Next Article in Journal
Hydrogen Storage Assessment in Depleted Oil Reservoir and Saline Aquifer
Next Article in Special Issue
Insights into Enhancing Electrochemical Performance of Li-Ion Battery Anodes via Polymer Coating
Previous Article in Journal
Numerical Investigation on the Impact of Exergy Analysis and Structural Improvement in Power Plant Boiler through Co-Simulation
Previous Article in Special Issue
Hard Carbon Reprising Porous Morphology Derived from Coconut Sheath for Sodium-Ion Battery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Micron-Sized LiNi0.8Co0.1Mn0.1O2 and Its Application in Bimodal Distributed High Energy Density Li-Ion Battery Cathodes

by
Chia-Hsin Lin
1,
Senthil-Kumar Parthasarathi
1,*,
Satish Bolloju
1,
Mozaffar Abdollahifar
2,3,
Yu-Ting Weng
4 and
Nae-Lih Wu
1,4,*
1
Department of Chemical Engineering, National Taiwan University, Taipei 10617, Taiwan
2
Institute for Particle Technology, Technische Universität Braunschweig, 38104 Braunschweig, Germany
3
Battery LabFactory Braunschweig (BLB), Technische Universität Braunschweig, Langer Kamp 19, 38106 Braunschweig, Germany
4
Advanced Research Center for Green Materials Science and Technology, National Taiwan University, Taipei 10617, Taiwan
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(21), 8129; https://doi.org/10.3390/en15218129
Submission received: 4 October 2022 / Revised: 19 October 2022 / Accepted: 28 October 2022 / Published: 31 October 2022
(This article belongs to the Special Issue Particle Design and Processing for Battery Production)

Abstract

:
The uniform and smaller-sized (~3 μm) LiNi0.8Co0.1Mn0.1O2 (SNCM) particles are prepared via a fast nucleation process of oxalate co-precipitation, followed by a two-step calcination procedure. It is found that the fast nucleation by vigorous agitation enables us to produce oxalate nuclei having a uniform size which then grow into micron-particles in less than a few minutes. The impacts of solution pH, precipitation time, calcination temperature, and surface modification with ZrO2 on the structural, morphological, and electrochemical properties of SNCM are systematically examined to identify the optimal synthetic conditions. A novel bimodal cathode design has been highlighted by using the combination of the SNCM particles and the conventional large (~10 μm) LiNi0.83Co0.12Mn0.05O2 (LNCM) particles to achieve the high volumetric energy density of cathode. The volumetric discharge capacity is found to be 526.6 mAh/cm3 for the bimodal cathode L80% + S20%, whereas the volumetric discharge capacity is found to be only 480.3 and 360.6 mAh/cm3 for L100% and S100% unimodal, respectively. Moreover, the optimal bi-modal cathode delivered higher specific energy (622.4 Wh/kg) and volumetric energy density (1622.6 Wh/L) than the L100% unimodal (596.1 Wh/kg and 1402.1 Wh/L) cathode after the 100th cycle. This study points to the promising utility of the SNCM material in Li-ion battery applications.

1. Introduction

In the past decades, stimulated by the increasing awareness of the energy crisis and eco-friendly conservation around the world, research with great physio-electrochemical energy storage devices such as lithium-ion batteries (LIBs), supercapacitors, etc., become a vital issue [1,2,3,4,5,6]. In comparison with the commercial LiCoO2 cathode, Ni-rich layer-structured Li(Ni, Co, Mn)O2 (NCM) has received noteworthy attention and is considered a suitable cathode for next-generation LIBs because of its lower cost, less toxicity, and high energy density. Abundant studies in the literature focused on challenges that obstruct the performance of the Ni-rich NCM cathodes, which include Li+/Ni2+ cation disordering during the synthesis, stress-induced microcrack formation triggered by substantial volume deviation through Li+ intercalation/extraction, and instability in air and moisture during storage. Significant efforts from the viewpoint of particle synthesis have been dedicated to enriching the structural parameters, as well as electrochemical properties of the Ni-rich NCM electrodes, including those focusing on bulk doping [7,8,9,10,11,12,13,14,15,16], surface modification [17,18,19], gradient concentration design [20,21,22], synthetic condition [23], calcination temperature [24,25], and calcination atmosphere [26]. Recently, Arun et al. [27] reported a review article based on the modification strategies used to improve the specific capacity and cycling stability of the NCM cathode materials in detail. So far, relatively less attention has been given to the issue of enhancing the volumetric energy and/or capacity density of the NCM cathode in the existing literature.
Currently, the hydroxide co-precipitation process in a continuously stirred tank reactor (CSTR) has most widely been applied for producing the precursors of commercial NCM cathode powders, which are secondary particles predominantly larger than 10 μm [28]. Although a large particle size offers the advantage of high tap density, it leads to inferior rate performance due to a long Li-ion diffusion path upon lithiation/delithiation. In addition, large particles are less capable of accommodating the stress arising from the volume variations associated with lithiation–delithiation, and the variation is known to increase with the Ni content of NCM [29]. As a result, microcracking along grain boundaries between primary grains within the interior has become one major fading mechanism for NCMs with Ni contents exceeding 80% [30]. The microcracking causes electrical isolation of part of the particle interior, leading to capacity fade.
Recently, there has been an increase in interest in the use of a few micron-sized single-crystal Ni-rich NCM particles, which are known to mitigate microcracking [31,32]. Moreover, the buildup of side reaction products and unpredicted phase transformations owing to the restricted electrolyte/electrode area can be alleviated by the sole structure of single crystal NCM [33]. However, the smaller size of around 3~4 µm is crucial for single crystal NCM particles to retain the rate capabilities due to the smaller surface area for Li insertion/exertion when compared with the conventional polycrystalline NCM. Moreover, the growth mechanism and methodology, such as the molten salt method, are not compatible with the existing synthesis processes, adding difficulties for industrialization. On the other hand, the formation of smaller polycrystalline NCM (SNCM) particles with a larger surface area will lead to the higher rate capability of the cathode via the shorter lithium diffusive pathway. Furthermore, the internal cracking on NCM particles will be minimized during the cycling mechanism.
Herein, a novel oxalate-based co-precipitation approach is developed to prepare uniform and micron-sized SNCM precursors by inducing fast nucleation. Other than the benefits of feasible preparation, the formation of metal constituents breaks down the precipitation rate and creates the nucleation, as well as the growth of the particles more controllable. The novel fast agitation mechanism is developed to control the uniformity and size of the nuclei formed during the oxalate-based co-precipitation approach for the first time. It is found that the fast nucleation by vigorous agitation enables producing oxalate nuclei having a uniform size, which then grow into micron-particles in less than a few minutes. The uniform and smaller particles with a diameter of around 3 μm are synthesized by optimizing the pH value, precipitation time, calcination temperature, and ZrO2. The results show that ideal calcination temperature provides enhanced cycle life, which is highly viable compared to the conventional hydroxide-based co-precipitation route. Then, the novel bimodal cathode distributions are created with the blending combination of SNCM particles and commercially available larger NCM (LNCM) particles to achieve the high volumetric capacity density of the cathode, where the smaller SNCM particles are embedded into the vacancies between the larger LNCM particles. Remarkably, from the electrochemical analysis, a synergistic effect is found in particular blended ratios that intensely encourage the rate capability and cycle life, and this is favorable for fast-charging batteries. This novel research thoroughly investigates the mechanism and principle in the processing of SNCM, which unlocks a new way for synthesizing bimodal layered Ni-rich cathode materials in LIBs.

