Next Article in Journal
Manufacturing a TiO2-Based Semiconductor Film with Nanofluid Pool Boiling and Sintering Processes toward Solar-Cell Applications
Next Article in Special Issue
Optically Controlled Terahertz Dynamic Beam Splitter with Adjustable Split Ratio
Previous Article in Journal
Changes in Number and Antibacterial Activity of Silver Nanoparticles on the Surface of Suture Materials during Cyclic Freezing
Previous Article in Special Issue
Research Progress on Sound Absorption of Electrospun Fibrous Composite Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Silkworm Protein-Derived Nitrogen-Doped Carbon-Coated Li[Ni0.8Co0.15Al0.05]O2 for Lithium-Ion Batteries

1
Department of Chemical Engineering (Integrated Engineering), College of Engineering, Kyung Hee University, 1732 Deogyeong-daero, Giheung, Yongin 17104, Gyeonggi, Korea
2
Center for the SMART Energy Platform, College of Engineering, Kyung Hee University, 1732 Deogyeong-daero, Giheung, Yongin 17104, Gyeonggi, Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2022, 12(7), 1166; https://doi.org/10.3390/nano12071166
Submission received: 11 March 2022 / Revised: 25 March 2022 / Accepted: 28 March 2022 / Published: 31 March 2022
(This article belongs to the Special Issue Nanotechnologies and Nanomaterials: Selected Papers from CCMR)

Abstract

:
Li[Ni0.8Co0.15Al0.05]O2 (NCA) is a cathode material for lithium-ion batteries and has high power density and capacity. However, this material has disadvantages such as structural instability and short lifespan. To address these issues, herein, we explore the impact of N-doped carbon wrapping on NCA. Sericin, an easily obtained carbon- and nitrogen-rich component of silk cocoons, is utilized as the precursor material. The electrochemical performance evaluation of N-doped carbon-coated NCA shows that the capacity retention of 0.3 NC@NCA at 1C current density is 69.83% after 200 cycles, which is about 19% higher than the 50.65% capacity retention of bare NCA. The results reveal that the sericin-resultant N-doped carbon surface wrapping improves the cycling stability of NC@NCA.

1. Introduction

With the expansion of various devices, there has been a growing demand for various energy storage devices that may be utilized as energy sources [1,2]. Lithium-ion batteries (LIBs) have become the most frequently utilized secondary batteries among energy storage systems due to their attractive properties such as good energy density, operating potential, and extended cycle life [3,4]. The cathode is a crucial part of a battery that influences energy density, power density, longevity, and potential of the battery; it also accounts for 40% of the overall battery price [5,6]. The cathode materials of LIBs including LiCoO2 (LCO), LiMn2O4 (LMO), and LiFePO4 (LFP) are those with layered, spinel, and olivine structures, respectively [7,8,9]. Recently, layer structured, high nickel content cathodes, such as Li[Ni0.8Co0.15Al0.05]O2 (NCA) and Li[Ni0.8Co0.1Mn0.1]O2 (NCM811), have obtained widespread consideration because of their high capacity, relatively inexpensive price, and eco-friendly nature [10,11]. Among them, NCA has been popularized in the battery industry for its low price, high capacity, and power density [11,12]. However, NCA has many problems including poor lifespan, and low thermal and structural stability. Additionally, the degradation of NCA caused by side reactions with the electrolyte is also a major drawback [13,14].
There have been various attempts to address the drawbacks of NCA. Among these attempts, the active material surface wrapping can form a barrier to effectively protect the active material from an unstable structure by suppressing the surface degradation of the material and side reactions with the electrolyte; this improves the structural stability of the material [15,16,17,18]. Depending on the coating material used, the surface wrapping approach can also improve the overall conductivity of the electrode material [19,20].
Carbonaceous materials with wide availability, low cost, high conductivity, and diverse structure have gained significant research interest. Among them, N-doped carbon is an effective material for surface modification of various active materials because it efficiently suppresses surface degradation, prevents by-products from unnecessary side reactions with electrolytes, and improves the electronic conductivity of the host material [21,22,23]. Recently, Feng et al. reported on polyacrylonitrile (PAN)-induced conductive carbon coating on NCA. The PAN-induced carbon coated NCA, with a 5 nm thick coating layer, offered faster electron movement and hence provided an enhanced cycling performance [24]. Ion-conductive polymer Nafion-coating on NCA was investigated by Yigitalp et al. to obtain superior electrochemical performance. They concluded that Nafion-coating could suppress the passive layer formation and provide improved cycling stability [25]. Park et al. explored the conformal graphene coating on NCA by a scalable Pickering emulsion method. The conformal graphene coating reduces the surface degradation of NCA and hence delivered an excellent life cycle [26]. In this study, we utilized sericin to coat N-doped carbon on active materials; it is a biomaterial that is a component of silk cocoons. Sericin, a gummy component that surrounds fibroin and maintains the cohesiveness of the silk cocoon, accounts for about 20–30% of the total mass of a cocoon [27,28]. Sericin is a hydrophilic protein composed of various amino acids such as glycine, alanine, and arginine. It contains a nitrogen content of approximately 17%, rendering it appropriate as a surface modification agent for electrode materials [28,29].
Herein, we propose a simple solid-state technique to prepare N-doped carbon wrapping on the NCA surface using sericin powder extracted by the degumming process. To optimize the suitable weight percentages of the N-doped carbon for effective surface wrapping, different weight percentages of sericin were used. The as-prepared N-doped carbon-coated NCA material was then physically and electrochemically characterized to identify the effect of N-doped carbon surface wrapping to overcome the disadvantages associated with NCA.

