Next Article in Journal
Local Structure and Magnetism of Fe2O3 Maghemite Nanocrystals: The Role of Crystal Dimension
Next Article in Special Issue
Binder Free and Flexible Asymmetric Supercapacitor Exploiting Mn3O4 and MoS2 Nanoflakes on Carbon Fibers
Previous Article in Journal
Preparation and Application of Bismuth/MXene Nano-Composite as Electrochemical Sensor for Heavy Metal Ions Detection
Previous Article in Special Issue
3-Dimensional Porous Carbon with High Nitrogen Content Obtained from Longan Shell and Its Excellent Performance for Aqueous and All-Solid-State Supercapacitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sn-Doping and Li2SnO3 Nano-Coating Layer Co-Modified LiNi0.5Co0.2Mn0.3O2 with Improved Cycle Stability at 4.6 V Cut-off Voltage

1
School of Physics and Electronic Science, Changsha University of Science and Technology, Changsha 410114, China
2
Qinyuan Jiazhi Institute Co. Ltd., Qingyuan 511517, China
3
School of Materials Science and Engineering, Changsha University of Science and Technology, Changsha 410114, China
4
School of Chemistry and Food Engineering, Changsha University of Science and Technology, Changsha 410114, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(5), 868; https://doi.org/10.3390/nano10050868
Submission received: 11 April 2020 / Revised: 26 April 2020 / Accepted: 27 April 2020 / Published: 30 April 2020
(This article belongs to the Special Issue Design and Synthesis of Nanomaterials for Energy Storage)

Abstract

:
Nickel-rich layered LiNi1−xyCoxMnyO2 (LiMO2) is widely investigated as a promising cathode material for advanced lithium-ion batteries used in electric vehicles, and a much higher energy density in higher cut-off voltage is emergent for long driving range. However, during extensive cycling when charged to higher voltage, the battery exhibits severe capacity fading and obvious structural collapse, which leads to poor cycle stability. Herein, Sn-doping and in situ formed Li2SnO3 nano-coating layer co-modified spherical-like LiNi0.5Co0.2Mn0.3O2 samples were successfully prepared using a facile molten salt method and demonstrated excellent cyclic properties and high-rate capabilities. The transition metal site was expected to be substituted by Sn in this study. The original crystal structures of the layered materials were influenced by Sn-doping. Sn not only entered into the crystal lattice of LiNi0.5Co0.2Mn0.3O2, but also formed Li+-conductive Li2SnO3 on the surface. Sn-doping and Li2SnO3 coating layer co-modification are helpful to optimize the ratio of Ni2+ and Ni3+, and to improve the conductivity of the cathode. The reversible capacity and rate capability of the cathode are improved by Sn-modification. The 3 mol% Sn-modified LiNi0.5Co0.2Mn0.3O2 sample maintained the reversible capacity of 146.8 mAh g−1 at 5C, corresponding to 75.8% of its low-rate capacity (0.1C, 193.7mAh g−1) and kept the reversible capacity of 157.3 mAh g−1 with 88.4% capacity retention after 100 charge and discharge cycles at 1C rate between 2.7 and 4.6 V, showing the improved electrochemical property.

1. Introduction

The lithium-ion battery (LIB) is one of the most promising power supply devices for portable electronic products and electric vehicles because of its high energy density and power density, long cycle lifetime and environmental benignity among various novel battery systems [1,2,3,4,5]. Because the commercialized LiCoO2 cathode material has the disadvantages of high cost and poor thermal safety, the aim of present research work is to develop the prospective alternatives for LiCoO2 toward better lithium batteries [6,7,8,9,10]. Nickel-rich layered LiMO2 is an important cathode material for LIB because of its superior theoretical discharge capacity compared with that of olivine or spinel materials. Herein, high nickel content in LiMO2 is beneficial to increase capacity, high manganese content enhances the structural stability and high cobalt content improves the rate performance [11]. Among these LiMO2 materials, LiNi0.5Co0.2Mn0.3O2 has been gradually used as a component of commercial lithium secondary batteries due to its low price, high capacity and improved safety. Nevertheless, LiNi0.5Co0.2Mn0.3O2 still suffers from several issues, including severe capacity degradation and limited high-rate capability, especially at high cut-off voltage, which is ascribed to the transition metal dissolution and surface structure transformation during cycling [11].
At present, surface coating has been implemented to improve the cyclic property of cathode materials. Coatings like metal oxides (such as Al2O3 [12,13], antimony doped tin oxide (ATO) [14], CeO2 [15], CuO [16], Cr8O21 [17], MoO3 [18], SiO2 [19], TiO2 [20], ZnO [21], and ZrO2 [22]), fluoride (AlF3 [23]), lithium salts (such as LiAlO2 [24], LiBO2 [25], Li2MnO3 [26], Li2MoO4 [27], Li2SiO3 [28], Li2TiO3 [29], Li3VO4 [30], and Li2ZrO3 [31]), and others (polypyrrole (PPy) [32], carbon nanotube (CNT) [33]) are proven to be effective for alleviating the transition metal dissolution and then improving the cyclic property of the cathode materials. On the other side, partial substitution with cations or anions is considered as a promising method to stabilize the crystalline structure of LiNi0.5Co0.2Mn0.3O2 materials and improve its high-rate capability, such as bulk doping with Al [24,34,35], K [36], Mo [37], Na [38,39], Nd [40], Ti [41,42], Zr [22,32,42], Y [43], F [44], and Cl [36]. In principle, researchers choose the ions which show a large ionic radius and high electronegativity to substitute for the transition metals in LiMO2 because these kinds of ions can expand the channel-like Li+ diffusion pathway and decrease the covalence characteristics of cation–oxygen bonds of LiMO2 materials. Ion substitution inhibits the release of oxygen and has little effect on the structure of LiNi0.5Co0.2Mn0.3O2 materials, so the host lattice of the materials will be well maintained. As we know, the electrochemical properties of LiMO2 materials can be greatly enhanced when combined with the advantages of doping and coating co-modification by LiAlO2-coating layer and Al-dopant [24]. Mo-coating and doping for LiNi0.5Co0.2Mn0.3O2 [45] have been studied by researchers and demonstrated enhanced electrochemical properties.
Among various doping cations, Sn4+ has the same ionic radius of 0.69 Å as Ni2+, close to that of Li+ (0.76 Å). Meanwhile, the bonding energy of Sn-O is 548 kJ mol−1, while those of Ni-O, Co-O and Mn-O are 391.6, 368 and 402 kJ mol−1, respectively. The high bonding energy of Sn-O is favorable to improve the crystalline structural stability of cathode materials. In addition, Sn4+ has a high electronegativity, leading to strong ionicity of the metal–oxygen bond [46]. Therefore, Sn-modification is expected to enhance the cyclic property and high-rate capability of LiMO2 materials. The rate performances of some layered LiMO2 materials have been enhanced by substituting stannum for transition metals. For example, LiNi3/8Co2/8/Mn3/8-xSnxO2 has enhanced the chemical diffusion coefficient DLi of Li-ion, leading to improved rate capability [47]. Sn-doped LiNi0.8Co0.2O2 has increased electronic conductivity because a free electron was released into the conduction band after doping [48]. The electrochemical properties of LiNi0.8Co0.1Mn0.1O2 have been enhanced by SnO2 at high voltage [49].
In this study, we report a facile synthesis of Sn-doping and Li2SnO3 in situ coating layer co-modified (Sn-modified) LiNi0.5Co0.2Mn0.3O2. The crystalline structures, morphologies, surface chemical states of cations and electrochemical performances of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples are characterized. As expected, the amount of Sn has a major impact on the modification treatment. Substituting a large amount of Sn for transition metals in LiNi0.5Co0.2Mn0.3O2 can form Li2SnO3 on the surface. Suitable Sn-substituting can relieve the cation mixing degree and provide a stable structure as well as form the Li+-conductive coating layer on the surface of the sample, leading to improved physical and electrochemical properties.

