholar] [CrossRef]
  25. Wang, C.-Y.; Liu, T.; Yang, X.-G.; Ge, S.; Stanley, N.V.; Rountree, E.S.; Leng, Y.; McCarthy, B.D. Fast charging of energy-dense lithium-ion batteries. Nature 2022, 611, 485–490. [Google Scholar] [CrossRef] [PubMed]
  26. Nuhu, B.A.; Bamisile, O.; Adun, H.; Abu, U.O.; Cai, D. Effects of transition metals for silicon-based lithium-ion battery anodes: A comparative study in electrochemical applications. J. Alloys Compd. 2023, 933, 167737. [Google Scholar] [CrossRef]
  27. Ping, W.; Yang, C.; Bao, Y.; Wang, C.; Xie, H.; Hitz, E.; Cheng, J.; Li, T.; Hu, L. A silicon anode for garnet-based all-solid-state batteries: Interfaces and nanomechanics. Energy Storage Mater. 2019, 21, 246–252. [Google Scholar] [CrossRef]
  28. Gou, L.; Jing, W.; Li, Y.; Wang, M.; Hu, S.; Wang, H.; He, Y.-B. Lattice-Coupled Si/MXene Confined by Hard Carbon for Fast Sodium-Ion Conduction. ACS Appl. Energy Mater. 2021, 4, 7268–7277. [Google Scholar] [CrossRef]
  29. Zhu, J.; Gladden, C.; Liu, N.; Cui, Y.; Zhang, X. Nanoporous silicon networks as anodes for lithium ion batteries. Phys. Chem. Chem. Phys. PCCP 2013, 15, 440–443. [Google Scholar] [CrossRef]
  30. Li, Y.; Yan, K.; Lee, H.-W.; Lu, Z.; Liu, N.; Cui, Y. Growth of conformal graphene cages on micrometre-sized silicon particles as stable battery anodes. Nat. Energy 2016, 1, 15029. [Google Scholar] [CrossRef]
  31. Song, T.; Jeon, Y.; Samal, M.; Han, H.; Park, H.; Ha, J.; Yi, D.K.; Choi, J.-M.; Chang, H.; Choi, Y.-M.; et al. A Ge inverse opal with porous walls as an anode for lithium ion batteries. Energy Environ. Sci. 2012, 5, 9028–9033. [Google Scholar] [CrossRef]
  32. Bogart, T.D.; Chockla, A.M.; Korgel, B.A. High capacity lithium ion battery anodes of silicon and germanium. Curr. Opin. Chem. Eng. 2013, 2, 286–293. [Google Scholar] [CrossRef]
  33. Liang, S.; Cheng, Y.-J.; Zhu, J.; Xia, Y.; Müller-Buschbaum, P. A Chronicle Review of Nonsilicon (Sn, Sb, Ge)-Based Lithium/Sodium-Ion Battery Alloying Anodes. Small Methods 2020, 4, 2000218. [Google Scholar] [CrossRef]
  34. Fugattini, S.; Gulzar, U.; Andreoli, A.; Carbone, L.; Boschetti, M.; Bernardoni, P.; Gjestila, M.; Mangherini, G.; Camattari, R.; Li, T.; et al. Binder-free nanostructured germanium anode for high resilience lithium-ion battery. Electrochim. Acta 2022, 411, 139832. [Google Scholar] [CrossRef]
  35. Wu, S.; Wu, H.; Zou, M.; Shi, X.; Yuan, Y.; Bai, W.; Cao, A. Short-range ordered graphitized-carbon nanotubes with large cavity as high-performance lithium-ion battery anodes. Carbon 2020, 158, 642–650. [Google Scholar] [CrossRef]
  36. Zhu, G.; Tang, C.; Jiang, M.; Du, A.; Zhang, H.; Yang, J. Regulating the interfacial behavior of carbon nanotubes for fast lithium storage. Electrochim. Acta 2021, 388, 138591. [Google Scholar] [CrossRef]
  37. Qiu, T.-C.; Shao, Z.-G.; Wang, C.-L.; Yang, L. QPHT graphene as a high-performance lithium ion battery anode materials with low diffusion barrier and high capacity. Phys. Lett. A 2022, 456, 128549. [Google Scholar] [CrossRef]
  38. Wang, B.; Ruan, T.; Chen, Y.; Jin, F.; Peng, L.; Zhou, Y.; Wang, D.; Dou, S. Graphene-based composites for electrochemical energy storage. Energy Storage Mater. 2020, 24, 22–51. [Google Scholar] [CrossRef]