2. Materials and Methods

2.1. Sericin Synthesis

Sericin was extracted through the degumming of the silkworm cocoon. Initially, 0.02 M Na2CO3 was dissolved in 100 mL distilled water and allowed to boil on a hot plate at 100 °C. Then, the cocoon was added to it and boiled for 2–3 h to extract the sericin content from the cocoon [30,31]. After extraction of the sericin content, the pale yellowish solution was filtered to eliminate the remaining silk fiber after the sericin was dissolved. Then, the solution was dried in a convection oven at 80 °C for 12 h, and the sericin powder was obtained after drying.

2.2. Synthesis of Nitrogen-Doped Carbon Coated NCA (NC@NCA)

Using a simple solid-state method, NC@NCAs were synthesized. Approximately 2 g of commercial NCA (EcoPro BM, Ltd., Cheongju, Korea) was ground with sericin powder for approximately 1 h for thorough mixing. Subsequently, the mixed powder was heated at 500 °C for 3h in the presence of argon gas to get NC@NCA. For comparison, different amounts of sericin were used, such as 0.2, 0.3, and 0.4 weight percent, to prepare NC@NCA and designated as 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA, respectively.

2.3. Physical Measurements

The crystal structure and phase of the as-synthesized samples were analyzed with an X-ray diffractometer (XRD, Bruker D8 Advance, Billerica, MA, USA, Cu-Kα radiation, λ = 1.5406 Å) in the 2θ range of 10°–70°. X-ray photoelectron spectroscopy (XPS, K-Alpha; Thermo Fisher, Waltham, MA, USA) was conducted to determine the constituent elements on the surface of the material. We also carried out high-resolution field-emission scanning electron microscopy (HR-FE-SEM, MERLIN-LEO SUPRA 55, Carl Zeiss, Jena, Germany) embedded with energy-dispersive X-ray spectroscopy (EDX, Ox-ford Instruments, Abingdon, UK) to determine the morphological characteristics and to conduct elemental analysis and elemental mapping.

2.4. Electrochemical Measurements

The obtained active materials were combined with Denka black as a conductive agent and polyvinylidene fluoride (PVDF) as a binder at a weight ratio of 80:10:10 to make the working electrode. Then, as a dispersion agent, an appropriate amount of N-methyl-2-pyrrolidone (NMP) was added to make a slurry. The prepared slurry was laminated on aluminum foil followed by room temperature drying for a day. The electrode was then dried in a convection oven at 120 °C for 5 h, lab-pressed, and punched into 14 mm diameter disks using a punching tool. Finally, the electrode was dried for 5 h at 100 °C in a vacuum oven. The as-produced electrode’s electrochemical performance was assessed in 2032-type coin cells, with lithium foil as the reference electrode, Celgard® 2320 membrane as the separator, and 1 M LiPF6 in ethylene carbonate (EC) and diethyl carbonate (DEC) (1:1 vol. percent) as the electrolyte. The average mass loading of the cathode was 6 mg cm−2. The CR2032-type half-cells used to evaluate electrochemical performance were made in a glove box filled with Ar gas and aged for several hours. The electrochemical performance of the prepared cells was evaluated within a voltage range of 3.0–4.3 V using a cycler (ETH cycler, Hwaseong, Korea). Electrochemical impedance spectroscopy (EIS) was also conducted; the NCA cells completed for 200 cycles at a current density of 1 C were used and measured by applying an amplitude of 10 mV and a frequency range of 0.5 mHz–100 kHz. EIS measurements were carried out using an electrochemical workstation (Iviumstat, Ivium Technologies, Eindhoven, The Netherlands).

