Next Article in Journal
The Effect of Direct and Pulsed Current in the Presence of Surfactants on the Electrodeposition of Zn–SiC Nanocomposite Coatings
Next Article in Special Issue
Strategies of Anode Materials Design towards Improved Photoelectrochemical Water Splitting Efficiency
Previous Article in Journal
Coating of Polyetheretherketone Films with Silver Nanoparticles by a Simple Chemical Reduction Method and Their Antibacterial Activity
Previous Article in Special Issue
Rational Construction of LaFeO3 Perovskite Nanoparticle-Modified TiO2 Nanotube Arrays for Visible-Light Driven Photocatalytic Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultrathin Al2O3 Coating on LiNi0.8Co0.1Mn0.1O2 Cathode Material for Enhanced Cycleability at Extended Voltage Ranges

1
College of Energy, Soochow Institute for Energy and Materials InnovationS, Soochow University, Suzhou 215006, China
2
Key Laboratory of Advanced Carbon Materials and Wearable Energy Technologies of Jiangsu Province, Soochow University, Suzhou 215006, China
3
Suzhou University of Science and Technology & Jiangsu Key Laboratory of Environmental Science and Engineering, Suzhou 215001, China
4
Department of Mechanical & Industrial Engineering, Louisiana State University, Baton Rouge, LA 70803, USA
5
Institute of Chemical Power Sources, Soochow University, Suzhou 215600, China
*
Authors to whom correspondence should be addressed.
Coatings 2019, 9(2), 92; https://doi.org/10.3390/coatings9020092
Submission received: 20 December 2018 / Revised: 23 January 2019 / Accepted: 31 January 2019 / Published: 3 February 2019
(This article belongs to the Special Issue Thin Films for Energy Harvesting, Conversion, and Storage)

Abstract

:
Ni-rich LiNi0.8Co0.1Mn0.1O2 oxide has been modified by ultrathin Al2O3 coatings via atomic layer deposition (ALD) at a growth rate of 1.12 Å/cycle. All characterizations results including TEM, SEM, XRD and XPS together confirm high conformality and uniformity of the resultant Al2O3 layer on the surface of LiNi0.8Co0.1Mn0.1O2 particles. Coating thickness of the Al2O3 layer is optimized at ~2 nm, corresponding to 20 ALD cycles to enhance the electrochemical performance of Ni-rich cathode materials at extended voltage ranges. As a result, 20 Al2O3 ALD-coated LiNi0.8Co0.1Mn0.1O2 cathode material can deliver an initial discharge capacity of 212.8 mAh/g, and an associated coulombic efficiency of 84.0% at 0.1 C in a broad voltage range of 2.7–4.6 V vs. Li+/Li in the first cycle, which were both higher than 198.2 mAh/g and 76.1% of the pristine LiNi0.8Co0.1Mn0.1O2 without the Al2O3 protection. Comparative differential capacity (dQ/dV) profiles and electrochemical impedance spectra (EIS) recorded in the first and 100th cycles indicated significant Al2O3 ALD coating effects on suppressing phase transitions and electrochemical polarity of the Ni-rich LiNi0.8Co0.1Mn0.1O2 core during reversible lithiation/delithiation. This work offers oxide-based surface modifications with precise thickness control at an atomic level for enhanced electrochemical performance of Ni-rich cathode materials at extended voltage ranges.

