Next Article in Journal
Dissipation Dynamic, Residue Distribution and Risk Assessment of Emamectin Benzoate in Longan by High-Performance Liquid Chromatography with Fluorescence Detection
Next Article in Special Issue
Research Progress of Hydrogen Production Technology and Related Catalysts by Electrolysis of Water
Previous Article in Journal
Optimization of Extraction Process and Dynamic Changes in Triterpenoids of Lactuca indica from Different Medicinal Parts and Growth Periods
Previous Article in Special Issue
The Progress of Hard Carbon as an Anode Material in Sodium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Outstanding Electrochemical Performance of Ni-Rich Concentration-Gradient Cathode Material LiNi0.9Co0.083Mn0.017O2 for Lithium-Ion Batteries

1
Guangxi Key Laboratory of Electrochemical and Magneto-Chemical Functional Materials, College of Chemistry and Bioengineering, Guilin University of Technology, Guilin 541004, China
2
College of Mechanical and Vehicle Engineering, Hunan University, Changsha 410082, China
3
School of Automobile Engineering, Guilin University of Aerospace Technology, Guilin 541004, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(8), 3347; https://doi.org/10.3390/molecules28083347
Submission received: 22 March 2023 / Revised: 8 April 2023 / Accepted: 9 April 2023 / Published: 10 April 2023
(This article belongs to the Special Issue Novel Electrode Materials for Rechargeable Batteries)

Abstract

:
The full-concentrationgradient LiNi0.9Co0.083Mn0.017O2 (CG-LNCM), consisting of core Ni-rich LiNi0.93Co0.07O2, transition zone LiNi1−x−yCoxMnyO2, and outmost shell LiNi1/3Co1/3Mn1/3O2 was prepared by a facile co-precipitation method and high-temperature calcination. CG-LNCM was then investigated with an X-ray diffractometer, ascanning electron microscope, a transmission electron microscope, and electrochemical measurements. The results demonstrate that CG-LNCM has a lower cation mixing of Li+ and Ni2+ and larger Li+ diffusion coefficients than concentration-constant LiNi0.9Co0.083Mn0.017O2 (CC-LNCM). CG-LNCM presents a higher capacity and a better rate of capability and cyclability than CC-LNCM. CG-LNCM and CC-LNCM show initial discharge capacities of 221.2 and 212.5 mAh g−1 at 0.2C (40 mA g−1) with corresponding residual discharge capacities of 177.3 and 156.1 mAh g−1 after 80 cycles, respectively. Even at high current rates of 2C and 5C, CG-LNCM exhibits high discharge capacities of 165.1 and 149.1 mAh g−1 after 100 cycles, respectively, while the residual discharge capacities of CC-LNCM are as low as 148.8 and 117.9 mAh g−1 at 2C and 5C after 100 cycles, respectively. The significantly improved electrochemical performance of CG-LNCM is attributed to its concentration-gradient microstructure and the composition distribution of concentration-gradient LiNi0.9Co0.083Mn0.017O2. The special concentration-gradient design and the facile synthesis are favorable for massive manufacturing of high-performance Ni-rich ternary cathode materials for lithium-ion batteries.

Graphical Abstract

1. Introduction

Compared with other secondary batteries such as lead acid, nickel–cadmium and nickel–metal hydride batteries, lithium-ion batteries (LIBs) have been widely used in portable electronic devices and electric vehicles (EVs) because of their higher energy density and longer life span [1,2,3,4,5,6]. With the increasing demand for LIBs with a high energy density and outstanding cyclability, it is urgent and imperative to develop high-energy density and cyclability of cathode materials for LIBs. Numerous efforts have been made to optimize cathode materials, and dozens of these have been well developed [7], but the Ni-based ternary cathode materials LiNi1−x−yCoxMnyO2 (NCM) and LiNi1−x−yCoxAlyO2 (NCA) have received extensive and intensive attention and have been extensively utilized [8,9,10,11,12,13]. Ni-rich ternary cathode materials supply a large capacity that rises proportionately with Ni content, even though they suffer from inferior cyclic stability during cycling and thermal instability [14,15,16,17,18]. These shortcomings mainly result from the microcracking of electroactive particles upon the charging/discharging processes, which tremendously enlarges the inner surface exposed to electrolyte attack [15,19].
To improve the performance of Ni-rich ternary cathode materials, diverse approaches such as doping heterogeneous ions [20,21,22], surface coating [23,24,25], the construction of a single crystal phase [26,27,28], core/shell microstructure [29,30] and particles with concentration-gradient composition [31,32,33] have been adopted to inhibit the formation of microcracking. Among the preceding approaches, building particles with concentration-gradient composition is one of the most effective ways to improve the cyclability of electroactive materials. Generally, core composition can provide high capacity, and shell composition can improve cyclability and thermal stability in the ideal concentration-gradient microstructure [34]. In the present investigation, full concentration-gradient LiNi0.9Co0.087Mn0.013O2, in which the composition varies gradiently from core composition LiNi0.93Co0.07O2 to outmost shell composition LiNi1/3Co1/3Mn1/3O2, was proposed and prepared by a facile co-precipitation method combined with high-temperature calcination in an oxygen atmosphere. Benefiting from the high capacity provided by Ni-rich core composition and the excellent cyclability of shell composition LiNi1/3Co1/3Mn1/3O2, the concentration-gradient LiNi0.9Co0.087Mn0.013O2 exhibits significantly improved electrochemical performance in comparison with the concentration-constant LiNi0.9Co0.087Mn0.013O2 prepared by the same synthesis procedure.