  39. Du, J.; Zhang, Y.; Deng, S.; Xu, N.; Xiao, Z.; She, J.; Wu, Z.; Cheng, H. Correlation between topographic structures and local field emission characteristics of graphene-sheet films. Carbon 2013, 61, 507–514. [Google Scholar] [CrossRef]
  40. Fan, Z.; Yan, J.; Ning, G.; Wei, T.; Zhi, L.; Wei, F. Porous graphene networks as high performance anode materials for lithium ion batteries. Carbon 2013, 60, 558–561. [Google Scholar] [CrossRef]
  41. Zeferino González, I.; Chiu, H.-C.; Gauvin, R.; Demopoulos, G.P.; Miki-Yoshida, M.; Valenzuela-Muñiz, A.M.; Verde-Gómez, Y. Graphene nanobuds as a novel anode design paradigm with superior Li-ion storage capacity and rate capability. Carbon 2022, 199, 486–496. [Google Scholar] [CrossRef]
  42. Meng, C.; Yuan, M.; Cao, B.; Lin, X.; Zhang, J.; Li, A.; Chen, X.; Jia, M.; Song, H. Laser-modified graphitic onion-like carbon as anode for lithium/potassium-ion batteries. Carbon 2022, 192, 347–355. [Google Scholar] [CrossRef]
  43. Khosla, N.; Narayan, J.; Narayan, R.; Sun, X.-G.; Paranthaman, M.P. Microstructure and defect engineering of graphite anodes by pulsed laser annealing for enhanced performance of lithium-ion batteries. Carbon 2023, 205, 214–225. [Google Scholar] [CrossRef]
  44. Zhou, Y.-Q.; Dong, X.-L.; Li, W.-C.; Hao, G.-P.; Yan, D.; Lu, A.-H. Millimeter-sized few-layer graphene sheets with aligned channels for fast lithium-ion charging kinetics. J. Energy Chem. 2021, 55, 62–69. [Google Scholar] [CrossRef]
  45. Li, R.; Huang, J.; Ren, J.; Cao, L.; Li, J.; Li, W.; Lu, G.; Yu, A. A sandwich-like porous hard carbon/graphene hybrid derived from rapeseed shuck for high-performance lithium-ion batteries. J. Alloys Compd. 2020, 818, 152849. [Google Scholar] [CrossRef]
  46. Ning, G.; Xu, C.; Cao, Y.; Zhu, X.; Jiang, Z.; Fan, Z.; Qian, W.; Wei, F.; Gao, J. Chemical vapor deposition derived flexible graphene paper and its application as high performance anodes for lithium rechargeable batteries. J. Mater. Chem. A 2013, 1, 408. [Google Scholar] [CrossRef]
  47. Chen, S.; Bao, P.; Xiao, L.; Wang, G. Large-scale and low cost synthesis of graphene as high capacity anode materials for lithium-ion batteries. Carbon 2013, 64, 158–169. [Google Scholar] [CrossRef]
  48. Liu, C.; Xiao, N.; Wang, Y.; Li, H.; Wang, G.; Dong, Q.; Bai, J.; Xiao, J.; Qiu, J. Carbon clusters decorated hard carbon nanofibers as high-rate anode material for lithium-ion batteries. Fuel Process. Technol. 2018, 180, 173–179. [Google Scholar] [CrossRef]
  49. Ma, C.; Zhao, Y.; Li, J.; Song, Y.; Shi, J.; Guo, Q.; Liu, L. Synthesis and electrochemical properties of artificial graphite as an anode for high-performance lithium-ion batteries. Carbon 2013, 64, 553–556. [Google Scholar] [CrossRef]
  50. Ma, C.; Zhao, Y.; Li, J.; Song, Y.; Shi, J.; Guo, Q.; Liu, L. The electrochemical performance of pitch coke anodes containing hollow carbon nanostructures and nickel nanoparticles for high-power lithium ion batteries. Electrochim. Acta 2013, 112, 394–402. [Google Scholar] [CrossRef]