2. Materials and Methods

2.1. Material Synthesis

2.1.1. Preparation of Ni0.8Co0.1Mn0.1C2O4·2H2O Precursor

The precursor (Ni0.8Co0.1Mn0.1C2O4·2H2O) particles were produced via an oxalate co-precipitation route by using a batch reactor. Initially, the stoichiometric ratio of NiSO4, CoSO4, and MnSO4 with a total amount of 0.2 M was dissolved in 200 mL of deionized (DI) water and denoted as Solution A. The base Solution B was separately made by mixing 0.22 M oxalic acid and 0.2 M NaOH with 0.05 M NH4OH in 200 mL of DI water. Solution A and Solution B were heated at 50 °C, and the base solution was quickly poured into the metal ions solution under vigorous stirring at 600 rpm for 20 s. Then the mixed solution was stirred under 60 rpm for 3 h to make sure that the complete co-precipitation reaction. The precursors were gradually precipitated from 3 min to 3 h, under the constant pH = 2.0. The precipitates were thoroughly washed with DI water to eliminate the residual ions and then dried at 100 °C for 12 h in a vacuum oven.

2.1.2. Preparation of Smaller-Sized LiNi0.8Co0.1Mn0.1O2 Cathode

LiNi0.8Co0.1Mn0.1O2 (SNCM) was prepared via the two-step solid-state calcination procedure. The as-synthesized Ni-rich precursors were stoichiometrically mixed with LiOH (3 wt.% excess) and were kept in calcination at 700 °C for 12 h in an O2-atmosphere horizontal quartz-tube furnace with a heating rate of 2 °C/min. Then, before the second-step calcination process, the oxide powders were ground well to avoid agglomeration of particles and then calcined at 750 °C for 12 h in an O2 atmosphere with a heating frequency of 5 °C/min. The powder was cooled down naturally to room temperature and then finely sieved with a 400-mesh screen.

2.1.3. Preparation of ZrO2-Treated and Bimodal Cathode

To make zirconium surface-treated SNCM, four different amounts of ZrO2 (2.0, 1.0, 0.5, and 0.25 wt.% vs. the precursor) are selected and labeled as NCM-Zr2.0%, NCM-Zr1.0%, NCM-Zr0.5%, NCM-Zr0.25%, respectively. The precursors of SNCM were mixed with the appropriate amount of ZrO2, and the mixtures were ground well for 10 min. Then the hand-mixed precursors were mechanically mixed for 20 min, using the Planetary Inter Mixer (Model: IMX-150). The well-mixed powder was transferred to the tube furnace, and the preparation temperature of 750 °C for 12 h was elected as the optimum temperature. The larger-sized commercial LiNi0.83Co0.12Mn0.05O2 (LNCM) was received from Industrial Technology Research Institute (ITRI). The different weight ratios of SNCM and LNCM (10:90, 20:80, 30:70, and 40:60 of SNCM to LNCM, denoted as L90% + S10%, L80% + S20%, L70% + S30%, and L60% + S40%) were blended well for at least 15 min to make the homogenous distribution of powders. The aforementioned process was performed in the Ar-filled glove box to avoid the effect of the ambient environment.

2.2. Physical Characterization

XRD measurements were performed on an X-ray diffraction instrument with Cu-Kα radiation (λ = 1.5418 Å) (Rigaku Ultima, Tokyo, Japan), with an accelerating voltage of 40 kV and at 40 mA of current. The scanning rate was typically fixed as 10 °/min in the range of 10 ° to 80 °. The inductively coupled plasma–mass spectrometry (ICP–MS, Agilent 7500ce, Santa Clara, CA, USA) was an analytical technique used for semi-quantitative analysis and isotope measurement. The surface morphology of the specimen was revealed by field-emission scanning electron microscopy (JSM) operated with 10 kV of accelerated voltage. Particle size distribution was performed by Horiba-Partica-Laser analyzer (Model: LA-950-V2, Horiba, Kyoto, Japan), and the detective range was from 40 nm to 2000 μm. A thermal analysis of SNCM-precursors was performed by using thermal gravimetric and differential scanning calorimetry (TGA–DSC) to realize the thermal behavior under an air atmosphere (Model: SDT-650, TA Discovery Instruments, New Castle, DE, USA) at a heating rate of 5 °C/min.

2.3. Electrochemical Characterizations

The electrode slurry was made of 90 wt.% SNCM powders, 5 wt.% super P (carbon black), and 5 wt.% PVDF binder (polyvinylidene fluoride) cast on the Al-foil current collector. After drying at 120 °C for 12 h, the dried electrodes were calendered by a roll-press machine to about 70% of the original value and then pierced into discs that are 12 mm in diameter. The average mass loading of the electrodes ranged from 8.5 to 15.2 mg/cm2. The coin cell CR-2032 consisted of the active materials, such as a cathode, a lithium foil as anode, a 16 μm–thick polypropylene (PP, ND416Z) as a separator, and a 1.2 M LiPF6 in ethylene carbonate/dimethyl carbonate (EC/DMC) in the proportion of 1:2 by vol.% with 4 wt.% fluoroethylene carbonate (FEC) as an electrolyte. The cells were fabricated in an Ar-filled glove box, of which the humidity and oxygen levels are lower than 0.01 and 0.5 ppm, respectively. Galvanostatic charge/discharge experiments, cycling stability, and rate performance profiles were performed from 2.8 to 4.3 V at ambient temperature on a Neware-battery analyzer.

3. Results and Discussion

3.1. Impact of pH Value

Figure 1a depicts the SEM image of the oxalate precursors under a pH value of 2. The SEM images and particle size values of another pH (ranging from 3 to 9) are given in Supplementary Figures S1 and S3a, respectively. To optimize the pH value, the constant co-precipitation time of 3 h was fixed for all the samples. As seen from the SEM photos, the particles have a uniform oval shape, with an average particle size of about 3–5 µm. There are no obvious differences between the shapes for the pH value below 6, whereas, beyond the pH value of 6, some irregular shapes appear with the increase of particle size, which can be ascribed to the slow reaction kinetics that leads to an inhomogeneous nucleation rate.