2. Results and Discussions

Figure 1 shows the X-ray diffraction (XRD) patterns of concentration-gradient precursors (CG-NCMOH), concentration-constant precursors (CC-NCMOH), the concentration-gradient products (CG-LNCM), and the concentration-constant products (CC-LNCM). Figure 1a and the Rietveld refinement results (Figure S1a,b in Supplementary File) demonstrate that no manganese and cobalt hydroxides or other oxides exist in the XRD patterns of hydroxide precursors CC-NCMOH and CG-NCMOH, indicating that Mn2+ and Co2+ are successfully doped into Ni sites of Ni(OH)2 to form Ni-Co-Mn ternary hydroxide. The diffraction peaks of CC-NCMOH and CG-NCMOH shift to a lower angle, which implies that doping of Mn2+ and Co2+ enlarges the layer distance of Ni(OH)2 according to the Bragg equation 2dsin θ = nλ. As observed in Figure 1b, the diffraction peaks of CC-LNCM and CG-LNCM are strong and sharp, suggesting that both CC-LNCM and CG-LNCM have high crystallinity, and all the diffractions can be well indexed by the R3m space group. The structural parameters for CC-LNCM and CG-LNCM are listed in Tables S1 and S2, and the results reveal that CG-LNCM has a larger cell volume than CC-LNCM, hinting that CG-LNCM may have better electrochemical performance in comparison with CC-LNCM because a larger cell volume favors the more rapid transport of Li+ in electroactive particles and hence the better electrochemical performance. The layered structure of the material could be judged by the splitting of characteristic peaks. The more obvious the splits of (006)/(102) and (018)/(110) are, the more they indicate a higher degree of layered structure [35]. In addition, the ratio of c/a for both CC-LNCM and CG-LNCM is bigger than 4.9, showing that both samples have a well-developed layered crystal structure [36,37]. Furthermore, the quota of intensity ratio of (003) to (104), I(003)/I(104) reflects the cation mixing of Li+ and Ni2+, and the larger value of I(003)/I(104) means the lower cation mixing. It was reported that if the value of I(003)/I(104) is bigger than 1.2, the materials will have lower cation mixing [38,39] and improved electrochemical performance. According to Tables S1 and S2, the comparisons of CC-LNCM and CG-LNCM in I(003)/I(104) and the occupation of Li and Ni in Wyckoff sites 3a and 3b reveal that CG-LNCM has a lower cation mixing of Li+ and Ni2+ than CC-LNCM, and the special concentration-gradient microstructure may be responsible for the lower cation mixing.
The scanning electron microscope (SEM) images of CC-NCMOH, CG-NCMOH, CC-LNCM and CG-LNCM are presented in Figure 2. As observed in Figure 2a,b, both two hydroxide precursors display sphere-like morphology and a rough surface and are composed of nanoplates agglomerated loosely together. The loose agglomeration of CC-NCMOH and CG-NCMOH is beneficial for rapid reaction with LiOH to form LiNi0.9Co0.083Mn0.017O2 with high crystallinity. Figure 2c,d demonstrate that the final products CC-LNCM and CG-LNCM present a similar sphere-like morphology to that of hydroxide precursors. However, unlike the precursors CC-NCMOH and CG-NCMOH, CC-LNCM and CG-LNCM consist of nanoparticles. Compared with the compact CC-LNCM, the CG-LNCM particles are loosely agglomerated, which is favorable for the permeation of the electrolyte and rapid diffusion of Li+ in the electroactive particles. The high-resolution transmission electron microscope (HRTEM) images of CC-LNCM and CG-LNCM are shown in Figure 2e,f, respectively, and d-spacing of about 0.2040 and 0.2042 nm is observed in the lattice fringes of CC-LNCM and CG-LNCM, respectively, corresponding to the lattice plane (104). The minor difference of the d-space may result from the different surface composition of LiNi0.9Co0.083Mn0.017O2 and LiNi1/3Co1/3Mn1/3O2 for CC-LNCM and CG-LNCM, respectively.
The SEM image of cross-section of CG-LNCM presented in Figure 3a further clearly demonstrates that the sphere-like secondary particles are composed of primary nanoparticles, and the corresponding EDX elemental mappings in Figure 3b–d indicate that the elements Ni, Co and Mn are evenly dispersive, similar to the concentration-constant LiNi0.8Co0.1Mn0.1O2 [36], implying that it is difficult to separate concentration-gradient samples from concentration-constant samples by elemental mappings. Figure 3e,f display the SEM images of CC-LNCM and CG-LNCM primary particles, respectively. The element contents of the selected area of particles are determined by energy dispersive X-ray spectroscopy (EDXS), and EDXS mappings of the selected area are shown in Figure S2. It can be observed that contents of Ni, Co, and Mn in the edge and interior of the primary particle are almost identical and close to the molar ratio of Ni, Co and Mn of the concentration-constant LiNi0.9Co0.083Mn0.017O2. The compositions are different in the different areas of the CG-LNCM, further confirming that composition varies gradiently from interior to edge and CG-LNCM is truly a concentration-gradient oxide.
To learn the effects of the concentration-gradient and concentration-constant microstructure on the oxidate state of Ni in LiNi0.9Co0.083Mn0.017O2, an X-ray photoelectron spectroscopy (XPS) measurement was conducted and the corresponding Ni 2p XPS spectra were shown in Figure 4. As can be observed in Figure 4a,b, the XPS spectra of both CC-LNCM and CG-LNCM consist of two satellite peaks and two main peaks, and the two peaks centered at about 855.96 and 873.5 eV are deconvoluted into 856.1 and 854.9 eV, and 873.9 and 872.4 eV, respectively. The Binding energy of 854.9 and 872.4 eV correspond to Ni 2p3/2 and Ni 2p1/2 of Ni(II) [40,41], respectively, and the binding energy of 856.1 and 873.9 eV match Ni 2p3/2 and Ni 2p1/2 of Ni(III) [40,42], respectively. The molar ratio of Ni2+/Ni3+ is calculated according to the ratio of the closed area of Ni 2p3/2 XPS spectra of Ni2+ to the closed area of Ni 2p3/2 XPS spectra of Ni3+. The molar ratios of Ni2+/Ni3+ are 0.48/0.52 and 0.58/0.42 for CC-LNCM and CG-LNCM, respectively. The different content of Ni2+ results from the different surface compositions of CC-LNCM and CG-LNCM, similar to the previous report that the molar ratio of Ni2+/Ni3+ is larger on the surface of LiNi1−x−yCoxMnyO2 with a lower content of Ni [43]. This finding further confirms the concentration-gradient microstructure of CG-LNCM, in which the composition varies from the core LiNi0.93Co0.07O2 to the outmost shell LiNi1/3Co1/3Mn1/3O2.
To investigate the effect of composition distribution of LiNi0.9Co0.083Mn0.017O2 on the electrochemical mechanism, cyclic voltammetry measurements were conducted on CC-LNCM and CG-LNCM at a scan rate of 0.1 mV s−1 in the potential scope of 3.0–4.3 V (vs. Li+/Li) at room temperature, and the first three cycles of cyclic voltammograms (CVs) of CC-LNCM and CG-LNCM electrodes are shown in Figure 5. The shape and the closed area of the first three CVs are almost unchanged, suggesting that both CC-LNCM and CG-LNCM have the better cyclability. As seen in Figure 5a,b, two CVs exhibit a similar shape, and three couples of redox peaks have close potential, indicating that CC-LNCM and CG-LNCM possess the identical electrochemical mechanism upon the charging and discharging processes. The three redox peaks are associated with the interconversion of the hexagonal phase to the monoclinic phase (H1↔M), the monoclinic phase to the hexagonal phase (M↔H2), and the hexagonal phase to the hexagonal phase (H2↔H3), respectively [44,45]. The accurate comparison of redox peak potentials of CC-LNCM and CG-LNCM demonstrates that CG-LNCM has lower oxidation peak potentials and larger reduction peak potentials than CC-LNCM, implying that CG-LNCM exhibits lower electrochemical polarization and better electrochemical reaction reversibility than CC-LNCM.