  51. Li, L.; Zhang, D.; Deng, J.; Gou, Y.; Fang, J.; Cui, H.; Zhao, Y.; Cao, M. Carbon-based materials for fast charging lithium-ion batteries. Carbon 2021, 183, 721–734. [Google Scholar] [CrossRef]
Figure 1. SEM images: (ae) The morphology of CNBs, (f) The selected area scanning test of (e), (g) The cross-section morphology of MPGFs, (h) The graphite fold of MPGFs.
Figure 1. SEM images: (ae) The morphology of CNBs, (f) The selected area scanning test of (e), (g) The cross-section morphology of MPGFs, (h) The graphite fold of MPGFs.
Coatings 13 01671 g001
Figure 2. TEM image of CNBs (a,b,d) and MPGFs (e), Electron diffraction spots of CNBs (c) and MPGFs (f).
Figure 2. TEM image of CNBs (a,b,d) and MPGFs (e), Electron diffraction spots of CNBs (c) and MPGFs (f).
Coatings 13 01671 g002
Figure 3. XRD pattern of CNBs (a) and MPGFs (b).
Figure 3. XRD pattern of CNBs (a) and MPGFs (b).
Coatings 13 01671 g003
Figure 4. (a) Cycle performance of MPGFs, (b) Rate capability of MPGFs, (c) Cycle performance of CNBs, and (d) Rate capability of CNBs.
Figure 4. (a) Cycle performance of MPGFs, (b) Rate capability of MPGFs, (c) Cycle performance of CNBs, and (d) Rate capability of CNBs.
Coatings 13 01671 g004
Figure 5. Volume–pore distribution diagram.
Figure 5. Volume–pore distribution diagram.
Coatings 13 01671 g005
Figure 6. (a) Cyclic voltammetry curves, and (b) Nyquist plots of MPGFs, (c) Cyclic voltammetry curves, and (d) Nyquist plots of CNBs, (e) Equivalent circuit of MPGFs, and (f) Equivalent circuit of CNBs.
Figure 6. (a) Cyclic voltammetry curves, and (b) Nyquist plots of MPGFs, (c) Cyclic voltammetry curves, and (d) Nyquist plots of CNBs, (e) Equivalent circuit of MPGFs, and (f) Equivalent circuit of CNBs.
Coatings 13 01671 g006
Table 1. Application of carbon materials in LIBs.
Table 1. Application of carbon materials in LIBs.
NumberAnode MaterialCurrent Density (A/g)Cycle NumberCapacity (mAh g−1)Reference
1Porous graphene21200142[44]
2Graphitized CNTs0.1150527[35]
3Hard-carbon/graphene0.1500623[45]
4graphene nanomeshes0.051001221[46]
5Graphene sheets0.0744200730[47]
6Graphitic onion-like carbon1300312[42]
7Carbon nanofibers11000323[48]
8CNBs0.2256596.2This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, L.; Liu, T.; Ji, S.; Yang, S.; Wang, H.; Yang, J. Electrochemical Properties of Carbon Nanobeads and Mesophase-Pitch-Based Graphite Fibers as Anodes for Rechargeable Lithium-Ion Batteries. Coatings 2023, 13, 1671. https://doi.org/10.3390/coatings13101671

AMA Style

Wang L, Liu T, Ji S, Yang S, Wang H, Yang J. Electrochemical Properties of Carbon Nanobeads and Mesophase-Pitch-Based Graphite Fibers as Anodes for Rechargeable Lithium-Ion Batteries. Coatings. 2023; 13(10):1671. https://doi.org/10.3390/coatings13101671

Chicago/Turabian Style

Wang, Liyong, Tiantian Liu, Shengsheng Ji, Shiwen Yang, Huiqi Wang, and Jinhua Yang. 2023. "Electrochemical Properties of Carbon Nanobeads and Mesophase-Pitch-Based Graphite Fibers as Anodes for Rechargeable Lithium-Ion Batteries" Coatings 13, no. 10: 1671. https://doi.org/10.3390/coatings13101671

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