Next Article in Journal
Curcumin Sustained Release with a Hybrid Chitosan-Silk Fibroin Nanofiber Containing Silver Nanoparticles as a Novel Highly Efficient Antibacterial Wound Dressing
Next Article in Special Issue
Nanoflower-like P-doped Nickel Oxide as a Catalytic Counter Electrode for Dye-Sensitized Solar Cells
Previous Article in Journal
Lattice Boltzmann Simulation of Coupling Heat Transfer between Solid and Gas Phases of Nanoporous Materials
Previous Article in Special Issue
Two-Dimensional Core-Shell Structure of Cobalt-Doped@MnO2 Nanosheets Grown on Nickel Foam as a Binder-Free Battery-Type Electrode for Supercapacitor Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stabilizing Li-Rich Layered Cathode Materials Using a LiCoMnO4 Spinel Nanolayer for Li-Ion Batteries

Department of Materials Science and Engineering, National Formosa University, Yunlin 632301, Taiwan
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(19), 3425; https://doi.org/10.3390/nano12193425
Submission received: 2 September 2022 / Revised: 23 September 2022 / Accepted: 23 September 2022 / Published: 29 September 2022

Abstract

:
Lithium–rich cathodes have excess lithium in the transition metal layer and exhibit an extremely high specific capacity and good energy density. However, they still have some disadvantages. Here, we propose LiCoMnO4, a new nanolayer coating material with a spinel structure, to modify the surface of lithium cathode oxide (Li7/6Mn1/2Ni1/6Co1/6O2) with a layered structure. The designed cathode with nanolayer spinel coating delivers an excellent reversible capacity, outstanding rate capability, and superior cycling ability whilst exhibiting discharge capacities of 300, 275, 220, and 166 mAh g−1 at rates of 0.1 C at 2.0−4.8 V formation and 0.1, 1, and 5 C, respectively, between 2.0 and 4.6 V. The cycling ability and voltage fading at a high operational voltage of 4.9 V were also investigated, with results showing that the nanolayer spinel coating can depress the surface of the lithium cathode oxide layer, leading to phase transformation that enhances the electrochemical performance.

1. Introduction

With the intensification of climate change and environmental pollution, lithium–ion batteries (LIBs) are being increasingly pursued to power electric vehicles (EVs) and store renewable energy [1,2,3]. A safe, long cycle life with higher energy density and good chemical performance is desired for LIBs. Increasing the capacity of the cathode and the charging voltage of the cells are potential methods for achieving this. A lithium–rich cathode with excess lithium in the transition metal layer shows an extraordinarily high specific capacity after an irreversible oxygen release activation process at approximately 4.5 V, which may be a promising solution for producing next–generation cathode materials. However, owing to the shortcomings of the low initial Coulombic efficiency, poor rate capability, and voltage attenuation upon cycling, lithium–rich cathodes are difficult to commercialize [4,5,6,7].
Recently, the disadvantages of lithium–rich cathodes were found to be associated with the poor mass transport of the rearranged surface once the activation of Li2MnO3 exceeds 4.5 V [8]. In addition, the elimination of the oxygen atom and Li+ vacancies from the layered lattice, which decreases the number of sites for insertion and extraction of Li+ in subsequent cycles, is the reason for the irreversible capacity [9]. These materials also possess fragile surface properties at high potentials, can experience electrolytic erosion, and may induce the dissolution of transition metal ions [10,11,12]. Recently, it was reported that surface modification and bulk doping could potentially overcome these issues.
The surface modification of cathode materials by metal oxide layers such as Al2O3, NbO5, Ta2O5, ZrO2, and ZnO [13,14,15,16,17,18] can decrease side reactions at the electrode–electrolyte interface, but this usually has a negative impact on the rate capability due to the slow Li+ or electron–conductivity of the coating layer [19,20]. Recent studies have suggested that coating cathode materials with a layer of electron conductors or Li+ conductors, such as carbon [21], AlF3 [22], LiNiPO4 [23], Li2ZrO3 [24], or LiMgPO4 [25], can enhance the rate capability to some extent, as these can promote charge transfer or Li+ diffusion from the surface of the electrode material.
Spinel–type LiNixMn2-xO4 (0 < x < 0.5) oxides exhibit great potential as high–power cathode materials for LIBs [17,26,27] because of their high Li+ ionic and electronic conductivities [28], high voltage, low cost, and low toxicity. Presently, spinel oxides are a considerable choice to encapsulate Ni–rich layered oxides for identical cubic close–packed arrays [29,30]. Yan et al. [31] designed a composition of a Ni–rich layered core and a LiNi0.5Mn1.5O4–like spinel shell. The heterostructure, with a low degree of Ni2+ and Li+ disordering, can afford a stable structure framework and high–efficiency paths for the diffusion of lithium ions. Min et al. [32] proposed novel particles, the spinel interphase of which is coherent with the surface of the LiNi0.91Co0.06Mn0.03O2 (NCM) cathode. The spinel–coated NCM structures outperformed the bare sample in terms of the capacity retention rate and rate capability: they conserved more than 90% of the initial capacity at 1C for up to 50 cycles. After cycling, the crack generation was significantly reduced in the secondary particles of the coated structures compared to that of the bare sample.
In addition, some studies have combined the advantages of lithium–rich cathodes and spinel materials to enhance electrochemical performance in terms of kinetics, energetics, and core material stability [33,34,35]. In a study by Choi et al. [36], Li–rich composite oxides were synthesized by doping LiMn2O4, LiNi0.5Mn1.5O4, and LiCo0.5Mn1.5O4. Of the various composites created, the LiNi0.5Mn1.5O4–embedded sample showed lower formation energy as determined by first–principles calculation [37,38]. Furthermore, the Li–rich oxide composite formed by doping the optimal spinel, LiNi0.5Mn1.5O4, exhibited the best cycling performance with little voltage decay when used in a cylindrical 18650–type full cell. Wu et al. [39] developed a lithium–rich cathode (Li1.2Mn0.6Ni0.2O2) coated with a 4 V Li1+xMn2O4 spinel membrane for high–energy and high–power Li–ion batteries. The modified cathode displayed significant enhancements in specific capacity and cycling ability compared to an unmodified cathode. It yielded a maximal capacity of 295.6 mAh g1, and a capacity of 280.0 mAh g1 by the 50th cycle, presenting excellent capacity retention at 94.7%. Yu et al. [40] reported an effective approach for synthesizing layered spinel–capped, nanotube–assembled 3D Li–rich hierarchitectures. The resulting material’s unique 3D hollow hierarchical structure significantly shortened electron and ion transfer pathways whilst maintaining reliable structural stability. As a result, the LiMn1.5Ni0.5O4/Li2MnO3–LiMn0.4Ni0.4Co0.2O2 composite components delivered a high capacity of 293 mAh g−1 at 0.1 C, showing strong capacity retention of 89.5% after 200 cycles at 1 C, and exhibited a high capacity of 202 mAh g−1 even at 5 C. To date, there have been few reports on the combined merits of layered and spinel materials to render a superior material in terms of kinetics, energetics, and stability [41,42,43]. Yin et al. [44] synthesized the high–voltage cathode of Li1.2Mn0.54Ni0.13Co0.13O2 spheres coated with thin LiNi0.5Mn1.5O4 layers via a simple coprecipitation method. This coated composite cathode recorded a superior rate capacity and stable cycle performance in the 2.0–4.8 V range. The Li1.2Mn0.54Ni0.13Co0.13O2/LiNi0.5Mn1.5O4 microsphere cathodes delivered a high capacity of 249.5 and 96 mAh g−1 at the rate of 0.2 C and 5 C, respectively. In addition, LiMn2O4/LiNi0.5Mn1.5O4 hollow microspheres have been developed by Liu et al. [45], with a facile solution–phase coating and subsequent solid–phase lithiation route in an atmosphere of air. Their double–shell LiMn2O4/LiNi0.5Mn1.5O4 hollow microspheres showed a high specific capacity of 120 mAh g1 at a 1 C rate and an excellent rate capability (90 mAh g1 at 10 C) over the 3.5–5 V range compared to Li/Li+, with a retention of 95% over 500 cycles.
In this experiment, a nanolayer of spinel LiCoMnO4 (LCMO) was used as a coating because of its voltage stability window (5.3 V vs Li/Li+) [46,47,48]. In this composition, lithium–rich oxide contributes to high energy density, while the spinel nanolayer of LCMO provides structural stability to maintain the surface of the layered lithium–rich cathode during high–voltage cycling. Most importantly, this high Li+ conductive nanolayer can rapidly transport Li+ between the electrolyte and the layered lithium–rich cathode. Owing to the high capacitance advantage of the lithium–rich cathode combined with the structural stability of the high–voltage spinel LCMO nanolayer, the electrochemical performance of lithium–rich cathode materials is expected to improve.

