Skip to content
BY 4.0 license Open Access Published by De Gruyter April 20, 2021

Improvement of long-term cycling performance of high-nickel cathode materials by ZnO coating

  • Fangfang Wang , Ruoyu Hong EMAIL logo , Xuesong Lu , Huiyong Liu , Yuan Zhu , Ying Zheng and David Hui
From the journal Nanotechnology Reviews

Abstract

The high-nickel cathode material of LiNi0.8Co0.15Al0.05O2 (LNCA) has a prospective application for lithium-ion batteries due to the high capacity and low cost. However, the side reaction between the electrolyte and the electrode seriously affects the cycling stability of lithium-ion batteries. In this work, Ni2+ preoxidation and the optimization of calcination temperature were carried out to reduce the cation mixing of LNCA, and solid-phase Al-doping improved the uniformity of element distribution and the orderliness of the layered structure. In addition, the surface of LNCA was homogeneously modified with ZnO coating by a facile wet-chemical route. Compared to the pristine LNCA, the optimized ZnO-coated LNCA showed excellent electrochemical performance with the first discharge-specific capacity of 187.5 mA h g−1, and the capacity retention of 91.3% at 0.2C after 100 cycles. The experiment demonstrated that the improved electrochemical performance of ZnO-coated LNCA is assigned to the surface coating of ZnO which protects LNCA from being corroded by the electrolyte during cycling.

Graphical abstract

1 Introduction

Since the first successful development and commercialization of lithium-ion batteries (LIBs) by Sony in Japan in 1991, LIB have received increasing attention [1]. Due to the high energy density, good cycling performance, and environmental friendliness [2], LIBs have become one of the most promising green secondary batteries in the twenty-first century [3]. As an important component of LIBs, the cathode material has much lower capacity than the anode material, accounting for the highest proportion of battery cost [4]. However, the initial Coulombic efficiency, rate capability, and cycling performance of these LiNiO2-based cathode materials are not satisfactory, especially at an elevated temperature [5]. Therefore, research on high-capacity cathode materials plays a crucial role in the development of LIBs [6].

Among the layered materials, LiNi0.8Co0.15Al0.05O2 is a promising cathode material, which is due to the combination of the advantages of high theoretical capacity and stable layered structure of LiNiO2 [7], LiCoO2 [8], and LiAlO2 [9]. As the precursor of LiNi0.8Co0.15Al0.05O2, Ni0.8Co0.15Al0.05(OH)2.05 is generally prepared by coprecipitation method [10]. However, the K sp of Al(OH)3 (4.57 × 10−33) is much less than that of Ni(OH)2 (2 × 10−15) and Co(OH)2 (1.58 × 10−15), meaning that the ions of Al3+ precipitate faster with the hydroxide, which inhibits the uniform growth of the precursor [11]. The generated precipitate will undergo a lattice change and structural collapse during the cycling [12]. In the traditional preparation process, the oxidation of Ni2+ was carried out during the calcination [13]. However, the agglomeration of materials could make it difficult for the internal Ni2+ to be oxidized to N3+, resulting in cation mixing [14]. The radius of Ni2+ is very close to Li+, and the 3b of Ni2+ will occupy the 3a of Li+ [15], which hinders the deintercalation of Li+ and further affects the cycling performance. In addition, the Ni-containing cathodes easily react with electrolyte and the transition metal ions dissolve due to HF corrosion [16]. The unstable Ni4+ ions in the charged state tend to transform to more stable NiO on the cathode surface, which results in high interfacial resistance and rapid capacity decay [17]. Moreover, the oxygen is released during the structural transition, causing poor safety performance of the battery [18]. Therefore, it is urgent for us to find a new strategy to solve these problems.

Doping aluminum has been found very effective to improve durability of LIBs because it stabilizes the charge-transfer impedance on the cathode. Much more effects have been paid to solve the problem of rapid precipitation of Al3+. Trease et al. [19] used a series of techniques to measure the aluminum distribution in layered LiNi0.8Co0.15Al0.05O2 (NCA) to investigate the influence of aluminum doping on layer stabilization. The experimental results showed that Ni3+ ions incurred a dynamic Jahn–Teller (JT) distortion, while the Al reduced the strain associated with JT distortion by prioritizing the electron ordering of the JT long bonds of Al3+ ions. A new aluminum source, NaAlO2, was employed by Liang et al. [20] to prepare a high-performance LNCA (LNCA-NaAlO2). NaAlO2 was hydrolyzed during the preparation of the precursor to avoid the rapid precipitation of Al3+ and the formation of flocculation precipitation. The obtained LNCA-NaAlO2 with uniform element distribution and orderly layered structure showed a high initial discharge capacity of 204.7 mA h g−1 at 0.1C and a good capacity retention of 74.1% after 200 cycles. Zhou et al. [21] also used AlO2 as Al source to synthesize the Ni-rich precursor, Ni0.9Co0.07Al0.03(OH)2. After optimizing the calcination conditions, LiNi0.9Co0.07Al0.03O2 demonstrated excellent electrochemical performances. Kim and Kim [22] used acetylacetone as a chelating agent to effectively control the Al reaction rate, improve the uniform distribution of Al, and more importantly, increase the average particle size and density of NCA. Both the volumetric energy and specific capacity of LIBs were increased.