1. Introduction

Lithium ion batteries have been dominantly applied in hybrid electric vehicles (HEVs) and pure electric vehicles (EVs) [1,2], which significantly enhances our living environment. In particular for electric transportation systems and future large-scale applications of stationary grid energy storage, lithium ion batteries are required to provide higher energy and power densities, together with excellent long-term service life and acceptable cost [3]. Performance of lithium ion batteries critically depends on the electrochemical properties and behaviors of electrode materials both in cathodes and anodes [4,5,6]. Because of the rapid development of practical Si–C composite anode materials, developing high-performance cathode materials with high specific capacity, high working potential, excellent cycle life, reliable safety, and low cost has become an urgent requirement [7,8,9]. Li-excess and Ni-rich layered oxides have been extensively investigated as high-energy cathode materials for superior lithium ion batteries [10,11]. Despite delivering an attractive discharge capacity of ~250 mAh/g, Li-excess layered cathode materials have been suffering from two crucial issues. One is the layered-to-spinel phase transition during prolonged electrochemical cycles, leading to continuous voltage fading. The other is low coulombic efficiency in the first cycle that results from the initial electrochemical activation of the Li+-inactive Li2MnO3 component [10,12,13]. These two problems impede practical applications of Li-excess cathode materials. Alternatively, Ni-rich layered oxides have been considered as feasible cathode materials by offering a reversible capacity of ~200 mAh/g and a stable working voltage at ~3.8 V vs. Li+/Li.
It is well-known that the resultant electrochemical performance of Ni-rich cathode materials is essentially determined on the specific composition of Ni, Co, and Mn elements within these ternary transition metal oxides, LiNi1−xyCoxMnyO2 (x + y < 0.5) [11,14,15]. Higher Ni content results in higher lithium storage capacity of LiNi1−xyCoxMnyO2, but leads to the undesired formation of Li+/Ni2+ cation mixing in the layered structure; the expense of safety in material usage and more complexity in the material preparation [16,17]. Mn and Co dopants have been demonstrated to enhance the structural robustness and cationic ordering for LiNi1−xyCoxMnyO2 variants, respectively. However, the specific capacity has to be sacrificed, when Mn and Co contents are increased. In order to rapidly promote practical applications of Ni-rich cathode materials for superior lithium ion batteries, poor cycling and thermal stabilities of LiNi1−xyCoxMnyO2 should be urgently addressed, especially for materials with high Ni content. Capacity fading can be exacerbated, when these Ni-rich LiNi1−xyCoxMnyO2 cathodes are cycled at broad voltage ranges with a high upper cut-off voltage above 4.3 V vs. Li+/Li. One reason is that the intensive oxidation capability of Ni3+/4+ redox could aggravate parasitic side reactions during lithiation/delithiation of LiNi1−xyCoxMnyO2 cathode materials, leading to undesired electrolyte decomposition and solid electrolyte interphase (SEI) formation at the surface of the cathode. The other reason is the induced phase transition from the initial layered structure to spinel-like and/or NiO-like rock-salt phases at the surface of LiNi1−xyCoxMnyO2 particles, because the formation of Li+ vacancies in high concentration at deep charge destabilizes the initial layered structure of LiNi1−xyCoxMnyO2 [18]. These unfavorable structural transformations are consecutive, and gradually overspread from the surface to bulk Ni-rich layered oxides [17,19]. In addition, the formation of impure phases definitely brings about structural degradation and increased kinetic barrier for lithium ion diffusion, causing capacity fading and inferior rate capability of LiNi1−xyCoxMnyO2 cathode materials [20,21].
Surface modifications, either on particles of active cathode materials or at the surface of the entire cathode, have been demonstrated to improve electrochemical performance for various cathode materials even in harsh operation conditions, such as elevated working temperatures and broad voltage ranges [22,23]. Oxides are considered common coating materials, because of their low material cost, simple synthetic procedure, and high effectiveness in protecting internal electrode materials. Various examples have been reported in the literature [24,25,26,27,28], involving Al2O3, ZrO2, ZnO, TiO2, Cr2O3, SnO2, etc. However, associated electrochemical performance of modified cathode materials is sensitively determined by qualities of the coated oxides in terms of thickness, conformality, uniformity, and integrity. Oxide coatings prepared by atomic layer deposition (ALD) have been demonstrated to considerably enhance electrochemical performance for various cathode materials, involving layered LiCoO2 [29,30], spinel LiMn2O4, and LiMn1.5Ni0.5O4 [31,32]; Ni-rich LiNi1−xyCoxMnyO2 and LiNi0.8Co0.15Al0.05O2 [33,34]; and Li-excess Li[Li1−xMx]O2 oxides [35]. The ALD technique can result in highly conformational and uniform oxide coating layers at the entire surface of the underlying substrates, and thus has been considered as an effective method for surface modifications of different cathode materials. Ni-rich cathode materials have been dominantly used for lithium ion batteries to power currently-developed electric automobiles, but still suffer from poor service life, especially in harsh operation conditions, such as at elevated working temperatures and/or in extended voltage ranges. As reported in the literature [33,34], Al2O3 ALD coatings contribute to enhanced cycling stabilities of LiNi0.5Co0.2Mn0.3O2, LiNi0.8Co0.1Mn0.1O2, and LiNi0.8Co0.15Al0.05O2 by protecting coated Ni-rich oxide particles from parasitic side reactions at the cathode’s surface. However, the insulating property of Al2O3 material makes it is necessary to precisely control the coating’s thickness, in order to maximize the electrochemical performance of Ni-rich cathode materials. In addition, it is also urgent to understand structural features and chemical environments between oxide coatings and Ni-rich oxide particles.
The ALD technique was employed in this work to coat ultrathin Al2O3 onto Ni-rich LiNi0.8Co0.1Mn0.1O2 target material (marked as n Al2O3 ALD NCM811, then responses to the number of ALD cycles related to the thickness control of Al2O3 coating layer). TEM observations were used to examine the structural characteristics of resultant Al2O3 coatings and determine the ALD growth rate of Al2O3 oxide. XPS spectra were used to study chemical environments of different n Al2O3 ALD NCM811 (n = 10, 20 and 50) materials. Electrochemical performance of three n Al2O3 ALD NCM811 cathode materials were evaluated and compared with the bare NCM811 cathode in a broad voltage range of 2.7–4.6 V vs. Li+/Li. This work offers a high-quality oxide coating for Ni-rich cathode materials by using the ALD method, which significantly contributes to improved electrochemical performance in extended voltage ranges for superior lithium ion batteries.

2. Materials and Methods

2.1. Synthesis of Ni-Rich Layered LiNi0.8Co0.1Mn0.1O2 Material

In a typical sol (solution)-gel route, stoichiometric amounts of CH3COOLi·2H2O, Ni(CH3COO)2·4H2O, Co(CH3COO)2·4H2O, and Mn(CH3COO)2·4H2O were dissolved in 50 mL of distilled water. An extra 3 mol.% of the lithium source was used to compensate for the volatilization of lithium at high heating temperatures. The green sol was obtained after the solvent was completely evaporated, and was then transferred to dry in vacuum at 120 °C for 12 h. Dried gel heat treatments were carried out at 850 °C for 5 h, under pure O2 flow at a temperature ramp of 5 °C/min. The resultant black LiNi0.8Mn0.1Co0.1O2 (marked as NCM811) powder was collected after cooling to room temperature.

2.2. Al2O3 Coating on LiNi0.8Co0.1Mn0.1O2 Particles Through Atomic Layer Deposition

The atomic layer deposition of Al2O3 coating on LiNi0.8Co0.1Mn0.1O2 particles was performed in a Savannah 100 ALD system (Cambridge NanoTech Inc., Waltham, MA, USA) at 120 °C, by applying Al(CH3)3 (Trimethylalumium, TMA) and H2O as precursors with exposure time of 0.01 s, waiting time of 5 s, and purge time of 40 s, respectively. Nitrogen gas was used as the flow gas, and the flow rate was set at 20 sccm. The vacuum condition in the reaction chamber was controlled at 10−3–10−2 torr, and the overpressure threshold for the coating processes was set at 250 torr. Two self-terminating reactions involved in the ALD growth of Al2O3 layer are described as follows:
AlOH* + Al(CH3)3 → Al–O–Al–(CH3)2* + CH4
Al–CH3* + H2O → AlOH* + CH4
LiNi0.8Co0.1Mn0.1O2 particles coated with the Al2O3 layer under different ALD cycles were marked as n Al2O3 ALD NCM811 (the n corresponds to the number of ALD cycles).

2.3. Characterizations

Crystallographic structures of bare NCM811 and Al2O3-coated NCM811 materials were examined by X-ray diffraction (XRD), using a Bruker D8 Advance automatic diffractometer (Bruker, Billerica, MA, USA) with Cu Kα radiation. Morphologies and energy dispersive spectroscopic (EDS) mappings of different samples were observed by field emission scanning electron microscopy (FESEM) on a Hitachi SU-8010 (Hitachi High-Tech Corp., Tokyo, Japan) and FEI Quanta FEG 250 microscopies (FEI, Hillsboro, OR, USA), respectively, with an acceleration voltage at 15/20 kV in a secondary electrons (SE) detection mode. The Everhart Thornley Detector (ETD) was used to collected electron signals. Detailed structural features of different specimens were captured by transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) coupled with selected area electron diffraction (SAED) on a FEI Tecnai G2 T20 equipment (FEI, Hillsboro, OR, USA), at an acceleration voltage of 200 kV. Chemical environments and valent states of cations within different materials were characterized by X-ray photoelectron spectroscopic (XPS) measurements on an ESCALAB 250Xi XPS device (Thermo Fisher Scientific, Waltham, MA USA). All XPS spectra were calibrated according to the binding energy of the C 1s peak at 284.6 eV.