2. Materials and Methods

2.1. Cathode Preparation

The synthesis of Li7/6Mn1/2Ni1/6Co1/6O2 was carried out via the coprecipitation of the precursor, Mn1/2Ni1/6Co1/6(CO3)5/6. A 2 M solution of CoSO4·7H2O, NiSO4·7H2O, and MnSO4·H2O was dropped into a continuous stirred–tank reactor. A 2 M Na2CO3 solution and an appropriate volume of NH4OH solution as the chelating agent were then separately pumped into the reactor. The solutions were maintained at 60 °C, with a stirring speed of 1200 rpm and a pH of 8. After the reaction, the Mn1/2Ni1/6Co1/6(CO3)5/6 precursor was filtered and air–dried.
The Mn1/2Ni1/6Co1/6(CO3)5/6 precursor was added to the batch reactor as a core, and 1%, 2.5%, and 5% concentrations of CoSO4·7H2O and MnSO4·H2O solutions at a molar ratio of Co2+:Mn2+ = 1:1 were pumped in. A stoichiometric volume of Na2CO3 and an appropriate volume of NH4OH were also added. After stirring for 12 h, the precursors of Mn1/2Ni1/6Co1/6(CO3)5/6 coated with 0%, 1%, 2.5%, and 5% CoMn(CO3)2 were obtained. After drying, the coated precursors with 5% excess LiOH were calcined for 14 h at 900 °C to obtain Li7/6Mn1/2Ni1/6Co1/6O2 cathodes coated with 0% (LR–Bare), 1% (LR–Sp1), 2.5% (LR–Sp2.5), and 5% (LR–Sp5) LiCoMnO4, respectively.

2.2. Characterization

Powder X–ray diffraction (XRD) measurements of the synthesized samples were performed using an X–ray diffractometer (D8A25, Bruker, USA) operated at 40 keV and 40 mA, using Cu Kα radiation. Microstructural investigation of the powders was performed via a scanning electron microscopy (SEM, JSM–7500F, JEOL, Japan) and a high–resolution transmission electron microscope (HR–TEM, JEM–2100Plus JEOL, Japan). The particle sizes of the powders were measured using a static light–scattering analyzer (Mastersizer 2000, Malvern Instruments Ltd., UK), resulting in a measurement range of 0.04–2000 μm. The density of the cathode was measured using gas pycnometry (AccuPyc II, ATA scientific pty ltd, Australia), a non–destructive technique that employs gas displacement, often helium, to measure the volume of a material. The local structure of the sample was studied by Raman spectroscopy (MRI) using a laser with a wavelength of 532 nm and a power of 100 mW. Chemical state and composition analyses were then performed by X–ray photoelectron spectroscopy (XPS) using a Physical Electronics instrument (PHI 5000 VersaProbe Ⅲ, ULVAC–PHI. Inc., Japan) with an Al X–ray source.

2.3. Electrochemical Measurements

The electrochemical properties were examined using a coin cell (CR–2032) with Li foil as the anode. Cathode electrodes were prepared by mixing 80 wt% as–prepared cathodes, 10 wt% acetylene carbon black, and 10 wt% polyvinylidene difluoride (PVDF) in an n–methyl–2–pyrrolidone solution. Next, the slurries were coated onto aluminum foil. The thickness of the loaded active materials was 31.6   ±   4.2 μm, whilst the mass loading was 7.9   ±   0.2 mg cm−2. An organic solution of 1 M LiPF6 in ethylene carbonate (EC) and dimethyl carbonate (DMC) with a 1:2 volume ratio was used as the electrolyte. The charge–discharge profiles were measured at a current density of 20 mA g−1 (0.1 C) in the 2–4.8 V range for the formation process and then continually tested from a current density of 20 mA g−1 (0.1 C) to 1000 mA g−1 (5 C) within the 2–4.6 V range. After one formation cycle, the cycle life was determined by charging at 0.2 C and subsequent discharging at 0.5 C from 2 to 4.6 V. A second cycle life test was conducted at a high voltage of 4.9 V, with the same current density. The electrochemical impedance spectroscopy (EIS) test of the cells in reference to the Li/Li+ couple was measured after 25 charge–discharge cycles at 2–4.6 V and then again at a charged state of 4.2 V, within a frequency range of 0.01–100,000 Hz with a perturbation amplitude of 10 mV.

3. Results

3.1. Morphological Characterization

The surfaces and morphologies of the precursors were examined using electron microscopy, as shown in Figure 1. For the LR–Bare precursors, the nucleating cores of MCO3 (M = Mn, Ni, Co) grew to form spherical primary particles with a diameter of 30 nm, as shown in Figure 1b. During the coprecipitation process, the primary particles aggregated into micro secondary particles with a spherical shape, as shown in Figure 1a, having the most stable thermodynamic state and minimal surface energy [49]. The primary particles of LR–Sp1 precursors with 1% nanolayer coverage had a size of 50–70 nm. When the coating concentration increased to 2.5%, the surface of the precursors was covered with dense square grains and a few flake particles. After a further increase to 5% coating, the flake particles on the surface of the precursors became larger.
The precursors were calcined with the required amount of LiOH at 900 °C to form LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 cathodes, and their morphologies are shown in Figure 2. The spherical primary particles of the LR–Bare cathode grew to 100 nm, whilst the primary particles of LR–Sp1, LR–Sp2.5, and LR–Sp5 also became spherical, with size growing alongside the increasing concentration, to eventually reach a diameter of approximately 200–300 nm. The secondary particle morphologies of all cathodes presented microsphere aggregation, and their mass median diameters (D50) and true densities increased with the coating concentration. The mass median diameters (D50) of secondary particles for LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 cathodes were 7.2, 7.8, 8.6, and 8.9 μm, respectively, as shown in Figure 3. Meanwhile, the true densities of these cathodes were 4.02, 4.12, 4.18, and 4.28 g cm−1, respectively.