The electrodes used to evaluate galvanostatic electrochemical performance at each given current rate were fresh CC-LNCM or CG-LNCM electrodes without an activation process. The initial charge/discharge profiles of CC-LNCM and CG-LNCM electrodes at various current rates (1C = 200 mA g−1) in the voltage range of 3.0–4.3 V (vs. Li+/Li) were shown in Figure 6a,b, and the big difference indicated between charge and discharge capacities is observed. The low initial Coulombic efficiency of fresh CC-LNCM and CG-LNCM electrodes, similar to that of LiNi0.9Co0.08Al0.02O2 [45], may be associated with properties of Ni-rich based oxide cathode materials. Low initial Coulombic efficiency is presented in rich Ni-based oxide cathode materials, especially nanoparticles. After several cycles, the Coulombic efficiency greatly increases, as observed in Figure S3, to about 97.8% in the fifth cycle, implying that the activation of electrodes can remarkably improve the Coulombic efficiency. It can be found that the discharge and charge capacities decrease with the increase of current rates due to a larger polarization at higher current rates. The initial discharge capacities of CC-LNCM are 212.5, 200.7, 183.9, 167.1 and 139.2 mAh g−1 at 0.2C, 0.5C, 1C, 2C and 5C, respectively, while CG-LNCM presents the somewhat higher initial discharge capacities of 221.2, 203.3, 185.0, 168.9 and 140.8 mAh g−1 at 0.2C, 0.5C, 1C, 2C and 5C, respectively. The initial discharge capacity of CC-LNCM at 0.2C is close to that of 210 mAh g−1 for LiNi0.9Co0.05Mn0.05O2 at 0.2C [37], higher than that of 203.8 mAh g−1 for LiNi0.91Co0.06Mn0.03O2 single crystal at 0.1C [28], 207 mAh g−1 for LiNi0.9Co0.08Al0.02O2 at 0.2C [46]. The initial discharge capacity of CG-LNCM at 0.2C is close to that of about 221 mAh g−1 for concentration-gradient LiNi0.84Co0.06Mn0.09Al0.01O2 at 0.1C [32], and larger than that of 200 mAh g−1 for concentration-gradient LiNi0.9Mn0.1O2 at 0.1C [47]. The initial discharge capacities of CG-LNCM at 0.5C, 1C and 2C are close to those of 200 mAh g−1 for LiNi0.92Co0.03Mn0.03Al0.02O2 at 0.5C [48], 180 mAh g−1 for TiO2-coated LiNi0.9Co0.08Al0.02O2 at 1C [47] and 165 mAh g−1 for TiO2-coated LiNi0.9Co0.08Al0.02O2 at 2C [45], respectively. The above-mentioned results demonstrate that CG-LNCM is competitive in the first discharge capacity with other Ni-based ternary cathode materials.
The cycle performance of CC-LNCM and CG-LNCM at 0.2C, 2C and 5C are depicted in Figure 6c,d. As observed in Figure 6c,d, discharge capacities of CC-LNCM and CG-LNCM at all current rates increase before the first number of specific cycles, which is ascribed to the activation process that originates from the insufficient contact between electrolyte and electroactive particles. After activation, the discharge capacity of CC-LNCM reaches the highest value of 218.3 mAh g−1 at 0.2C and decreases to 156.1 mAh g−1 after 80 cycles. The ratio of the residual discharge capacity to the highest discharge capacity corresponds to 71.4%. While CG-LNCM exhibits the highest capacity of 223.4 mAh g−1 in the third cycle and higher residual capacity of 177.3 mAh g−1 after 80 cycles at 0.2C, the ratio of the residual discharge capacity to the highest discharge capacity corresponds to 79.4%. The residual capacity of CG-LNCM at 0.2C is close to that of about 174 mAh g−1 for LiNi0.9Co0.08Al0.02O2 [46], about 170 mAh g−1 for TiO2-coated LiNi0.9Co0.08Al0.02O2 [45], and 169 mAh g−1 for LiTaO3 modified LiNi0.9Co0.06Mn0.04O2 [49] after 80 cycles at 0.2C. The residual discharge capacities of CC-LNCM are 148.8 and 117.9 mAh g−1 at 2C and 5C after 100 cycles, respectively, which are much smaller than the corresponding residual discharge capacities of CG-LNCM. The special concentration-gradient microstructure of CG-LNCM may be responsible for the improved electrochemical performance. The residual discharge capacity of CG-LNCM at 2C is 165.1 mAh g−1 after 100 cycles, which is larger than that of 130 mAh g−1 for TiO2-coated LiNi0.9Co0.08Al0.02O2 [45], 134.4 mAh g−1 for LiNi0.9Co0.05Mn0.05O2 [50], and 156.9 mAh g−1 for Li2SiO3 coated LiNi0.9Co0.05Mn0.05O2 [50] at 2C after 100 cycles. Even at the high current rate of 5C, CG-LNCM presents the high discharge capacity of 149.1 mAh g−1 after 100 cycles, significantly higher than that of 117.9 mAh g−1 for CC-LNCM. It is noted that both CC-LNCM and CG-LNCM show better cycle performance at higher current rates than the cycle performance of these two cathodes at 0.2C, and this phenomenon is similar to that of LiNi0.8Co0.1Mn0.1O2 [51]. This phenomenon may be ascribed to the lower tolerance of great deformation of CC-LNCM and CG-LNCM with loose microstructure and nanosized primary particles. Compared with cycles at higher current rates, extraction of more Li+ from electroactive material LiNi0.9Co0.083Mn0.017O2 and insertion of more Li+ into delithiated LiNi0.9Co0.083Mn0.017O2 occur at lower current rates during charging and discharging processes, resulting in huge deformation, structure instability and the resultant inferior cyclability at lower current rates. In summary, the aforementioned electrochemical tests demonstrate that CG-LNCM has a higher capacity and better rate capability and cyclability than CC-LNCM, which can be attributed to the advantages of the concentration-gradient microstructure that the Ni-rich core LiNi0.93Co0.07O2 provides. Furthermore, the high capacity and shell composition of LiNi1/3Co1/3Mn1/3O2 supplies excellent structural stability, and the multiple core/shell structure of concentration-gradient composition is favorable for alleviation of microcracks upon cycling.
To further understand the difference in electrochemical properties of concentration-gradient and concentration-constant LiNi0.9Co0.083Mn0.017O2, electrochemical impedance spectroscopy (EIS) was carried out on the fresh CG-LNCM and CC-LNCM electrodes and other electrodes that had been cycled for 100 cycles at 5 C, and the corresponding Nyquist plots are shown in Figure 7. As observed in Figure 7a, the Nyquist plots of fresh electrodes are composed of a pressed semicircle and a sloped line, which is different from the Nyquist plots (Figure 7b) of cycled electrodes that consist of two compressed semicircles and an inclined line. The equivalent circuit models for the different Nyquist plots are depicted in the insert of Figure 7a,b, respectively. In equivalent circuit models, Re, Rf and Rct represent internal resistance of cell, charge transfer resistance, and resistance of the solid electrolyte interface (SEI) [52,53], respectively, while CPE and Wo stand for double layer capacitance and capacity of the surface layer, and Warburg impedance, respectively. The fitting results are listed in Table S3 and indicate that values of Rf and Rct of CC-LNCM electrodes are larger than those of CG-LNCM electrodes at the corresponding states. The smaller Rct of CG-LNCM electrodes suggests that CG-LNCM possesses better electrochemical kinetics and hence better electrochemical performance in comparison with CC-LNCM. In addition, the smaller Rf of cycled CG-LNCM electrodes demonstrates that the resistance of SEI of a cycled CG-LNCM electrode is smaller than that of a cycled CG-LNCM electrode, implying that CG-LNCM has a lower polarization than CC-LNCM during cycling, favorable for improvement of cyclability.
It is well-known that the diffusion of Li+ in electroactive particles is a rate-determining step of the electrochemical reaction of electroactive materials, and the Li+ diffusion coefficient, DLi, of electrode materials is a key parameter to evaluate the kinetics of electrochemical reaction. DLi was calculated according to the following Equation (1) [54].
D L i = 4 π τ ( m B V m M B S ) 2 ( Δ E s Δ E τ ) 2
The meanings of symbols in Equation (1) are the same as in our previous report, and the detailed calculation procedure is also akin to our previous report [55]. To compare DLi of CC-LNCM and CG-LNCM, the galvanostatic intermittent titration technique (GITT) measurement was carried out and the value of DLi was estimated by Equation (1). The GITT curves of CC-LNCM and CG-LNCM are shown in Figure 8a,b. The calculated values of DLi of CC-LNCM and CG-LNCM at charged and discharged states are depicted in Figure 8c,d, demonstrating that CG-LNCM has higher values of DLi in both charged and discharged states than CC-LNCM. The differences of DLi result from the discrepancies of CG-LNCM and CC-LNCM. In the charging process, the values of DLi of CG-LNCM vary from 3.73 × 10−11 to 6.14 × 10−10 cm2 s−1, while DLi of CC-LNCM lies in the scope of 1.00 × 10−11 to 1.96 × 10−10 cm2 s−1. During discharging, the CG-LNCM and CC-LNCM range in DLi from 1.20 × 10−12 to 5.34 × 10−10 and 4.15 × 10−13 to 2.16 × 10−10 cm2 s−1, respectively. The comparison of DLi in CG-NCM and CC-NCM suggests that the electrochemical reaction rate of CG-LNCM is faster than that of CC-LNCM, which can explain why CG-LNCM has a better electrochemical performance than CC-LNCM. The larger diffusion coefficients of Li+ benefit from the concentration-gradient composition and microstructure, and result in the significantly improved electrochemical performance of LiNi0.9Co0.083Mn0.017O2.