3. Results and Discussion

3.1. Physico-Chemical Characterizations

The XRD patterns of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA are shown in Figure 1, and their equivalent lattice parameters are exhibited in Table 1; these diffraction patterns and lattice parameters are used to identify differences in crystal structure and crystallinity. In Figure 1, the XRD patterns of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA exhibited no differences, and all the peaks are well matched to JCPDS NO. 87-1562, which is indexed to α-NaFeO2 structure with R3m space group [32]. Furthermore, it is confirmed that the change of the lattice parameter did not appear significantly before and after coating, as indicated in Table 1. From these points, the N-doped carbon derived from sericin have no effect on the crystal structure of the bare NCA [33]. However, in terms of c/a values, there is little difference between the materials. The higher the ratio of c/a values, the higher the crystallinity and stability of NCA having a layered structure [34,35]. In this regard, it may be evident that 0.3 NC@NCA and 0.4 NC@NCA have higher crystallinity and stability than bare NCA. However, 0.2 NC@NCA has a lower c/a value than that of bare NCA, which means lower crystallinity and stability.
XPS was conducted to study the chemical valence states and elemental composition of 0.3 NC@NCA. As shown in Figure 2a, the 0.3 NC@NCA XPS spectra exhibited the peaks of Ni, Co, Al, O, C, and N. The Ni 2p, Co 2p, Al 2p, and O 1s peaks fit well with the previous reports of the NCA [36,37,38]. Moreover, the C 1s and N 1s peaks are also observed, indicated by the N-doped carbon. Figure 2f displays the C 1s spectrum, which is fitted with peaks of C–C, C–O–C, and O–C=O bonds at binding energies of 284.0, 286.2, and 288.3 eV [39,40]. In Figure 2g, the N 1s spectra are fitted into two peaks for pyridinic N and pyrrolic N at binding energies of 397.8 and 399.8 eV [21,41]. The pyridinic N and pyrrolic N observed in 0.3 NC@NCA generate numerous active sites, which facilitate the rapid diffusion of lithium ions [42,43]. Consequently, the N-doped carbon layer contains pyridinic N and pyrrolic N, leading to improved electrochemical performance without disturbing the chemical states of NCA.
Figure 3a–i show HR-FE-SEM images of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA. The surface of bare NCA, investigated by the HR-FE-SEM image, is observed as a rough surface. After the addition of sericin as the coating agent, the original shape of the materials is preserved for NC@NCA. Additionally, N-doped carbon on the surface of NC@NCA is clearly evident in the high-resolution images. In particular, in 0.3 NC@NCA, N-doped carbon particles are visible on the surface, and in 0.4 NC@NCA, more particles are noticeable on NCA. This indicates that sericin changes the surface of the NCA but does not affect the original spherical shape of the NCA.
Figure 4a–d show the EDS elemental mapping and spectra mapping to further investigate the existence of various elements of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA. It is confirmed that Ni, Co, and Al are evenly and densely spread on the surface of them through each elemental mapping image. For 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA, it is observed that nitrogen and carbon are also evenly distributed; an increase in the nitrogen content is noticeable as the weight percent of sericin increased, and the weight and atomic ratio of nitrogen for the materials are given in Table 2. The EDS mapping results confirmed that NCA is coated with N-doped carbon; as the weight percent of sericin increased, the thickness of the wrapping layer also grew.

3.2. Electrochemical Characterizations

The galvanostatic charge/discharge behavior analysis of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA was conducted using a cycler, as shown in Figure 5a–d. The evaluation process was conducted within a voltage range of 3.0–4.3 V with a current density of 0.2 C. Figure 5a exhibits that the bare NCA delivers the highest initial specific capacity among its counterparts. It is considered that the N-doped carbon wrapping would have impeded the ability of lithium ions to penetrate the NCA structure. For 50 cycles, it is verified that the decrease in specific capacity of the bare NCA is greater than its counterparts. Therefore, it is considered that the N-doped carbon wrapping minimizes side reactions with the electrolyte and polarization.
Figure 6a–d show the differential capacity (dQ/dV) vs. voltage curve of the 2nd and 50th cycles at a current density of 0.2 C for bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA. In the oxidation curve for the second cycle, the main sharp peak results from the formation of solid electrolyte interphase (SEI) on the NCA surface [44]. Furthermore, three pairs of clear peaks are visible, which are due to the multiple phase change from hexagonal (H1) to monoclinic (M), monoclinic (M) to hexagonal (H2) and hexagonal (H2) to hexagonal (H3) during cycling [45,46]. In the second cycle, the main oxidation peaks are 3.6431, 3.6679, 3.6677, and 3.6635 V, whereas, in the 50th cycle, the main oxidation peaks are shifted to 3.7595, 3.7205, 3.697, and 3.7118 V, respectively. For the bare NCA, the main oxidation peak shifted by 0.1164 V, and for the NC@NCA, the shifts in the main oxidation peaks are 0.0526, 0.0293, and 0.0483 V. Additionally, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA showed smaller shifts in oxidation peak compared with bare NCA; in particular, 0.3 NC@NCA experienced the smallest shift. The smaller shift in the oxidation peak means that the degree of reversibility during the cycling process improved [47]. That is, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA had better cycling reversibility than bare NCA; 0.3 NC@NCA, in particular, exhibited the best cycling reversibility.
The improved cycling reversibility was verified in Figure 6e–h; these figures exhibited the cycling performance of bare and N-doped carbon-coated NCA within a 3.0–4.3 V voltage range with a current density of 0.2 C. In the first cycle, bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA exhibited specific discharge capacities of 183.95, 176.02, 175.31, and 171.78 mAh g−1, respectively. However, after the 50 cycles, the specific discharge capacity of bare NCA is 160.27 mAh g−1, whereas those of 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA are 157.76, 161.63, and 155.10 mAh g−1, respectively. When investigated by capacity retention, bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA are 87.13, 89.62, 92.20, and 90.29%, respectively; 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA exhibited higher capacity retention compared with bare NCA. These outcomes imply that the N-doped carbon wrapping minimizes side reactions between the NCA and the electrolyte, improving the structural stability and cycling reversibility of NCA.
Figure 7a shows the specific discharge capacity when 200 cycles of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA were conducted at a high current density of 1 C. In the initial cycle, the specific discharge capacity of bare NCA is higher than 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA; however, after 200 cycles, the capacity retentions of 0.3 NC@NCA and 0.4 NC@NCA are 69.83% and 62.01%, respectively. This is higher than 0.2 NC@NCA with a capacity retention of 40.86% and a bare NCA of 50.65%. This low cycling stability in 0.2 wt.% N-C@NCA is attributed to the lower amount of sericin utilized. Therefore, it is considered that using a low amount of the coating agent could not form a stable coating layer to provide good cycling at a high current density, behaving instead as a minor variation. In addition, it is also considered that the low crystallinity and stability of 0.2 NC@NCA also contributed to the low-capacity retention at 1.0 C. For 0.4 NC@NCA, the capacity retention is higher than that of bare NCA; however, compared with 0.3 NC@NCA, the thicker coating layer impeded the movement of lithium ions and resulted in poor capacity retention.
Figure 7b shows the rate capability of bare NCA and 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA. The process was conducted at various current densities from 0.2 C to 2.0 C. Bare NCA at a current density of 0.2 C exhibited the highest capacity, while at 0.5 C, it exhibited almost the same capacity as 0.3 NC@NCA. At current densities of 1.0 and 2.0 C, 0.4 NC@NCA exhibited a lower capacity than 0.3 NC@NCA. These results indicate that the 0.3 weight percent N-doped carbon wrapping is an effective coating layer capable of sufficiently protecting NCA during cycling at a high current rate.
To further analyze the effect of N-doped carbon on electrochemical performance, EIS was carried out for bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA after 200 cycles at a 1 C current rate. Figure 7c shows that the semicircle in the high-frequency range, displayed in the Nyquist plot, is related to the charge transfer resistance (Rct), and the straight line in the low-frequency region is related to the Warburg resistance [48]. There is a large difference in charge transfer resistance, when comparing the bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA. In numerical terms, the charge transfers resistances of the bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA are 2091, 3049, 411, and 604 Ω, respectively. The charge transfer resistance is lowest for 0.3 NC@NCA, implying that this material showed the greatest improvement in ion and electron movement among them. This means that if 0.3 weight percent of sericin is used as the coating agent, the coating layer may serve as a protective layer for NCA and improve the movement of ions and electrons [49]. A table of comparison of previously reported similar studies with current work is presented in Table 3.