3.2. Effect of Co-Precipitation Time

Apart from altering the pH value, the oxalate precipitates are collected at different periods of co-precipitation time (3, 5, 10, 20, 30, 60, 120, and 180 min), as given in Supplementary Figure S2. To optimize the co-precipitation time, the constant pH value of 2 is fixed for all the samples. The SEM image of the oxalate precursors with the co-precipitation time of 3 h is shown in Figure 1b. At the initial stage (3 min), almost all precipitates reveal an identical shape to the optimized condition (pH = 2) in Supplementary Figure S2a, and the subsequent results also display a similar trend (from 5 to 180 min). Since the resemblance of average particle size at every testing stage (Supplementary Figure S3b), the co-precipitation time of 3 h is fixed to obtain a higher yield fraction. It is found that the fast nucleation by vigorous agitation enables the production of oxalate nuclei that have a uniform size (Figure 1c,e) and can grow into micron-particles in less than a few minutes, whereas the slow agitation leads to the formation of nuclei with an irregular shape with larger sizes (Figure 1d,f). Hence, considering the uniform morphology with smaller particle size and the high yield, the optimal synthetic parameter for the oxalate precursor is pH = 2, and co-precipitation time is 3 h.

3.3. Effect of Calcination Temperature

The DSC–TGA analysis of the as-prepared SNCM precursor in the temperature range from room temperature to 600 °C under air atmosphere is revealed in Figure 2a. As can be seen, TGA curves have two significant weight losses around 180 and 270 °C. The former indicates the release of crystal water, while the latter implies the oxidation of oxalate, which transforms into carbon dioxide. The DSC curve also displays a similar trend at 180 and 270 °C, which suggests an endothermic (release of crystal water) and exothermic (oxidation of oxalate) phenomenon, respectively. The entire reaction is demonstrated below:
Ni 0.8 Co 0.1 Mn 0.1 C 2 O 4 · 2 H 2 O Δ Ni 0.8 Co 0.1 Mn 0.1 O x + 2 H 2 O + 2 CO 2
Consequently, according to the DSC–TGA result, the heating rate for the sintering process from 180 to 350 °C is adopted as 0.2 °C/min to decompose crystal water, as well as to oxidize oxalate in a much slower manner to prevent the internal force generated from particles, resulting in the breakage of particles. Moreover, owing to the reaction between LiOH and SNCM precursor (Figure 2b) to form a layered structure, the heating rate from 350 to 500 °C is also reduced to 0.5 °C/min to prevent local nucleation problems. Then the rise of temperature to 700 °C for 12 h is used to enhance the particle strength. After cooling down, the SNCM powder will be calcined to obtain a well-ordered layered structure.
Six different temperatures were chosen from 725 to 850 °C (with an increment of 25 °C), for 12 h, to study the effect of calcination temperature on SNCM materials. The stoichiometric compositions of synthesized SNCM powders are well-matched with the experimental scheme obtained from the ICP–MS result (Supplementary Table S1). The observed lower-stoichiometric-level lithium in the SNCM calcined at 825 °C and 850 °C are attributed to the desertion of lithium during high-temperature synthesis.
Figure 3a exhibits the powder XRD patterns of SNCM cathodes synthesized at various calcination temperatures from 725 to 850 °C. The corresponding samples are denoted as NCM-725, NCM-750, NCM-775, NCM-800, NCM825, and NCM-850. The characteristic diffraction pattern of the phase-pure NCM was detected without any notable secondary phases. The recorded XRD patterns could be matched to the α-NaFeO2 layered structure with space group   R 3 ¯ m , except for NCM-725. It is found that the lower calcination temperature of 725 °C is not sufficient for the complete formation of the fcc lattice structure. The perfect splitting of (006)/(012) and (108)/(110) planes in Figure 3b,c (indicated by arrows) specifies that a formation of well-arranged hexagonal-lattice structure [34,35,36]. The above results confirmed that the main crystallinity of layer-SNCM is formed at calcination temperature above 725 °C. The integrated intensity ratio [I(003)/I(104)] is a significant factor to decide the degree of cation ordering owing to the resemblance between ionic radii of Ni2+ (0.69 Å) and Li+ (0.76 Å). Generally, the value of I(003)/I(104) higher than 1.2 guarantees lower cation mixing [37,38,39], and SNCM calcined between 750 and 850 °C displays similar values of around 1.5. The cation mixing problem is minimized, and the intensity ratio slightly increases at higher sintering temperatures due to their high crystallinity.
In addition, with the rise of synthesis temperature, the value of FWHM (full width at half maximum) of (003) plane drops, indicating a growth of crystalline-size SNCM particles. The crystalline size is calculated from the Debye–Scherer formula and given in Table 1. The average crystallite size is found to be increased from 37 to 213 nm for the sample calcined at 725 °C to 850 °C, respectively. The larger growth in grain size will prolong the length of the Li+/electron diffusive pathway, resulting in inferior rate performance and electrochemical kinetics, as discussed later. The N2 desorption/adsorption curves and pore-size-distribution curves are displayed in Figure S4a,b, respectively. The SEM images of the SNCM powders prepared from 725 to 850 °C are given in Supplementary Figure S5. The higher calcination temperature inclines to give a smaller specific surface area on account of the larger primary particles, as well as an agglomeration of secondary particles. The observed trend is quite consistent with BET results, and the relevant results are tabulated in Table 1.
As given in Figure 4a, the primary oxidation-reduction curves at 0.1 C in the voltage window of 2.8–4.3 V for all the SNCM cells exhibit similar capacities and voltage profiles. The specific discharge capacities are found to be 197.8, 194.8, 200.2, 196.6, and 199.3 mAh/g for NCM-750, NCM-775, NCM-800, NCM-825, and NCM-850, respectively. Moreover, an initial coulombic efficiency is found to be 85.83, 82.98, 82.94, 82.37, and 81.98% for NCM-750, NCM-775, NCM-800, NCM-825, and NCM-850, respectively. The results validate that a higher annealing temperature provides better crystallinity, which could be the probable reason for the slight increase in initial capacity. Among all the samples, NCM-750 displays preferably high coulombic efficiency due to the high BET and larger pore-size distribution, which induces efficient Li+ diffusion.
All the SNCM cells show clear variations in terms of cycle number vs. specific discharge capacity at 0.3 C, as shown in Figure 4b. It is of great interest that the observed capacity retention at the end of 200 cycles decreases constantly with the raising of the preparation temperature. The specific discharge capacities are found to be 112.9, 126.9, 118.9, 107.8, 88.5, and 86.1 mAh/g for cells made with NCM-725, NCM-750, NCM-775, NCM-800, NCM-825, and NCM-850, respectively. The capacity retentions are found to be 72.84, 75.76, 70.90, 63.11, 52.27, and 50.68% for cells made with NCM-725, NCM-750, NCM-775, NCM-800, NCM-825, and NCM-850, respectively. Moreover, the cells made with the samples calcined at 750 and 775 °C display notable reversible capacity during the redox process and are still proficient in providing discharge capacities of 126.9 and 118.9 mAh/g at the end of 200 cycles, making it eye-catching for large-scale utilization. Comprehensive cycling data are demonstrated in Supplementary Table S2. The observed serious capacity fading above NCM-775 should be ascribed to the rise of electrode resistance attributed to the inner stress generated during the repetitive oxidation/reduction process, which results in harming the electrode materials. The calcination temperature of 750 °C is selected as the ideal temperature for the preparation of SNCM-active materials based on the specific reversible capacity and cycle stability.