2.4. Electrochemical Measurements

The working electrodes were composed of 80 wt.% active cathode materials, 10 wt.% super p as the conductive carbon, and 10 wt.% polyvinylidenefluoride (PVDF) as the binder. These cathodes were assembled into two-electrode CR2025-type coin cells (Shenzhen Kejing Star Technology Co., LTD., Shenzhen, China), in an Ar-filled glove box (MBraun, Garching, Germany) for electrochemical measurements, with the metallic lithium foil as the reference, and counter electrodes with the glass microfiber (Whatman, Grade GF/B, Little Chalfont, UK) as the separator. The electrolyte was a 1 M LiPF6 solution dissolved in ethylene carbonate (EC) and dimethyl carbonate (DMC) at a volumetric ratio of 1:1. The real loading of active material was ~3.5 mg in the cathode with a diameter of 12 mm and a thickness of ~50 mm for all electrode materials. The electrolyte volume used during the coin cell assembly was ~0.5 mL per unit. Galvanostatic charge and discharge were performed at different current densities in a voltage range of 2.7–4.6 V vs. Li+/Li on a LAND battery testing system (Jinnuo, Wuhan, China). The current density corresponding to 1 C was 200 mA/g. Electrochemical impedance spectroscopy (EIS) was conducted in a frequency range of 10 mHz–100 kHz with an AC amplitude of 5 mV on the Autolab electrochemical workstation (PGSTAT302N, Metrohm Autolab, Utrecht, The Netherlands).

3. Results and Discussion

3.1. Characteristics of Al2O3 Layer Deposited on NCM811 Particles via ALD Coating

As shown in the schematic illustrations in Figure 1a, the ALD growth of Al2O3 coating at the surface of NCM811 particles was subjected to a sequence of chemisorption and self-terminating surface reactions (Equations (1) and (2) in Section 2.2). Detailed deposition mechanisms for the different oxides, involving Al2O3, ZnO and ZrO2, can be found in our previous work [12,36,37,38,39,40]. As a result, ultrathin oxide coatings under ALD growth always reveal desirable conformality, uniformity, robustness, and pinhole-free features. Furthermore, thickness of the resultant oxide coatings can be precisely controlled at the Angstrom or atomic monolayer level, which is determined by the number of ALD cycles. Accordingly, morphologic and structural differences of pristine NCM811, 20 Al2O3 ALD NCM811, and 50 Al2O3 ALD NCM811 materials are compared in Figure 1b–j. The pristine NCM811 material (Figure 1b) reveals a distinct aggregation of tremendous primary nanoparticles, which have smooth facets and sharp edges with an average particle size of ~500 nm. Such morphologic features generally indicate the high crystallinity and purity of the prepared NCM811 particles. By contrast, FESEM images of the 20 Al2O3 ALD NCM811 particles showed apparent roughness at the entire surface after coating the Al2O3 layer in 20 ALD cycles (Figure 1c), implying the conformal and uniform surface modification of the NCM811 substrate. The elemental energy dispersive spectroscopic (EDS) mapping was captured on the selected 20 Al2O3 ALD NCM811 particles (Figure 1d), which revealed even distributions of Al, Ni, Co, Mn, and O elements, and thus again demonstrated desirable conformality and uniformity of Al2O3 coatings at the surface of NCM811 particles. In order to comprehensively understand structural characteristics of the resultant Al2O3 layer, 50 Al2O3 ALD NCM811 particles with increased coating thickness were examined by TEM observations, and compared to bare NCM811 and the 20 Al2O3 ALD NCM811 materials. As shown in Figure 1e, TEM images of NCM811 particles showed consistent contrast throughout the whole observation area, revealing a single solid structure without any surface modifications. HRTEM images of the selected region in Figure 1e, marked by a blue square, showed perfect lattice fringes with a d-space of 0.47 nm (Figure 1h), corresponding to (003) planes in the layered structure of NCM811 oxide. By contrast, TEM images of the 20 Al2O3 ALD NCM811 and 50 Al2O3 ALD NCM811 particles captured in Figure 1f,g gave rise to distinct contrast differences between the dark NCM811 core and light Al2O3 coating shells at the surface, respectively. It was noticeable that the ALD coating contributed to satisfactory uniformity and conformality of the resultant Al2O3 layer, which was in accordance with that coated at the surface of different cathode materials reported in the literature [36,38]. The precise thickness was 5.6 nm as shown in the HRTEM image of the 50 Al2O3 ALD NCM811 (Figure 1j), corresponding to an Al2O3 ALD growth rate of 1.12 Å/cycle. Accordingly, 20 ALD cycles resulted in an Al2O3 coating with 2.2 nm thickness as shown in Figure 1i for the 20 Al2O3 ALD NCM811 materials. Furthermore, these two HRTEM observations indicated an amorphous feature of the Al2O3 coating layer, which was distinguishable from the internal crystalline NCM811 particles. As a result, SEM and TEM images demonstrated the conformational and uniform coating of the Al2O3 layer for the surface modification of Ni-rich NCM811 particles through ALD processes, together with thickness control at the sub-nano level.
Al2O3 coating effects on changing the crystal structure of Ni-rich NCM811 core were evaluated using XRD patterns of different coated materials, coupled with quantitative analyses from Rietveld refinements. As shown in Figure 2a, the XRD pattern of pristine NCM811 material was well indexed to a layered α-NaFeO2 structure with a R-3m space group (JCPDS No. 74-0919). It also showed a desirable intensity ratio of characteristic 003 over 104 reflections, i.e., I003/I104 = 1.48 and distinct peak splitting of two 006/102 and 108/110 doublets, indicating outstanding structural ordering of the layered NCM811 material. The Rietveld refinement on the XRD pattern of pristine NCM811 resulted in lattice parameters of a = 2.8746 Å, c = 14.2102 Å, and V = 101.72 Å3, together with a low estimated Li+/Ni2+ cation mixing of 3.68% based on a low reliability-factor of Rwp = 6.62% (Table 1). These XRD features all indicate an excellent layered structure of the prepared Ni-rich NCM811 material, in accordance with HRTEM images (Figure 1h). XRD patterns and the associated refinement results of NCM811 particles coated with 10, 20, and 50 Al2O3 ALD layers are displayed in Figure 2b–d, respectively. The calculated lattice parameters and cation mixings of these three coated materials, with gradually-increasing coating thickness of Al2O3, are summarized in Table 1 compared to bare NCM811, together with values of standard deviations (s.d.) for the calculations. The Al2O3 phase was not detectable even in the 50 Al2O3 ALD NCM811 samples, owing to the amorphous phase. However, the Al2O3 coating slightly suppressed the crystal structural of the internal NCM811 host. Coatings thicker than 2 nm, i.e., 20 ALD cycles (20 cycle* 1.12 Å/cycle = 2.24 nm), resulted in a detectable decrease in a, c, and V values, according to the refinement results of the 20 Al2O3 ALD NCM811 sample, while the 10 Al2O3 ALD NCM811 showed little change in lattice parameters as compared with that in the pristine NCM811 (Table 1). Although the thick Al2O3 coating reduced a and c values of the 50 Al2O3 ALD NCM811 material, its c/a ratio was almost the same with the NCM811, again indicating high conformality and uniformity of surface modifications via oxide ALD coatings. On the other hand, XRD patterns of the three Al2O3-coated NCM811 materials revealed I003/I104 values equal to 1.54, 1.46, and 1.45 after 10, 20, and 50 ALD cycles (Figure 2b–d), respectively, indicating little effects of Al2O3 ALD coating on the structural ordering of internal NCM811 material. The estimated cation mixings within the layered structures of these three samples wre all in an acceptable range of 3%–4%, consistent with that of the pristine NCM811 material (Table 1).
In order to confirm Al2O3 coating formation and its possible impacts on the chemical environments of transition metal cations at the surface of NCM811 particles, characteristic Al 2p, Ni 2p, Co 2p, and Mn 2p XPS peaks of bare NCM811 and three Al2O3-coated NCM811 materials were measured and compared. As shown in Figure 3a, 10 Al2O3 ALD NCM811, 20 Al2O3 ALD NCM811, and 50 Al2O3 ALD NCM811 samples all showed sharp Al 2p XPS peaks, verifying successful Al2O3 coating at the surface of NCM811 particles. These collected Al 2p peaks slightly shifted to higher binding energies at 75.4, 75.6, and 75.7 eV, respectively, compared to 74.6 eV of pure Al2O3 oxide reported in NIST XPS Database, implying possible chemical bonding between the Al2O3 coating layer and NCM811 particles. As reported in the literature [36,37,38,39,40], the oxide coating and substrate in a typical ALD growth can result in a favorable chemical bond of substrate–O–MxOy (M = Al, Zn, Zr, Sn, etc.), this process is illustrated in Figure 1a. However, it is unexpected to see that all the characteristic of Ni 2p, Co 2p, and Mn 2p XPS peaks captured on three Al2O3-coated NCM811 particles also gave rise to higher bonding energies compared to those of pristine NCM811 particles, when the coating thickness of Al2O3 was gradually increased (Figure 3b–d). These interesting XPS results may be attributed to the tremendous –OH groups that were left at the surface of Al2O3 coating at the end of ALD growth (Figure 1a), which are highly electronegative, and thus probably induce reduced electron cloud densities of Al, Ni, Co, and Mn cations. The thicker Al2O3 coating resulted in increased surface area of Al2O3-coated NCM811 particles as well as the quantity of residual –OH groups at the most outside surface; hence, the inductive effect was enhanced by the increased number of –OH groups, leading to binding energy shifts of different cations to higher values. XPS results indicate a probable chemical bonding between Al2O3 and NCM811, which may facilitate charge transfer and lithium ion diffusion through the oxide coating layer.