4. Conclusions

The N-doped carbon derived from sericin was successfully prepared using the solid-state method, and it was used to provide a coating layer for the NCA cathode material of LIBs. The influence of surface modification on the electrochemical characteristics of the NCA was characterized, and it was confirmed that this coating material is effective in improving electrochemical performance, particularly cycling stability. Specifically, 0.3 NC@NCA demonstrated the greatest improvements in cycling stability at current rates of 0.2 and 1.0 C as well as in rate capability. The measurement of charge transfer resistance using EIS confirmed that 0.3 NC@NCA showed the lowest charge transfer resistance. Based on these results, when 0.3 weight percent of sericin is used for the coating on NCA, it effectively suppresses side reactions with the electrolyte and surface degradation. Therefore, it could improve the structural stability of NCA, resulting in enhanced electrochemical performance, particularly in terms of cycling stability and rate capability.

Author Contributions

Conceptualization, methodology, validation, investigation, writing—original draft preparation, G.S.S. and N.S.; validation, formal analysis, P.S.; formal analysis, J.W.P., C.W.H., M.N. and H.K.K.; writing—review and editing, supervision, project administration, funding acquisition, C.W.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Next Generation Engineering Researcher Program of National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (No. 2017H1D8A2031138), and by the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (No. 2020K1A3A1A48111073). It was also supported by the Korea Institute for Advancement of Technology (KIAT) funded by the Ministry of Trade, Industry & Energy (MOTIE) of Korea (No. P0017363).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. AL Shaqsi, A.Z.; Sopian, K.; Al-Hinai, A. Review of energy storage services, applications, limitations, and benefits. Energy Rep. 2020, 6, 288–306. [Google Scholar] [CrossRef]
  2. Manthiram, A. A reflection on lithium-ion battery cathode chemistry. Nat. Commun. 2020, 11, 1550. [Google Scholar] [CrossRef] [PubMed]
  3. Nitta, N.; Wu, F.; Lee, J.T.; Yushin, G. Li-ion battery materials: Present and future. Mater. Today 2015, 18, 252–264. [Google Scholar] [CrossRef]
  4. Pender, J.P.; Jha, G.; Youn, D.H.; Ziegler, J.M.; Andoni, I.; Choi, E.J.; Heller, A.; Dunn, B.S.; Weiss, P.S.; Penner, R.M.; et al. Electrode Degradation in Lithium-Ion Batteries. ACS Nano 2020, 14, 1243–1295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Mohamed, N.; Allam, N.K. Recent advances in the design of cathode materials for Li-ion batteries. RSC Adv. 2020, 10, 21662–21685. [Google Scholar] [CrossRef]
  6. Xi, Z.; Wang, Z.; Peng, W.; Guo, H.; Wang, J. Effect of copper and iron substitution on the structures and electrochemical properties of LiNi0.8Co0.15Al0.05O2 cathode materials. Energy Sci. Eng. 2020, 8, 1868–1879. [Google Scholar] [CrossRef] [Green Version]
  7. Manthiram, A.; Kim, J. Low Temperature Synthesis of Insertion Oxides for Lithium Batteries. Chem. Mater. 1998, 10, 2895–2909. [Google Scholar] [CrossRef]
  8. Chen, Z.; Zhang, W.; Yang, Z. A review on cathode materials for advanced lithium ion batteries: Microstructure designs and performance regulations. Nanotechnology 2020, 31, 012001. [Google Scholar] [CrossRef]
  9. Guan, P.; Zhou, L.; Yu, Z.; Sun, Y.; Liu, Y.; Wu, F.; Jiang, Y.; Chu, D. Recent progress of surface coating on cathode materials for high-performance lithium-ion batteries. J. Energy Chem. 2020, 43, 220–235. [Google Scholar] [CrossRef] [Green Version]
  10. Do, S.J.; Santhoshkumar, P.; Kang, S.H.; Prasanna, K.; Jo, Y.N.; Lee, C.W. Al-Doped Li[Ni0.78Co0.1Mn0.1Al0.02]O2 for High Performance of Lithium Ion Batteries. Ceram. Int. 2019, 45, 6972–6977. [Google Scholar] [CrossRef]