3.4. Effect of ZrO2 Modification

The powder XRD patterns of bare (NCM-750) and ZrO2-surface-treated SNCM samples are presented in Figure 5a. All the characteristic planes can be matched with the typical structure of hexagonal α-NaFeO2 (space group:   R 3 ¯ m ) without a secondary phase. The obvious doublet of (006)/(012) and (108)/(110) planes show that all of the samples have a well-organized layered crystalline structure. The slight deviation of the (003) and (104) diffracted peaks to the lower angle side is due to the larger ionic radii of Zr4+ (0.72 Å) compared to the Ni2+/Ni3+ (0.69 Å/0.56 Å), Co3+ (0.545 Å), and Mn4+ (0.53 Å). Furthermore, from the value of I(003) to I(104), we can calculate the degree of cation ordering, which is 1.49 for the NCM-pristine and 1.56 for the NCM-Zr0.5%, meaning that Li+/Ni2+ cation mixing decreases after the modification of ZrO2. In terms of charge compensation, the minor parts of Ni3+ might be reduced to Ni2+ due to the incorporation of zirconium ions. However, they may act as a pillar in the lithium layer owing to the similar ionic radii between Li+ (0.76 Å) and Zr4+ (0.72 Å). It alleviates the subsequent relocation of Ni2+ to the Li layer and, thus, reduces the cation mixing and enhances the overall structural properties.
The existence of ZrO2 on the particle surface is further identified by TEM, as shown in Supplementary Figure S6. An obvious ZrO2 layer with a thickness around 5 nm can be observed in NCM-Zr0.5% in comparison with NCM-pristine. From Supplementary Figure S7, we can see that the Ni-2p3/2 spectra display two characteristic peaks, which are allotted to Ni3+ and Ni2+. The broad and bigger area of the Ni2+ peak in NCM-Zr0.5% cathode compared to that of NCM-pristine determines the higher content of Ni2+ on the NCM-Zr0.5% cathode’s surface. That is favorable for the layered crystal structure of NCM, which also confirms the results of XRD analysis. Otherwise, a characteristic peak of Zr-3d5/2 appears at 180.98 eV, along with a Zr-3d3/2 peak detected at 183.34 eV in the Zr- spectra, standing for the valence state of Zr, which is presented as +4 [9,40], further certifying that the ZrO2 protective layers are effectively coated on the outer surface of the NCM-Zr0.5% cathodes.
The first cycle charge–discharge curves of cells made with pristine and ZrO2-modified SNCM cathodes are shown in Figure 5b. The initial discharge capacity values are found to be 197.8 and 194.2 mAh/g for the pristine and NCM-Zr0.5% cathode, respectively. The long-term cyclability of cells made with bare and ZrO2-surface modified cathodes is tested and given in Figure 5c. It shows that all the modified cells have a significant improvement in capacity retention except for NCM-Zr2.0%, which shows a reversible capacity of only 126.73 mAh/g with 79.02% of retention. Meanwhile, the reversible capacities and capacity retention are found to be 146.75, 166.26, and 159.37 mAh/g and 85.86, 89.01, and 88.03%, for cells made with NCM-Zr1.0%, NCM-Zr0.5%, and NCM-Zr0.25% cathodes, respectively. On the contrary, the NCM-pristine sample shows the reversible specific capacity of 139.51 mAh/g with 83.44% of retention at the end of the 100th cycle. The observed electrochemical data are compared with the literature reports, and the comparative table is demonstrated in Supplementary Table S3. Notably, it is concluded that the amount of ZrO2 should be carefully optimized to avoid the insulating properties that would potentially deteriorate the performance of NCM cells. Based on the above results, the 0.5 wt.% ZrO2–modified cathode is chosen as optimized and used for the investigations of bimodal studies.

3.5. Synergistic Effect of Bimodal Behavior

The preparation of bimodal cathode materials is carried out by directly mixing with the different blend ratios of SNCM and LNCM, as discussed in the experimental section. The schematic for the preparation of the bimodal cathode is given in Supplementary Figure S8a, in which the SNCM and LNCM particle sizes are found to be 3–4 μm and 10 μm, respectively. The SEM photos of other blended cathodes are given in Supplementary Figure S9. To investigate the variation of particle size of unimodal and bimodal materials, a particle-size-distribution analysis was applied to examine the diversification of L100% and one of the representatives of blended samples, L80% + S20%. From Supplementary Figure S8b, we can observe the broad peaks at about 14.94 μm for L100%, while for blended cathodes (L80% + S20%), the peak at 7.08 μm is obtained owing to the combination of different size NCM powders.
The SEM images of the electrode plane and cross-section mode are also observed to verify the homogeneity of blended materials after the preparation of slurry, as well as serve as the target for the measurements of electrode density. Figure 6 displays the normal and cross-sectional images of L100% and L80% + S20%, and it can be seen that the particles are well blended with carbon conductive and binder (Figure 6a,c). The electrode thickness is found to be 65.25 μm and 50.63 μm for the cell L100% and L80% + S20%, respectively (Figure 6b,d). As seen, the electrode thickness is decreased abruptly when even a trace amount of SNCM is blended with LNCM. The L100% shows an electrode density of 2.33 g/cm3, whereas the value gradually enhances and reaches up to 2.61 g/cm3 for the L80% + S20% sample, indicating that suitable amounts of blended cathodes would have a positive effect on increasing the electrode density.
Figure 7a reveals the long cycle behavior (0.3 C at room temperature) according to specific and volumetric discharge capacity, and the relevant specific/volumetric profiles are illustrated in Supplementary Figure S10. Because of the identical trend of specific and volumetric capacities, the latter explanation is given in detail. The initial volumetric capacity of L100% (black) is 443.5 mAh/cm3 and then decreases to 366.1 mAh/cm3 in the 100th cycle, with the retention equal to 82.56%, whereas the volumetric capacity retention of the L80% + S20% blended sample is found to be 88.09%, which uncovers the possibility of a bimodal cathode that goes through a different declination rate through the interaction with the surroundings. The difference in volumetric capacity is ascribed to the intrinsic nature between large and small particles, in which the larger particle provides a higher tap density, as well as electrode density, but the smaller one guarantees the ability of cell operation under high rates.
As demonstrated in Figure 7b, it is obvious that a similar specific capacity can be acquired regardless of pure or blended cathodes, and the deviation is approximately less than 10% from 0.1 to 5 C, standing for the identical ability for the cell during “discharge” process. Furthermore, at 5 C, the specific discharge capacities overlap, which have 135.2, and, 138.1 mAh/g for L100%, and L80% + S20%, respectively. The fact causing rise in capacity at elevated C-rate is attributed to the ability of SNCM for rapid charge/discharge owing to the shorter diffusion track of lithium-ion compared to LNCM. The volumetric discharge capacity possesses 526.6, 498.9, 476.2, 454.4, 427.3, and 363.8 mAh/cm3 in comparison with L100%, which only exhibits 480.3, 462.2, 435.9, 419.1, 396.6, and 340.3 mAh/cm3 under different C-rates from 0.1 to 5 C. The superior performance of these bimodal materials could be explained by the optimized blended ratio that allows the small particle to insert into the space of large particles, thus helping to facilitate the overall performance.
To understand the prevailing advantage of bimodal design, the volumetric charge capacity versus time under 5 C (high rate) is further shown in Figure 8a. Interestingly, the optimal blended ratio (L80% + S20% and L70% + S10%) displays a higher value than others. It demonstrates an extraordinary performance in terms of mass loading and electrode thickness. To achieve the rate capability under higher C-rates for commercialization, the pellet density should be supposed to reduce to its 75% original value for large particles since their overall resistance originated from the bulk electrolyte diffusion onto the particle surface. Nevertheless, the introduction of a bimodal system that embeds the small particle into the space of a large particle could not only enhance the electrode density and volumetric capacity but simultaneously increase the rate performance. The blend of SNCM with LNCM has a porous structure, promoting the infiltration of electrolytes and high rates of performance. The histogram illustrated in Figure 8b represents the available increase in volumetric capacity, using the bimodal method. Among these materials, L80% + S20% also demonstrates the best volumetric energy density after cycling (1622.6 Wh/L) compared to those L100%, mainly because of its higher electrode density (2.61 g/cm3), as well as better capacity retention (88.09%) that is considered to offer the largest amount of energy density, which make it preferable for commercialization utilization.