3. Materials and Methods

3.1. Synthesis of Materials

The concentration-gradient precursor Ni0.9Co0.083Mn0.017(OH)2 was synthesized by a facile co-precipitation method using NiSO4·6H2O, CoSO4·7H2O, MnSO4·H2O, NaOH and NH3·H2O as raw materials, and the corresponding schematic illustration is depicted in Figure 9. To avoid the oxidation of Ni2+, Co2+ and Mn2+ during the synthesis of precursor NixCoyMn1−x−y(OH)2 precipitate, both solution and reaction are in a nitrogen atmosphere. The total concentration of transition metal ions for both the aqueous solution A (tank 1) and B (tank 2) is 0.075 and 1.425 mol L−1, respectively, and the molar ratios of Mn2+:Co2+:Ni2+ for solution A and B are 1:1:1 and 0:0.07:0.93, respectively. Furthermore, solution A is equal to solution B in volume. At the start of the co-precipitation procedure, Ni-rich solution (tank 2) was firstly fed at a constant rate of 100 mL h−1 into the reactor containing a certain amount of distilled water and NH3·H2O at 50 °C. Simultaneously, the solution A (tank 1) was pumped into tank 2 at a constant rate of 50 mL h−1. At the same time, the mixture solution containing 1.66 mol L−1 NaOH and adequate quantity of NH3·H2O was added to the reactor at a reasonable rate for adjusting pH of the reaction solution to 11.5~11.8, which is suitable for formation of Ni1−x−yCoxMny(OH)2 precipitate. With the continuous addition of solution A to solution B, the molar ratio of Mn2+:Co2+:Ni2+ of solution B changes gradiently from 0:0.07:0.93 to 1:1:1; thus the concentration distribution of Ni, Co and Mn of the target precursor Ni1−x−yCoxMny(OH)2 precipitate will be gradient. The final concentration-gradient precipitate Ni1−x−yCoxMny(OH)2 consists of a Ni0.93Co0.07(OH)2 core, a transitional zone containing a series of Ni1−x−yCoxMny(OH)2 (0.07 < x < 1/3, 0 < y < 1/3) precipitate and a Ni1/3Co1/3Mn1/3(OH)2 shell. After complete addition of solutions A and B, the precipitate was aged for 15 h. Subsequently, the suspension was filtered and washed three times to get the dark green precipitate. Finally, the dark green precipitate was dried for 24 h at 110 °C in a vacuum oven. The molar ratio of total transition metal ions for solution A and solution B was 5:95 (0.075:1.425), so the chemical formula of the concentration-gradient precipitate can be simply expressed as Ni0.9Co0.083Mn0.017(OH)2 (CG-NCMOH).
The concentration-gradient LiNi0.9Co0.083Mn0.017O2 (CG-LNCM) was prepared by calcination of the mixture of LiOH·H2O and the dried concentration-gradient precursor Ni0.9Co0.083Mn0.017(OH)2 with a molar ratio of 1.04:1 under an oxygen atmosphere in a tube furnace in which the mixed reactants were heated with an adequate rate of 5 °C per minute to 480 °C and maintained for 5 h, and then subsequently heated to 750 °C and held for 13.5 h, and finally cooled naturally to room temperature to get concentration-gradient LiNi0.9Co0.083Mn0.017O2. For comparison, the control concentration-constant LiNi0.9Co0.083Mn0.017O2 (CC-LNCM) was prepared by the same calcination of the stochiometric mixture of concentration-constant precursor Ni0.9Co0.083Mn0.017(OH)2 (CC-NCMOH) and LiOH·H2O with a molar ratio of 1:1.04 under oxygen atmosphere, and the concentration-constant precursor Ni0.9Co0.083Mn0.017(OH)2 was prepared by the same co-precipitation method without tank 1, and the molar ratio of Ni2+, Co2+ and Mn2+ is fixed to 0.9:0.083:0.017.