3.2. Structural Characterization

The structural characterization XRD patterns for all samples are plotted in Figure 4. The main peaks in the diffraction patterns of LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 could be identified as that for hexagonal α–NaFeO2, with the space group of R 3 ¯ m. On the other hand, the low–intensity superlattice peaks in the 2θ range of 20–25° could be attributed to the Li2MnO3–like component, with a space group of C2/m and the ordering of Li, Ni, and Mn atoms in the transition metal layers [50,51]. In the coated samples, the spinel structure of LiCoMnO₄ (LCMO) was not identified because of the cubic close–packing structures. This presented similar peaks with the XRD patterns of layered and spinel lithium metal oxides, making it difficult to distinguish the layered and spinel components from coated samples with low spinel content.
To further discern these spinel and layered structures, Raman spectroscopy was used to determine the short–range local structures of the materials. Figure 5 displays the Raman spectra of the four samples. The Raman bands located at 590, 471, and 425 cm−1 were observed in all samples, presenting typical characteristics of layered lithium–rich oxides, indicating that the layered structure was maintained after the deposition of the spinel nanolayer coating. There were two vibrational modes for the lithium–rich oxides with hexagonal R 3 ¯ m symmetry: Eg with symmetrical deformation and A1g with the symmetrical stretching of M–O (M = Mn, Ni, and Co) [52]. The weak band at 425 cm−1 was the characteristic vibration of Li2MnO3, which has monoclinic C2/m symmetry and a lower local symmetry than R 3 ¯ m [40]. Meanwhile, two additional shoulder bands at around 635 and 551 cm−1 and another peak at 481 cm−1 in LR–Sp1, LR–Sp2.5, and LR–Sp5, as indicated by the blue arrows, were typical peaks of the spinel–phase with Fd3m symmetry [53,54]. Overall, the Raman spectra presented mixed characteristics of the spinel and layered phases for the coated samples.
To investigate the surface chemical states of the samples, X–ray photoelectron spectroscopy (XPS) measurements were obtained, as shown in Figure 6. The Ni 2p3/2 peaks at 854.5 and 860.9 eV indicated the coexistence of Ni2+ and Ni3+, as shown in Figure 6a [55]. In addition, the binding energy of 779.9 eV for the Co 2p3/2 peak was assigned to the Co3+ species, shown in Figure 6b [56]. Figure 6c shows that Mn p3/2 and Mn 2p1/2 had binding energies of 641.9 and 653.2 eV, respectively, which are typical values for Mn4+ [57,58]. The Mn, Ni, and Co XPS peaks for LR–Bare and LR−Sp2.5 presented the same binding energies. Figure 6d compares the O 1s peaks; there were two pronounced peaks at 528.9 and 530.6 eV, which corresponded to lattice oxygen atoms (O–M–O) and oxygen vacancies, respectively. The O 1s peak areas of LR–Sp2.5, especially the peak of oxygen vacancies, were smaller than those for LR–Bare, which indicated that the LCMO spinel nanolayer was successfully constructed on the surface of the layered lithium–rich oxides, thus exhibiting good structural compatibility.