Many methods have been reported to improve such instable structural of high-nickel cathode material via doping and surface coating [23,24]. Liang et al. [25] found that the SiO2 coating onto the surface of LiNi0.8Co0.1Mn0.1O2 can stabilize the layered structure, effectively reduce the corrosion of the cathode material by the electrolyte, and hence, improve the cycling performance. Liu et al. [26] used the pillaring effect of inactive Mg in the crystal structure to modify LiNi0.8Co0.1Mn0.1O2 by doping Mg and obtained the improved cycling stability (high-capacity retention of 81% over 350 cycles at 0.5C).

In this work, we report the synthesis, structural characterization, and electrochemical of ZnO-coated LiNi0.8Co0.15Al0.05O2 cathodes. The synergistic action of Ni2+ preoxidation by (NH4)2S2O8 and the optimization of calcination temperature reduced the cation mixing of LiNi0.8Co0.15Al0.05O2. The solid-phase Al-doping improved the uniformity of element distribution and the orderliness of the layered structure. In addition, the surface of LiNi0.8Co0.15Al0.05O2 was coated with ZnO [27,28] which improved the structural stability and cycling performance. The preparation process of ZnO-coated LiNi0.8Co0.15Al0.05O2 is shown in Figure 1.

Figure 1 
               Schematic view of the synthesis process of ZnO-coated LiNi0.8Co0.15Al0.05O2.
Figure 1

Schematic view of the synthesis process of ZnO-coated LiNi0.8Co0.15Al0.05O2.

2 Experimental

2.1 Preparation of LiNi0.8Co0.15Al0.05O2

The NiSO4·6H2O and CoSO4·7H2O (molar ratio of 0.80: 0.15) were dissolved in deionized water to form a mixed salt solution at a concentration of 1.9 mol L−1 and added a certain amount of (NH4)2S2O8 solution. 4 mol L−1 NaOH solution and 1.5 mol L−1 NH3·H2O solution are served as the pH control agent and the chelating agent, respectively. The NH3·H2O solution is first added to a 5 L reactor, and then the salt solution and the NaOH solution are pumped together into the reactor at a flow rate ratio of 2:1 to react at 50°C for 30 h; the pH was controlled at 11.00 ± 0.02 by an online pH meter. Then, the Ni0.8Co0.15(OH)2 was obtained after washing and drying. Subsequently, solid-phase Al-doping is used to prepare the LiNi0.8Co0.15Al0.05O2. The Li2CO3, Ni0.8Co0.15(OH)1.9, and Al(NO3)3·9H2O (molar ratio of Li:Ni:Co:Al = 1.08:0.8:0.15:0.05) were mixed and ball-milled for 5 h, then calcined at 500°C for 5 h under the oxygen atmosphere in a tube furnace to obtain the intermediate product. After that, the intermediate product is cooled to room temperature, ball-milled for 5 h, placed in the tube furnace again, and calcined at 750°C for 12 h in the oxygen atmosphere to form LiNi0.8Co0.15Al0.05O2.

2.2 Preparation of ZnO-coated LiNi0.8Co0.15Al0.05O2

A certain amount of Zn(AC)2·2H2O was dissolved in 40 mL of ethanol, and after ultrasonic dissolution, 2 g of the LiNi0.8Co0.15Al0.05O2 obtained in the above step was added under vigorous stirring for 4 h, and then increased to 65°C to evaporate the ethanol. The mixture was dried at 120°C for 12 h, and finally calcined at 450°C for 5 h to obtain ZnO-coated LiNi0.8Co0.15Al0.05O2 powder which could be used as cathode material. The amount of Zn(AC)2·2H2O added to prepare ZnO-coated LiNi0.8Co0.15Al0.05O2 were 1, 2, 3 wt%, and the prepared samples were recorded as Z1, Z2, and Z3, respectively, and the pristine LiNi0.8Co0.15Al0.05O2 as Z0.

2.3 Materials characterization

The material structure was analyzed by X-ray powder diffractometer (XRD, Ultima type, Japan Science) at the scanning speed of 0.2° s−1 with the scanning range of 10–80°. Thermogravimetric (TG) measurements (Netzsch-STA 449C) were conducted from room temperature to 900°C at a heating rate of 10°C min−1 in air. The surface morphology was analyzed by scanning electron microscope (SEM, Zeiss supra 55, Zeiss, Germany) and transmission electron microscopy (TEM, JEM-2100HR, Japan).

2.4 Electrochemical measurements

The working electrodes were prepared by mixing active material, acetylene black, and PVDF at a mass ratio of 8:1:1 in N-methylpyrrolidinone solution. The slurry was coated onto the aluminum foil and dried at 110°C for 5 h. Then the coated foil was pressed and punched into 14 mm diameter disks. The electrodes were transferred to a glove box (Super 1220/750, Mikrouna Co., Ltd. China) to assemble in 2025 coin-type cells with a lithium foil counter electrode. Celgard-2400 was used as the separator and 1 M LiPF6 dissolved in ethylene carbonate (EC) and diethyl carbonate (DEC) (1:1 in volume ratio) as the electrolyte solution.

The galvanostatic charge and discharge cycle were conducted on CT2001A (LAND, Bart Rui Tech. Co., Ltd.) battery test system between 2.8 and 4.3 V (vs Li/Li+) under different current densities. Cyclic voltammetry (CV) test was recorded on Gamry electrochemical system (Interface1010E) with a potential range of 3.0–4.3 V (vs Li/Li+) at a scanning rate of 0.1 mV s−1. Electrochemical impedance spectrum (EIS, 100 kHz to 0.01 Hz) was performed using the same electrochemical system.