3.2. Enhanced Electrochemical Performance of Ni-Rich NCM811 Cathode Materials by Al2O3 ALD Coating

Effects of Al2O3 ALD coating on improving electrochemical performance of Ni-rich NCM811 cathode material are evaluated in Figure 4, where different cathode materials were subjected to a broad voltage range of 2.7–4.6 V vs. Li+/Li. Figure 4a compares resultant cycling performance of the Al2O3-coated NCM811 cycled at 0.1 C with bare NCM811 material. The coating thickness of Al2O3 layer was optimized at ~2 nm, corresponding to 20 ALD cycles. As a result, the 20 Al2O3 ALD NCM811 cathode delivered an impressive discharge capacity of 212.8 mAh/g and an associated initial columbic efficiency of 84.0% in the first cycle, which were both higher than 198.2 mAh/g and 76.1% of the pristine NCM811 without Al2O3 protection. Accordingly, the charge/discharge curves of these two cathode materials cycled at 1st and 100th cycles are shown in Figure 4b. The improved electrochemical reversibility with increased specific capacity of the 20 Al2O3 ALD NCM811 in the first cycle can be attributed to the conformal surface modification from the Al2O3 coating (Figure 1a), which effectively suppressed detrimental side reactions even when the NCM811 cathode material was cycled in an extended voltage range up to 4.6 V. In addition to protecting the internal Ni-rich core material from the HF attack in electrolytes as reported in the literature [29,41], the conformational and uniform Al2O3 coating also reduced the electrolyte decomposition induced by high oxidation of Ni3+/Ni4+ redox, especially under the deeply-charged condition at a high cut-off voltage of 4.6 V. As shown in Figure 4a,b, the pristine NCM811 cathode material suffered from distinct capacity decay to 127.5 mAh/g after 100 electrochemical cycles, while the 20 Al2O3 ALD NCM811 retained a much higher capacity of 157.2 mAh/g in the 100th cycle. Furthermore, the NCM811 also encountered an obvious working voltage fading from 3.80 initially to 3.76 V in the 100th cycle. By contrast, electrochemical polarity was effectively inhibited in the 20 Al2O3 ALD NCM811 cathode material, resulting in a well-preserved working voltage from 3.84 V in the first cycle to 3.83 V after 100 cycles. The 10 Al2O3 ALD NCM811 delivered moderate cycling performance as shown in Figure 4a, between the NCM811 and 20 Al2O3 ALD NCM811, which was attributed to the thinner Al2O3 coating around 1 nm. On the other hand, the much thicker coating of Al2O3 layer (>5 nm) after 50 ALD cycles resulted in satisfactory cycling stability, but the lithium storage capacity had to be sacrificed (Figure 4a), owing to the inferior conductivity of insulating Al2O3 oxide [42,43]. According to reports in the literature [13,44], Ni-rich cathode materials undergo inevitable phase transitions among three hexagonal (H1, H2, and H3) and one monoclinic (M) phases in their layered structures during lithiation/delithiation, which further aggravate undesired capacity fading in long-term electrochemical cycles. In order to examine possible phase transitions, related differential capacity (dQ/dV) profiles are shown in Figure 4c, which were transferred from charge/discharge curves in Figure 4b. The initial anodic peak of pristine NCM811 showed the much higher intensity by contrast with 20 Al2O3 ALD NCM811, indicating severe parasitic side reactions that occurred during the first phase transformation from the hexagonal H1 to monoclinic M phases. Subsequent M to H2, and H2 to H3 took place around at 4.0 and 4.2 V as marked in Figure 4c, respectively, when the Ni-rich cathode materials were charged to a high cutoff voltage. In successive discharge processes, phase transitions from H3 to H2 to M were reduced, resulting in dominant cathodic peaks around 3.75 V both in NCM811 and 20 Al2O3 ALD NCM811 cathodes. However, two redox pairs at higher voltages around 4.0 and 4.2 V were detected in the 100th cycle of pristine NCM811 cathode (peak voltages of detected anodic and cathodic peaks are marked by underlines), while the 20 Al2O3 ALD NCM811 gave rise to the smoother dQ/dV curves above 3.9 V, revealing significantly suppressed phase transitions for its enhanced cycling stability after the optimal Al2O3 ALD coating. Furthermore, total resistance changes of these two cathodes after 100 cycles were determined by EIS spectra as shown in Figure 4d. These measurements were conducted on their fully-discharged states after prolonged electrochemical cycles. EIS spectra of fresh cells were also recorded at open circuit voltage states for the comparison. It is worth noting that two fresh cells of NCM811 and the 20 Al2O3 ALD NCM811 cathodes both showed small resistances around 25 Ω. However, two Nyquist circles with larger diameters were captured in both cathodes after 100 cycles. The selected equivalent circuit, as the inset, was used to calculate the associated resistances based on fitting solid lines. For the pristine NCM811, one Nyquist circle in high frequency may be attributed to impure phases, such as spinel-like and/or rock-salt phases formed during prolonged cycles as reported in the literature [45,46], and others in low frequency were assigned to the solid electrolyte interphase (SEI) film. By contrast, two Nyquist circles in 20 Al2O3 ALD NCM811 resulted from the Al2O3 coating and SEI film at the surface of cycled cathode material, respectively. It is noticeable that introduction of 2 nm Al2O3 coating led to acceptable resistance at 29 Ω, which was smaller than the 35 Ω caused by the secondary impure structure that formed at the surface of NCM811 without surface modification. Moreover, the 20 Al2O3 ALD NCM811 resulted in considerable decrease (160 Ω) corresponding to SEI film resistance, while the pristine NCM811 brought about larger SEI resistance (250 Ω). Electrochemical performance demonstrated significant effects of the optimized Al2O3 ALD coating on reducing electrochemical polarity and improving reversibility of Ni-rich cathode materials for superior lithium ion batteries.