3.3. Electrochemical Characterization

The charge–discharge rate performances of the samples at 0.1 C over 2.0–4.8 V during the formation process and at 0.1, 1, and 5 C between 2.0 and 4.6 V are presented in Figure 7. LR–Bare had the lowest rate capability, alongside a dramatic capacity drop with an increase in C–rate; this was because of the fast Li+ insertion and extraction, which damaged the fragile surface of the layered structure at a high rate. The C–rate results for LR–Sp1 were similar to those for LR–Bare because of the low concentration of the spinel nanolayer. LR–Sp2.5 and LR–Sp5, especially the former, exhibited superior discharge capacities and C–rate performances compared to LR–Bare. These results indicate that the Li–rich layered oxide with LCMO spinel nanolayer provided the combined advantages of high capacity from the Li–rich layered oxide and high rate capability from the spinel nanolayer. The discharge capacities of LR–Bare were 271, 254, 193, and 144 mAh g−1 at 0.1 C and 2.0–4.8 V during formation and at rates of 0.1, 1, and 5 C across 2.0−4.6 V, respectively. Under the same conditions, LR–Sp2.5 exhibited excellent discharge capacities of 300, 275, 220, and 166 mAh g−1, respectively. Thus, it is believed that an optimal concentration of the LCMO spinel nanolayer has good structural compatibility with Li–rich layered oxides and can provide relatively high working C–rates owing to 3D interstitial spaces [59]; this accelerates lithium–ion insertion and extraction kinetics [60,61].
After formation at 2.0–4.8 V, the cycle performance of the samples was tested for 100 cycles at 0.2 C/0.5 C charge–discharge between 2.0 and 4.6 V, with these results shown in Figure 8. When cycling for 60 cycles, the capacities of all samples were stable at approximately 200 mAh g−1; however, after 60 cycles, the capacity of LR–Bare began to decrease rapidly. After 100 cycles, the discharge capacity of LR–Bare was 150.5 mAh g−1. In contrast, remarkably improved cycling stability was observed in the samples with the LCMO spinel nanolayers. Even after 100 cycles, the discharge capacities of LR–Sp1, LR–Sp2.5, and LR–Sp5 remained at 183.7, 193.6, and 192.1 mAh g−1, respectively. This is because the spinel nanolayer prevented phase transition on the Li–rich layered oxide surface during cycling, thus improving the cycle performance. Therefore, it prevented the generation of stress induced by the anisotropic volume change of the primary crystallites in secondary particles, further preventing the generation of cracks [58]. Overall, samples with spinel nanolayers displayed better cycle performance [62].
To explain the electrochemical properties of LR–Bare and LR–Sp2.5, electrochemical impedance spectroscopy (EIS) was operated after 25 charge–discharge cycles at 2.0–4.6 V and then in a charged state at 4.2 V. The impedance spectra shown in Figure 9 contained the bulk resistance of the electrolyte (Re), the resistance at the contact of the composite cathode (RSEI), the charge transfer resistance at the interface of the cathode–electrolyte (Rct), and a line inclined at a constant angle to the real axis in the low–frequency range corresponding to Li+ diffusion in the materials [5,55,58]. The Rct and lithium ion diffusion coefficient (DLi+) were 23.1 Ω and 2.96 × 10−12 cm2 s−1 for LR–Bare and 20.0 Ω and 3.21 × 10−12 cm2 s−1 for LR–Sp2.5, respectively. This shows that the sample with a spinel nanolayer had the fastest lithium–ion insertion and extraction kinetics at the electrolyte–electrolyte interface.
Transmission electron microscopy (TEM) was further utilized to study the detailed morphological structures of LR–Bare and LR–Sp2.5 samples. LR–Bare is an agglomerated secondary particle with nanosized primary grains, as shown in Figure 10a. In Figure 10b, LR–Bare presents a smooth edge line without any other layer on the surface of the grain. During long–term cycling, partial cation ordering in the layered structure of LR–Bare was locally rearranged into an inactive phase that contained transition metal ions in both the lithium and transition metal layers [63]. The phase transformation accompanied the lattice strain in the grains; consequently, nanosized domains were identified in LR–Bare grains, which developed into a polycrystalline structure after 100 cycles, as shown in Figure 10c,d. After modification with the LCMO spinel nanolayer, LR–Sp2.5 was covered by a thin film with a width of 5–8 nm, as shown in Figure 10e. Throughout 100 cycles, LR–Sp2.5 exhibited polycrystalline and intact grains, as shown in Figure 10f,g. According to Gu [64] and Xu et al. [12], the migration of transmission metal ions to Li layers leads to a layered–to–inactive spinel phase transformation, which only occurs at the surface of the particles in the beginning. During long–term cycling, the inactive spinel phase transformation extends to the inner parts of the particles. In our study, after 100 cycles, LR–Sp2.5 presented mosaic grains coated with a spinel nanolayer (Figure 10f) and undamaged particles without a coating (Figure 10g). Therefore, we presume the former was likely located on the outer surface, and the latter was present in the center of the microsphere aggregates of the LR–Sp2.5 particle. This proves that the LCMO spinel nanolayer could effectively stabilize the structure and relieve the particle damage caused by phase transformation, thereby protecting the inner particles, and allowing them to function normally.
To further understand the utility of the LCMO spinel nanolayer under high voltage on the Li–rich layered oxide, the range of the charge–discharge voltage was raised to 2.0–4.9 V. The cycle performances of the samples at 0.2 C and 0.5 C and charge–discharge at 4.9 V are shown in Figure 11a. In this case, the capacity of LR–Bare decayed very quickly with an increasing number of cycles owing to the high voltage, and the battery completely failed after 42 cycles. On the contrary, the samples with the LCMO spinel nanolayers showed better cycling performance, with LR–Sp2.5 having the best and most stable capacity retention. Figure 11b–e shows the discharge profiles of LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 for cycles 10, 20, 30, and 40, respectively. In Figure 11b, LR–Bare delivered a capacity of 228.5 mAh g−1 at the 10th cycle, but a capacity of only 28.1 mAh g−1 remained after 40 cycles, alongside significant voltage fading. For LR–Sp1 and LR–Sp5 in Figure 11c,e, the capacity was approximately 230 mAh g−1 at the 10th cycle. However, capacity and voltage had rapidly decayed by the 20th cycle before stabilizing during later cycles. However, LR–Sp1 completely failed after 45 cycles because of the low concentration of the LCMO spinel nanolayer. In Figure 11d, LR–Sp2.5 demonstrated significant improvements in both specific capacity and voltage maintenance with increasing cycles at 2.0–4.9 V. LCMO is a high–voltage cathode material with an operating voltage of approximately 5.3 V [48]; therefore, its electrochemical performance is stable at 4.9 V. When LCMO is used as a coating layer, it can yield good structural compatibility to stabilize the surface of the Li–rich layered oxide and avoid the generation of oxygen vacancies, which is induced by the phase transformation from the layered structure to the inactive spinel structure at the surface. In addition, LCMO can suppress the side reactions caused by the electrolyte at high voltages. These features enhance the diffusion kinetics of Li ions, improve cycling performance, and alleviate voltage decay.
The differential capacity vs. voltage (dQ/dV) curves provide a clear perspective on the electrochemical reactions that occur during cycling. Figure 12 shows the dQ/dV curves of LR–Bare and LR–Sp2.5 cathodes cycled at 0.2 C and 0.5 C across 2.0–4.9 V. There were three redox reactions corresponding to Mn, Ni, Co, and O. For LR–Bare, two anodic peaks of Mn3+/4+ were observed; the first one, at 3.1 V, was induced by Mn in an inactive spinel structure and the other one, at 3.3 V, was related to Mn in the Li–rich layered oxides [65,66]. A prominent oxygen–activating peak also appeared at approximately 4.5 V, which was related to the layered–to–inactive spinel phase transformation of Li–rich layered oxides [67,68,69,70], as shown in Figure 12a. In Figure 12b, the divided peak of Mn was not observed, and the intensity of the O2− anodic peak for LR–Sp2.5 was much lower than those of LR–Bare. In addition, during the cycling, LR–Sp2.5 presented greater overlap among the cathodic peaks of Mn compared with LR–Bare. This indicates that LR–Sp2.5 exhibited inconspicuous polarization at a higher cutoff voltage of 4.9 V than LR–Bare. After cycling at high voltage, the LCMO spinel nanolayer maintained the structural stability of the lithium–rich oxide surface, thereby suppressing phase transformation and crystallographic distortions.

4. Conclusions

Lithium–rich oxides, Li7/6Mn1/2Ni1/6Co1/6O2, were successfully modified with a LiCoMnO4 (LCMO) spinel nanolayer on the surface using a coprecipitation method. The LCMO spinel nanolayer effectively prevented side reactions between the lithium–rich oxide and organic electrolytes, thereby promoting ion and charge transfer on the surface of the lithium–rich oxides. Meanwhile, the LCMO nanolayer could effectively restrain the migration of transmission metal ions and the generation of oxygen vacancies induced by the phase transformation from the layered structure to the inactive spinel structure at the surface of lithium–rich oxides. As a result, the obtained LR–Sp2.5 showed excellent cycle and rate performances during cycling at normal (4.6 V) and high (4.9 V) voltages.

Author Contributions

H.-F.L.: Conceptualization, Methodology, Validation, Data curation, Writing—original draft, Formal analysis, Writing—review & editing, Visualization, Supervision, Project administration. S.-T.C.: Investigation, Formal analysis, Data curation, Formal analysis, Project administration. D.-Z.C.: Investigation, Formal analysis. N.-Y.W.: Investigation, Formal analysis. Z.-X.J.: Data curation. C.-T.C.: Data curation. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Ministry of Science and Technology of the Republic of China (MOST 110–2637–E–150–010 and MOST 110–3116–F–011–002).