3 Results and discussion

3.1 Effect of different calcination temperatures on LiNi0.8Co0.15Al0.05O2

As known, the crystal structures and microstructures of LiNi0.8Co0.15Al0.05O2 are also affected by the calcination temperature. Therefore, to determine the calcination temperature for preparing LiNi0.8Co0.15Al0.05O2, the thermogravimetric analysis (TGA) was conducted in air for the mixture Li2CO3, Ni0.8Co0.15(OH)1.9, and Al(NO3)3·9H2O (molar ratio of Li:Ni:Co:Al = 1.08:0.8:0.15:0.05) in the temperature range from 25 to 900°C, as shown in Figure 2. The differential curve is the relationship between the temperature and the first derivative of time in the TG curve, which represents the weight loss rate. It can be seen that there is a slow weight loss from room temperature to 90°C, corresponding to a small endothermic peak (a) on the differential curve, which is due to weight loss of absorbed water on surface and crystal water of Al(NO3)3. A dramatic decline occurs from 190 to 270°C, corresponding to the maximum endothermic peak of the differential curve (b), mainly ascribed to the process of Ni0.8Co0.15(OH)1.9 forming Ni0.8Co0.15O2. From 270 to 500°C, there is a relatively slow weight loss, corresponding to the broad endothermic peak (c) on the differential curve, which is due to the thermal decomposition of Li2CO3 to produce Li2O and release CO2. With the further increase in temperature, Al2O3 was formed by decomposition of Al(NO3)3, and finally LiNi0.8Co0.15Al0.05O2 was synthesized. The analytical curve shows that the thermal reaction of the mixture is carried out in stages. In order to obtain a positive cathode material with excellent performance, it is necessary to strictly control the calcination conditions. In addition, when the temperature is around 850°C, there is a slow weight loss, due to the volatilization of lithium, thereby, the calcination temperature cannot exceed 850°C.

Figure 2 
                  TG–DTG curve of the mixture of Li2CO3, Ni0.8Co0.15(OH)1.9, and Al(NO3)3·9H2O.
Figure 2

TG–DTG curve of the mixture of Li2CO3, Ni0.8Co0.15(OH)1.9, and Al(NO3)3·9H2O.

Figure 3 shows the XRD patterns of the samples prepared at different calcination temperatures. All peaks are well-indexed to the R3m group and consistent with the layered structure of α-NaFeO2 [29]. Two pairs of splitting peaks, (006/102) and (108/110), in the XRD patterns correspond to the typical structural features of hexagonal layered materials [30]. None of the four samples show any impurity peaks, proving that the pure phase of LiNi0.8Co0.15Al0.05O2 for cathode material was synthesized. As the temperature rises, the intensity of the (108/110) peak and (006/102) peak increases, indicating that the increase in the calcination temperature improves the crystallinity of the samples. The lattice parameters of the four samples are shown in Table 1. It can be seen that the calcination temperature has little effect on the lattice parameters of a, c, and V, but has a certain influence on the ratio of I 003/I 104. In the ternary cathode material, the ratio of I 003/I 104 corresponds to the internal cation mixing, and the sample calcined at 750°C has the largest ratio of I 003/I 104, and the corresponding cation mixing is the smallest. Therefore, the calcination temperature of 750°C is the most suitable.

Figure 3 
                  XRD pattern of LiNi0.8Co0.15Al0.05O2 prepared at different calcination temperatures.
Figure 3

XRD pattern of LiNi0.8Co0.15Al0.05O2 prepared at different calcination temperatures.

Table 1

Lattice parameters of LiNi0.8Co0.15Al0.05O2 prepared at different calcination temperatures

Sample/°C a c c/a V (Å × 103) I 003/I 104
700 2.86112 14.16892 4.9433 101.31 1.51
750 2.86232 14.16877 4.9421 101.23 1.79
800 2.86332 14.16965 4.9488 101.87 1.52
850 2.86375 14.16433 4.9367 101.53 1.74

Figure 4 shows the cycling performance of LiNi0.8Co0.15Al0.05O2 prepared at different calcination temperatures in the voltage range of 2.8–4.3 V at 0.2C. It can be seen that the discharge capacity generally improves with the increasing calcination temperature at first, and then tends to decline. The sample synthesized at 750°C has a higher initial discharge-specific capacity of 186.8 mA h g−1 with a capacity retention of 83.6% after 50 cycles, which is better than that of the samples prepared at 700°C (183.8 mA h g−1, 80.5%) and 800°C (185.8 mA h g−1, 78.5%). When the temperature reached 850°C, a quick capacity fading is observed (from 175.3 to 137.1 mA h g−1). In summary, the experimental results prove that the sample obtained at the calcination temperature of 750°C has the best cycling performance and higher discharge-specific capacity, which is consistent with the XRD results.

Figure 4 
                  Cycling performance of LiNi0.8Co0.15Al0.05O2 prepared at different calcination temperatures.
Figure 4

Cycling performance of LiNi0.8Co0.15Al0.05O2 prepared at different calcination temperatures.