  11. Purwanto, A.; Yudha, C.S.; Ubaidillah, U.; Widiyandari, H.; Ogi, T.; Haerudin, H. NCA cathode material: Synthesis methods and performance enhancement efforts. Mater. Res. Express 2018, 5, 122001. [Google Scholar] [CrossRef]
  12. Wang, D.; Liu, W.; Zhang, X.; Huang, Y.; Xu, M.; Xiao, W. Review of Modified Nickel-Cobalt Lithium Aluminate Cathode Materials for Lithium-Ion Batteries. Int. J. Photoenergy 2019, 2019, 2730849. [Google Scholar] [CrossRef] [Green Version]
  13. Xu, J.; Lin, F.; Doeff, M.M.; Tong, W. A review of Ni-based layered oxides for rechargeable Li-ion batteries. J. Mater. Chem. A 2017, 5, 874–901. [Google Scholar] [CrossRef] [Green Version]
  14. Xia, Y.; Zheng, J.; Wang, C.; Gu, M. Designing principle for Ni-rich cathode materials with high energy density for practical applications. Nano Energy 2018, 49, 434–452. [Google Scholar] [CrossRef]
  15. Yang, S.; Zhou, X.; Zhang, J.; Liu, Z. Morphology-controlled solvothermal synthesis of LiFePO4 as a cathode material for lithium-ion batteries. J. Mater. Chem. 2010, 20, 8086–8091. [Google Scholar] [CrossRef]
  16. Chen, J.; Wang, X.L.; Jin, E.M.; Moon, S.G.; Jeong, S.M. Optimization of B2O3 coating process for NCA cathodes to achieve long-term stability for application in lithium ion batteries. Energy 2021, 222, 119913. [Google Scholar] [CrossRef]
  17. Senthil, C.; Vediappan, K.; Nanthagopal, M.; Seop Kang, H.; Santhoshkumar, P.; Gnanamuthu, R.; Lee, C.W. Thermochemical conversion of eggshell as biological waste and its application as a functional material for lithium-ion batteries. Chem. Eng. J. 2019, 372, 765–773. [Google Scholar] [CrossRef]
  18. Sim, G.S.; Santhoshkumar, P.; Park, J.W.; Ho, C.W.; Shaji, N.; Kim, H.K.; Nanthagopal, M.; Lee, C.W. Chitosan-derived nitrogen-doped carbon on Li2ZnTi3O8/TiO2 composite as an anode material for lithium-ion batteries. Ceram. Int. 2021, 47, 33554–33562. [Google Scholar] [CrossRef]
  19. Park, K.S.; Son, J.T.; Chung, H.T.; Kim, S.J.; Lee, C.H.; Kang, K.T.; Kim, H.G. Surface modification by silver coating for improving electrochemical properties of LiFePO4. Solid State Commun. 2004, 129, 311–314. [Google Scholar] [CrossRef]
  20. Moskon, J.; Dominko, R.; Cerc-Korosec, R.; Gaberscek, M.; Jamnik, J. Morphology and electrical properties of conductive carbon coatings for cathode materials. J. Power Sources 2007, 174, 683–688. [Google Scholar] [CrossRef]
  21. Nanthagopal, M.; Santhoshkumar, P.; Shaji, N.; Praveen, S.; Kang, H.S.; Senthil, C.; Lee, C.W. Nitrogen-doped carbon-coated Li[Ni0.8Co0.1Mn0.1]O2 cathode material for enhanced lithium-ion storage. Appl. Surf. Sci. 2019, 492, 871–878. [Google Scholar] [CrossRef]
  22. Lee, D.J.; Ryou, M.H.; Lee, J.N.; Kim, B.G.; Lee, Y.M.; Kim, H.W.; Kong, B.S.; Park, J.K.; Choi, J.W. Nitrogen-doped carbon coating for a high-performance SiO anode in lithium-ion batteries. Electrochem. Commun. 2013, 34, 98–101. [Google Scholar] [CrossRef]
  23. Zhao, L.; Hu, Y.S.; Li, H.; Wang, Z.; Chen, L. Porous Li4Ti5O12 coated with N-doped carbon from ionic liquids for Li-ion batteries. Adv. Mater. 2011, 23, 1385–1388. [Google Scholar] [CrossRef] [PubMed]
  24. Feng, D.; Liu, Q.; Hu, T.; Chen, Y.; Zeng, T. Boosting cyclability performance of the LiNi0.8Co0.15Al0.05O2 cathode by a polyacrylonitrile-induced conductive carbon surface coating. Ceram. Int. 2021, 47, 12706–12715. [Google Scholar] [CrossRef]
  25. Yiğitalp, A.; Taşdemir, A.; Alkan Gürsel, S.; Yürüm, A. Nafion-coated LiNi0.80Co0.15Al0.05O2 (NCA) cathode preparation and its influence on the Li-ion battery cycle performance. Energy Storage 2020, 2, e154. [Google Scholar] [CrossRef]
  26. Park, K.Y.; Lim, J.M.; Luu, N.S.; Downing, J.R.; Wallace, S.G.; Chaney, L.E.; Yoo, H.; Hyun, W.J.; Kim, H.U.; Hersam, M.C. Concurrently Approaching Volumetric and Specific Capacity Limits of Lithium Battery Cathodes via Conformal Pickering Emulsion Graphene Coatings. Adv. Energy Mater. 2020, 10, 2001216. [Google Scholar] [CrossRef]