2. Materials and Methods

2.1. Materials Preparation

Layered Sn-modified LiNi0.5Co0.2Mn0.3O2 samples were synthesized using a facile molten salt method. Molten salt (0.76LiOH·H2O–0.12Li2CO3), commercial Ni0.5Co0.2Mn0.3(OH)2 precursors and nano-sized Sn powder were used as raw materials and mixed completely by mortar and pestle with the appropriate amount of ethyl alcohol. LiOH·H2O with a purity of 98% was bought from Xilong Chemical Co., Ltd. (Shantou, China). Li2CO3 with a purity of 99.5% was provided by Sichuan Tianqi Lithium Co., Ltd (Chengdu, China). Ni0.5Co0.2Mn0.3(OH)2 with a transition metal element content of 62.27% was purchased from Chongqing Teri battery materials Co., Ltd. (Chongqing, China). Sn powder with a purity of 99.9% was bought from Shanghai Chaowei Nano Technology Co., Ltd. The molar ratio of the Li and M in LiMO2 was 1.05:1. The mixture was pre-heated at 480 °C for 120 min followed by calcination at 880 °C for 720 min in air atmosphere. Finally, the obtained samples were ground for 30 min for physical and electrochemical property tests. Here, Sn-modified LiNi0.5Co0.2Mn0.3O2 compounds, in which certain amounts of transition metals were substituted by Sn, were marked as MS1 (1 mol% Sn), MS3 (3 mol% Sn) and MS5 (5 mol% Sn), and were prepared via the above-mentioned approaches. The pristine LiNi0.5Co0.2Mn0.3O2 compounds were obtained through the same method and labeled as M523.

2.2. Characterizations

The crystalline structures of synthesized LiNi0.5Co0.2Mn0.3O2 materials were characterized using X-ray diffraction (XRD, D/Max 2000/PC, Rigaku, Tokyo, Japan) with Cu Kα radiation (λ = 1.54056 Å) from 10° to 90° with a scan rate of 5° per min. The morphologies of modified samples were characterized by scanning electron microscopy (SEM, Sirion200, FEI Ltd., Eindhoven, The Netherlands). The microstructure of the sample surface was analyzed using transmission electron microscopy (TEM, TECNAI G2 F20, FEI Company, Hillsboro, USA). The element distributions were determined using energy dispersive X-ray spectroscopy (EDS, Model 7426, Oxford, UK). The surfaces of the samples were examined using X-ray photoelectron spectroscopy (XPS, K-Alpha 1063, Thermo Fisher Scientific, Waltham, MA, USA) with AlKα line (1486.6 eV) as the source of X-ray.
The CR2025 cell assembly process, the electrochemical charge and discharge tests and the electrochemical impedance spectroscopy (EIS) tests were conducted according to the experimental section of our recently published article [36].