4. Conclusions

In summary, uniform SNCM particles are prepared by optimal synthesis conditions (pH = 2 and co-precipitation time of 3 h) through a fast nucleation process during the oxalate co-precipitation process. For the first time, the novel fast agitation mechanism is developed to control the uniformity and size of the nuclei formed during the oxalate-based co-precipitation approach. It is found that the fast nucleation by vigorous agitation enables producing oxalate nuclei with a uniform size that then grow into micron-particles in less than a few minutes. The SNCM cathode materials prepared at 750 °C for 12 h are found to be the optimal calcination condition in terms of intact morphology, smaller primary particle, specific capacity, and long-term cycle stability. The existence of bulk Zr4+ doping, as well as surface ZrO2 coating on SNCM particles, is further confirmed by XRD and XPS analyses. After 100 cycles at 0.3 C, the capacity withholding is 89.01% for NCM-Zr0.5%, as opposed to 83.44% of NCM-pristine material. Moreover, the discharge capability at 5 C can exhibit nearly 150 mAh/g for NCM-Zr0.5%, which is nearly twice as much as NCM-pristine. The novel bimodal cathode distributions were successfully created with the blending combination of as-prepared smaller NCM and commercial larger NCM. The bimodal cathode behavior between SNCM and LNCM particles was investigated in detail. Owing to the inset of smaller particles into the void open space of larger particles, the electrode density and the volumetric capacity were increased dramatically in bimodal samples. More importantly, from the observation of voltage profiles, the outstanding performance of smaller particles under the 5 C rate further supports the fast charging capability for bimodal cathodes. This uncomplicated but important analysis unveils the feasibility of promoting specific/volumetric energy density for practical utility, shedding light on the possibility of commercialization, and simultaneously opens up an innovative concept of using a bimodal cathode system.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/en15218129/s1. Figure S1: SEM images of SNCM precursors with different pH value (a) pH = 2.0, (b) pH = 3.0, (c) pH = 4.0, (d) pH = 5.0, (e) pH = 6.0, (f) pH = 7.0, (g) pH = 8.0, and (h) pH = 9.0. Figure S2: SEM micrographs of SNCM precursors with different co-precipitation time (a) 3 min, (b) 5 min, (c) 10 min, (d) 20 min, (e) 30 min, (f) 60 min, (g) 120 min, and (h) 180 min. Figure S3: (a) The pH value vs. average particle size, and (b) co-precipitation time vs. average particle size. Figure S4: (a) The N2 adsorption/desorption isotherms and (b) pore size distribution curves of SNCM prepared at different temperatures. Figure S5: SEM micrograph of SNCM particles prepared at different calcination temperature: (a) 725 °C, (b) 750 °C, (c) 775 °C, (d) 800 °C, (e) 825 °C, (f) 850 °C. Figure S6: TEM micrographs of (a) SNCM-pristine, and (b) SNCM-Zr0.5% cathodes. Figure S7: XPS spectra of (a) Ni 2p, (c) Zr 3d for SNCM-pristine and SNCM-Zr0.5% cathodes. Figure S8: (a) Schematic picture of bimodal NCM materials, and (b) Particle size distribution data of L100% and (L80% + S20%) bimodal NCM cathode. Figure S9: SEM morphology of bimodal cathode powders at different blended ratio: (a) L90% + S10%, (b) L80% + S20%, (c) L70% + S30%, and (d) L60% + S40%. Figure S10: Specific charge/discharge profiles, and Volumetric charge/discharge profiles of (a,b) L100%, and (c,d) L80% + S20% cathodes under 0.3 C between 2.8 and 4.3 V at room temperature. Table S1: Li/Ni/Co/Mn molar ratios of SNCM samples calcined at different temperatures. Table S2: Electrochemical parameters for cells made with SNCM cathodes. Table S3: Comparison of electrochemical parameters for ZrO2-modified NCM cathode [41,42,43,44].

Author Contributions

Conceptualization, N.-L.W.; data curation, C.-H.L.; formal analysis, S.-K.P., S.B., M.A. and Y.-T.W.; funding acquisition, N.-L.W.; methodology, C.-H.L.; supervision, N.-L.W.; writing—original draft, C.-H.L.; writing—review and editing, S.-K.P., S.B., M.A., Y.-T.W. and N.-L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work is financially supported by the “Advanced Research Center for Green Materials Science and Technology” from the Featured Area Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (111L9006) and the National Science and Technology Council in Taiwan (NSTC-111-2634-F-002-016), and also of NSTC-110-2221-E-002-015-MY3 and NSTC 111-2923-E-011-001.

Data Availability Statement

All the reported data are included in the manuscript.

Acknowledgments

The authors also thank S. J. Ji and C. Y. Chien (NSTC; NTU) for their assistance in performing electron microscopy analyses.

Conflicts of Interest

The authors declare that there is no conflict of interest.