3.2 Effect of ZnO coating on LiNi0.8Co0.15Al0.05O2

Since ZnO has relatively high electron mobility and low cost, the as-prepared LiNi0.8Co0.15Al0.05O2 was coated with ZnO to improve the structural stability and cycling performance. Figure 5a is the XRD patterns of the Z0, Z1, Z2, and Z3, showing that the XRD patterns of the four samples are basically the same. Compared with the standard XRD pattern of ZnO (Figure 5b), there is no characteristic peak of ZnO appeared in the XRD pattern of Z1, Z2, and Z3 because the amount of ZnO used in modification is very small and beyond the XRD resolution. The lattice parameters of the four samples are listed in Table 2. There is no significant change in the lattice constants for pristine (Z0) and coated samples (Z1, Z2 and Z3), indicating that the Zn2+ does not enter the host structure of LiNi0.8Co0.15Al0.05O2. The ratio of I 003/I 104 for the coated samples is larger than that of pristine sample, indicating that ZnO-coated layer reduces the cation mixing of LiNi0.8Co0.15Al0.05O2 [31].

Figure 5 
                  (a) XRD patterns of the pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2, (b) standard XRD pattern of ZnO.
Figure 5

(a) XRD patterns of the pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2, (b) standard XRD pattern of ZnO.

Table 2

Lattice parameters of pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 samples

Sample a c c/a V3) I 003/I 104
Z0 2.866232 14.16877 4.9433 101.31 1.79
Z1 2.866239 14.16253 4.9430 101.37 1.81
Z2 2.86210 14.17036 4.9435 101.46 1.89
Z3 2.86256 14.17508 4.9440 101.57 1.83

Figure 6 shows the cycling performance of Z0, Z1, Z2, and Z3 at 0.2C in the voltage range of 2.8–4.3 V. It can be seen that the discharge capacity of pristine sample fades from 186.8 to 146.5 mA h g−1 after 100 cycles with a capacity retention of 78.4%. In contrast, the ZnO-coated samples remain 91.3, 90.2, and 86.6% for Z1, Zn2, and Z3 after 100 cycles, respectively, which demonstrates that the ZnO-coated samples have higher cyclic stability than pristine samples. The reason for the improvement of cycling performance is that ZnO-coated layer works as an isolating layer, which effectively reduces the side reactions between electrode and electrolyte, inhibits the internal collapse of the structure caused by the dissolution of transition metals [32], improves the structural stability of the material, and further improves the cycling performance of LiNi0.8Co0.15Al0.05O2 cathode material. In addition, the thin ZnO-coated layer can improve the diffusion rate of Li+ and the conductivity of the electrode. This result will be verified by CV and EIS. However, the cycling performance of Z2 and Z3 is slightly lower than that of Z1 after 100 cycles, indicating that the thick coating will hinder the diffusion of Li+ to a certain extent.

Figure 6 
                  Cycling performance of the pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 samples at 0.2C.
Figure 6

Cycling performance of the pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 samples at 0.2C.

Figure 7 is the SEM image of pristine and 1 wt% ZnO-coated LiNi0.8Co0.15Al0.05O2. The pristine sample is spherical particles with a clean surface (Figure 7a), and the secondary particles are formed by primary particle aggregation (Figure 7b). In contrast, the surface of the ZnO-coated LiNi0.8Co0.15Al0.05O2 spheres became obviously rough, and there were many needle-liked nano-rods and channels on the surface (Figure 4c and d), which expand the tunnels for Li+ diffusion and also stabilize the LiNi0.8Co0.15Al0.05O2 frame, proving that the ZnO successfully attached to the LiNi0.8Co0.15Al0.05O2 surface.

Figure 7 
                  SEM images of the (a and b) pristine and (c and d) 1 wt% ZnO-coated LiNi0.8Co0.15Al0.05O2 samples.
Figure 7

SEM images of the (a and b) pristine and (c and d) 1 wt% ZnO-coated LiNi0.8Co0.15Al0.05O2 samples.

More detailed structural features of the pristine and 1 wt% ZnO-coated LiNi0.8Co0.15Al0.05O2 were explored by high-resolution TEM. The pristine sample shows smooth surface and clear lattice fringes, indicating high crystallization (Figure 8a). The inter planar distance of pristine sample is about 0.47 nm (Figure 8b), corresponding to the planar distance of the (003) plane in XRD patterns [33]. As shown in Figure 8c and d, it is obvious that the ZnO nanoparticles are attached to the surface of the LiNi0.8Co0.15Al0.05O2, which is consistent with the result of the SEM image. As anticipated, this coating layer should improve electrochemical performances of LiNi0.8Co0.15Al0.05O2 by protecting its surface from electrolyte corrosion and HF attack.

Figure 8 
                  TEM images of (a and b) pristine and (c and d) ZnO-coated LiNi0.8Co0.15Al0.05O2.
Figure 8

TEM images of (a and b) pristine and (c and d) ZnO-coated LiNi0.8Co0.15Al0.05O2.

The rate performance tests of pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 were conducted under different current rates (0.2, 0.5, 1, 3, 5, 10C) in the voltage range of 2.8–4.3 V. As shown in Figure 9, at a low current rate of 0.2C, the discharge capacity of the ZnO-coated LiNi0.8Co0.15Al0.05O2 is not significantly improved. However, as the current density increases, the rate performance of ZnO-coated LiNi0.8Co0.15Al0.05O2 is obviously better than that of the pristine one. At the current rate of 10C, ZnO-coated LiNi0.8Co0.15Al0.05O2 shows a higher discharge capacity of 135 mA h g−1 than that of the pristine sample (115 mA h g−1). In general, the ZnO-coated LiNi0.8Co0.15Al0.05O2 has ZnO protective layer, which can reduce the interfacial impedance, inhibit the electrochemical polarization, and hence improve the rate performance.