3. Results and Discussion

The crystalline structures of the Sn-modified LiNi0.5Co0.2Mn0.3O2 samples were studied using XRD in order to characterize the effects of Sn-substitution on the crystal, and the typical diffraction patterns of all samples are demonstrated in Figure 1. The XRD patterns of well-crystallized pristine and Sn-modified LiNi0.5Co0.2Mn0.3O2 samples were all indexed to a hexagonal α-NaFeO2 layered structure (R-3m space group) with sharp and clear diffraction peaks. The obvious splitting of diffraction peaks of (006)/(102) and (108)/(110) reflects the highly ordered hexagonal structure. However, there are some impurities in the patterns of samples MS3 and MS5 near the 2θ of 35 and 43°, which are identified as Li2SnO3. It is obvious that the formation of Li2SnO3 phase is related to the amount of dopant. To identify the effects of tin substitution on the structures of Sn-modified samples, the crystallographic data of samples are demonstrated in Table 1. Even though the doping amount was small, the cell parameters of all samples changed, showing that Sn-modification affected the main structure of the host. All the crystallographic data changed, which suggests that the substituting element entered into the crystal lattice. All the crystallographic data ratios c/a are higher than 4.899, showing the highly ordered crystal structure. The I003/I104 ratios (R) of modified samples are larger than the value of 1.2, indicating that Sn-substituting can relieve the cation mixing degree. That is to say, Sn4+ helps to stabilize the crystal structure of LiNi0.5Co0.2Mn0.3O2 during the Li+ intercalation and de-intercalation process because Sn-O has a higher bonding energy than those of transition metals and oxygen. The 3 mol% substituting sample showed the largest intensity ratio R’((I006+ I102)/I101) and crystal volume, which may have resulted in the best electrochemical performance.
The SEM images of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples and the EDS images are displayed in Figure 2. As is shown, there were no significant differences in the grain sizes from the pristine and Sn-modified samples. All the compounds showed a spherical-like morphology with a particle size from 4 to 6 μm, which is made up by lots of fine primary particles with a length range of 0.5–1 μm. The sample surface was not only compact but also provided enough surface area to make full contact between the cathode and the electrolyte. According to the EDS measurements showing in Figure 2k, it can be obviously seen that stannum and transition metals were uniformly distributed on the surface of the MS3 compound.
To reveal the in situ formation of the Li2SnO3 on the surface of the samples, the microstructure of Sn-modified LiNi0.5Co0.2Mn0.3O2 sample MS3 was examined using TEM as shown in Figure 3. It can be seen from Figure 3a that a nano-sized coating layer was obtained on the particle surface of MS3. Three of the coating sites (Figure 3b–d) were enlarged in order to observe the thin layer more clearly. The surface coating layer, which uniformly adhered to the bulk of MS3 particles, had a thickness within the range of 5–10 nm. In addition, Figure 3b,c clearly indicates the crystalline interplanar spacing of 0.495 and 0.228 nm, which can be indexed as (002) and (–221) facets of Li2SnO3 (JCPDS no. 31-0761), respectively. The (003) facet of bulk MS3 with the crystalline interplanar spacing of 0.478 nm is demonstrated in Figure 3d. These results were consistent with the XRD result and further identified the formation of Li2SnO3 nano-coating layer.
To understand the surface chemical composition of the transition metal elements (Ni, Co, Mn) and Sn, the pristine M523 and Sn-modified sample MS3 were examined using XPS. The XPS patterns are shown in Figure 4 and Figure S1. Compared to the pristine M523, the peak positions of Co 2p3/2 and Mn 2p3/2 in 3 mol% Sn-modified sample MS3 had no obvious shift, showing that the surface chemical states of the transition metals did not change. To further clarify the effect of Sn-modification on the chemical states of cations, the peak positions and mole fractions of transition metal ions and Sn4+ in the crystal of M523 and MS3 compounds deduced from XPS fittings are listed in Table 2. According to the corresponding binding energies of Ni 2p3/2, Co 2p3/2 and Mn 2p3/2, we can ascertain that the chemical valences of Ni are Ni2+ (853.6 and 854.7 eV) and Ni3+(856.2 eV), while those of Co and Mn are Co3+ (779.8 eV) and Mn4+ (642.4 eV), respectively. The results show that the oxidation valences of Ni, Co and Mn in the Sn-modified samples are still the same as those of the pristine one, only that the mole fraction ratio of Ni2+/Ni3+ increased from 72.27%/27.73% to 74.88%/25.12% after Sn-modification, which indicates that MS3 has better structure stability. Sn3d peaks appear at 486.5 and 494.9 eV, showing that Sn exists in +4 chemical state [47]. Additionally, Sn-modification has a great influence on the chemical state of O1s. The peak at 529.35 eV was caused by the interaction of transition metal ions and oxygen in the crystal structure, and the peak at 531.54 eV is related to formation of lithium carbonate at the sample surface [50]. The peak intensities of the two characteristic peaks of O1s occur in deflection, which indicates that the lattice oxygen increased and the adsorbed oxygen on the surface decreased after modification. It is beneficial to keep the layered structure stable and reduce the formation of impurities on the sample surface.
The curves of electrochemical performance are characterized in Figure 5. The initial charge and discharge capacities were tested at 0.1 C rate at room temperature. We can determine from Figure 5a that the initial discharge capacities for the M523, MS1, MS3 and MS5 samples are 203.9, 196.8, 193.7 and 188.0 mAh g−1, and the corresponding coulombic efficiencies are 79.2%, 82.6%, 84.9% and 84.0%, respectively. As discussed previously, a large amount of Sn-doping and Li2SnO3 impurity existed in MS5, which probably led to the lowest initial discharge capacity. The coulombic efficiency values of the Sn-modified samples are all higher than that of the pristine one. This should be attributed to the Sn-substituting, which can relieve the cation mixing degree and is favorable for Li+ transfer.
The rate performances of LiNi0.5Co0.2Mn0.3O2 samples are compared in Figure 5b, in which the charge–discharge cycle was successively taken from 0.1 to 5 C at 2.7–4.6 V for every five cycles. The Sn-modified samples displayed more enhanced rate performance than the pristine M523 at high rates. The MS3 sample presented a reversible capacity of 146.8 mAh g−1 at 5 C, corresponding to 75.8% of its initial capacity (0.1C, 193.7mAh g−1). However, the pristine M523 kept a reversible capacity of 116.0 mAh g−1, just 56.9% of its initial capacity (0.1C, 203.9 mAh g−1). This can be attributed to the fact that the bonding energy of Sn–O is higher than those of the transition metal and oxygen in LiNi0.5Co0.2Mn0.3O2 samples. It can be seen from the previous XRD results that Sn-substituting can relieve the cation mixing degree and benefit the Li+ intercalation/de-intercalation, even in high current density. Furthermore, the formed Li+-conductive Li2SnO3 nano-coating layer prevents the side reaction at the cathode and the electrolyte interface and accelerates the transport of lithium ions as well.
The cyclic stability properties of Sn-modified LiNi0.5Co0.2Mn0.3O2 at 1C are illustrated in Figure 5c. It is observed that the Sn-modified sample MS3 exhibited excellent capacity retention with a capacity of 157.3mAh g−1 and discharge capacity retention of 88.4% at the 100th cycle, while the pristine sample M523 only kept a capacity of 124.9 mAh g−1 and discharge capacity retention of 73.2%. These results indicate that Sn-modification is favorable for keeping the structural stability of the pristine materials and obtaining enhanced cycle performance. Doping can improve the conductivity of the material, and the increase of conductivity after a small amount of doping is reflected in the increase of capacity; however, with the increase of doping amount, the active material decreases, resulting in the loss of electrochemical capacity. Therefore, there is a lack of continuous changes for data shown in Figure 5b,c with the trend of MS5 < MS3 < MS1 or MS1 < MS3 < MS5.
To better understand the effect of Sn-modification on the electrochemical properties of cathode materials, EIS analysis was carried out. Figure 5d demonstrates the EIS profiles of the Sn-modified LiNi0.5Co0.2Mn0.3O2 cathodes after the 100th cycle at 1C. According to the equivalent circuit [40] in the inset in which Rsei stands for the resistance of solid electrolyte interface (SEI) layer, Rct is the charge transfer impedance at the interface of electrolyte–electrode, and Zw is the Li ion diffusion in the crystal lattice, the EIS spectra were analyzed using Zview-2 software. The values of the total resistance of the Rsei and Rct are 306.0, 233.1, 171.0 and 152.6 Ω for M523, MS1, MS3 and MS5, respectively. It is shown that the Rsei and Rct of the modified samples are significantly smaller than those of the pristine one. The MS3 and MS5 samples show lower total resistance, which may be relative to the formed Li+-conductive Li2SnO3. This is helpful for the intercalation/de-intercalation of Li ions during the charge/discharge process. To investigate the influence of bulk performance of LiNi0.5Co0.2Mn0.3O2 modified by Sn-doping on electrochemical performance, the relationships between ω−1/2 and Z’ based on the experimental results are shown in Figure S2. The apparent Li+ diffusion coefficient was calculated via a widely used method [8], and it was 1.64 × 10−10, 1.75 × 10−10 2.11 × 10−10 and 1.82 × 10−10 cm2 S−1 for M523, MS1, MS3 and MS5, respectively. Hence, it could be claimed that Sn-modification contributes to decreasing the charge transfer impedance and improving the Li+ diffusion, resulting in better capacity reversibility.