4. Conclusions

Enhanced cycling stability of Ni-rich LiNi0.8Co0.1Mn0.1O2 cathode material in an extended voltage range has been realized by surface modification of ultrathin Al2O3 coating through atomic layer deposition. The oxide ALD growth leads to impressive conformality and uniformity of resultant Al2O3 layer, and the coating thickness can be precisely controlled at 1.12 Å per ALD cycle. The optimal thickness of the Al2O3 coating for LiNi0.8Co0.1Mn0.1O2 cathode material is in 2 nm level through performing 20 ALD cycles. Either the thicker or thinner coating thickness leads to decreased lithium storage capacity together with inferior cycling performance of different Al2O3 coated LiNi0.8Co0.1Mn0.1O2 cathodes. The optimized 20 Al2O3 ALD modified LiNi0.8Co0.1Mn0.1O2 cathode material distinctly shows increased discharge capacity and columbic efficiency in the first cycle, as well as improved capacity retention after successive 100 cycles in comparison with the pristine material. Electrochemical features in the first and 100th cycle of cathode materials, with and without Al2O3 protection were compared, involving charge/discharge curves, differential capacity profiles, and electrochemical impedance spectra. The Al2O3 coating has been demonstrated to suppress parasitic side reactions and phase transitions during prolonged lithiation/delithiation of the LiNi0.8Co0.1Mn0.1O2 cathode, resulting in reduced electrochemical polarity and enhanced cycling stability for Ni-rich cathode materials cycled in broad voltage ranges.

Author Contributions

Conceptualization, J.Z.; Methodology, W.Z., T.L., Z.X., and X.H.; Software, L.B.; Data Curation, W.Z., X.H., L.B., and K.T.; Writing—Original Draft Preparation, W.Z.; Writing—Review and Editing, Y.W., H.W., J.Z., and L.G.; Project Administration, L.G. and J.Z.; Funding Acquisition, L.G. and J.Z.

Funding

This research was funded by the National Natural Science Foundation of China (Nos. 21703147 and U1401248); the Jiangsu Provincial Natural Science Foundations for the Young Scientist (No. BK20170338); the General Financial Grant from the China Postdoctoral Science Foundation (No. 2016M601876); and the Open Fund of Jiangsu Key Laboratory of Materials and Technology for Energy Conversion (No. MTEC-2017M01).