3.2. Material Characterizations

The phases of the concentration-gradient and concentration-constant samples were investigated by X-ray diffraction (XRD, PANalytical, X’Pert3 powder) using Cu kα radiation in the 2θ range of 10–80°. The morphologies of the prepared materials were observed by scanning with an electron microscope (SEM, HITACHI, SU5000) and a high-resolution transmission electron microscope (HRTEM, JEOL, JEM-2100F). X-ray photoelectron spectroscopy (XPS, Ulvac-Phi, PHI 5000 VersaProbe III) with monochromatic Al kα radiation was applied to determine the oxidation state of the elements in the prepared materials.

3.3. Electrochemical Properties Characterizations

The 2016-type coin cell was used to evaluate the electrochemical performance of concentration-gradient and concentration-constant LiNi0.9Co0.083Mn0.017O2 electrodes by electrochemical tests. The coin cell consists of a lithium foil-counter electrode, a working electrode, a Cellguard 2500 film separator and an electrolyte of 1 mol/L LiPF6 in the mixed solvent of ethylene carbonate (EC), ethyl methyl carbonate (EMC) and diethyl carbonate (DEC) with a volume ratio of 4:2:4. The dried working electrodes are composed of 80 wt.% of electroactive material CG-LNCM or CC-LNCM, 10 wt.% of acetylene black and 10 wt.% of polyvinylidene fluoride, and the loading of electroactive materials is about 1.8 mg cm−2. The galvanostatic charge/discharge tests were carried out in the voltage range of 3.0 to 4.3 V (vs. Li+/Li) at 25 °C. The cyclic voltammetry (CV) was conducted on a CHI660E electrochemical workstation at 0.1 mV s−1 in the potential range of 3.0 to 4.3 V. To compare the diffusion coefficients of Li+, DLi, of CG-LNCM and CC-LNCM, the galvanostatic intermittent titration technique (GITT) was conducted on the electrodes cycled at 0.1C for three times. The GITT measurement was performed at a pulse current of 20 mA g−1 for 10 min, followed by a relaxation of 30 min.

4. Conclusions

The full concentration-gradient LiNi0.9Co0.083Mn0.017O2, in which composition varies from the Ni-rich core LiNi0.93Co0.07O2 to the outmost shell LiNi1/3Co1/3Mn1/3O2, was prepared by a facile co-precipitation method combined with high-temperature calcination under an inert atmosphere. Benefiting from the special functions of a concentration-gradient microstructure for which the Ni-rich core provides a high capacity, the shell supplies the excellent structural stability of the surface, and concentration-gradient distribution of compositions alleviates the formation of microcrack. Moreover, concentration-gradient LiNi0.9Co0.083Mn0.017O2 possesses a higher capacity and a better rate capability and cyclability in comparison with the concentration-constant LiNi0.9Co0.083Mn0.017O2 prepared by the same method. The combination of the concentration-gradient design and the rapid co-precipitation synthesis may provide an effective strategy for large-scale production of Ni-based ternary cathode materials with high-performance for lithium-ion batteries.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28083347/s1, Figure S1: Refine XRD patterns of (a) CC-NCMOH, (b) CG-NCMOH, (c) CC-LNCM and (d) CG-LNCM; Figure S2: The EDX graph of (a) area A and (b) area B of CC-LNCM, (c) area C and (d) area D of CG-LNCM; Figure S3: Activated charge/discharge profiles of (a) CC-LNCM and (b) CG-LNCM at various current rates; Table S1: Cell parameters of as-prepared samples; Table S2: Atomic parameters of as-prepared samples; Table S3: The fitting results of fresh electrodes and electrodes after cycled for 100 cycles at 5C.

Author Contributions

H.L.: Conceptualization, methodology, validation, formal analysis, writing-original draft, and editing; Y.G.: Conceptualization, methodology, software, investigation.; Y.C.: Supervision, conceptualization, methodology, data curation, formal analysis, writing-review and editing; N.G.: Conceptualization, methodology, software, investigation, data curation; R.S.: Conceptualization, methodology, software, investigation; Y.L.: Conceptualization, methodology, software, investigation; Q.C.: Supervision, project administration, funding acquisition, conceptualization, review, and editing. 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 (Grant No. 52164027) and the Natural Science Foundation of Guangxi Province (Grant No. 2021GXNSFAA220063).

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.

Sample Availability

Not applicable.