References

  1. Song, Z.; Miao, L.; Ruhlmann, L.; Lv, Y.; Zhu, D.; Li, L.; Gan, L.; Liu, M. Self-assembled carbon superstructures achieving ultra-stable and fast proton-coupled charge storage kinetics. Adv. Mater. 2021, 33, 2104148. [Google Scholar] [CrossRef] [PubMed]
  2. Dunn, B.; Kamath, H.; Tarascon, J.M. Electrical energy storage for the grid: A battery of choices. Science 2011, 334, 928–935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Liu, C.; Li, F.; Ma, L.P.; Cheng, H.M. Advanced materials for energy storage. Adv. Mater. 2010, 22, E28–E62. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, J.; Zhang, J.G.; Yang, Z.; Lemmon, J.P.; Imhoff, C.; Graff, G.L.; Li, L.; Hu, J.; Wang, C.; Xiao, J.; et al. Materials science and materials chemistry for large scale electrochemical energy storage: From transportation to electrical grid. Adv. Funct. Mat. 2013, 23, 929–946. [Google Scholar] [CrossRef]
  5. Song, Z.; Miao, L.; Ruhlmann, L.; Lv, Y.; Zhu, D.; Li, L.; Gan, L.; Liu, M. Lewis pair interaction self-assembly of carbon superstructures harvesting high-energy and ultralong-life zinc-ion storage. Adv. Funct. Mater. 2022, 33, 2208049. [Google Scholar] [CrossRef]
  6. Mizushima, K.; Jones, P.; Wiseman, P.; Goodenough, J.B. LixCoO2 (0 < x < −1): A new cathode material for batteries of high energy density. Mater. Res. Bull. 1980, 15, 783–789. [Google Scholar]
  7. Li, H.; Zhou, P.; Liu, F.; Li, H.; Cheng, F.; Chen, J. Stabilizing nickel-rich layered oxide cathodes by magnesium doping for rechargeable lithium-ion batteries. Chem. Sci. 2019, 10, 1374–1379. [Google Scholar] [CrossRef] [Green Version]
  8. 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]
  9. Li, X.; Zhang, K.; Wang, M.; Liu, Y.; Qu, M.; Zhao, W.; Zheng, J. Dual functions of zirconium modification on improving the electrochemical performance of Ni-rich LiNi0.8Co0.1Mn0.1O2. Sustain. Energy Fuels 2018, 2, 413–421. [Google Scholar] [CrossRef]
  10. Zhang, D.; Liu, Y.; Wu, L.; Feng, L.; Jin, S.; Zhang, R.; Jin, M. Effect of Ti ion doping on electrochemical performance of Ni-rich LiNi0.8Co0.1Mn0.1O2 cathode material. Electrochim. Acta 2019, 328, 135086. [Google Scholar] [CrossRef]
  11. Wu, F.; Li, Q.; Chen, L.; Lu, Y.; Su, Y.; Bao, L.; Chen, R.; Chen, S. Use of Ce to reinforce the interface of Ni-rich LiNi0.8Co0.1Mn0.1O2 cathode materials for lithium-ion batteries under high operating voltage. ChemSusChem 2019, 12, 935–943. [Google Scholar] [CrossRef] [PubMed]
  12. Shang, G.; Tang, Y.; Lai, Y.; Wu, J.; Yang, X.; Li, H.; Peng, C.; Zheng, J.; Zhang, Z. Enhancing structural stability unto 4.5 V of Ni-rich cathodes by tungsten-doping for lithium storage. J. Power Sources 2019, 423, 246–254. [Google Scholar] [CrossRef]
  13. Dong, J.; He, H.; Zhang, D.; Chang, C. Boron improved electrochemical performance of LiNi0.8Co0.1Mn0.1O2 by enhancing the crystal growth with increased lattice ordering. J. Mater. Sci. Mater. Electron. 2019, 30, 18200–18210. [Google Scholar] [CrossRef]
  14. Zhang, M.; Zhao, H.; Tan, M.; Liu, J.; Hu, Y.; Liu, S.; Shu, X.; Li, H.; Ran, Q.; Cai, J.; et al. Yttrium modified Ni-rich LiNi0.8Co0.1Mn0.1O2 with enhanced electrochemical performance as high energy density cathode material at 4.5 V high voltage. J. Alloys Compd. 2019, 774, 82–92. [Google Scholar] [CrossRef]
  15. Wu, K.; Jia, G.; Shangguan, X.; Yang, G.; Zhu, Z.; Peng, Z.; Zhuge, Q.; Li, F.; Cui, X. Improved high rate performance and cycle stability for LiNi0.8Co0.2O2by doping of the high valence state ion Nb5+ into Li+ sites. J. Alloys Compd. 2018, 765, 700–709. [Google Scholar] [CrossRef]
  16. Yuan, A.; Tang, H.; Liu, L.; Ying, J.; Tan, L.; Sun, R. High performance of phosphorus and fluorine co-doped LiNi0.8Co0.1Mn0.1O2 as a cathode material for lithium ion batteries. J. Alloys Compd. 2020, 844, 156210. [Google Scholar] [CrossRef]
  17. Jung, R.; Metzger, M.; Maglia, F.; Stinner, C.; Gasteiger, H.A. Oxygen release and its effect on the cycling stability of LiNixMnyCozO2 (NMC) cathode materials for Li-ion batteries. J. Electrochem. Soc. 2017, 164, A1361–A1377. [Google Scholar] [CrossRef]
  18. Zhao, R.; Liang, J.; Huang, J.; Zeng, R.; Zhang, J.; Chen, H.; Shi, G. Improving the Ni-rich LiNi0.5Co0.2Mn0.3O2 cathode properties at high operating voltage by double coating layer of Al2O3 and AlPO4. J. Alloys Compd. 2017, 724, 1109–1116. [Google Scholar] [CrossRef]
  19. Yao, L.; Liang, F.; Jin, J.; Chowdari, B.V.R.; Yang, J.; Wen, Z. Improved electrochemical property of Ni-rich LiNi0.6Co0.2Mn0.2O2 cathode via in-situ ZrO2 coating for high energy density lithium ion batteries. Chem. Eng. J. 2020, 389, 124403. [Google Scholar] [CrossRef]
  20. Lim, B.; Yoon, S.J.; Park, K.J.; Yoon, C.S.; Kim, S.J.; Lee, J.J.; Sun, Y.K. Advanced concentration gradient cathode material with two-slope for high-energy and safe lithium batteries. Adv. Funct. Mater. 2015, 25, 4673–4680. [Google Scholar] [CrossRef]
  21. Park, K.J.; Choi, M.J.; Maglia, F.; Kim, S.J.; Kim, K.H.; Yoon, C.S.; Sun, Y.K. High-capacity concentration gradient Li[Ni0.865Co0.120Al0.015]O2 cathode for lithium-ion batteries. Adv. Energy Mater. 2018, 8, 1703612. [Google Scholar] [CrossRef]
  22. Sun, Y.K.; Chen, Z.; Noh, H.J.; Lee, D.J.; Jung, H.G.; Ren, Y.; Wang, S.; Yoon, C.S.; Myung, S.T.; Amine, K. Nanostructured high-energy cathode materials for advanced lithium batteries. Nat. Mater. 2012, 11, 942–947. [Google Scholar] [CrossRef] [PubMed]