Acknowledgments

The authors acknowledge Suzhou Key Laboratory for the Advanced Carbon Materials and Wearable Energy Technologies, Suzhou 215006, China.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Poizot, P.; Laruelle, S.; Grugeon, S.; Grugeon, L.; Tarascon, J.M. Nano-sized transition-metal oxides as negative-electrode materials for Lithium-ion batteries. Nature 2000, 407, 496–499. [Google Scholar] [CrossRef]
  2. Mizushima, K.; Jones, P.C.; Wiseman, P.J.; 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] [CrossRef]
  3. 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]
  4. Tang, Y.; Deng, J.; Li, W.; Malyi, O.I.; Zhang, Y.; Zhou, X.; Pan, S.; Wei, J.; Cai, Y.; Chen, Z. Water-soluble sericin protein enabling stable solid-electrolyte interphase for fast charging high voltage battery electrode. Adv. Mater. 2017, 29, 1701828. [Google Scholar] [CrossRef]
  5. Tang, Y.; Zhang, Y.; Malyi, O.I.; Bucher, N.; Xia, H.; Xi, S.; Zhu, Z.; Lv, Z.; Li, W.; Wei, J. Identifying the origin and contribution of surface storage in TiO2(B) nanotube electrode by in situ dynamic valence state monitoring. Adv. Mater. 2018, 36, 1802200. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Malyi, O.I.; Tang, Y.; Wei, J.; Zhu, Z.; Xia, H.; Li, W.; Guo, J.; Zhou, X.; Chen, Z. Reducing the charge carrier transport barrier in functionally layer-graded electrodes. Angew. Chem. 2017, 56, 14847. [Google Scholar] [CrossRef]
  7. Yi, R.; Dai, F.; Gordin, M.L.; Chen, S.; Wang, D. Lithium-ion batteries: micro-sized Si–C composite with interconnected nanoscale building blocks as high-performance anodes for practical application in Lithium-ion batteries. Adv. Energy Mater. 2013, 3, 273. [Google Scholar] [CrossRef]
  8. Fergus, J.W. Recent developments in cathode materials for Lithium ion batteries. J. Power Sources 2010, 195, 939–954. [Google Scholar] [CrossRef]
  9. Lee, J.; Kitchaev, D.A.; Kwon, D.H.; Lee, C.W.; Papp, J.K.; Liu, Y.S.; Lun, Z.; Clément, R.J.; Shi, T.; Mccloskey, B.D. Reversible Mn2+/Mn4+ double redox in Lithium-excess cathode materials. Nature 2018, 556, 185–190. [Google Scholar] [CrossRef]
  10. Jiang, K.C.; Wu, X.L.; Yin, Y.X.; Lee, J.S.; Kim, J.; Guo, Y.G. Superior hybrid cathode material containing Lithium-excess layered material and graphene for Lithium-ion Batteries. ACS Appl. Mater. Interfaces 2012, 4, 4858–4863. [Google Scholar] [CrossRef]
  11. Zhao, J.; Zhang, W.; Huq, A.; Misture, S.T.; Zhang, B.; Guo, S.; Wu, L.; Zhu, Y.; Chen, Z.; Amine, K. In situ probing and synthetic control of cationic ordering in Ni-Rich layered oxide cathodes. Adv. Energy Mater. 2017, 7, 1601266. [Google Scholar] [CrossRef]
  12. Zhao, J.; Aziz, S.; Wang, Y. Hierarchical functional layers on high-capacity Lithium-excess cathodes for superior Lithium ion batteries. J. Power Sources 2014, 247, 95–104. [Google Scholar] [CrossRef]
  13. Zhao, J.; Huang, R.; Gao, W.; Zuo, J.M.; Zhang, X.F.; Misture, S.T.; Chen, Y.; Lockard, J.V.; Zhang, B.; Guo, S. An Ion-exchange promoted phase transition in a Li-excess layered cathode material for high-performance Lithium ion batteries. Adv. Energy Mater. 2015, 5, 1401937. [Google Scholar] [CrossRef]
  14. Martha, S.K.; Sclar, H.; Framowitz, Z.S.; Kovacheva, D.; Saliyski, N.; Gofer, Y.; Sharon, P.; Golik, E.; Markovsky, B.; Aurbach, D. A comparative study of electrodes comprising nanometric and submicron particles of LiNi0.5Mn0.5O2, LiNi0.33Mn0.33Co0.33O2, and LiNi0.4Mn0.4Co0.2O2 layered compounds. J. Power Sources 2009, 189, 248–255. [Google Scholar] [CrossRef]
  15. Li, X.; Cheng, F.; Guo, B.; Chen, J. Template-synthesized LiCoO2, LiMn2O4, and LiNi0.8Co0.2O2 nanotubes as the cathode materials of Lithium ion batteries. J. Phys. Chem. B 2005, 109, 14017–14024. [Google Scholar] [CrossRef]
  16. Liu, W.; Oh, P.; Liu, X.; Lee, M.J.; Cho, W.; Chae, S.; Kim, Y.; Cho, J. Nickel-rich layered Lithium transition-metal oxide for high-energy Lithium-ion batteries. Angew. Chem. 2015, 54, 4440–4457. [Google Scholar] [CrossRef]
  17. Zhong, S.; Lai, M.; Yao, W.; Li, Z. Synthesis and electrochemical properties of LiNi0.8CoxMn0.2−xO2 positive-electrode material for Lithium-ion batteries. Electrochim. Acta 2016, 212, 343–351. [Google Scholar] [CrossRef]
  18. Zou, Y.; Yang, X.; Lv, C.; Liu, T.; Xia, Y.; Shang, L.; Waterhouse, G.I.; Yang, D.; Zhang, T. Multishelled Ni-Rich Li(NixCoyMnz)O2 hollow fibers with low cation mixing as high-performance cathode materials for Li-ion batteries. Adv. Sci. 2017, 4, 1600262. [Google Scholar] [CrossRef]
  19. Zhang, X.; Jiang, W.J.; Mauger, A.; QILU; Gendron, F.; Julien, C.M. Minimization of the cation mixing in Li1+x(NMC)1−xO2 as cathode material. J. Power Sources 2009, 195, 1292–1301. [Google Scholar] [CrossRef]
  20. Manthiram, A.; Song, B.; Li, W. A perspective on nickel-rich layered oxide cathodes for Lithium-ion batteries. Energy Storage Mater. 2017, 6, 125–139. [Google Scholar] [CrossRef]
  21. Zheng, J.; Kan, W.H.; Manthiram, A. Role of Mn content on the electrochemical properties of nickel-rich layered LiNi(0.8−x)Co(0.1)Mn(0.1+x)O2 (0.0 ≤ x ≤ 0.08) cathodes for lithium-ion batteries. ACS Appl. Mater. Interfaces 2015, 7, 6926–6934. [Google Scholar] [CrossRef]
  22. Oh, P.; Oh, S.M.; Li, W.; Myeong, S.; Cho, J.; Manthiram, A. High-performance heterostructured cathodes for Lithium-ion batteries with a Ni-rich layered oxide core and a Li-rich layered oxide shell. Adv. Sci. 2016, 3, 1600184. [Google Scholar] [CrossRef]
  23. Song, B. A facile cathode design combining Ni-rich layered oxides with Li-rich layered oxides for Lithium-ion batteries. J. Power Sources 2016, 325, 620–629. [Google Scholar] [CrossRef]
  24. Schipper, F.; Bouzaglo, H.; Dixit, M.; Erickson, E.M.; Weigel, T.; Talianker, M.; Grinblat, J.; Burstein, L.; Schmidt, M.; Lampert, J. From surface ZrO2 coating to bulk Zr doping by high temperature annealing of nickel-rich lithiated oxides and their enhanced electrochemical performance in Lithium ion batteries. Adv. Energy Mater. 2018, 8, 1701682. [Google Scholar] [CrossRef]
  25. Tan, L.; Liu, H. Influence of ZnO coating on the structure, morphology and electrochemical performances for LiNi1/3Co1/3Mn1/3O2 material. Russ. J. Electrochem. 2011, 47, 156–160. [Google Scholar] [CrossRef]
  26. Li, J.; Fan, M.; He, X.; Zhao, R.; Jiange, C.; Wan, C. TiO2 coating of LiNi1/3Co1/3Mn1/3O2 cathode materials for Li-ion batteries. Ionics 2006, 12, 215–218. [Google Scholar] [CrossRef]
  27. Cheng, C.; Yi, H.; Chen, F. Effect of Cr2O3 Coating on LiNi1/3Co1/3Mn1/3O2 as cathode for Lithium-ion batteries. J. Electron. Mater. 2014, 43, 3681–3687. [Google Scholar] [CrossRef]
  28. Hudaya, C.; Ji, H.P.; Lee, J.K.; Choi, W. SnO2-coated LiCoO2 cathode material for high-voltage applications in Lithium-ion batteries. Solid State Ionics 2014, 256, 89–92. [Google Scholar] [CrossRef]
  29. Li, X.; Liu, J.; Meng, X.; Tang, Y.; Banis, M.N.; Yang, J.; Hu, Y.; Li, R.; Cai, M.; Sun, X. Significant impact on cathode performance of Lithium-ion batteries by precisely controlled metal oxide nanocoatings via atomic layer deposition. J. Power Sources 2014, 247, 57–69. [Google Scholar] [CrossRef]
  30. Wang, Z.; Liu, L.; Chen, L.; Huang, X. Structural and electrochemical characterizations of surface-modified LiCoO2 cathode materials for Li-ion batteries. Solid State Ionics 2002, 148, 335–342. [Google Scholar] [CrossRef]
