Next Article in Journal
Tuning Transport Phenomena in Agarose Gels for the Control of Protein Nucleation Density and Crystal Form
Previous Article in Journal
Crystallization Process and Site-Selective Excitation of Nd3+ in LaF3/NaLaF4 Sol–Gel-Synthesized Transparent Glass-Ceramics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Ni2+ on Lithium-Ion Diffusion in Layered LiNi1−x−yMnxCoyO2 Materials

1
Institute of High Energy Physics, Chinese Academy of Sciences (CAS), Beijing 100049, China
2
Spallation Neutron Source Science Center, Dongguan 523808, China
3
School of Nuclear Science and Technology, University of Chinese Academy of Sciences, Beijing 100049, China
4
Chongqing Institute of Green and Intelligent Technology, Chinese Academy of Sciences (CAS), Chongqing 400714, China
5
Collaborative Innovation Center of Extreme Optics, Shanxi University, Taiyuan 030006, China
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(5), 465; https://doi.org/10.3390/cryst11050465
Submission received: 17 March 2021 / Revised: 10 April 2021 / Accepted: 13 April 2021 / Published: 22 April 2021

Abstract

:
LiNi 1 x y Mn x Co y O 2 materials are a typical class of layered cathode materials with excellent electrochemical performance in lithium-ion batteries. Molecular dynamics simulations are performed for LiNi 1 x y Mn x Co y O 2 materials with different transition metal ratios. The Li/Ni exchange ratio, ratio of anti-site Ni 2 + to total Ni 2 + , and diffusion coefficient of Li ions in these materials are calculated. The results show that the Li-ion diffusion coefficient strongly depends on the ratio of anti-site Ni 2 + to total Ni 2 + because their variation tendencies are similar. In addition, the local coordination structure of the Li/Ni anti-site is analyzed.

1. Introduction

Lithium-ion batteries (LIBs) are promising chemical batteries, which have wide applications in mobile electronic devices and electric vehicles. It is well known that the working principle of LIBs is based on the intercalation/de-intercalation of Li ions in positive and negative electrodes during charge/discharge. Therefore, the diffusion of Li ions in the cathode and anode materials plays a dominant role in the charging and discharging ratio of LIBs [1,2,3,4]. Among the various cathode materials, layered transition metal oxides LiNi 1 x y Mn x Co y O 2 (NMC) have attracted wide attention. Compared with LiCoO 2 , LiMn 2 O 4 , and   LiNiO 2 , NMC materials have excellent electrochemical properties, such as high energy density, high reversibility, good environmental compatibility, and an excellent charge-discharge rate [5,6,7,8,9,10,11,12,13]. In previous studies, these electrochemical properties of NMC materials have been investigated for different transition metal (TM) ratios, including LiNi 1 / 3 Mn 1 / 3 Co 1 / 3 O 2 (333) [8], LiNi 0.4 Mn 0.4 Co 0.2 O 2 (442) [9], LiNi 0.5 Mn 0.3 Co 0.2 O 2 (532) [10], LiNi 0.6 Mn 0.2 Co 0.2 O 2 (622) [11], LiNi 0.7 Mn 0.15 Co 0.15 O 2 (71515), [12] and LiNi 0.8 Mn 0.1 Co 0.1 O 2 (811) [13]. Among these electrochemical properties, the Li-ion diffusion coefficient, which is an important parameter for LIB cathode materials, directly determines the charge and discharge rate capability [14]. In particular, the rapid transportation of Li ions in cathode materials is a key factor for high-power density applications. Therefore, many experimental and theoretical studies have been conducted on the diffusion properties of Li ions in NMC materials [15,16,17,18,19]. For example, Yang et al. used the galvanostatic intermittent titration technique (GITT) and electrochemical impedance spectroscopy (EIS) to study the variation of the Li-ion diffusion coefficient with Li concentration in 532 [15]. In addition, Noh et al. investigated the influence of Ni content on the Li-ion diffusion coefficient of Li [ Ni x Co y Mn z ] O 2 (x = 1/3, 0.5, 0.6, 0.7, 0.8, 0.85) [16].
In NMC materials, Ni, Mn, and Co are distributed in the TM layer. Mn and Co primarily stabilize the interlayer structure during charging and discharging [20,21], and Ni participates in redox behavior [22,23,24]. The Ni content and valence directly affect the diffusion coefficient of Li ions. Generally, Li/Ni exchange (anti-site) is formed in this system. The exchange between Li + and Ni 2 + at the 3a and 3b sites, respectively, impedes the migration path of Li ions. The anti-site Ni 2 + narrows the Li layer space, which reduces the Li-ion diffusion [25,26]. Moreover, the anti-site Ni 2 + migrates to the material surface and destroys the stability of the structure. This results in rate performance degradation and capacity fading [27,28]. However, previous studies have reported that the high Ni 2 + concentration in the TM layer is conducive to Li-ion diffusion [29,30]. As mentioned above, Li-ion diffusion can be tuned by Ni 2 + in the TM layer and the anti-site Ni 2 + in the Li layer. Thus, it is important to understand the relationship between Ni 2 + concentration and the Li-ion diffusion coefficient to optimize the layered NMC materials. Molecular dynamics (MD) can track the movement of   Li + , which is useful for the study of Li-ion diffusion [31,32,33,34].
In this paper, we focus on Ni-rich NMC materials because there is a significant Li/Ni exchange in them, which will impact the crystal structure and electrochemical performance of materials. We perform MD simulations of NMC materials with different TM ratios, which are LiNi 0.8 Mn x Co 0.2 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2), LiNi 0.7 Mn x Co 0.3 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3), LiNi 0.6 Mn x Co 0.4 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35), and LiNi 0.5 Mn x Co 0.5 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4), to research the influence of Ni 2 + content on Li-ion diffusion. In Section 2, the details of the MD simulation are presented. In Section 3, the effect of Ni 2 + content on Li-ion diffusion is analyzed. In the last part, the local coordination structure of the Li/Ni anti-site is discussed.