Figure 9 
                  Rate performance of pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 samples.
Figure 9

Rate performance of pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 samples.

Figure 10 shows the CV of the pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 after 10 cycles with a scanning rate of 0.1 mV s−1 in a voltage range of 3.0–4.3 V. Those two curves show the unique redox couples of Nickel-based compounds [34]. The three pairs of redox peaks from left to right correspond to Ni2+/Ni3+, Ni3+/Ni4+, and Co3+/Co4+ [35]. In the anodic process, the peak located at about 3.76 V corresponds to phase transition from the hexagonal phase (H1) to the monoclinic phase (M), the peak at about 4.00 V corresponds to the phase transition from M to a new hexagonal phase (H2), and the peak at about 4.18 V corresponds to the phase transition from the H2 to another hexagonal phase (H3) [36]. There is no other peak in the curves of the ZnO-coated LiNi0.8Co0.15Al0.05O2, indicating that the modification of ZnO does not affect the structure of LiNi0.8Co0.15Al0.05O2 cathode material.

Figure 10 
                  CV curves of pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 samples at a scanning rate of 0.1 mV s−1 in the voltage of 3.0–4.3 V.
Figure 10

CV curves of pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 samples at a scanning rate of 0.1 mV s−1 in the voltage of 3.0–4.3 V.

To further investigate the reasons for the improvement of rate performance, the CV measurements were performed at various scanning rates. The potential intervals (ΔV) between the anodic and cathodic peaks indicate the reversibility of Li+ insertion/extraction and electrode polarization [37]. As shown in Figure 11a and b, ΔV increases with the increase of the scanning rate, and when the scanning rate is 0.8 mV s−1, the ΔV of the main peak (Ni3+/Ni4+, located at about 3.8 V) of ZnO-coated LiNi0.8Co0.15Al0.05O2 is 0.37 V, smaller than that of the pristine sample (0.43 V), which indicated that the modification of ZnO can inhibit the electrode polarization and improve the electrochemical reversibility [38]. As shown in Figure 11c and d, the linear relationship between the redox peak current (i p) and the square root of scanning rate (ν 1/2) is employed to calculate the Li+ diffusion coefficient ( D Li + ) of the main peak (Ni3+/Ni4+, located at about 3.7 V), which is the same method used in the literature [39]. The D Li + is calculated by the following equation:

i p = 2.69 × 10 5 n 3 / 2 A D Li 1 / 2 C v 1 / 2

where, the n is charge-transfer number, and C* is the concentration of Li+ in the active material. The D Li + at the oxidation peak and the reduction peak of the initial sample are 1.89 × 10−10 and 1.47 × 10−10 cm2 s−1, respectively, which is smaller than that of ZnO-coated LiNi0.8Co0.15Al0.05O2 (2.21 × 10−10 and 1.71 × 10−10 cm2 s−1). The diffusion control dominates the electrochemical reaction during the Li+ insertion/extraction progress. Therefore, the modification of ZnO improves the rate performance.

Figure 11 
                  Cyclic voltammogram of (a) pristine and (b) ZnO-coated LiNi0.8Co0.15Al0.05O2 at various scanning rates; linear relationship between the peak currents (i
                     p) and square root of scan rate (ν
                     1/2): (c) pristine sample, (d) ZnO-coated LiNi0.8Co0.15Al0.05O2 sample.
Figure 11

Cyclic voltammogram of (a) pristine and (b) ZnO-coated LiNi0.8Co0.15Al0.05O2 at various scanning rates; linear relationship between the peak currents (i p) and square root of scan rate (ν 1/2): (c) pristine sample, (d) ZnO-coated LiNi0.8Co0.15Al0.05O2 sample.

Figure 12 shows the Nyquist plots of the pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 samples before and after 10 and 100 cycles. The equivalent circuit model of the studied system is also shown in Figure 9d. R s represents the resistance of the electrolyte, R f means the surface film resistance, R ct is the charge-transfer resistance [40], and W is the Warburg impedance that reflects the diffusion of Li+ in the solid [41].

Figure 12 
                  Nyquist plots of (a) pristine and (b) ZnO-coated LiNi0.8Co0.15Al0.05O2 samples after different cycles, (c) equivalent circuit model.
Figure 12

Nyquist plots of (a) pristine and (b) ZnO-coated LiNi0.8Co0.15Al0.05O2 samples after different cycles, (c) equivalent circuit model.

All the fitted EIS parameters are listed in Table 3. It is noted that these two samples both show a large R ct value before cycling, which is caused by the inactivation of the electrodes [42]. As the cycling test begins, the R ct value decreases rapidly during the internal activation process [43]. However, after 100 cycles, the R ct value of the pristine sample quickly increased to 172.5 Ω, which was much higher than that of after 10th cycle. However, the value of R ct for ZnO-coated LiNi0.8Co0.15Al0.05O2 after 100 cycles is only about 62.8 Ω, much smaller than that of the pristine one. It can be concluded from the experimental results that the ZnO-coated layer decreases the charge-transfer resistance, increases the electronic conductivity of LiNi0.8Co0.15Al0.05O2, and further improves the rate performance.