Acknowledgments

The authors appreciate the technical support from the Industrial Technology Research Institute and Common Laboratory for Micro/Nano Science and Technology of National Formosa University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goodenough, J.B.; Park, K. −S. The Li−ion rechargeable battery: A perspective. J. Am. Chem. Soc. 2013, 135, 1167–1176. [Google Scholar] [CrossRef] [PubMed]
  2. Thackeray, M.M.; Wolverton, C.; Isaacs, E.D. Electrical energy storage for transportation−approaching the limits of, and going beyond, lithium−ion batteries. Energy Environ. Sci. 2012, 5, 7854–7863. [Google Scholar] [CrossRef]
  3. Yu, R.; Wang, X.; Fu, Y.; Wang, L.; Cai, S. Effect of magnesium doping on properties of lithium−rich layered oxide cathodes based on a one−step co−precipitation strategy. J. Mater. Chem. A 2016, 4, 4941–4951. [Google Scholar] [CrossRef]
  4. Pang, W.K.; Lin, H.-F.; Peterson, V.K.; Lu, C.-Z.; Liu, C.-E.; Liao, S.-C.; Chen, J.M. Effects of Fluorine and Chromium Doping on the Performance of Lithium−Rich Li1+xMO2 (M = Ni, Mn, Co) Positive Electrode. Chem. Mater. 2017, 29, 10299–10311. [Google Scholar] [CrossRef]
  5. Lin, H.-F.; Guo, H.-L.; Hsiao, S.-C. Electrochemical performances of lithium rich cathodes prepared via co−precipitation method with different precipitator and atmosphere. J. Mater. Sci. Mater. Electron. 2018, 29, 10427–10436. [Google Scholar] [CrossRef]
  6. Yang, X.; Wang, D.; Yu, R.; Bai, Y.; Shu, H.; Ge, L.; Guo, H.; Wei, Q.; Liu, L.; Wang, X. Suppressed capacity/voltage fading of high−capacity lithium−rich layered materials via the design of heterogeneous distribution in the composition. J. Mater. Chem. A 2014, 2, 3899–3911. [Google Scholar] [CrossRef]
  7. Wang, G.; Wang, X.; Yi, L.; Yu, R.; Liu, M.; Yang, X. Preparation and performance of 0.5Li2MnO3·0.5LiNi1/3Co1/3Mn1/3O2 with a fusiform porous micro−nano structure. J. Mater. Chem. A 2016, 4, 15929–15939. [Google Scholar] [CrossRef]
  8. Li, J.; Camardese, J.; Glazier, S.; Dahn, J.R. Structural and Electrochemical Study of the Li–Mn–Ni Oxide System within the Layered Single Phase Region. Chem. Mater. 2014, 26, 7059–7066. [Google Scholar] [CrossRef]
  9. Martha, S.K.; Nanda, J.; Veith, G.M.; Dudney, N.J. Electrochemical and rate performance study of high−voltage lithium−rich composition: Li1.2Mn0.525Ni0.175Co0.1O2. J. Power Source 2012, 199, 220–226. [Google Scholar] [CrossRef]
  10. Thackeray, M.M.; Kang, S.-H.; Johnson, C.S.; Vaughey, J.T.; Benedek, R.; Hackney, S.A. Li2MnO3–stabilized LiMO2 (M= Mn, Ni, Co) electrodes for lithium–ion batteries. J. Mater. Chem. 2007, 17, 3112. [Google Scholar] [CrossRef]
  11. Wang, D.; Belharouak, I.; Zhou, G.K. Amine Nanoarchitecture Multi–Structural Cathode Materials for High Capacity Lithium Batteries. Adv. Funct. Mater. 2012, 23, 1070. [Google Scholar] [CrossRef]
  12. Xu, B.; Fell, C.R.; Chi, M.; Meng, Y.S. Identifying surface structural changes in layered Li–excess nickel manganeseoxides in high voltage lithium ion batteries: A joint experimental and theoretical study. Energy Environ. Sci. 2011, 4, 2223. [Google Scholar] [CrossRef]
  13. Kottegoda, I.R.M.; Kadoma, Y.; Ikuta, H.; Uchimoto, Y.; Wakihara, M. Enhancement of Rate Capability in Graphite Anode by Surface Modification with Zirconia. Electrochem. Solid State Lett. 2002, 5, A275. [Google Scholar] [CrossRef]
  14. Eftekhari, A. Surface Modification of Thin–Film Based LiCoPO4 5 V Cathode with Metal Oxide. J. Electrochem. Soc. 2004, 151, A1456. [Google Scholar] [CrossRef]
  15. Van Landschoot, N.; Kelder, E.M.; Kooyman, P.J.; Kwakernaak, C.; Schoonman, J. Electrochemical performance of Al2O3–coated Fe doped LiCoVO4. J. Power Source 2004, 138, 262. [Google Scholar] [CrossRef]
  16. Sun, Y.-K.; Yoon, C.S.; Oh, I.-H. Surface structural change of ZnO–coated LiNi0.5Mn1.5O4 spinel as 5 V cathode materials at elevated temperatures. Electrochim. Acta 2002, 48, 503. [Google Scholar]
  17. Miyashiro, H.; Yamanaka, A.; Tabuchi, M.; Seki, S.; Nakayama, M.; Ohno, Y.; Kobayashi, Y.; Mita, Y.; Usami, A.; Wakihara, M. Improvement of degradation at elevated temperature and at high state–of–charge storage by ZrO2 coating on LiCoO2. J. Electrochem. Soc. 2006, 153, A348. [Google Scholar] [CrossRef]
  18. Lin, Y.-M.; Wu, H.-C.; Yen, Y.-C.; Guo, Z.-Z.; Yang, M.-H.; Chen, H.-M.; Sheu, H.-S.; Wu, N.-L. Enhanced High–Rate Cycling Stability of LiMn2O4 Cathode by ZrO2 Coating for Li–Ion Battery. J. Electrochem. Soc. 2005, 152, A1526. [Google Scholar] [CrossRef]
  19. Zhao, T.; Li, L.; Chen, R.; Wu, H.; Zhang, X.; Chen, S.; Xie, M.; Wu, F.; Lu, J. Amine Design of surface protective layer of LiF/FeF3 nanoparticles in Li–rich cathode for high–capacity Li–ion batteries. Nano Energy 2015, 15, 164–176. [Google Scholar] [CrossRef]
  20. Zheng, J.; Gu, M.; Xiao, J.; Polzin, B.J.; Yan, P.; Chen, X.; Wang, C.; Zhang, G.-J. Functioning Mechanism of AlF3 Coating on the Li– and Mn–Rich Cathode Materials. Chem. Mater. 2014, 26, 6320–6327. [Google Scholar] [CrossRef]
  21. Xia, Q.; Zhao, X.; Xu, M.; Ding, Z.; Liu, J.; Chen, L.; Ivey, D.G.; Wei, W. A Li–rich Layered–Spinel–Carbon heterostructured cathode material for high capacity and high rate lithium–ion batteries fabricated via an in situ synchronous carbonization–reduction method. J. Mater. Chem. A 2015, 3, 3995–4003. [Google Scholar] [CrossRef]