  31. Kim, J.W.; Kim, D.H.; Oh, D.Y.; Lee, H.; Kim, J.H.; Lee, J.H.; Jung, Y.S. Surface chemistry of LiNi0.5Mn1.5O4 particles coated by Al2O3 using atomic layer deposition for Lithium-ion batteries. J. Power Sources 2015, 274, 1254–1262. [Google Scholar] [CrossRef]
  32. Guan, D.; Wang, Y. Ultrathin surface coatings to enhance cycling stability of LiMn2O4 cathode in Lithium-ion batteries. Ionics 2013, 19, 1–8. [Google Scholar] [CrossRef]
  33. Mohanty, D.; Dahlberg, K.; King, D.M.; David, L.A.; Sefat, A.S.; Wood, D.L.; Daniel, C.; Dhar, S.; Mahajan, V.; Lee, M. Modification of Ni-rich FCG NMC and NCA cathodes by atomic layer deposition: preventing surface phase transitions for high-voltage Lithium-ion batteries. Sci. Rep. 2016, 6, 26532. [Google Scholar] [CrossRef]
  34. Shi, Y.; Zhang, M.; Qian, D.; Meng, Y.S. Ultrathin Al2O3 coatings for improved cycling performance and thermal stability of LiNi0.5Co0.2Mn0.3O2 cathode material. Electrochim. Acta 2016, 203, 154–161. [Google Scholar] [CrossRef]
  35. Zhang, X.; Belharouak, I.; Li, L.; Lei, Y.; Elam, J.W.; Nie, A.; Chen, X.; Yassar, R.S.; Axelbaum, R.L. Structural and electrochemical study of Al2O3 and TiO2 Coated Li1.2Ni0.13Mn0.54Co0.13O2 cathode material using ALD. Adv. Energy Mater. 2013, 3, 1299–1307. [Google Scholar] [CrossRef]
  36. Zhao, J.; Wang, Y. Atomic layer deposition of epitaxial ZrO2 coating on LiMn2O4 nanoparticles for high-rate lithium ion batteries at elevated temperature. Nano Energy 2013, 2, 882–889. [Google Scholar] [CrossRef]
  37. Aziz, S.; Zhao, J.; Cain, C.; Wang, Y. Nanoarchitectured LiMn2O4/Graphene/ZnO composites as electrodes for Lithium ion batteries. J. Mater. Sci. Technol. 2014, 30, 427–433. [Google Scholar] [CrossRef]
  38. Zhao, J.; Wang, Y. Surface modifications of Li-ion battery electrodes with various ultrathin amphoteric oxide coatings for enhanced cycleability. J. Solid State Electrochem. 2013, 17, 1049–1058. [Google Scholar] [CrossRef]
  39. Zhao, J.; Ying, W. Ultrathin surface coatings for improved electrochemical performance of Lithium ion battery electrodes at elevated temperature. J. Phys. Chem. C 2012, 116, 11867–11876. [Google Scholar] [CrossRef]
  40. Zhao, J.; Wang, Y. High-capacity full Lithium-ion cells based on nanoarchitectured ternary manganese-nickel-cobalt carbonate and its lithiated derivative. J. Mater. Chem. A 2014, 2, 14947–14956. [Google Scholar] [CrossRef]
  41. Aykol, M.; Kirklin, S.; Wolverton, C. Thermodynamic aspects of cathode coatings for Lithium-ion batteries. Adv. Energy Mater. 2015, 4, 1400690. [Google Scholar] [CrossRef]
  42. Zhao, F.; Tang, Y.; Wang, J.; Tian, J.; Ge, H.; Wang, B. Vapor-assisted synthesis of Al2O3-coated LiCoO2 for high-voltage lithium ion batteries. Electrochim. Acta 2015, 174, 384–390. [Google Scholar] [CrossRef]
  43. Araki, K.; Taguchi, N.; Sakaebe, H.; Tatsumi, K.; Ogumi, Z. Electrochemical properties of LiNi1/3Co1/3 Mn1/3O2 cathode material modified by coating with Al2O3 nanoparticles. J. Power Sources 2014, 269, 236–243. [Google Scholar] [CrossRef]
  44. Kalluri, S.; Yoon, M.; Jo, M.; Liu, H.K.; Dou, S.X.; Cho, J.; Guo, Z. Feasibility of cathode surface coating technology for high-energy Lithium-ion and beyond-Lithium-ion batteries. Adv. Mater. 2017, 29, 1605807. [Google Scholar] [CrossRef]
  45. Liu, S.; Xiong, L.; He, C. Long cycle life lithium ion battery with lithium nickel cobalt manganese oxide (NCM) cathode. J. Power Sources 2014, 261, 285–291. [Google Scholar] [CrossRef]
  46. Jung, S.K.; Gwon, H.; Hong, J.; Park, K.Y.; Seo, D.H.; Kim, H.; Hyun, J.; Yang, W.; Kang, K. Understanding the degradation mechanisms of LiNi0.5Co0.2Mn0.3O2 cathode material in Lithium ion batteries. Adv. Energy Mater. 2014, 4, 1300787. [Google Scholar] [CrossRef]
Figure 1. (a) Schematics illustration of the Al2O3 ALD growth at the surface of NCM811 particles; FESEM images of (b) bare NCM811, and (c) 20 Al2O3 ALD NCM811; EDS mapping of (d) 20 Al2O3 ALD NCM811; TEM images of (e) bare NCM811, (f) 20 Al2O3 ALD NCM811, and (g) 50 Al2O3 ALD NCM811; and HRTEM images of (h) bare NCM811, (i) 20 Al2O3 ALD NCM811, and (j) 50 Al2O3 ALD NCM811.
Figure 1. (a) Schematics illustration of the Al2O3 ALD growth at the surface of NCM811 particles; FESEM images of (b) bare NCM811, and (c) 20 Al2O3 ALD NCM811; EDS mapping of (d) 20 Al2O3 ALD NCM811; TEM images of (e) bare NCM811, (f) 20 Al2O3 ALD NCM811, and (g) 50 Al2O3 ALD NCM811; and HRTEM images of (h) bare NCM811, (i) 20 Al2O3 ALD NCM811, and (j) 50 Al2O3 ALD NCM811.
Coatings 09 00092 g001
Figure 2. XRD patterns of (a) bare NCM8111; (b) 10 Al2O3 ALD NCM8111; (c) 20 Al2O3 ALD NCM8111; and (d) 50 Al2O3 ALD NCM8111 materials, coupled with rietveld refinement results.
Figure 2. XRD patterns of (a) bare NCM8111; (b) 10 Al2O3 ALD NCM8111; (c) 20 Al2O3 ALD NCM8111; and (d) 50 Al2O3 ALD NCM8111 materials, coupled with rietveld refinement results.
Coatings 09 00092 g002
Figure 3. Characteristics XPS spectra captured on the 10 Al2O3 ALD NCM811, 20 Al2O3 ALD NCM811, and 50 Al2O3 ALD NCM811 materials compared to the bare NCM811: (a) Al 2p, (b) Ni 2p, (c) Co 2p, and (d) Mn 2p peaks.
Figure 3. Characteristics XPS spectra captured on the 10 Al2O3 ALD NCM811, 20 Al2O3 ALD NCM811, and 50 Al2O3 ALD NCM811 materials compared to the bare NCM811: (a) Al 2p, (b) Ni 2p, (c) Co 2p, and (d) Mn 2p peaks.
Coatings 09 00092 g003
Figure 4. Electrochemical performance of different cathode materials in a broad voltage range of 2.7–4.6 V vs. Li+/Li: (a) cycling performance at 0.1 C for 100 cycles; (b) charge/discharge curves of bare NCM811 and 20 Al2O3 ALD NCM811 in the first and 100th cycles at 0.1 C; (c) dQ/dV profiles of bare NCM811 and 20 Al2O3 ALD NCM811 in the first and 100th cycles; and (d) EIS spectra captured on fresh cells (solid patterns) at open circuit voltage states and after 100 cycles at fully-discharged states (hollow patterns).
Figure 4. Electrochemical performance of different cathode materials in a broad voltage range of 2.7–4.6 V vs. Li+/Li: (a) cycling performance at 0.1 C for 100 cycles; (b) charge/discharge curves of bare NCM811 and 20 Al2O3 ALD NCM811 in the first and 100th cycles at 0.1 C; (c) dQ/dV profiles of bare NCM811 and 20 Al2O3 ALD NCM811 in the first and 100th cycles; and (d) EIS spectra captured on fresh cells (solid patterns) at open circuit voltage states and after 100 cycles at fully-discharged states (hollow patterns).
Coatings 09 00092 g004
Table 1. Crystal parameters of different Al2O3-coated NCM811 materials in comparison to bare NCM811, according to XRD patterns and refinement results as shown in Figure 2.
Table 1. Crystal parameters of different Al2O3-coated NCM811 materials in comparison to bare NCM811, according to XRD patterns and refinement results as shown in Figure 2.
Samplesa
(Å)
s.d.c
(Å)
s.d.Volume (Å3)s.d.c/aLi+ at
3b Sites
s.d.Rwp
(%)
Rp
(%)
NCM8112.87463 × 10−514.21021.8 × 10−3101.722.87 × 10−34.94320.96322.203 × 10−36.624.81
10 Al2O3 ALD NCM8112.87453 × 10−514.20953.1 × 10−3101.713.1 × 10−34.94330.96342.351 × 10−36.095.12
20 Al2O3 ALD NCM8112.87432 × 10−514.20943.6 × 10−3101.653.05 × 10−34.94360.96971.936 × 10−36.425.18
50 Al2O3 ALD NCM8112.87394 × 10−514.20715.2 × 10−3101.497.93 × 10−34.94350.96667.725 × 10−37.035.58