References

  1. Tarascon, J.M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359. [Google Scholar] [CrossRef] [PubMed]
  2. Palacin, M.R. Recent advances in rec hargeable battery materials: A chemist’s perspective. Chem. Soc. Rev. 2009, 38, 2565–2575. [Google Scholar] [CrossRef]
  3. Sun, Y.-K.; Myung, S.-T.; Park, B.-C.; Prakash, J.; Belharouak, I.; Amine, K. High-energy cathode material for long-life and safe lithium batteries. Nat. Mater. 2009, 8, 320. [Google Scholar] [CrossRef] [PubMed]
  4. Scrosati, B.; Hassoun, J.; Sun, Y.-K. Lithium-ion batteries. A look into the future. Energy Environ. Sci. 2011, 4, 3287–3295. [Google Scholar] [CrossRef]
  5. Zhao, J.; Hong, M.; Ju, Z.; Yan, X.; Gai, Y.; Liang, Z. Durable lithium metal anodes enabled by interfacial layers based on mechanically interlocked networks capable of energy dissipation. Angew. Chem. Int. Ed. Engl. 2022, 61, e202214386. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, S.J.; Yao, N.; Jiang, F.N.; Xie, J.; Sun, S.Y.; Chen, X.; Yuan, H.; Cheng, X.B.; Huang, J.Q.; Zhang, Q. Thermally stable polymer-rich solid electrolyte interphase for safe lithium metal pouch cells. Angew. Chem. Int. Ed. Engl. 2022, 61, e202214545. [Google Scholar] [CrossRef]
  7. Miao, Y.; Hynan, P.; von Jouanne, A.; Yokochi, A. Current Li-ion battery technologies in electric vehicles and opportunities for advancements. Energies 2019, 12, 1074. [Google Scholar] [CrossRef] [Green Version]
  8. Shao, Y.; Huang, B.; Liu, Q.; Liao, S. Preparation and modification of Ni-Co-Mn ternary cathode materials. Prog. Chem. 2018, 30, 410–419. [Google Scholar] [CrossRef]
  9. Bai, X.; Ban, L.; Zhuang, W. Research progress on coating and doping modification of nickel rich ternary cathode materials. J. Inorg. Mater. 2020, 35, 972–986. [Google Scholar] [CrossRef] [Green Version]
  10. Liu, L.; Li, M.; Chu, L.; Jiang, B.; Lin, R.; Zhu, X.; Cao, G. Layered ternary metal oxides: Performance degradation mechanisms as cathodes, and design strategies for high-performance batteries. Prog. Mater. Sci. 2020, 111, 100655. [Google Scholar] [CrossRef]
  11. Song, L.; Du, J.; Xiao, Z.; Jiang, P.; Cao, Z.; Zhu, H. Research progress on the surface of high-nickel nickel-cobalt-manganese ternary cathode materials: A mini review. Front. Chem. 2020, 8, 761. [Google Scholar] [CrossRef]
  12. Choi, J.U.; Voronina, N.; Sun, Y.-K.; Myung, S.-T. Recent progress and perspective of advanced high-energy Co-less Ni-rich cathodes for Li-ion batteries: Yesterday, today, and tomorrow. Adv. Energy Mater. 2020, 10, 2002027. [Google Scholar] [CrossRef]
  13. Xu, R.; Xu, W.; Wang, J.; Liu, F.; Sun, W.; Yang, Y. A review on regenerating materials from spent lithium-ion batteries. Molecules 2022, 27, 2285. [Google Scholar] [CrossRef]
  14. Wang, X.; Ding, Y.-L.; Deng, Y.-P.; Chen, Z. Ni-rich/Co-poor layered cathode for automotive Li-ion batteries: Promises and challenges. Adv. Energy Mater. 2020, 10, 1903864. [Google Scholar] [CrossRef]
  15. Jiang, M.; Danilov, D.L.; Eichel, R.-A.; Notten, P.H.L. A review of degradation mechanisms and recent achievements for Ni-rich cathode-based Li-ion batteries. Adv. Energy Mater. 2021, 11, 2103005. [Google Scholar] [CrossRef]
  16. Kim, J.; Lee, H.; Cha, H.; Yoon, M.; Park, M.; Cho, J. Prospect and reality of Ni-rich cathode for commercialization. Adv. Energy Mater. 2018, 8, 1702028. [Google Scholar] [CrossRef]
  17. Butt, A.; Ali, G.; Kubra, K.T.; Sharif, R.; Salman, A.; Bashir, M.; Jamil, S. Recent advances in enhanced performance of Ni-rich cathode materials for Li-ion batteries: A review. Energy Technol. 2022, 10, 2100775. [Google Scholar] [CrossRef]
  18. Liao, C.; Li, F.; Liu, J. Challenges and modification strategies of Ni-rich cathode materials operating at high-voltage. Nanomaterials 2022, 12, 1888. [Google Scholar] [CrossRef]
  19. Yin, S.; Deng, W.; Chen, J.; Gao, X.; Zou, G.; Hou, H.; Ji, X. Fundamental and solutions of microcrack in Ni-rich layered oxide cathode materials of lithium-ion batteries. Nano Energy 2021, 83, 105854. [Google Scholar] [CrossRef]
  20. Ahaliabadeh, Z.; Kong, X.; Fedorovskaya, E.; Kallio, T. Extensive comparison of doping and coating strategies for Ni-rich positive electrode materials. J. Power Sources 2022, 540, 231633. [Google Scholar] [CrossRef]
  21. Tao, J.; Mu, A.; Geng, S.; Xiao, H.; Zhang, L.; Huang, Q. Influences of direction and magnitude of Mg2+ doping concentration gradient on the performance of full concentration gradient cathode material. J. Solid State Electrochem. 2021, 25, 1959–1974. [Google Scholar] [CrossRef]
  22. Li, Q.; Li, Z.; Wu, S.; Wang, Z.; Liu, X.; Li, W.; Li, N.; Wang, J.; Zhuang, W. Utilizing diverse functions of zirconium to enhance the electrochemical performance of Ni-rich layered cathode materials. ACS Appl. Energy Mater. 2020, 3, 11741–11751. [Google Scholar] [CrossRef]
  23. Wang, Y.; Liu, K.; Wang, B. Coating strategies of Ni-rich layered cathode in LIBs. Chem. Res. Chin. Univ. 2021, 42, 1514–1529. [Google Scholar] [CrossRef]
  24. Wang, X.; Ruan, X.; Du, C.-F.; Yu, H. Developments in surface/interface engineering of Ni-rich layered cathode materials. Chem. Rec. 2022, 22, e202200119. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, H.; Ge, W.; Li, W.; Wang, F.; Liu, W.; Qu, M.Z.; Peng, G. Facile fabrication of ethoxy-functional polysiloxane wrapped LiNi0.6Co0.2Mn0.2O2 cathode with improved cycling performance for rechargeable Li-ion battery. ACS Appl. Mater. Interfaces 2016, 8, 18439–18449. [Google Scholar] [CrossRef] [PubMed]