  22. Sun, Y.-K.; Lee, M.-J.; Yoon, C.S.; Hassoun, J.; Amine, K.; Scrosati, B. The Role of AlF3 Coatings in Improving Electrochemical Cycling of Li–Enriched Nickel–Manganese Oxide Electrodes for Li–Ion Batteries. Adv. Mater. 2012, 24, 1192. [Google Scholar] [CrossRef] [PubMed]
  23. Croy, J.R.; Balasubramanian, M.; Kim, D.; Kang, J.-S.; Thackeray, M.M. Designing High–Capacity, Lithium–Ion Cathodes Using X–ray Absorption Spectroscopy. Chem. Mater. 2011, 23, 5415. [Google Scholar] [CrossRef]
  24. Zhang, X.; Sun, S.; Wu, Q.; Wan, N.; Pan, D.; Bai, Y. Improved electrochemical and thermal performances of layered Li[Li0.2Ni0.17Co0.07Mn0.56]O2 via Li2ZrO3 surface modification. J. Power Source 2015, 282, 378–384. [Google Scholar] [CrossRef]
  25. Liu, W.; Oh, P.; Liu, X.; Myeong, S.; Cho, W.; Cho, J. Countering Voltage Decay and Capacity Fading of Lithium–Rich Cathode Material at 60 °C by Hybrid Surface Protection Layers. Adv. Energy Mater. 2015, 5, 1500274. [Google Scholar] [CrossRef]
  26. Lee, S.; Cho, Y.; Song, H.-K.; Lee, K.T.; Cho, J. Carbon–Coated Single–Crystal LiMn2O4 Nanoparticle Clusters as Cathode Material for High–Energy and High–Power Lithium–Ion Batteries. Angew. Chem. Int. Ed. 2012, 51, 8748–8752. [Google Scholar] [CrossRef]
  27. Zhou, L.; Zhao, D.; Lou, X.W. LiNi0.5Mn1.5O4 Hollow Structures as High–Performance Cathodes for Lithium–Ion Batteries. Angew. Chem. Int. Ed. 2012, 51, 239. [Google Scholar] [CrossRef]
  28. Thackeray, M.M. Lithium–ion batteries: An unexpected conductor. Nat. Mater. 2002, 1, 81. [Google Scholar] [CrossRef]
  29. 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]
  30. Su, Y.; Chen, G.; Chen, L.; Li, Q.; Lu, Y.; Bao, L.; Li, N.; Chen, S.; Wu, F. Advances and Prospects of Surface Modification on Nickel–Rich Materials for Lithium–Ion Batteries. Chin. J. Chem. 2019, 37, 1817–1831. [Google Scholar] [CrossRef]
  31. Huang, Y.; Zhang, X.; Yu, R.; Jamil, S.; Cao, S.; Fang, S.; Wang, Y.; Tang, K.; Chen, G.; Luo, Z.; et al. Preparation and Performance of the Heterostructured Material with a Ni–Rich Layered Oxide Core and a LiNi0.5Mn1.5O4–like Spinel Shell. ACS Appl. Mater. Inter. 2019, 11, 16556–16566. [Google Scholar] [CrossRef] [PubMed]
  32. Min, K.; Jung, C.; Ko, D.-S.; Kim, K.; Jang, J.; Park, K.; Cho, E. High–Performance and Industrially Feasible Ni–Rich Layered Cathode Materials by Integrating Coherent Interphase. ACS Appl. Mater. Inter. 2018, 10, 20599–20610. [Google Scholar] [CrossRef] [PubMed]
  33. Zheng, Y.; Chen, L.; Su, Y.; Tan, J.; Bao, L.; Lu, Y.; Wang, J.; Chen, R.; Chen, S.; Wu, F. An interfacial framework for breaking through the Li−ion transport barrier of Li−rich layered cathode materials. J. Mater. Chem. A 2017, 5, 24292. [Google Scholar] [CrossRef]
  34. Chen, S.; Zheng, Y.; Lu, Y.; Su, Y.; Bao, L.; Li, N.; Li, Y.; Wang, J.; Chen, R.; Wu, F. Enhanced Electrochemical Performance of Layered Lithium−Rich Cathode Materials by Constructing Spinel−Structure Skin and Ferric Oxide Islands. Appl. Mater. Interfaces 2017, 9, 8669–8678. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, X.; Shi, J.; Liang, J.; Yin, Y.; Zhang, J.; Yu, X.; Guo, Y. Suppressing Surface Lattice Oxygen Release of Li−Rich Cathode Materials via Heterostructured Spinel Li4Mn5O12 Coating. Adv. Mater. 2018, 30, 1801751. [Google Scholar] [CrossRef] [PubMed]
  36. Park, K.; Yeon, D.; Kim, J.H.; Park, J.H.; Doo, S.; Choi, B. Spinel–embedded lithium–rich oxide composites for Li–ion batteries. J. Power Source 2017, 360, 453–459. [Google Scholar] [CrossRef]
  37. Freitas, R.R.Q.; de Brito Mota, F.; Rivelino, R.; de Castilho, C.M.C.; Kakanakova–Georgieva, A.; Gueorguiev, G.K. Spin–orbit–induced gap modification in buckled honeycomb XBi and XBi3 (X  =  B, Al, Ga, and In) sheets. J. Phys. Condens. Matter 2015, 27, 485306. [Google Scholar] [CrossRef]
  38. Bakoglidis, K.D.; Palisaitis, J.; dos Santos, R.B.; Rivelino, R.; Persson, P.O.Å.; Gueorgui, K.; Gueorguiev Hultman, L. Self–Healing in Carbon Nitride Evidenced As Material Inflation and Superlubric Behavior. ACS Appl. Mater. Interfaces 2018, 10, 16238–16243. [Google Scholar] [CrossRef] [PubMed]
  39. Wu, F.; Li, N.; Su, Y.; Zhang, L.; Bao, L.; Wang, J.; Chen, L.; Zheng, Y.; Dai, L.; Peng, J. Ultrathin Spinel Membrane−Encapsulated Layered Lithium−Rich Cathode Material for Advanced Li−Ion Batteries. Nano Lett. 2014, 14, 3550–3555. [Google Scholar]
  40. Yu, F.-D.; Que, L.-F.; Wang, Z.-B.; Zhang, Y.; Xue, Y.; Liu, B.-S.; Gu, D.-M. Layered−spinel capped nanotube assembled 3D Li−rich hierarchitectures for high performance Li−ion battery cathodes. Mater. Chem. A 2016, 4, 18416. [Google Scholar] [CrossRef]
  41. Cho, Y.; Lee, S.; Lee, Y.; Hong, T.; Cho, J. Spinel–Layered Core–Shell Cathode Materials for Li–Ion Batteries. Adv. Energy Mater. 2011, 1, 821. [Google Scholar] [CrossRef]
  42. Park, S.-H.; Kang, S.-H.; Johnson, C.-S.; Amine, K.; Thackeray, M.M. Lithium–manganese–nickel–oxide electrodes with integrated layered–spinel structures for lithium batteries. Electrochem. Commun. 2007, 9, 262. [Google Scholar] [CrossRef]
  43. Lee, E.-S.; Huq, A.; Chang, H.-Y.; Manthiram, A. High–Voltage, High–Energy Layered–Spinel Composite Cathodes with Superior Cycle Life for Lithium–Ion Batteries. Chem. Mater. 2012, 24, 600. [Google Scholar] [CrossRef]
  44. Yin, H.; Ji, S.; Gu, M.; Zhang, L.; Liu, J. Scalable Synthesis of Li1.2Mn0.54Ni0.13Co0.13O2/LiNi0.5Mn1.5O4 Sphere Composites as Stable and High Capacity Cathode for Li–Ion. RSC Adv. 2015, 5, 84673–84679. [Google Scholar] [CrossRef]