2. Simulation Method

MD simulations were carried out using the DL_POLY program developed by Todorov and Smith [35]. Model building and atomic visualization were realized using CrystalMaker (CrystalMaker Software Limited, Oxfordshire, UK) [36]. In NMC materials, the Ni, Mn and Co coexist at the 3b Wyckoff position. The Ni exists as both Ni2+ and Ni3+, while Co and Mn stay as Co3+ and Mn4+, respectively. Although there is no primitive cell of NMC material in the cell database, the structure of NMC is the same as that of LiNiO2. In modeling, we first obtained a 35 × 35 × 12 supercell of LiNiO2 which contains 176,400 atoms in total, and then randomly substituted Ni atoms with Mn and Co atoms. The specific contents of Ni2+, Ni3+, Mn4+, and Co3+ in different NMC materials are listed in Table S1. Most NMC materials simultaneously consist of two valence states, Ni2+ and Ni3+, to keep the system electrically neutral [37,38,39]. For instance, when establishing the initial configuration of 622, 8820 Mn4+ and 8820 Co3+ are used to substitute the equal number of Ni atoms. As a result, the TM layer of 622 contains 8820 Ni2+, 17640 Ni3+, 8820 Mn4+ and 8820 Co3+. The stoichiometry of 622 can be written as LiNi 0.2 2 + Ni 0.4 3 + Mn 0.2 4 + Co 0.2 3 + O 2 [37]. The Ewald method and Verlet algorithm were used to calculate the Coulomb interactions and atomic motions, respectively [40,41]. The NVT ensemble was used in all simulations. After a lot of tests, the time required to reach equilibration was 15 ps, and the runtime of the entire simulation was 80 ps. For comparison, 442, 532, 622, and 71515 were simulated at the same temperature and state of charge (SOC) as in the work by Wei et al. [30]. For LiNi 0.8 Mn x Co 0.2 x O 2 (x = 0, 0.05, 0.1 0.15, 0.2), LiNi 0.7 Mn x Co 0.3 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3), LiNi 0.6 Mn x Co 0.4 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35), and LiNi 0.5 Mn x Co 0.5 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4), the temperature was maintained at 1500 K, which is approximately the sintering temperature [42,43]. In addition, the temperature of 333 was maintained at 1500 K to study the local structure of the Li/Ni anti-site.
Because NMC materials are Li-doped multicomponent oxides, it is appropriate to consider the pairwise potential with partial charges for our simulations. The Morse function is often used to build bond interactions in covalent systems, which fits our potential model very well. The specific functional form is given by Equation (1) [44]
U ( r ) = Z i Z j e 2 / r + D i j [ { 1 e a i j ( r r 0 ) } 2 1 ] + C i j / r 12 ,
where Z i and Z j are the atomic charges, e is the elementary electric charge, r is the interatomic distance, D i j represents the dissociation energy that controls the depth of the potential well, a i j is the parameter that controls the width of the potential well, and C i j is the parameter representing the finite distance at which the inter-particle potential is zero. In Equation (1), the first term refers to the Coulomb interaction, the second term to the Morse potential, and the third term to the long-range repulsive interaction. In our previous work [45], we have established a good potential field model for transition metal (TM)/lithium–oxygen interactions. It is convenient to transplant them to the Li-TM-O system of cathode materials. We used the empirical fitting method implemented in the GULP package [46] to derive the interatomic potential parameters [44]. In our strategy, the parameters are derived to reproduce the experimental crystal structures (lattice parameters and atomic positions). Table 1 lists the final potential-field parameters used in our simulations. In addition, we calculated the lattice parameters of these NMC material as shown in Figure S1, which is in good agreement with the experiment.