4. Conclusions

Uniform near-spherical Sn-doping and Li2SnO3 co-modified LiNi0.5Co0.2Mn0.3O2 were obtained using a facile molten salt method with 0.76LiOH·H2O-0.12Li2CO3, commercial Ni0.5Co0.2Mn0.3(OH)2 and Sn nano-powders as the raw materials. The crystal structures, morphologies and electrochemical properties were investigated in detail. The results of the analyses indicate that suitable stannum-modified samples exhibit low cation mixing degrees, enhanced rate capabilities and excellent cyclic performances. Notably, the MS3 sample with 3 mol % Sn-modification aimed at the transition metal site maintained a capacity of 146.8 mAh g−1 at the current density of 5C, corresponding to 75.8% of its low rate capacity (0.1C, 193.7mAh g−1), while the pristine one kept the capacity of 116.0 mAh g−1, just 56.9% of its initial capacity (0.1C, 203.9 mAh g−1). The pristine sample also kept the reversible capacity of 157.3 mAh g−1 as well as a favorable capacity retention of 88.4% after 100 cycles (2.7–4.6 V, 1C), which is 15.2% higher than that of the pristine M523 (124.9 mAh g−1, 73.2%). The MS3 sample exhibited a lower mole fraction of Ni3+, implying less structural transition during the charge–discharge cycles. The improvement of the electrochemical properties can be attributed to the suitable Sn-substituting and formed Li+-conductive Li2SnO3, which can relieve the cation mixing degree, offer more stable crystalline structure for the fast Li+-intercalation/de-intercalation during repeated cycles and improve the conductivity to obtain enhanced high-rate reversibility and cycle stability. These results illustrate that Sn-modified LiNi0.5Co0.2Mn0.3O2 is an excellent cathode material for increasingly wide utilization in the fields of electric vehicles and energy storage systems.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/5/868/s1, Figure S1: XPS spectra of the synthesized materials: Co 2p3/2 (a), Mn 2p3/2 (c) spectra of M523, Co 2p3/2 (b), Mn 2p3/2 (d) spectra of MS3; Figure S2: The relationships between ω−1/2 and Z’.