  26. Ni, L.; Zhang, S.; Di, A.; Deng, W.; Zou, G.; Hou, H.; Ji, X. Challenges and strategies towards single-crystalline Ni-rich layered cathodes. Adv. Energy Mater. 2022, 12, 2201510. [Google Scholar] [CrossRef]
  27. Pang, P.; Tan, X.; Wang, Z.; Cai, Z.; Nan, J.; Xing, Z.; Li, H. Crack-free single-crystal LiNi0.83Co0.10Mn0.07O2 as cycling/thermal stable cathode materials for high-voltage lithium-ion batteries. Electrochim. Acta 2021, 365, 137380. [Google Scholar] [CrossRef]
  28. Lee, S.-H.; Sim, S.-J.; Jin, B.-S.; Kim, H.-S. High performance well-developed single crystal LiNi0.91Co0.06Mn0.03O2 cathode via LiCl-NaCl flux method. Mater. Lett. 2020, 270, 127615. [Google Scholar] [CrossRef]
  29. Ran, Q.; Zhao, H.; Hu, Y.; Hao, S.; Liu, J.; Li, H.; Liu, X. Enhancing surface stability of LiNi0.8Co0.1Mn0.1O2 cathode with hybrid core-shell nanostructure induced by high-valent titanium ions for Li-ion batteries at high cut-off voltage. J. Alloys Compd. 2020, 834, 155099. [Google Scholar] [CrossRef]
  30. Sun, Y.K.; Myung, S.T.; Kim, M.H.; Prakash, J.; Amine, K. Synthesis and characterization of Li[(Ni0.8Co0.1Mn0.1)(0.8)(Ni0.5Mn0.5)(0.2)]O2 with the microscale core-shell structure as the positive electrode material for lithium batteries. J. Am. Chem. Soc. 2005, 127, 13411–13418. [Google Scholar] [CrossRef]
  31. Lim, B.-B.; Myung, S.-T.; Yoon, C.S.; Sun, Y.-K. Comparative Study of Ni-rich layered cathodes for rechargeable lithium batteries: Li[Ni0.85Co0.11Al0.04]O2 and Li[Ni0.84Co0.06Mn0.09Al0.01]O2 with two-step full concentration gradients. ACS Energy Lett. 2016, 1, 283–289. [Google Scholar] [CrossRef]
  32. Park, G.-T.; Ryu, H.-H.; Noh, T.-C.; Kang, G.-C.; Sun, Y.-K. Microstructure-optimized concentration-gradient NCM cathode for long-life Li-ion batteries. Mater. Today 2022, 52, 9–18. [Google Scholar] [CrossRef]
  33. Park, N.Y.; Ryu, H.H.; Park, G.T.; Noh, T.C.; Sun, Y.K. Optimized Ni-rich NCMA cathode for electric vehicle batteries. Adv. Energy Mater. 2021, 11, 200377. [Google Scholar] [CrossRef]
  34. Hou, P.; Zhang, H.; Zi, Z.; Zhang, L.; Xu, X. Core-shell and concentration-gradient cathodes prepared via co-precipitation reaction for advanced lithium-ion batteries. J. Mater. Chem. A 2017, 5, 4254–4279. [Google Scholar] [CrossRef]
  35. Tsai, S.-Y.; Fung, K.-Z. Synthesis routes on electrochemical behavior of Co-free layered LiNi0.5Mn0.5O2 cathode for Li-ion batteries. Molecules 2023, 28, 794. [Google Scholar] [CrossRef] [PubMed]
  36. Jiang, Y.; Liu, Z.; Zhang, Y.; Hu, H.; Teng, X.; Wang, D.; Gao, P.; Zhu, Y. Full-gradient structured LiNi0.8Co0.1Mn0.1O2 cathode material with improved rate and cycle performance for lithium ion batteries. Electrochim. Acta 2019, 309, 74–85. [Google Scholar] [CrossRef]
  37. Tan, Z.; Li, Y.; Xi, X.; Yang, J.; Xu, Y.; Xiong, Y.; Wang, S.; Liu, S.; Zheng, J. Lattice engineering to alleviate microcrack of LiNi0.9Co0.05Mn0.05O2 cathode for optimization their Li+ storage functionalities. Electrochim. Acta 2022, 401, 139482. [Google Scholar] [CrossRef]
  38. 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]
  39. Zhou, H.; Zhou, F.; Shi, S.; Yang, W.; Song, Z. Influence of working temperature on the electrochemical characteristics of Al2O3-coated LiNi0.8Co0.1Mn0.1O2 cathode materials for Li-ion batteries. J. Alloys Compd. 2020, 847, 156412. [Google Scholar] [CrossRef]
  40. Lian, K.K.; Kirk, D.W.; Thorpe, S.J. Investigation of a “two-state” tafel phenomenon for the oxygen evolution reaction on an amorphous Ni-Co alloy. J. Electrochem. Soc. 1995, 142, 3704. [Google Scholar] [CrossRef]
  41. Mansour, A.N. Characterization of NiO by XPS. Surf. Sci. Spectra 1994, 3, 231–238. [Google Scholar] [CrossRef]
  42. Mansour, A.N. Characterization of LiNiO2 by XPS. Surf. Sci. Spectra 1996, 3, 279–286. [Google Scholar] [CrossRef]
  43. Kosova, N.V.; Devyatkina, E.T.; Kaichev, V.V. Optimization of Ni2+/Ni3+ ratio in layered Li(Ni,Mn,Co)O2 cathodes for better electrochemistry. J. Power Sources 2007, 174, 965–969. [Google Scholar] [CrossRef]
  44. Yang, J.; Xia, Y. Suppressing the phase transition of the layered Ni-rich oxide cathode during high-voltage cycling by introducing low-content Li2MnO3. ACS Appl. Mater. Interfaces 2016, 8, 1297–1308. [Google Scholar] [CrossRef] [PubMed]
  45. Li, W.-W.; Zhang, X.-J.; Si, J.-J.; Yang, J.; Sun, X.-Y. TiO2-coated LiNi0.9Co0.08Al0.02O2 cathode materials with enhanced cycle performance for Li-ion batteries. Rare Met. 2021, 40, 1719–1726. [Google Scholar] [CrossRef]
  46. Huang, J.; Duan, J.; Du, K.; Cao, Y.; Peng, Z.; Hu, G. Enhanced cycling performance of LiNi0.9Co0.08Al0.02O2 via Co-rich surface. JOM 2020, 72, 738–744. [Google Scholar] [CrossRef]
  47. Mo, Y.; Guo, L.; Jin, H.; Du, B.; Cao, B.; Chen, Y.; Li, D.; Chen, Y. Building nickel-rich cathodes with large concentration gradient for high performance lithium-ion batteries. J. Power Sources 2020, 468, 228405. [Google Scholar] [CrossRef]
  48. Zhang, C.; Wan, J.; Li, Y.; Zheng, S.; Zhou, K.; Wang, D.; Wang, D.; Hong, C.; Gong, Z.; Yang, Y. Restraining the polarization increase of Ni-rich and low-Co cathodes upon cycling by Al-doping. J. Mater. Chem. A 2020, 8, 6893–6901. [Google Scholar] [CrossRef]