  27. Wang, H.Y.; Wang, Y.J.; Zhou, L.X.; Zhu, L.; Zhang, Y.Q. Isolation and bioactivities of a non-sericin component from cocoon shell silk sericin of the silkworm Bombyx mori. Food Funct. 2012, 3, 150–158. [Google Scholar] [CrossRef]
  28. Kunz, R.I.; Brancalhão, R.M.C.; Ribeiro, L.D.F.C.; Natali, M.R.M. Silkworm Sericin: Properties and Biomedical Applications. BioMed Res. Int. 2016, 2016, 8175701. [Google Scholar] [CrossRef] [Green Version]
  29. Das, G.; Shin, H.S.; Campos, E.V.R.; Fraceto, L.F.; del Pilar Rodriguez-Torres, M.; Mariano, K.C.F.; de Araujo, D.R.; Fernández-Luqueño, F.; Grillo, R.; Patra, J.K. Sericin based nanoformulations: A comprehensive review on molecular mechanisms of interaction with organisms to biological applications. J. Nanobiotechnol. 2021, 19, 30. [Google Scholar] [CrossRef]
  30. Abdolmaleki, A.; Mahmoudian, M. Use of biomass sericin as matrices in functionalized graphene/sericin nanocomposites for the removal of phenolic compounds. Heliyon 2020, 6, e04955. [Google Scholar] [CrossRef]
  31. Sim, G.S.; Nanthagopal, M.; Santhoshkumar, P.; Park, J.W.; Ho, C.W.; Shaji, N.; Kim, H.K.; Lee, C.W. Biomass-derived nitrogen-doped carbon on LiFePO4 material for energy storage applications. J. Alloys Compd. 2022, 902, 163720. [Google Scholar] [CrossRef]
  32. Park, K.J.; Hwang, J.Y.; Ryu, H.H.; Maglia, F.; Kim, S.J.; Lamp, P.; Yoon, C.S.; Sun, Y.K. Degradation Mechanism of Ni-Enriched NCA Cathode for Lithium Batteries: Are Microcracks Really Critical? ACS Energy Lett. 2019, 4, 1394–1400. [Google Scholar] [CrossRef] [Green Version]
  33. Shan, W.; Huang, S.; Zhang, H.; Hou, X. Surface coating for high-nickel cathode materials to achieve excellent cycle performance at elevated temperatures. J. Alloy. Compd. 2021, 862, 158022. [Google Scholar] [CrossRef]
  34. Wu, F.; Wang, M.; Su, Y.; Chen, S.; Xu, B. Effect of TiO2-coating on the electrochemical performances of LiCo1/3Ni1/3Mn1/3O2. J. Power Sources 2009, 191, 628–632. [Google Scholar] [CrossRef]
  35. Song, C.; Wang, W.; Peng, H.; Wang, Y.; Zhao, C.; Zhang, H.; Tang, Q.; Lv, J.; Du, X.; Dou, Y. Improving the electrochemical performance of LiNi0.80Co0.15Al0.05O2 in lithium ion batteries by LiAlO2 surface modification. Appl. Sci. 2018, 8, 378. [Google Scholar] [CrossRef] [Green Version]
  36. Li, P.; Jiao, Y.; Yao, S.; Wang, L.; Chen, G. Dual role of nickel foam in NiCoAl-LDH ensuring high-performance for asymmetric supercapacitors. New J. Chem. 2019, 43, 3139–3145. [Google Scholar] [CrossRef]
  37. Gao, P.; Jiang, Y.; Zhu, Y.; Hu, H. Improved cycle performance of nitrogen and phosphorus co-doped carbon coatings on lithium nickel cobalt aluminum oxide battery material. J. Mater. Sci. 2018, 53, 9662–9673. [Google Scholar] [CrossRef]
  38. Jamil, S.; Ran, Q.; Yang, L.; Huang, Y.; Cao, S.; Yang, X.; Wang, X. Improved high-voltage performance of LiNi0.87Co0.1Al0.03O2 by Li+-conductor coating. Chem. Eng. J. 2021, 407, 126442. [Google Scholar] [CrossRef]
  39. Jabeen, M.; Ishaq, M.; Song, W.; Xu, L.; Deng, Q. Synthesis of Ni/Co/Al-layered triple hydroxide@brominated graphene hybrid on nickel foam as electrode material for high-performance supercapacitors. RSC Adv. 2017, 7, 46553–46565. [Google Scholar] [CrossRef] [Green Version]
  40. Genieser, R.; Ferrari, S.; Loveridge, M.; Beattie, S.D.; Beanland, R.; Amari, H.; West, G.; Bhagat, R. Lithium ion batteries (NMC/graphite) cycling at 80 °C: Different electrolytes and related degradation mechanism. J. Power Sources 2018, 373, 172–183. [Google Scholar] [CrossRef]
  41. Chulliyote, R.; Hareendrakrishnakumar, H.; Raja, M.; Gladis, J.M.; Stephan, A.M. Sulfur-Immobilized Nitrogen and Oxygen Co–Doped Hierarchically Porous Biomass Carbon for Lithium-Sulfur Batteries: Influence of Sulfur Content and Distribution on Its Performance. ChemistrySelect 2017, 2, 10484–10495. [Google Scholar] [CrossRef]