Table 3

EIS parameters of pristine and ZnO-coated LiNi0.8Co0.15Al0.05O2 after different cycles

Sample Cycle number R s (Ω) R f (Ω) R ct (Ω)
LiNi0.8Co0.15Al0.05O2 0 5.825 15.456 233.31
10 3.164 12.146 65
100 4.163 16.428 172.5
ZnO-coated LiNi0.8Co0.15Al0.05O2 0 4.125 14.943 151
10 2.823 6.962 28
100 3.426 10.944 62.8

4 Conclusion

In conclusion, we designed a novel method for the preparation of high-capacity LiNiO2-based cathode materials. First of all, the addition of (NH4)2S2O8 reduces the cation mixing of LiNi0.8Co0.15Al0.05O2; solid-phase Al-doping improves the element distribution and layered structure. Then, optimizing the calcination temperature enhances the crystallinity of LiNi0.8Co0.15Al0.05O2, and the ZnO was coated on the surface of LiNi0.8Co0.15Al0.05O2 to improve the cycling performance. The electrochemical tests demonstrated a high reversible discharge capacity of 172.1 mA h g−1 at 0.2C with the capacity retention of 91.3% after 100 cycles and a good rate performance at high current density. The above-mentioned improvement indicates the potential application of ZnO-coated LiNi0.8Co0.15Al0.05O2 as the cathode material for advanced LIBs.

  1. Funding information: This research was financially supported by Minjiang Scholarship of Fujian Province (No. Min-Gaojiao[2010]-117), Central-government Guided Fund for Local Economic Development (No. 830170778), R&D Fund for Strategic Emerging Industry of Fujian Province (No. 82918001), and International Cooperation Project of Fujian Science and Technology Department (No. 830170771).

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: David Hui, who is the coauthor of this article, is a current Editorial Board member of Nanotechnology Reviews. This fact did not affect the peer-review process. The authors declare no other conflict of interest.

References

[1] Lin D, Liu Y, Cui Y. Reviving the lithium metal anode for high-energy batteries. Nat Nanotechnol. 2017;12:194–206.10.1038/nnano.2017.16Search in Google Scholar PubMed

[2] Cano ZP, Banham D, Ye S, Hintennach A, Lu J, Fowler M, et al. Batteries and fuel cells for emerging electric vehicle markets. Nat Energy. 2018;3:279–89.10.1038/s41560-018-0108-1Search in Google Scholar

[3] Pfleging W. A review of laser electrode processing for development and manufacturing of lithium-ion batteries. Nanotechnol Rev. 2018;7(3):549–73.10.1515/nanoph-2017-0044Search in Google Scholar

[4] Lee MJ, Noh M, Park MH, Jo M, Kim H, Nam H, et al. The role of nanoscale-range vanadium treatment in LiNi0.8Co0.15Al0.05O2 cathode materials for Li-ion batteries at elevated temperatures. J Mater Chem. 2015;3:13450–53.10.1039/C5TA01571ESearch in Google Scholar

[5] Gao S, Cheng YT, Shirpour M. Effects of cobalt deficiency on nickel-rich layered LiNi0.8Co0.1Mn0.1O2 positive electrode materials for lithium-ion batteries. ACS Appl Mater Inter. 2019;11:982–9.10.1021/acsami.8b19349Search in Google Scholar PubMed

[6] Novikova SA, Yaroslavtsev AB. Cathode materials based on olivine lithium iron phosphates for lithium-ion batteries. Rev Adv Mater Sci. 2017;49(2):129–39.Search in Google Scholar

[7] Augustyn V, Therese S, Turner TC, Manthiram A. Nickel-rich layered LiNi1−xMxO2 (M = Mn, Fe, and Co) electrocatalysts with high oxygen evolution reaction activity. J Mater Chem. 2015;3:16604–12.10.1039/C5TA04637HSearch in Google Scholar

[8] Shi JL, Xiao DD, Zhang XD, Yin YX, Guo YG, Gu L, et al. Improving the structural stability of Li-rich cathode materials via reservation of cations in the Li-slab for Li-ion batteries. Nano Res. 2017;12:1–9.10.1007/s12274-017-1489-3Search in Google Scholar

[9] Li LJ, Chen ZY, Zhang QB, Xu M, Zhou X, Zhu HL, et al. A hydrolysis-hydrothermal route for the synthesis of ultrathin LiAlO2-inlaid LiNi0.5Co0.2Mn0.3O2 as a high-performance cathode material for lithium ion batteries. J Mater Chem. 2015;3:894–904.10.1039/C4TA05902FSearch in Google Scholar

[10] Makimura Y, Sasaki T, Nonaka T, Nishimura YF, Uyama T, Okuda C, et al. Factors affecting cycling life of LiNi0.8Co0.15Al0.05O2 for lithium-ion batteries. J Mater Chem. 2016;4:8350–58.10.1039/C6TA01251ESearch in Google Scholar

[11] Zhang HZ, Liu C, Song DW, Zhang LQ, Bie LJ. A new synthesis strategy towards enhancing the structure and cycle stabilities of the LiNi0.80Co0.15Al0.05O2 cathode material. J Mater Chem. 2017;5:835–41.10.1039/C6TA08084GSearch in Google Scholar