Author Contributions

Methodology and Conceptualization, H.Z., Y.T., and Z.C.; Resources, S.Z. and J.L.; Formal analysis, H.Z., R.S., X.Y., and Z.F.; Writing—original draft preparation, H.Z.; Writing—review and editing, Z.C.; Funding acquisition, H.Z., L.S., and Z.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Nos. 51604042, 21501015, 51874048), the Scientific Research Foundation of Hunan Provincial Education Department (Nos. 19B010, 19A003).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tarascon, J.-M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef]
  2. Lu, H.; Zhu, Y.; Zheng, B.; Du, H.; Zheng, X.; Liu, C.; Yuan, Y.; Fang, J.; Zhang, K. A hybrid ionic liquid-based electrolyte for high-performance lithium–sulfur batteries. New J. Chem. 2020, 44, 361–368. [Google Scholar] [CrossRef]
  3. Zheng, J.; Yao, Y.; Mao, G.; Chen, H.; Li, H.; Cao, L.; Ou, X.; Yu, W.; Ding, Z.; Tong, H. Iron–zinc sulfide Fe2Zn3S5/Fe1−xS@C derived from a metal–organic framework as a high performance anode material for lithium-ion batteries. J. Mater. Chem. A 2019, 7, 16479–16487. [Google Scholar] [CrossRef]
  4. Wang, X.; Tan, Y.; Liu, Z.; Fan, Y.; Li, M.; Younus, H.A.; Duan, J.; Deng, H.; Zhang, S. New insight into the confinement effect of microporous carbon in Li/Se battery chemistry: A cathode with enhanced conductivity. Small 2020, 2000266. [Google Scholar] [CrossRef]
  5. Qu, Y.; Guo, M.; Wang, X.; Yuan, C. Novel nitrogen-doped ordered mesoporous carbon as high-performance anode material for sodium-ion batteries. J. Alloys Compd. 2019, 791, 874–882. [Google Scholar] [CrossRef]
  6. Manthiram, A.; Knight, J.C.; Myung, S.-T.; Oh, S.-M.; Sun, Y.-K. Nickel-rich and lithium-rich layered oxide cathodes: Progress and perspectives. Adv. Energy Mater. 2016, 6, 1501010. [Google Scholar] [CrossRef]
  7. Chen, Z.; Xu, M.; Du, B.; Zhu, H.; Xie, T.; Wang, W. Morphology control of lithium iron phosphate nanoparticles by soluble starch-assisted hydrothermal synthesis. J. Power Sources 2014, 272, 837–844. [Google Scholar] [CrossRef]
  8. Liu, Q.; Zhu, H.; Liu, J.; Liao, X.; Tang, Z.; Zhou, C.; Yuan, M.; Duan, J.; Li, L.; Chen, Z. High-performance lithium-rich layered oxide material: Effects of preparation methods on microstructure and electrochemical properties. Materials 2020, 13, 334. [Google Scholar] [CrossRef] [Green Version]
  9. Liu, J.; Liu, Q.; Zhu, H.; Lin, F.; Ji, Y.; Li, B.; Duan, J.; Li, L.; Chen, Z. Effect of different composition on voltage attenuation of Li-rich cathode material for lithium-ion batteries. Materials 2020, 13, 40. [Google Scholar] [CrossRef] [Green Version]
  10. Ji, Y.; Zhou, C.; Lin, F.; Li, B.; Yang, F.; Zhu, H.; Duan, J.; Chen, Z. Submicron-sized Nb-doped lithium garnet for high ionic conductivity solid electrolyte and performance of quasi-solid-state lithium battery. Materials 2020, 13, 560. [Google Scholar] [CrossRef] [Green Version]
  11. Jung, S.-K.; Gwon, H.; Hong, J.; Park, K.-Y.; Seo, D.-H.; Kim, H.; Hyun, J.; Yang, W.; Kang, K. Understanding the degradation mechanisms of LiNi0.5Co0.2Mn0.3O2 cathode material in lithium ion batteries. Adv. Energy Mater. 2013, 4, 201300787. [Google Scholar]
  12. Su, Y.; Cui, S.; Zhuo, Z.; Yang, W.; Wang, X.; Pan, F. Enhancing the high-voltage cycling performance of LiNi0.5Co0.2Mn0.3O2 by retarding its interfacial reaction with an electrolyte by atomic-layer-deposited Al2O3. ACS Appl. Mater. Interfaces 2015, 7, 25105–25112. [Google Scholar] [CrossRef] [PubMed]
  13. Shi, Y.; Zhang, M.; Qian, D.; Meng, Y.S. Ultrathin Al2O3 coatings for improved cycling performance and thermal stability of LiNi0.5Co0.2Mn0.3O2 cathode material. Electrochim. Acta 2016, 203, 154–161. [Google Scholar] [CrossRef]
  14. Tao, T.; Chen, C.; Qi, W.; Liang, B.; Yao, Y.; Lu, S.-G. Antimony doped tin oxide-coated LiNi0.5Co0.2Mn0.3O2 cathode materials with enhanced electrochemical performance for lithium-ion batteries. J. Alloys Compd 2018, 765, 601–607. [Google Scholar] [CrossRef]
  15. Liu, K.; Yang, G.-L.; Dong, Y.; Shi, T.; Chen, L. Enhanced cycling stability and rate performance of Li[Ni0.5Co0.2Mn0.3]O2 by CeO2 coating at high cut-off voltage. J. Power Sources 2015, 281, 370–377. [Google Scholar] [CrossRef]
  16. Liu, T.; Zhao, S.-X.; Wang, K.; Nan, C.-W. CuO-coated Li[Ni0.5Co0.2Mn0.3]O2 cathode material with improved cycling performance at high rates. Electrochim. Acta 2012, 85, 605–611. [Google Scholar] [CrossRef]
  17. Yang, J.; Yu, Z.; Yang, B.; Liu, H.; Hao, J.; Yu, T.; Chen, K. Electrochemical characterization of Cr8O21 modified LiNi0.5Co0.2Mn0.3O2 cathode material. Electrochim. Acta 2018, 266, 342–347. [Google Scholar] [CrossRef]
  18. Wu, F.; Tian, J.; Su, Y.; Guan, Y.; Jin, Y.; Wang, Z.; He, T.; Bao, L.; Chen, S. Lithium-active molybdenum trioxide coated LiNi0.5Co0.2Mn0.3O2 cathode material with enhanced electrochemical properties for lithium-ion batteries. J. Power Sources 2014, 269, 747–754. [Google Scholar] [CrossRef]
  19. Chen, C.; Tao, T.; Qi, W.; Zeng, H.; Wu, Y.; Liang, B.; Yao, Y.; Lu, S.; Chen, Y. High-performance lithium ion batteries using SiO2-coated LiNi0.5Co0.2Mn0.3O2 microspheres as cathodes. J. Alloys Compd. 2017, 709, 708–716. [Google Scholar] [CrossRef]