  45. Liu, W.; Liu, J.; Chen, K.; Ji, S.; Wan, Y.; Zhou, Y.; Xue, D.; Hodgson, P.; Li, Y. Enhancing the Electrochemical Performance of the LiMn2O4 Hollow Microsphere Cathode with a LiNi0.5Mn1.5O4 Coated Layer. Chem. Eur. J. 2014, 20, 824–830. [Google Scholar] [CrossRef] [PubMed]
  46. Kawai, H.; Nagata, M.; Kageyama, H.; Tukamoto, A.R.; West, A. New Lithium Cathode LiCoMnO4: Toward Practical 5 V Lithium Batteries. Electrochem. Solid State Lett. 1998, 1, 212. [Google Scholar] [CrossRef]
  47. Hu, M.; Tian, Y.; Su, L.; Wei, J.; Zhou, Z. Preparation and Ni−doping effect of nanosized truncated octahedral LiCoMnO4 as cathode materials for 5 V li−ion batteries. ACS Appl. Mater. Interfaces 2013, 5, 12185–12189. [Google Scholar] [CrossRef]
  48. Lin, H.-F.; Tsai, Y.R.; Cheng, C.-H.; Cheng, S.T.; Chen, D.Z.; Wu, N.-Y. Structural and electrochemical properties of LiCoMnO4 doped with Mg, La, and F as a high-voltage cathode material for lithium ion batteries. Electrochim. Acta 2022, 427, 140904. [Google Scholar] [CrossRef]
  49. Yang, Z.; Zhang, W.; Wang, Q.; Song, X.; Qian, Y. Synthesis of porous and hollow microspheres of nanocrystalline Mn2O3. Chem. Phys. Lett. 2006, 418, 46–49. [Google Scholar] [CrossRef]
  50. Li, Q.; Li, G.; Fu, C.; Luo, D.; Fan, J.; Xie, D.; Li, L. Balancing stability and specific energy in Li–rich cathodes for lithium ion batteries: A case study of a novel Li–Mn–Ni–Co oxide. J. Mater. Chem. A 2015, 3, 10592–10602. [Google Scholar] [CrossRef]
  51. Feng, X.; Yang, Z.; Tang, D.; Kong, Q.; Gu, L.; Wang, Z.; Chen, L. Performance improvement of Li–rich layer–structured Li1.2 Mn0.54Ni0.13Co0.13O2 by integration with spinel LiNi0.5Mn1.5O4. Phys. Chem. Chem. Phys. 2015, 17, 1257–1264. [Google Scholar]
  52. Hong, J.; Seo, D.H.; Kim, S.W.; Gwon, H.; Oh, S.T.; Kang, K. Structural evolution of layered Li1.2Ni0.2Mn0.6O2 upon electrochemical cycling in a Li rechargeable battery. J. Mater. Chem. 2010, 20, 10179–10186. [Google Scholar] [CrossRef]
  53. Yu, H.; Zhou, H. High–Energy Cathode Materials (Li2MnO3–LiMO2) for Lithium–Ion Batteries. J. Phys. Chem. Lett. 2013, 4, 1268–1280. [Google Scholar] [CrossRef] [PubMed]
  54. Baddour-Hadjean, R.; Pereira, J.P. Raman Microspectrometry Applied to the Study of Electrode Materials for Lithium Batteries. Chem. Rev. 2010, 110, 1278–1319. [Google Scholar] [CrossRef]
  55. Yi, L.; Liu, Z.; Yu, R.; Zhao, C.; Peng, H.; Liu, M.; Wu, B.; Chen, M.; Li-Rich, X.W. Layered/Spinel Heterostructured Special Morphology Cathode Material with High Rate Capability for Li-Ion Batteries. ACS Sustain. Chem. Eng. 2017, 5, 11–11005. [Google Scholar] [CrossRef]
  56. Moses, A.W.; Flores, H.G.G.; Kim, J.G.; Langell, A.M. Surface properties of LiCoO2, LiNiO2 and LiNi1-xCoxO2. Appl. Surf. Sci. 2007, 253, 4782–4791. [Google Scholar] [CrossRef]
  57. Treuil, N.; Labrugère, C.; Menetrier, M.; Portier, J.; Campet, G.; Deshayes, A.; Frison, J.-C.; Hwang, S.-J.; Song, S.-W.; Choy, J.-H. Relationship between Chemical Bonding Nature and Electrochemical Property of LiMn2O4 Spinel Oxides with Various Particle Sizes: “Electrochemical Grafting” Concept. J. Phys. Chem. B 1999, 103, 2100–2106. [Google Scholar] [CrossRef]
  58. Peng, H.; Zhao, S.-X.; Huangv, C.; Yu, L.Q.; Fang, Z.-Q.; Guo-Dan, W. In Situ Construction of Spinel Coating on the Surface of a Lithium- Rich Manganese-Based Single Crystal for Inhibiting Voltage Fade. ACS Appl. Mater. Interfaces 2020, 12, 11579–11588. [Google Scholar] [CrossRef]
  59. Thackeray, M.M.; David, W.I.F.; Bruce, P.G.; Goodenough, B.J. Lithium insertion into manganese spinels. Mater. Res. Bull. 1983, 18, 461–472. [Google Scholar] [CrossRef]
  60. Armstrong, A.R.; Holzapfel, M.; Novak, P.; Johnson, C.S.; Kang, S.-H.; Thackeray, M.M.; Bruce, P.G. Demonstrating Oxygen Loss and Associated Structural Reorganization in the Lithium Battery Cathode Li[Ni0.2Li0.2Mn0.6]O2. J. Am. Chem. Soc. 2006, 128, 8694–8698. [Google Scholar] [CrossRef]
  61. Lim, J.-H.; Bang, H.; Lee, K.-S.; Amine, K.; Sun, Y.-K. Electrochemical characterization of Li2MnO3–Li[Ni1/3Co1/3Mn1/3]O2–LiNiO2 cathode synthesized via co–precipitation for lithium secondary batteries. J. Power Source 2009, 189, 571–575. [Google Scholar]
  62. Zheng, J.; Gu, M.; Xiao, J.; Zuo, P.; Wangv, C.; Zhang, J.-G. Corrosion/fragmentation of layered composite cathode and related capacity/voltage fading during cycling process. Nano Lett. 2013, 13, 3824. [Google Scholar] [CrossRef] [PubMed]
  63. Zheng, J.; Xiao, J.; Gu, M.; Zuo, P.; Wang, C.; Zhang, J.G. Interface modifications by anion receptors for high energy lithium ion batteries. J. Power Source 2014, 250, 313–318. [Google Scholar] [CrossRef]
  64. Gu, M.; Belharouak, I.; Zheng, J.; Wu, H.; Xiao, J.; Genc, A.; Amine, K.; Thevuthasan, S.; Baer, D.R.; Zhang, J.-G.; et al. Formation of the Spinel Phase in the Layered Composite Cathode Used in Li-Ion Batteries. ACS Nano 2013, 7, 760–767. [Google Scholar] [CrossRef] [PubMed]
  65. Mohanty, D.; Kalnaus, S.; Meisner, R.A.; Rhodes, K.J.; Li, J.L.; Payzant, E.A.; Wood, D.L.; Daniel, C. Structural transformation of a lithium–rich Li1.2Co0.1Mn0.55Ni0.15O2 cathode during high voltage cycling resolved by in situ X–ray diffraction. J. Power Sources 2013, 229, 239. [Google Scholar] [CrossRef]