Share and Cite

MDPI and ACS Style

Zhu, W.; Huang, X.; Liu, T.; Xie, Z.; Wang, Y.; Tian, K.; Bu, L.; Wang, H.; Gao, L.; Zhao, J. Ultrathin Al2O3 Coating on LiNi0.8Co0.1Mn0.1O2 Cathode Material for Enhanced Cycleability at Extended Voltage Ranges. Coatings 2019, 9, 92. https://doi.org/10.3390/coatings9020092

AMA Style

Zhu W, Huang X, Liu T, Xie Z, Wang Y, Tian K, Bu L, Wang H, Gao L, Zhao J. Ultrathin Al2O3 Coating on LiNi0.8Co0.1Mn0.1O2 Cathode Material for Enhanced Cycleability at Extended Voltage Ranges. Coatings. 2019; 9(2):92. https://doi.org/10.3390/coatings9020092

Chicago/Turabian Style

Zhu, Wenchang, Xue Huang, Tingting Liu, Zhiqiang Xie, Ying Wang, Kai Tian, Liangming Bu, Haibo Wang, Lijun Gao, and Jianqing Zhao. 2019. "Ultrathin Al2O3 Coating on LiNi0.8Co0.1Mn0.1O2 Cathode Material for Enhanced Cycleability at Extended Voltage Ranges" Coatings 9, no. 2: 92. https://doi.org/10.3390/coatings9020092

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