  23. Ren, D.; Shen, Y.; Yang, Y.; Shen, L.; Levin, B.D.A.; Yu, Y.; Muller, D.A.; Abruna, H.D. Systematic optimization of battery materials: Key parameter optimization for the scalable synthesis of uniform, high-energy, and high stability LiNi0.6Mn0.2Co0.2O2 cathode material for lithium-ion batteries. ACS Appl. Mater. Interfaces 2017, 9, 35811–35819. [Google Scholar] [CrossRef] [PubMed]
  24. Zheng, J.; Yan, P.; Estevez, L.; Wang, C.; Zhang, J.G. Effect of calcination temperature on the electrochemical properties of nickel-rich LiNi0.76Mn0.14Co0.10O2 cathodes for lithium-ion batteries. Nano Energy 2018, 49, 538–548. [Google Scholar] [CrossRef]
  25. Iqbal, A.; Li, D. Systematic study of the effect of calcination temperature and Li/M molar ratio on high performance Ni-rich layered LiNi0.9Co0.1O2 cathode materials. Chem. Phys. Lett. 2019, 720, 97–106. [Google Scholar] [CrossRef]
  26. Shim, J.H.; Kim, C.Y.; Cho, S.W.; Missiul, A.; Kim, J.K.; Ahn, Y.J.; Lee, S. Effects of heat-treatment atmosphere on electrochemical performances of Ni-rich mixed-metal oxide (LiNi0.80Co0.15Mn0.05O2) as a cathode material for lithium ion battery. Electrochim. Acta 2014, 138, 15–21. [Google Scholar] [CrossRef]
  27. Arun, M.D.; Manickam, M.; Parisa, A.B.; Shashi, P.; Pooja, K.; Anoop, M.D.; Krishna, N.M. Rational design on materials for developing next generation lithium-ion secondary battery. Prog. Solid State Chem. 2021, 62, 100298. [Google Scholar]
  28. Entwistle, T.; Perez, E.S.; Murray, G.J.; Anthonisamy, N.; Cussen, A. Co-precipitation synthesis of nickel-rich cathodes for Li-ion batteries. Energy Rep. 2022, 8, 67–73. [Google Scholar] [CrossRef]
  29. Xu, Z.; Rahman, M.; Mu, L.; Liu, Y.; Lin, F. Chemomechanical behaviors of layered cathode materials in alkali metal ion batteries. J. Mater. Chem. A 2018, 6, 21859–21884. [Google Scholar] [CrossRef]
  30. Yan, P.; Zheng, J.; Liu, J.; Wang, B.; Cheng, X.; Zhang, Y.; Sun, X.; Wang, C.; Zhang, J. Tailoring grain boundary structures and chemistry of Ni-rich cathodes for enhanced cycle stability of lithium-ion batteries. Nat. Energy 2018, 3, 600–605. [Google Scholar] [CrossRef]
  31. Fan, X.; Hu, G.; Zhang, B.; Ou, X.; Zhang, J.; Zhao, W.; Jia, H.; Zou, L.; Li, P.; Yang, Y. Crack-free single-crystalline Ni-rich layered NCM cathode enable superior cycling performance of lithium-ion batteries. Nano Energy 2020, 70, 104450. [Google Scholar] [CrossRef]
  32. Trevisanello, E.; Ruess, R.; Conforto, G.; Ritcher, F.H.; Janek, J. Polycrystalline and single Crystalline NCM cathode materials—Quantifying particle cracking, active surface area, and lithium diffusion. Adv. Energy Mater. 2021, 11, 2003400. [Google Scholar] [CrossRef]
  33. Li, F.; King, L.; Sun, Y.; Jin, Y.; Hou, P. Micron-sized monocrystalline LiNi1/3Co1/3Mn1/3O2 as high-volumetric-energy-density cathode for lithium-ion batteries. J. Mater. Chem. A 2018, 6, 1234–12352. [Google Scholar] [CrossRef]
  34. Kumar, P.S.; Sakunthala, A.; Rao, R.P.; Adams, S.; Chowdari, B.V.R.; Reddy, M.V. Layered Li1+x(Ni0.33Co0.33Mn0.33)O2 cathode material prepared by the microwave-assisted solvothermal method for lithium-ion batteries. Mater. Res. Bull. 2017, 93, 381–390. [Google Scholar] [CrossRef]
  35. Kumar, P.S.; Sakunthala, A.; Reddy, M.V.; Shanmugam, S.; Prabu, M. Correlation between the structural, electrical, and electrochemical performance of layered Li(Ni0.33Co0.33Mn0.33)O2 for lithium-ion battery. J. Solid State Electrochem. 2016, 20, 1865–1876. [Google Scholar] [CrossRef]
  36. Zhang, B.; Dong, P.; Tong, H.; Yao, Y.; Zheng, J.; Yu, W.; Zhang, J.; Chu, D. Enhanced electrochemical performance of LiNi0.8Co0.1Mn0.1O2 with lithium-reactive Li3VO4 coating. J. Alloys Compd. 2017, 706, 198–204. [Google Scholar] [CrossRef]
  37. Lee, S.H.; Kim, H.S.; Jin, B.S. Recycling of Ni-rich Li(Ni0.8Co0.1Mn0.1)O2 cathode materials by a thermomechanical method. J. Alloys Compd. 2019, 803, 1032–1036. [Google Scholar] [CrossRef]
  38. Yabuuchi, N.; Ohzuku, T. Novel lithium insertion material of LiCo01/3Ni1/3Mn1/3O2 for advanced lithium-ion batteries. J. Power Sources 2003, 119, 171–174. [Google Scholar] [CrossRef]
  39. Ngala, J.K.; Chernova, N.A.; Ma, M.; Mamak, M.; Zavalij, P.Y.; Whittingham, M.S. The synthesis, characterization and electrochemical behavior of the layered LiNi0.8Co0.2Mn0.4O2 compound. J. Mater. Chem. 2004, 14, 214–220. [Google Scholar] [CrossRef]
  40. Nisar, U.; Amin, R.; Essehli, R.; Shakoor, R.A.; Kahraman, R.; Kim, D.K.; Khaleel, M.A.; Belharouak, I. Extreme fast charging characteristics of zirconia modified LiNi0.5Mn1.5O4 cathode for lithium ion batteries. J. Power Sources 2018, 396, 774–781. [Google Scholar] [CrossRef]
  41. Gao, S.; Zhan, X.; Cheng, Y.T. Structural, electrochemical and Li-ion transport properties of Zr-modified LiNi0.8Co0.1Mn0.1O2 positive electrode materials for Li-ion batteries. J. Power Sources 2019, 410–411, 45–52. [Google Scholar] [CrossRef]
  42. 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 electrochemical performance in lithium-ion batteries. Adv. Energy Mat. 2018, 8, 1701682–1701691. [Google Scholar] [CrossRef]
  43. Ho, V.C.; Jeong, S.; Yim, T.; Mun, J. Crucial role of thioacetamide for ZrO2 coating on the fragile surface of Ni-rich layered cathode in lithium-ion batteries. J. Power Sources. 2020, 450, 227625–227635. [Google Scholar] [CrossRef]