  66. Fell, C.R.; Chi, M.; Meng, Y.S.; Jones, J.L. In situ X–ray diffraction study of the lithium excess layered oxide compound Li[Li0.2Ni0.2Mn0.6]O2 during electrochemical cycling. Solid State Ion. 2012, 207, 44. [Google Scholar] [CrossRef]
  67. Qiu, B.; Zhang, M.H.; Xia, Y.G.; Liu, Z.P.; Meng, Y.S. Understanding and Controlling Anionic Electrochemical Activity in High-Capacity Oxides for Next Generation Li-Ion Batteries. Chem. Mater. 2017, 29, 908–915. [Google Scholar] [CrossRef]
  68. Yan, P.F.; Nie, A.; Zheng, J.M.; Zhou, Y.G.; Lu, D.P.; Zhang, X.F.; Xu, R.; Belharouak, I.; Zu, X.A.; Xiao, J.; et al. Evolution of Lattice Structure and Chemical Composition of the Surface Reconstruction Layer in Li1.2Ni0.2Mn0.6O2 Cathode Material for Lithium Ion Batteries. Nano Lett. 2015, 15, 514–522. [Google Scholar] [CrossRef]
  69. Ye, D.L.; Zeng, G.; Nogita, K.; Ozawa, K.; Hankel, M.; Searles, D.J.; Wang, L.Z. Understanding the Origin of Li2MnO3 Activation in Li-Rich Cathode Materials for Lithium-Ion Batteries. Adv. Funct. Mater. 2015, 25, 7488–7496. [Google Scholar] [CrossRef]
  70. Wu, F.; Liu, J.R.; Li, L.; Zhang, X.X.; Luo, R.; Ye, Y.S.; Chen, R.J. Surface Modification of Li-Rich Cathode Materials for Lithium-Ion Batteries with a PEDOT:PSS Conducting Polymer. ACS Appl. Mater. Interfaces 2016, 8, 23095–23104. [Google Scholar] [CrossRef]
Figure 1. SEM images of (a,b) LR–Bare, (c,d) LR–Sp1, (e,f) LR–Sp2.5, and(g,h) LR–Sp5 precursors.
Figure 1. SEM images of (a,b) LR–Bare, (c,d) LR–Sp1, (e,f) LR–Sp2.5, and(g,h) LR–Sp5 precursors.
Nanomaterials 12 03425 g001
Figure 2. SEM images of (a,b) LR–Bare, (c,d) LR–Sp1,(e,f) LR–Sp2.5, and (g,h) LR–Sp2.5 cathodes.
Figure 2. SEM images of (a,b) LR–Bare, (c,d) LR–Sp1,(e,f) LR–Sp2.5, and (g,h) LR–Sp2.5 cathodes.
Nanomaterials 12 03425 g002
Figure 3. Size distribution of each cathode.
Figure 3. Size distribution of each cathode.
Nanomaterials 12 03425 g003
Figure 4. X–ray diffraction (XRD) patterns of the LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 cathodes.
Figure 4. X–ray diffraction (XRD) patterns of the LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 cathodes.
Nanomaterials 12 03425 g004
Figure 5. Raman spectra of the LR-Bare, LR-Sp1, LR-Sp2.5, and LR-Sp5 cathodes.
Figure 5. Raman spectra of the LR-Bare, LR-Sp1, LR-Sp2.5, and LR-Sp5 cathodes.
Nanomaterials 12 03425 g005
Figure 6. XPS spectra of (a) Ni 2p, (b) Co 2p, (c) Mn 2p, and (d) O 1s for LR–Bare and LR–Sp2.5 cathodes.
Figure 6. XPS spectra of (a) Ni 2p, (b) Co 2p, (c) Mn 2p, and (d) O 1s for LR–Bare and LR–Sp2.5 cathodes.
Nanomaterials 12 03425 g006
Figure 7. Charge and discharge performance at (a) 0.1 C/0.1 C between 2–4.8 V, (b) 0.1 C/0.1 C, (c) 0.2 C/1 C and (d) 0.2 C/5 C between 2–4.6 V for the LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 cathodes.
Figure 7. Charge and discharge performance at (a) 0.1 C/0.1 C between 2–4.8 V, (b) 0.1 C/0.1 C, (c) 0.2 C/1 C and (d) 0.2 C/5 C between 2–4.6 V for the LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 cathodes.
Nanomaterials 12 03425 g007
Figure 8. Cycle life performance of LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 cathodes at 2–4.6 V.
Figure 8. Cycle life performance of LR–Bare, LR–Sp1, LR–Sp2.5, and LR–Sp5 cathodes at 2–4.6 V.
Nanomaterials 12 03425 g008
Figure 9. Electrochemical impedance spectra of LR–Bare and LR-Sp2.5 after 25 charge–discharge cycles.
Figure 9. Electrochemical impedance spectra of LR–Bare and LR-Sp2.5 after 25 charge–discharge cycles.
Nanomaterials 12 03425 g009
Figure 10. Transmission electron microscopy (TEM) images of (a,b) LR–Bare before cycle and (c,d) after 100 cycles; and (e) LR–Sp2.5 before cycle and (f,g) after 100 cycles.
Figure 10. Transmission electron microscopy (TEM) images of (a,b) LR–Bare before cycle and (c,d) after 100 cycles; and (e) LR–Sp2.5 before cycle and (f,g) after 100 cycles.
Nanomaterials 12 03425 g010
Figure 11. (a) Cycle life performance and discharge profiles for the 10th, 20th, 30th, and 40th cycles for (b) LR–Bare, (c) LR–Sp1, (d) LR–Sp2.5, and (e) LR–Sp5 cathodes at 2.0–4.9 V.
Figure 11. (a) Cycle life performance and discharge profiles for the 10th, 20th, 30th, and 40th cycles for (b) LR–Bare, (c) LR–Sp1, (d) LR–Sp2.5, and (e) LR–Sp5 cathodes at 2.0–4.9 V.
Nanomaterials 12 03425 g011aNanomaterials 12 03425 g011b
Figure 12. Differential capacity curves of (a) LR–Bare and (b) LR–Sp2.5 cathodes for the 10th, 20th, and 30th cycles.
Figure 12. Differential capacity curves of (a) LR–Bare and (b) LR–Sp2.5 cathodes for the 10th, 20th, and 30th cycles.
Nanomaterials 12 03425 g012aNanomaterials 12 03425 g012b
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, H.-F.; Cheng, S.-T.; Chen, D.-Z.; Wu, N.-Y.; Jiang, Z.-X.; Chang, C.-T. Stabilizing Li-Rich Layered Cathode Materials Using a LiCoMnO4 Spinel Nanolayer for Li-Ion Batteries. Nanomaterials 2022, 12, 3425. https://doi.org/10.3390/nano12193425

AMA Style

Lin H-F, Cheng S-T, Chen D-Z, Wu N-Y, Jiang Z-X, Chang C-T. Stabilizing Li-Rich Layered Cathode Materials Using a LiCoMnO4 Spinel Nanolayer for Li-Ion Batteries. Nanomaterials. 2022; 12(19):3425. https://doi.org/10.3390/nano12193425

Chicago/Turabian Style

Lin, Hsiu-Fen, Si-Ting Cheng, De-Zhen Chen, Nian-Ying Wu, Zong-Xiao Jiang, and Chun-Ting Chang. 2022. "Stabilizing Li-Rich Layered Cathode Materials Using a LiCoMnO4 Spinel Nanolayer for Li-Ion Batteries" Nanomaterials 12, no. 19: 3425. https://doi.org/10.3390/nano12193425

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