[12] Wu NT, Wu H, Yuan W, Liu SJ, Liao JY, Zhang Y, et al. Facile synthesis of one-dimensional LiNi0.9Co0.07Al0.03O2 microrods as advanced cathode materials for lithium ion batteries. J Mater Chem. 2015;3:13648–52.10.1039/C5TA02767ESearch in Google Scholar

[13] Song X, Liu GX, Yue HF, Luo L, Yang SY, Huang YY, et al. A novel low-cobalt long-life LiNi0.88Co0.06Mn0.03Al0.03O2 cathode material for lithium ion batteries. Chem Eng J. 2020;407:126301–10.10.1016/j.cej.2020.126301Search in Google Scholar

[14] Liu W, Oh P, Liu X, Lee MJ, Cho W, Chae S, et al. Nickel-rich layered lithium transition-metal oxide for high-energy lithium-ion batteries. Angew Chem Int Ed. 2015;127:4518–36.10.1002/ange.201409262Search in Google Scholar

[15] Chen M, Zhao EY, Chen DF, Wu MM, Han SB, Huang QZ, et al. Decreasing Li/Ni disorder and improving the electrochemical performances of Ni-Rich LiNi0.8Co0.1Mn0.1O2 by Ca doping. Inorg Chem. 2017;56:8355–62.10.1021/acs.inorgchem.7b01035Search in Google Scholar PubMed

[16] Jo M, Noh M, Oh P, Kim Y, Cho J. A new high power LiNi0.81Co0.1Al0.09O2 cathode material for lithium-ion batteries. Adv Energy Mater. 2014;4:1301583.10.1002/aenm.201301583Search in Google Scholar

[17] Chen WH, Li YY, Yang D, Feng XM, Guan XX, Mi LW, et al. Controlled synthesis of spherical hierarchical Li–Ni1−x−yCoxAlyO2 (0 < x, y < 0.2) via a novel cation exchange process as cathode materials for high-performance lithium batteries. Electrochim Acta. 2016;190:932–38.10.1016/j.electacta.2016.01.024Search in Google Scholar

[18] Bak SM, Nam KW, Chang WY, Yu XQ, Hu EY, Hwang S, et al. Correlating structural changes and gas evolution during the thermal decomposition of charged LixNi0.8Co0.15Al0.05O2 cathode materials. Chem Mater. 2013;25:337–51.10.1021/cm303096eSearch in Google Scholar

[19] Trease NM, Seymour ID, Radin MD, Liu H, Hao L, Hy S, et al. Identifying the distribution of Al3+ in LiNi0.8Co0.15Al0.05O2. Chem Mater. 2016;28(22):870–80.10.1021/acs.chemmater.6b02797Search in Google Scholar

[20] Liang M, Song D, Zhang H. Improved performances of LiNi0.8Co0.15Al0.05O2 material employing NaAlO2 as a new aluminium source. ACS Appl Mater Inter. 2017;9:7b12306.10.1021/acsami.7b12306Search in Google Scholar PubMed

[21] Zhou P, Meng H, Zhang Z, Chen CC, Lu Y, Cao J, et al. Stable layered Ni-rich LiNi0.9Co0.07Al0.03O2 microspheres assembled with nanoparticles as high-performance cathode materials for lithium-ion batteries. J Mater Chem A Mater Energ Sus. 2017;5(6):2724–31.10.1039/C6TA09921ASearch in Google Scholar

[22] Kim Y, Kim D. Synthesis of high-density nickel cobalt aluminum hydroxide by continuous coprecipitation method. ACS Appl Mater Inter. 2012;4(2):586.10.1021/am201585zSearch in Google Scholar PubMed

[23] Zhang W, Chi ZX, Mao WX, Lv RW, Cao AM, Wan LJ, et al. One-nanometer-precision control of Al2O3 nanoshell through solution-based synthesis route. Angew Chem Int Ed. 2014;53:12776–80.10.1002/anie.201406856Search in Google Scholar PubMed

[24] Zhu B, Liu N, Mcdowell M, Jin Y, Cui Y, Zhu J, et al. Interfacial stabilizing effect of ZnO on Si anodes for lithium ion battery. Nano Energy. 2015;13:620–25.10.1016/j.nanoen.2015.03.019Search in Google Scholar

[25] Liang L, Hu G, Jiang F, Cao Y. Electrochemical behaviours of SiO2-coated LiNi0.8Co0.1Mn0.1O2 cathode materials by a novel modification method. J Alloy Compound. 2016;657:570–81.10.1016/j.jallcom.2015.10.177Search in Google Scholar

[26] Liu X, Wang S, Wang L, Wang K, Wu XZ, Zhou PF, et al. Stabilizing the high-voltage cycle performance of LiNi0.8Co0.1Mn0.1O2 cathode material by Mg doping. J Power Sources. 2019;438(31):227017–24.10.1016/j.jpowsour.2019.227017Search in Google Scholar

[27] Yu W, Huang L, Yang D, Fu P, Zhou L, Zhang J, et al. Efficiency exceeding 10% for inverted polymer solar cells with a ZnO/ionic liquid combined cathode interfacial layer. J Mater Chem. 2015;A310:660–65.10.1039/C5TA00930HSearch in Google Scholar