  20. Liu, W.; Wang, M.; Gao, X.; Zhang, W.; Chen, J.; Zhou, H.; Zhang, X. Improvement of the high-temperature, high-voltage cycling performance of LiNi0.5Co0.2Mn0.3O2 cathode with TiO2 coating. J. Alloys Compd. 2012, 543, 181–188. [Google Scholar] [CrossRef]
  21. Kong, J.-Z.; Ren, C.; Tai, G.-A.; Zhang, X.; Li, A.-D.; Wu, D.; Li, H.; Zhou, F. Ultrathin ZnO coating for improved electrochemical performance of LiNi0.5Co0.2Mn0.3O2 cathode material. J. Power Sources 2014, 266, 433–439. [Google Scholar] [CrossRef]
  22. Schipper, F.; Bouzaglo, H.; Dixit, M.; Erickson, E.M.; Weigel, T.; Talianker, M.; Grinblat, J.; Burstein, L.; Schmidt, M.; Lampert, J.; et al. From Surface ZrO2 coating to bulk Zr doping by high temperature annealing of nickel-rich lithiated oxides and their enhanced electrochemical performance in lithium ion batteries. Adv. Energy Mater. 2018, 8, 1701682. [Google Scholar] [CrossRef]
  23. Yang, K.; Fan, L.-Z.; Guo, J.; Qu, X. Significant improvement of electrochemical properties of AlF3-coated LiNi0.5Co0.2Mn0.3O2 cathode materials. Electrochim. Acta 2012, 63, 363–368. [Google Scholar] [CrossRef]
  24. Li, L.; Zhang, Z.; Fu, S.; Liu, Z.; Liu, Y. Co-modification by LiAlO2-coating and Al-doping for LiNi0.5Co0.2Mn0.3O2 as a high-performance cathode material for lithium-ion batteries with a high cutoff voltage. J. Alloys Compd 2018, 768, 582–590. [Google Scholar] [CrossRef]
  25. Hu, W.; Zhang, C.; Jiang, H.; Zheng, M.; Wu, Q.-H.; Dong, Q. Improving the electrochemistry performance of layer LiNi0.5Co0.2Mn0.3O2 material at 4.5 V cutoff potential using lithium metaborate. Electrochim. Acta 2017, 243, 105–111. [Google Scholar] [CrossRef]
  26. Yang, X.; Wang, X.; Hu, L.; Zou, G.; Su, S.; Bai, Y.; Shu, H.; Wei, Q.; Hu, B.; Ge, L.; et al. Layered Li[Ni0.5Co0.2Mn0.3]O2–Li2MnO3 core–shell structured cathode material with excellent stability. J. Power Sources 2013, 242, 589–596. [Google Scholar] [CrossRef]
  27. Zhang, M.; Hu, G.; Liang, L.; Peng, Z.; Du, K.; Cao, Y. Improved cycling performance of Li2MoO4-inlaid LiNi0.5Co0.2Mn0.3O2 cathode materials for lithium-ion battery under high cutoff voltage. J. Alloys Compd. 2016, 673, 237–248. [Google Scholar] [CrossRef]
  28. Hu, G.; Zhang, M.; Wu, L.; Peng, Z.; Du, K.; Cao, Y. Effects of Li2SiO3 coating on the performance of LiNi0.5Co0.2Mn0.3O2 cathode material for lithium ion batteries. J. Alloys Compd. 2017, 690, 589–597. [Google Scholar] [CrossRef]
  29. Wang, J.; Yu, Y.; Li, B.; Fu, T.; Xie, D.; Cai, J.; Zhao, J. Improving the electrochemical properties of LiNi0.5Co0.2Mn0.3O2 at 4.6 V cutoff potential by surface coating with Li2TiO3 for lithium-ion batteries. Phys. Chem. Chem. Phys. 2015, 17, 32033–32043. [Google Scholar] [CrossRef]
  30. Huang, Y.; Jin, F.-M.; Chen, F.-J.; Chen, L. Improved cycle stability and high-rate capability of Li3VO4-coated Li[Ni0.5Co0.2Mn0.3]O2 cathode material under different voltages. J. Power Sources 2014, 256, 1–7. [Google Scholar] [CrossRef]
  31. Xu, Y.; Liu, Y.; Lu, Z.; Wang, H.; Sun, D.; Yang, G. The preparation and role of Li2ZrO3 surface coating LiNi0.5Co0.2Mn0.3O2 as cathode for lithium-ion batteries. App. Surf. Sci. 2016, 361, 150–156. [Google Scholar] [CrossRef]
  32. Song, H.; Yuan, H.; Chen, H.; Tang, A.; Xu, G.; Liu, L.; Zhang, Z.; Kuang, Q. Synthesis of TiO2/S@PPy composite for chemisorption of polysulfides in high performance Li-S batteries. J. Solid State Electr. 2020. [Google Scholar] [CrossRef]
  33. Wu, Z.; Han, X.; Zheng, J.; Wei, Y.; Qiao, R.; Shen, F.; Dai, J.; Hu, L.; Xu, K.; Lin, Y.; et al. Depolarized and fully active cathode based on Li(Ni0.5Co0.2Mn0.3)O2 embedded in carbon nanotube network for advanced batteries. Nano Lett. 2014, 14, 4700–4706. [Google Scholar] [CrossRef] [PubMed]
  34. Li, G.; Zhang, Z.; Wang, R.; Huang, Z.; Zuo, Z.; Zhou, H. Effect of trace Al surface doping on the structure, surface chemistry and low temperature performance of LiNi0.5Co0.2Mn0.3O2 cathode. Electrochim. Acta 2016, 212, 399–407. [Google Scholar] [CrossRef]
  35. Dixit, M.; Markovsky, B.; Aurbach, D.; Major, D.T. Unraveling the effects of Al doping on the electrochemical properties of LiNi0.5Co0.2Mn0.3O2 using first principles. J. Electrochem. Soc. 2017, 164, A6359–A6365. [Google Scholar] [CrossRef]
  36. Chen, Z.; Gong, X.; Zhu, H.; Cao, K.; Liu, Q.; Liu, J.; Li, L.; Duan, J. High performance and structural stability of K and Cl co-doped LiNi0.5Co0.2Mn0.3O2 cathode materials in 4.6 voltage. Front. Chem. 2018, 6, 643. [Google Scholar] [CrossRef]
  37. Li, Y.; Su, Q.; Han, Q.; Li, P.; Li, L.; Xu, C.; Cao, X.; Cao, G. Synthesis and characterization of Mo-doped LiNi0.5Co0.2Mn0.3O2 cathode materials prepared by a hydrothermal process. Ceram. Int. 2017, 43, 3483–3488. [Google Scholar] [CrossRef]
  38. Hua, W.; Zhang, J.; Zheng, Z.; Liu, W.; Peng, X.; Guo, X.D.; Zhong, B.; Wang, Y.J.; Wang, X. Na-doped Ni-rich LiNi0.5Co0.2Mn0.3O2 cathode material with both high rate capability and high tap density for lithium ion batteries. Dalton Trans. 2014, 43, 14824–14832. [Google Scholar] [CrossRef]
  39. Zhao, R.; Yang, Z.; Liang, J.; Lu, D.; Liang, C.; Guan, X.; Gao, A.; Chen, H. Understanding the role of Na-doping on Ni-rich layered oxide LiNi0.5Co0.2Mn0.3O2. J. Alloys Compd. 2016, 689, 318–325. [Google Scholar] [CrossRef]