  49. Su, Y.; Chen, G.; Chen, L.; Shi, Q.; Lv, Z.; Lu, Y.; Bao, L.; Li, N.; Chen, S.; Wu, F. Roles of fast-ion conductor LiTaO3 modifying Ni-rich cathode material for Li-ion batteries. Chemsuschem 2021, 14, 1955–1961. [Google Scholar] [CrossRef]
  50. Qian, R.; Liu, Y.; Cheng, T.; Li, P.; Chen, R.; Lyu, Y.; Guo, B. Enhanced surface chemical and structural stability of Ni-rich cathode materials by synchronous lithium-ion conductor coating for lithium-ion batteries. ACS Appl. Mater. Interfaces 2020, 12, 13813–13823. [Google Scholar] [CrossRef]
  51. Hao, S.; Zhang, D.; Li, Y.; Xi, X.; Wang, S.; Li, X.; Shen, X.; Liu, S.; Zheng, J. Multifunctionality of cerium decoration in enhancing the cycling stability and rate capability of a nickel-rich layered oxide cathode. Nanoscale 2021, 13, 20213–20224. [Google Scholar] [CrossRef]
  52. Park, J.-S.; Hong, Y.J.; Kim, J.H.; Kang, Y.C. Carbon-templated strategy toward the synthesis of dense and yolk-shell multi-component transition metal oxide cathode microspheres for high-performance Li ion batteries. J. Power Sources 2020, 461, 15. [Google Scholar] [CrossRef]
  53. Yang, X.; Tang, Y.; Shang, G.; Wu, J.; Lai, Y.; Li, J.; Qu, Y.; Zhang, Z. Enhanced cyclability and high-rate capability of LiNi0.88Co0.095Mn0.025O2 cathodes by homogeneous Al3+ doping. ACS Appl. Mater. Interfaces 2019, 11, 32015–32024. [Google Scholar] [CrossRef]
  54. Weppner, W.; Huggins, R.A. Determination of the kinetic parameters of mixed-conducting flectrodes and application to the system Li3Sb. J. Electrochem. Soc. 1977, 124, 1569. [Google Scholar] [CrossRef]
  55. Chen, Q.; Qiao, X.; Wang, Y.; Zhang, T.; Peng, C.; Yin, W.; Liu, L. Electrochemical performance of Li3−xNaxV2(PO4)3/C composite cathode materials for lithium ion batteries. J. Power Sources 2012, 201, 267–273. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) CG-NCMOH and CC-NCMOH; (b) CG-LNCM and CC-LNCM.
Figure 1. XRD patterns of (a) CG-NCMOH and CC-NCMOH; (b) CG-LNCM and CC-LNCM.
Molecules 28 03347 g001
Figure 2. SEM images of (a) CC-NCMOH, (b) CG-NCMOH, (c) CC-LNCMO and (d) CG-LNCM; HRTEM images of (e) CC-NCMOH and (f) CG-NCMOH.
Figure 2. SEM images of (a) CC-NCMOH, (b) CG-NCMOH, (c) CC-LNCMO and (d) CG-LNCM; HRTEM images of (e) CC-NCMOH and (f) CG-NCMOH.
Molecules 28 03347 g002aMolecules 28 03347 g002b
Figure 3. SEM image of cross-section of (a) CG-LNCM and the corresponding EDX elemental mappings of (b) Ni, (c) Co and (d) Mn; SEM images of primary particles and compositions of the selected area of (e) CC-LNCM and (f) CG-LNCM.
Figure 3. SEM image of cross-section of (a) CG-LNCM and the corresponding EDX elemental mappings of (b) Ni, (c) Co and (d) Mn; SEM images of primary particles and compositions of the selected area of (e) CC-LNCM and (f) CG-LNCM.
Molecules 28 03347 g003aMolecules 28 03347 g003b
Figure 4. Ni 2p XPS spectra of (a) CC-LNCM and (b) CG-LNCM.
Figure 4. Ni 2p XPS spectra of (a) CC-LNCM and (b) CG-LNCM.
Molecules 28 03347 g004
Figure 5. The first three cycles of cyclic voltammograms of (a) CC-LNCM and (b) CG-LNCM at 0.1 mV s−1 in the potential range of 3.0–4.3 V (vs. Li+/Li).
Figure 5. The first three cycles of cyclic voltammograms of (a) CC-LNCM and (b) CG-LNCM at 0.1 mV s−1 in the potential range of 3.0–4.3 V (vs. Li+/Li).
Molecules 28 03347 g005
Figure 6. Initial charge/discharge profiles of (a) CC-LNCM and (b) CG-LNCM at various current rates, cycle performance of (c) CC-LNCM and (d) CG-LNCM at various current rates.
Figure 6. Initial charge/discharge profiles of (a) CC-LNCM and (b) CG-LNCM at various current rates, cycle performance of (c) CC-LNCM and (d) CG-LNCM at various current rates.
Molecules 28 03347 g006
Figure 7. Nyquist plots of (a) fresh electrodes and (b) electrodes after cycling for 100 cycles at 5C.
Figure 7. Nyquist plots of (a) fresh electrodes and (b) electrodes after cycling for 100 cycles at 5C.
Molecules 28 03347 g007
Figure 8. GITT curves of (a) CC-LNCM and (b) CG-LNCM, relationship between voltage and Li+ diffusion coefficient of CC-LNCM and CG-LNCM at (c) charged state and (d) discharged state.
Figure 8. GITT curves of (a) CC-LNCM and (b) CG-LNCM, relationship between voltage and Li+ diffusion coefficient of CC-LNCM and CG-LNCM at (c) charged state and (d) discharged state.
Molecules 28 03347 g008
Figure 9. Schematic illustration of the preparation of concentration-gradient precursor Ni0.9Co0.083Mn0.017(OH)2.
Figure 9. Schematic illustration of the preparation of concentration-gradient precursor Ni0.9Co0.083Mn0.017(OH)2.
Molecules 28 03347 g009
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

Li, H.; Guo, Y.; Chen, Y.; Gao, N.; Sun, R.; Lu, Y.; Chen, Q. Outstanding Electrochemical Performance of Ni-Rich Concentration-Gradient Cathode Material LiNi0.9Co0.083Mn0.017O2 for Lithium-Ion Batteries. Molecules 2023, 28, 3347. https://doi.org/10.3390/molecules28083347

AMA Style

Li H, Guo Y, Chen Y, Gao N, Sun R, Lu Y, Chen Q. Outstanding Electrochemical Performance of Ni-Rich Concentration-Gradient Cathode Material LiNi0.9Co0.083Mn0.017O2 for Lithium-Ion Batteries. Molecules. 2023; 28(8):3347. https://doi.org/10.3390/molecules28083347

Chicago/Turabian Style

Li, Hechen, Yiwen Guo, Yuanhua Chen, Nengshuang Gao, Ruicong Sun, Yachun Lu, and Quanqi Chen. 2023. "Outstanding Electrochemical Performance of Ni-Rich Concentration-Gradient Cathode Material LiNi0.9Co0.083Mn0.017O2 for Lithium-Ion Batteries" Molecules 28, no. 8: 3347. https://doi.org/10.3390/molecules28083347

Article Metrics

Back to TopTop