  44. Hu, G.; Zhang, Z.; Li, T.; Gan, Z.; Du, K.; Peng, Z.; Xia, J.; Tao, Y.; Cao, Y. In situ surface modification for improving the electrochemical performance of Ni-rich cathode materials by using ZrP2O7. Chem. Sus. Chem. 2019, 12, 1–11. [Google Scholar] [CrossRef]
Figure 1. SEM images of SNCM precursor for (a) pH value of 2 (with a fixed co-precipitation time of 3 h), (b) co-precipitation time of 180 min (with a fixed pH of 2), (c) the co-precipitation time of 3 min with vigorous agitation, (d) the co-precipitation time of 3 min without vigorous agitation, and (e,f) enlarged portion of the co-precipitation time of 3 min with/without vigorous agitation.
Figure 1. SEM images of SNCM precursor for (a) pH value of 2 (with a fixed co-precipitation time of 3 h), (b) co-precipitation time of 180 min (with a fixed pH of 2), (c) the co-precipitation time of 3 min with vigorous agitation, (d) the co-precipitation time of 3 min without vigorous agitation, and (e,f) enlarged portion of the co-precipitation time of 3 min with/without vigorous agitation.
Energies 15 08129 g001
Figure 2. DSC–TGA analyses of (a) oxalate SNCM precursor without LiOH and (b) oxalate SNCM precursor with LiOH.
Figure 2. DSC–TGA analyses of (a) oxalate SNCM precursor without LiOH and (b) oxalate SNCM precursor with LiOH.
Energies 15 08129 g002
Figure 3. Powder XRD patterns of (a) SNCM at different calcination temperature; and enlarged XRD patterns of (b) (006)/(012) and (c) (108)/(110) diffraction peaks.
Figure 3. Powder XRD patterns of (a) SNCM at different calcination temperature; and enlarged XRD patterns of (b) (006)/(012) and (c) (108)/(110) diffraction peaks.
Energies 15 08129 g003
Figure 4. (a) Initial charge and discharge curves at 0.1 C for SNCM cells and (b) cycling performance of SNCM cells at 0.3 C rate (temperature, RT; voltage window, 2.8–4.3 V).
Figure 4. (a) Initial charge and discharge curves at 0.1 C for SNCM cells and (b) cycling performance of SNCM cells at 0.3 C rate (temperature, RT; voltage window, 2.8–4.3 V).
Energies 15 08129 g004
Figure 5. (a) XRD patterns of pristine and ZrO2-modified SNCM cathodes, (b) first cycle charge–discharge curves of pristine and ZrO2-modified SNCM cells at 0.1 C, and (c) cycling performance of ZrO2-modified SNCM cells at 0.3 C rate (temperature, RT; voltage window, 2.8–4.3 V).
Figure 5. (a) XRD patterns of pristine and ZrO2-modified SNCM cathodes, (b) first cycle charge–discharge curves of pristine and ZrO2-modified SNCM cells at 0.1 C, and (c) cycling performance of ZrO2-modified SNCM cells at 0.3 C rate (temperature, RT; voltage window, 2.8–4.3 V).
Energies 15 08129 g005
Figure 6. SEM micrographs of normal and cross-section mode for (a,b) L100% and (c,d) L80% + S20% electrodes.
Figure 6. SEM micrographs of normal and cross-section mode for (a,b) L100% and (c,d) L80% + S20% electrodes.
Energies 15 08129 g006
Figure 7. (a) Cyclic stability plot of cells made with single and bimodal cathodes at 0.3 C rate (temperature, RT; voltage window, 2.8–4.3 V) and (b) rate performance in terms of specific and volumetric discharge capacity.
Figure 7. (a) Cyclic stability plot of cells made with single and bimodal cathodes at 0.3 C rate (temperature, RT; voltage window, 2.8–4.3 V) and (b) rate performance in terms of specific and volumetric discharge capacity.
Energies 15 08129 g007
Figure 8. (a) Volumetric charge capacity vs. time under the current density of 5 C, and the histogram of (b) volumetric energy density for uni- and bimodal NCM cathodes at 1st, 50th, and 100th cycle under CC mode.
Figure 8. (a) Volumetric charge capacity vs. time under the current density of 5 C, and the histogram of (b) volumetric energy density for uni- and bimodal NCM cathodes at 1st, 50th, and 100th cycle under CC mode.
Energies 15 08129 g008
Table 1. Crystallographic data and BET surface area of SNCM cathodes.
Table 1. Crystallographic data and BET surface area of SNCM cathodes.
Temperature725 °C750 °C775 °C800 °C825 °C850 °C
Grain size (nm)36.341.968.695.9170.7213.6
I(003)/I(104)1.231.491.461.561.581.57
BET (m2/g)4.113.232.961.491.290.83
Pore volume (cm3/g)0.013040.012540.010840.00390.00290.0018
Pore diameter (nm)10.7315.0114.4414.6411.2822.97
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lin, C.-H.; Parthasarathi, S.-K.; Bolloju, S.; Abdollahifar, M.; Weng, Y.-T.; Wu, N.-L. Synthesis of Micron-Sized LiNi0.8Co0.1Mn0.1O2 and Its Application in Bimodal Distributed High Energy Density Li-Ion Battery Cathodes. Energies 2022, 15, 8129. https://doi.org/10.3390/en15218129

AMA Style

Lin C-H, Parthasarathi S-K, Bolloju S, Abdollahifar M, Weng Y-T, Wu N-L. Synthesis of Micron-Sized LiNi0.8Co0.1Mn0.1O2 and Its Application in Bimodal Distributed High Energy Density Li-Ion Battery Cathodes. Energies. 2022; 15(21):8129. https://doi.org/10.3390/en15218129

Chicago/Turabian Style

Lin, Chia-Hsin, Senthil-Kumar Parthasarathi, Satish Bolloju, Mozaffar Abdollahifar, Yu-Ting Weng, and Nae-Lih Wu. 2022. "Synthesis of Micron-Sized LiNi0.8Co0.1Mn0.1O2 and Its Application in Bimodal Distributed High Energy Density Li-Ion Battery Cathodes" Energies 15, no. 21: 8129. https://doi.org/10.3390/en15218129

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