[28] Huang XR, Zeng LJ, Li RF, Xi ZJ, Li YC. Manipulating conductive network formation via 3D T-ZnO: A facile approach for a CNT-reinforced nanocomposite. Nanotechnol Rev. 2020;9:534–42.10.1515/ntrev-2020-0043Search in Google Scholar

[29] Fu CC, Li GS, Luo D, Li Q, Fan JM, Li LP, et al. Nickel-rich layered microspheres cathodes: lithium/nickel disordering and electrochemical performance. ACS Appl Mater Inter. 2014;6:15822–31.10.1021/am5030726Search in Google Scholar PubMed

[30] Yu HJ, Qian YM, Otani M, Tang DM, Guo SH, Zhu YB, et al. Study of the lithium/nickel ions exchange in the layered LiNi0.42Mn0.42Co0.16O2 cathode material for lithium ion batteries: experimental and first-principles calculations. Energy Environ Sci. 2014;7:1068–78.10.1039/c3ee42398kSearch in Google Scholar

[31] Wu F, Tian J, Su Y, Wang J, Zhang C, Bao L, et al. Effect of Ni(2+) content on lithium/nickel disorder for Ni-rich cathode materials. ACS Appl Mater Inter. 2015;7(14):7702–08.10.1021/acsami.5b00645Search in Google Scholar PubMed

[32] Zhao EY, Liu XF, Zhao H, Xiao XL, Hu ZB. Ion conducting Li2SiO3-coated lithium-rich layered oxide exhibiting high rate capability and low polarization. Chem Commun. 2015;51:9093–96.10.1039/C5CC00383KSearch in Google Scholar

[33] Zhao E, Chen M, Chen D, Xiao X, Hu Z. A versatile coating strategy to highly improve the electrochemical properties of layered oxide LiMO2 (M = Ni0.5Mn0.5 and Ni1/3Mn1/3Co1/3). ACS Appl Mater Inter. 2015;7(49):27096–105.10.1021/acsami.5b08777Search in Google Scholar PubMed

[34] Qiu Z, Zhang Y, Dong P, Xia S, Yao Y. A facile method for synthesis of LiNi0.8Co0.15Al0.05O2 cathode material. Solid State Ionics. 2017;307:73–8.10.1016/j.ssi.2017.04.011Search in Google Scholar

[35] Li LJ, Li XH, Wang ZX, Guo HJ, Yue P, Chen W, et al. Synthesis, structural and electrochemical properties of LiNi0.79Co0.1Mn0.1Cr0.01O2 via fast co-precipitation. J Alloys Compd. 2010;507(1):172–77.10.1016/j.jallcom.2010.07.148Search in Google Scholar

[36] Noh HJ, Youn S, Yoon CS, Sun YK. Comparison of the structural and electrochemical properties of layered Li[NixCoyMnz]-O2 (x = 1/3, 0.5, 0.6, 0.7, 0.8 and 0.85) cathode material for lithium-ion batteries. J Power Sources. 2013;233:121–30.10.1016/j.jpowsour.2013.01.063Search in Google Scholar

[37] Xie H, Hu G, Du K, Peng Z, Cao Y. An improved continuous co-precipitation method to synthesize LiNi0.8Co0.15Al0.05O2 cathode material. J Alloy Compound. 2016;666:84–7.10.1016/j.jallcom.2016.01.064Search in Google Scholar

[38] Rui XH, Ding N, Liu J, Li C, Chen CH. Analysis of the chemical diffusion coefficient of lithium ions in Li3V2(PO4)3 cathode material. Electrochim Acta. 2010;55:2384–90.10.1016/j.electacta.2009.11.096Search in Google Scholar

[39] Chen CC, Huang YN, An CH, Zhang H, Wang YJ, Jiao LF, et al. Copper-doped dual phase Li4Ti5O12-TiO2 nanosheets as high-rate and long cycle life anodes for high-power lithium-ion batteries. ChemSusChem. 2015;8:114–22.10.1002/cssc.201402886Search in Google Scholar PubMed

[40] Stenina IA, Kulova TL, Skundin AM, Yaroslavtsev AB. Carbon composites as anode materials for lithium-ion batteries. Rev Adv Mater Sci. 2017;49:140–49.Search in Google Scholar

[41] Reddy MV, Sakunthala A, Pandian SS, Chowdari BVR. Preparation, comparative energy storage properties, and impedance spectroscopy studies of environmentally friendly cathode, Li(MMn11/6)O4 (M = Mn1/6, Co1/6, (Co1/12Cr1/12)). J Phys Chem C. 2013;117:9056–64.10.1021/jp309180kSearch in Google Scholar

[42] Zhang Q, Sha ZF, Cui X, Qiu SQ, He CG, Zhang JL, et al. Incorporation of redox-active polyimide binder into LiFePO4 cathode for high-rate electrochemical energy storage. Nanotechnol Rev. 2020;9:1350–58.10.1515/ntrev-2020-0092Search in Google Scholar

[43] Levi MD, Aurbach D. Impedance of a single intercalation particle and of non-homogeneous, multilayered porous composite electrodes for Li-ion batteries. J Phys Chem B. 2004;108:11693–703.10.1021/jp0486402Search in Google Scholar

Received: 2021-03-17
Revised: 2021-03-31
Accepted: 2021-04-01
Published Online: 2021-04-20

© 2021 Fangfang Wang et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 23.4.2024 from https://www.degruyter.com/document/doi/10.1515/ntrev-2021-0020/html
Scroll to top button