  40. Mo, Y.; Guo, L.; Cao, B.; Wang, Y.; Zhang, L.; Jia, X.; Chen, Y. Correlating structural changes of the improved cyclability upon Nd-substitution in LiNi0.5Co0.2Mn0.3O2 cathode materials. Energy Storage Mater. 2019, 18, 260–268. [Google Scholar] [CrossRef]
  41. Zhang, Y.; Wang, Z.-B.; Lei, J.; Li, F.-F.; Wu, J.; Zhang, X.-G.; Yu, F.-D.; Ke, K. Investigation on performance of Li(Ni0.5Co0.2Mn0.3)1−xTixO2 cathode materials for lithium-ion battery. Ceram. Int. 2015, 41, 9069–9077. [Google Scholar] [CrossRef]
  42. Chen, Y.; Li, Y.; Li, W.; Cao, G.; Tang, S.; Su, Q.; Deng, S.; Guo, J. High-voltage electrochemical performance of LiNi0.5Co0.2Mn0.3O2 cathode material via the synergetic modification of the Zr/Ti elements. Electrochim. Acta 2018, 281, 48–59. [Google Scholar] [CrossRef]
  43. Zhao, X.; Liang, G.; Liu, H.; Liu, Y. Improved conductivity and electrochemical properties of LiNi0.5Co0.2Mn0.3O2 materials via yttrium doping. RSC Adv. 2018, 8, 4142–4152. [Google Scholar] [CrossRef] [Green Version]
  44. Wang, D.; Wang, Z.; Li, X.; Guo, H.; Xu, Y.; Fan, Y.; Pan, W. Effect of surface fluorine substitution on high voltage electrochemical performances of layered LiNi0.5Co0.2Mn0.3O2 cathode materials. Appl. Surf. Sci. 2016, 371, 172–179. [Google Scholar] [CrossRef]
  45. Zhang, Y.; Wang, Z.-B.; Yu, F.-D.; Que, L.-F.; Wang, M.-J.; Xia, Y.-F.; Xue, Y.; Wu, J. Studies on stability and capacity for long-life cycle performance of Li(Ni0.5Co0.2Mn0.3)O2 by Mo modification for lithium-ion battery. J. Power Sources 2017, 358, 1–12. [Google Scholar] [CrossRef]
  46. Qiao, Q.; Qin, L.; Li, G.; Wang, Y.; Gao, X. Sn-stabilized Li-rich layered Li(Li0.17Ni0.25Mn0.58)O2 oxide as a cathode for advanced lithium-ion batteries. J. Mater. Chem. A 2015, 3, 17627–17634. [Google Scholar] [CrossRef]
  47. Li, J.; He, X.; Zhao, R.; Wan, C.; Jiang, C.; Xia, D.; Zhang, S. Stannum doping of layered LiNi3/8Co2/8Mn3/8O2 cathode materials with high rate capability for Li-ion batteries. J. Power Sources 2006, 158, 524–528. [Google Scholar] [CrossRef]
  48. Ma, X.; Wang, C.; Cheng, J.; Sun, J. Effects of Sn doping on the structural and electrochemical properties of LiNi0.8Co0.2O2 cathode materials. Solid State Ion. 2007, 178, 125–129. [Google Scholar] [CrossRef]
  49. Song, L.; Li, A.; Xiao, Z.; Chi, Z.; Cao, Z.; Zhu, H. Enhanced electrochemical properties of Ni-rich LiNi0.8Co0.1Mn0.1O2 by SnO2 coating under high cutoff voltage. Ionics 2020. [Google Scholar] [CrossRef]
  50. Liu, X.; Kou, L.; Shi, T.; Liu, K.; Chen, L. Excellent high rate capability and high voltage cycling stability of Y2O3-coated LiNi0.5Co0.2Mn0.3O2. J. Power Sources 2014, 267, 874–880. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples (∗ represents Li2SnO3).
Figure 1. XRD patterns of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples (∗ represents Li2SnO3).
Nanomaterials 10 00868 g001
Figure 2. SEM images of MS1 (a,b), MS3 (c,d), MS5 (e,f), and the corresponding EDS mappings of image MS3 (c): (g)-O, (h)-Ni, (i)-Co, (j)-Mn, (k)-Sn.
Figure 2. SEM images of MS1 (a,b), MS3 (c,d), MS5 (e,f), and the corresponding EDS mappings of image MS3 (c): (g)-O, (h)-Ni, (i)-Co, (j)-Mn, (k)-Sn.
Nanomaterials 10 00868 g002
Figure 3. TEM images of MS3 (a) and the selected areas in Figure 3a (bd).
Figure 3. TEM images of MS3 (a) and the selected areas in Figure 3a (bd).
Nanomaterials 10 00868 g003
Figure 4. XPS spectra of the synthesized materials: Ni 2p3/2 (a) spectra of M523, Ni 2p3/2 (b), Sn3d (c) spectra of MS3, O1s (d) spectra of M523 and MS3.
Figure 4. XPS spectra of the synthesized materials: Ni 2p3/2 (a) spectra of M523, Ni 2p3/2 (b), Sn3d (c) spectra of MS3, O1s (d) spectra of M523 and MS3.
Nanomaterials 10 00868 g004
Figure 5. Electrochemical performance of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples: (a) Initial charge–discharge curves at 0.1C, (b) rate performance from 0.1 to 5C, (c) cyclic ability at 1C, and (d) electrochemical impedance spectroscopy (EIS) plots of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples after the 100th cycle.
Figure 5. Electrochemical performance of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples: (a) Initial charge–discharge curves at 0.1C, (b) rate performance from 0.1 to 5C, (c) cyclic ability at 1C, and (d) electrochemical impedance spectroscopy (EIS) plots of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples after the 100th cycle.
Nanomaterials 10 00868 g005
Table 1. Lattice constants of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples.
Table 1. Lattice constants of Sn-modified LiNi0.5Co0.2Mn0.3O2 samples.
Samplea (Å)c (Å)c/aR(I003/I104)R’((I006+ I102)/I101)V (Å3)
M5232.867314.21034.9561.3190.428101.18
MS12.869414.22864.9591.4600.474101.46
MS32.871014.22234.9541.4250.526101.53
MS52.871514.14994.9281.8210.508101.04
Table 2. Peak positions and mole fractions of the metal elements for M523 and MS3 samples obtained from XPS fittings.
Table 2. Peak positions and mole fractions of the metal elements for M523 and MS3 samples obtained from XPS fittings.
SampleElements
Ni2+Ni3+Co3+Mn4+Sn4+
Peak position/eVM523853.6854.7856.2779.8642.4--
MS3853.6854.7856.2779.8642.4486.4
Mole fraction/%M52372.2727.73100.0100.0--
MS374.8825.12100.0100.0100.0