  42. Xing, Z.; Ju, Z.; Zhao, Y.; Wan, J.; Zhu, Y.; Qiang, Y.; Qian, Y. One-pot hydrothermal synthesis of Nitrogen-doped graphene as high-performance anode materials for lithium ion batteries. Sci. Rep. 2016, 6, 26146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zhu, H.; Deng, W.; Chen, L.; Zhang, S. Nitrogen doped carbon layer of Li2MnSiO4 with enhanced electrochemical performance for lithium ion batteries. Electrochim. Acta 2019, 295, 956–965. [Google Scholar] [CrossRef]
  44. Zhou, P.; Zhang, Z.; Meng, H.; Lu, Y.; Cao, J.; Cheng, F.; Tao, Z.; Chen, J. SiO2-coated LiNi0.915Co0.075Al0.01O2 cathode material for rechargeable Li-ion batteries. Nanoscale 2016, 8, 19263–19269. [Google Scholar] [CrossRef]
  45. Han, C.J.; Yoon, J.H.; Cho, W., II; Jang, H. Electrochemical properties of LiNi0.8Co0.2−xAlxO2 prepared by a sol-gel method. J. Power Sources 2004, 136, 132–138. [Google Scholar] [CrossRef]
  46. Huang, B.; Li, X.; Wang, Z.; Guo, H.; Xiong, X. Synthesis of Mg-doped LiNi0.8Co0.15Al0.05O2 oxide and its electrochemical behavior in high-voltage lithium-ion batteries. Ceram. Int. 2014, 40, 13223–13230. [Google Scholar] [CrossRef]
  47. Chen, T.; Li, X.; Wang, H.; Yan, X.; Wang, L.; Deng, B.; Ge, W.; Qu, M. The effect of gradient boracic polyanion-doping on structure, morphology, and cycling performance of Ni-rich LiNi0.8Co0.15Al0.05O2 cathode material. J. Power Sources 2018, 374, 1–11. [Google Scholar] [CrossRef]
  48. Santhoshkumar, P.; Prasanna, K.; Jo, Y.N.; Sivagami, I.N.; Kang, S.H.; Lee, C.W. A facile and highly efficient short-time homogenization hydrothermal approach for the smart production of high-quality α-Fe2O3 for rechargeable lithium batteries. J. Mater. Chem. A 2017, 5, 16712–16721. [Google Scholar] [CrossRef]
  49. Nanthagopal, M.; Santhoshkumar, P.; Shaji, N.; Sim, G.S.; Park, J.W.; Senthil, C.; Lee, C.W. An encapsulation of nitrogen and sulphur dual-doped carbon over Li[Ni0.8Co0.1Mn0.1]O2 for lithium-ion battery applications. Appl. Surf. Sci. 2020, 511, 145580. [Google Scholar] [CrossRef]
  50. Yu, J.; Li, H.; Zhang, G.; Li, X.; Huang, J.; Li, C.; Wei, C.; Xiao, C. Carbon nanotubes coating on LiNi0.8Co0.15Al0.05O2 as cathode materials for lithium battery. Int. J. Electrochem. Sci. 2017, 12, 11892–11903. [Google Scholar] [CrossRef]
  51. Chung, Y.; Ryu, S.H.; Ju, J.H.; Bak, Y.R.; Hwang, M.J.; Kim, K.W.; Cho, K.K.; Ryu, K.S. A surfactant-based method for carbon coating of LiNi0.8Co0.15Al0.05O2 cathode in Li ion batteries. Bull. Korean Chem. Soc. 2010, 31, 2304–2308. [Google Scholar] [CrossRef] [Green Version]
  52. Yoon, S.; Jung, K.N.; Yeon, S.H.; Jin, C.S.; Shin, K.H. Electrochemical properties of LiNi0.8Co0.15Al0.05O2-graphene composite as cathode materials for lithium-ion batteries. J. Electroanal. Chem. 2012, 683, 88–93. [Google Scholar] [CrossRef]
  53. Liu, Z.; Wang, Z.; Lu, T.; Dai, P.; Gao, P.; Zhu, Y. Modification of LiNi0.8Co0.15Al0.05O2 using nanoscale carbon coating. J. Alloy. Compd. 2018, 763, 701–710. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA.
Figure 1. XRD patterns of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA.
Nanomaterials 12 01166 g001
Figure 2. XPS spectra for 0.3 NC@NCA: (a) survey; (b) Ni 2p; (c) Co 2p; (d) Al 2p; (e) O 1s; (f) C 1s; and (g) N 1s.
Figure 2. XPS spectra for 0.3 NC@NCA: (a) survey; (b) Ni 2p; (c) Co 2p; (d) Al 2p; (e) O 1s; (f) C 1s; and (g) N 1s.
Nanomaterials 12 01166 g002
Figure 3. HR-FE-SEM images of (ac) bare NCA; (df) 0.2 NC@NCA; (gi) 0.3 NC@NCA; and (jl) 0.4 NC@NCA.
Figure 3. HR-FE-SEM images of (ac) bare NCA; (df) 0.2 NC@NCA; (gi) 0.3 NC@NCA; and (jl) 0.4 NC@NCA.
Nanomaterials 12 01166 g003
Figure 4. HR-FE-SEM images, elemental mapping spectra, and map scan spectra of (a (iv)) bare NCA; (b (ivii)) 0.2 NC@NCA; (c (ivii)) 0.3 NC@NCA; and (d (ivii)) 0.4 NC@NCA.