Share and Cite

MDPI and ACS Style

Zhu, H.; Shen, R.; Tang, Y.; Yan, X.; Liu, J.; Song, L.; Fan, Z.; Zheng, S.; Chen, Z. Sn-Doping and Li2SnO3 Nano-Coating Layer Co-Modified LiNi0.5Co0.2Mn0.3O2 with Improved Cycle Stability at 4.6 V Cut-off Voltage. Nanomaterials 2020, 10, 868. https://doi.org/10.3390/nano10050868

AMA Style

Zhu H, Shen R, Tang Y, Yan X, Liu J, Song L, Fan Z, Zheng S, Chen Z. Sn-Doping and Li2SnO3 Nano-Coating Layer Co-Modified LiNi0.5Co0.2Mn0.3O2 with Improved Cycle Stability at 4.6 V Cut-off Voltage. Nanomaterials. 2020; 10(5):868. https://doi.org/10.3390/nano10050868

Chicago/Turabian Style

Zhu, Huali, Rui Shen, Yiwei Tang, Xiaoyan Yan, Jun Liu, Liubin Song, Zhiqiang Fan, Shilin Zheng, and Zhaoyong Chen. 2020. "Sn-Doping and Li2SnO3 Nano-Coating Layer Co-Modified LiNi0.5Co0.2Mn0.3O2 with Improved Cycle Stability at 4.6 V Cut-off Voltage" Nanomaterials 10, no. 5: 868. https://doi.org/10.3390/nano10050868

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