Figure 4. HR-FE-SEM images, elemental mapping spectra, and map scan spectra of (a (iv)) bare NCA; (b (ivii)) 0.2 NC@NCA; (c (ivii)) 0.3 NC@NCA; and (d (ivii)) 0.4 NC@NCA.
Nanomaterials 12 01166 g004
Figure 5. The galvanostatic charge/discharge potential profiles of (a) bare NCA; (b) 0.2 NC@NCA; (c) 0.3 NC@NCA; and (d) 0.4 NC@NCA.
Figure 5. The galvanostatic charge/discharge potential profiles of (a) bare NCA; (b) 0.2 NC@NCA; (c) 0.3 NC@NCA; and (d) 0.4 NC@NCA.
Nanomaterials 12 01166 g005
Figure 6. Differential capacity for (a) bare NCA, (b) 0.2 NC@NCA, (c) 0.3 NC@NCA, and (d) 0.4 NC@NCA; cycling performance profiles showing specific capacity and its coulombic efficiency at a current density of 0.2 C for (e) bare NCA, (f) 0.2 NC@NCA, (g) 0.3 NC@NCA, and (h) 0.4 NC@NCA.
Figure 6. Differential capacity for (a) bare NCA, (b) 0.2 NC@NCA, (c) 0.3 NC@NCA, and (d) 0.4 NC@NCA; cycling performance profiles showing specific capacity and its coulombic efficiency at a current density of 0.2 C for (e) bare NCA, (f) 0.2 NC@NCA, (g) 0.3 NC@NCA, and (h) 0.4 NC@NCA.
Nanomaterials 12 01166 g006
Figure 7. (a) Cycling performance at a current density of 1.0 C; (b) rate capability; (c) EIS analysis after 200 cycles at a current density of 1.0 C for bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA.
Figure 7. (a) Cycling performance at a current density of 1.0 C; (b) rate capability; (c) EIS analysis after 200 cycles at a current density of 1.0 C for bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA.
Nanomaterials 12 01166 g007
Table 1. Lattice parameters of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA.
Table 1. Lattice parameters of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA.
Bare NCA0.2 NC@NCA0.3 NC@NCA0.4 NC@NCA
a (Å)2.86362.86362.86252.8614
c (Å)14.153214.150214.164314.1456
c/a4.94254.94144.94824.9436
Table 2. Nitrogen weight ratio and nitrogen atomic ratio of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA.
Table 2. Nitrogen weight ratio and nitrogen atomic ratio of bare NCA, 0.2 NC@NCA, 0.3 NC@NCA, and 0.4 NC@NCA.
Bare NCA0.2 NC@NCA0.3 NC@NCA0.4 NC@NCA
Nitrogen Weight Ratio (wt.%)00.610.751.04
Nitrogen Atomic Ratio (%)01.131.301.94
Table 3. Comparison of previously reported similar studies with current work.
Table 3. Comparison of previously reported similar studies with current work.
MaterialCarbon SourceDischarge Capacity
(mAh g−1)
Capacity Retention
(%)
Ref.
Carbon nanotube coating on NCA
(CNT-NCA)
Carbon nanotubes205 (at 0.1 C)91 (at 0.1 C)[50]
Carbon-coated LNCAO
(LNCAO/C)
Sodium dodecyl sulfate183 (at 0.1 C)93 (at 0.1 C)[51]
NCA-grapheneGraphene nanoplatelets180 (at 0.5 C)97 (at 0.5 C)[52]
Nanoscale carbon coating on NCA (C@NCA)Glucose260 (at 1 C)88 (at 1.0 C)[53]
N-doped carbon-coated NCA
(NC@NCA)
Sericin161 (at 0.2 C)92 (at 0.2 C)*This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sim, G.S.; Shaji, N.; Santhoshkumar, P.; Park, J.W.; Ho, C.W.; Nanthagopal, M.; Kim, H.K.; Lee, C.W. Silkworm Protein-Derived Nitrogen-Doped Carbon-Coated Li[Ni0.8Co0.15Al0.05]O2 for Lithium-Ion Batteries. Nanomaterials 2022, 12, 1166. https://doi.org/10.3390/nano12071166

AMA Style

Sim GS, Shaji N, Santhoshkumar P, Park JW, Ho CW, Nanthagopal M, Kim HK, Lee CW. Silkworm Protein-Derived Nitrogen-Doped Carbon-Coated Li[Ni0.8Co0.15Al0.05]O2 for Lithium-Ion Batteries. Nanomaterials. 2022; 12(7):1166. https://doi.org/10.3390/nano12071166

Chicago/Turabian Style

Sim, Gyu Sang, Nitheesha Shaji, P. Santhoshkumar, Jae Woo Park, Chang Won Ho, Murugan Nanthagopal, Hong Ki Kim, and Chang Woo Lee. 2022. "Silkworm Protein-Derived Nitrogen-Doped Carbon-Coated Li[Ni0.8Co0.15Al0.05]O2 for Lithium-Ion Batteries" Nanomaterials 12, no. 7: 1166. https://doi.org/10.3390/nano12071166

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