In this study, the diffusion dynamics and local molecular structure of Li ions are investigated by analyzing the diffusion coefficient and defect energy of the Li/Ni anti-site. From MD simulations, we can obtain the mean square displacement (MSD) of Li-ion diffusion, as shown in Equation (2). Based on MSD, the diffusion coefficient of Li ions can be calculated using the Einstein relation as follows [47].
M S D ( t ) = | r ( t ) r ( t 0 ) | 2 ¯ ,
D L i = 1 / 6 lim t d [ | r ( t ) r ( t 0 ) | 2 ] ¯ / d t ,
where r (t) is the position vector of a Li ion at time t.
To understand the local coordination structure of the Li/Ni anti-site, the defect energies were calculated with different coordination structures using the Mott−Littleton method, which was implemented in GULP. The defect energy is the difference between the total energy of the defective lattice and the perfect lattice, which is defined as:
E d e f e c t = E d E p + E ,
where E d and E p are the total energies of the defective and perfect lattices, respectively. E represents the energy correction of any interstitial or vacancy at infinite separation from the lattice. It is easier to form a Li/Ni anti-site with a smaller defect energy E d e f e c t [45]. There have been some recent studies on the calculation of the Li/TM anti-site defect [48,49,50,51,52,53], and we used the same method to calculate the Li/Ni anti-site defects in our work. In this method, a defect center is defined. Two spherical regions are divided surrounding the defect center. The ions in the inner sphere are assumed to be strongly disturbed by the defects and therefore are relaxed explicitly with respect to their Cartesian coordinates. The ions in the outer sphere are assumed to be weakly perturbed, and they can be treated implicitly as a dielectric continuum. As the radius of the inner sphere increases, these approximations become more effective. Thus, an important stage of the defect calculation is to ensure sufficient convergence of the defect energy relative to the region radius. According to the research of Read and Jackson [54], the inner sphere radius larger than 11 Å is the most suitable, and the energy convergence reaches an optimal value when the radius of the outer sphere is twice that of the inner sphere. In our calculation, the radii of the inner and outer spheres are 13 Å and 26 Å, respectively.

3. Results and Discussion

3.1. Effect of N i 2 + on Li-Ion Diffusion

The space group of LiNi 1 x y Mn x Co y O 2 is R-3m. Li and O atoms are located at the 3a and 6c sites, respectively. TM ion (Ni, Mn, and Co) are occupied at the 3b site. The Li, O, and TM layers are stacked layer by layer consecutively in the ABC form (O3-type), and the structure is shown in Figure 1. There are two migration pathways for Li ions in NMC materials. In one, the Li ion directly hops from one octahedral site to the next through the center of the oxygen dumbbell, called oxygen dumbbell hopping (ODH). In the other, Li ion first hops from one octahedral site to a transition state tetrahedral site and then hops to the next octahedral site, called tetrahedral site hopping (TSH) [30,55,56]. Previous studies have reported the effect owing to the type of TM ions and their valence states ( Ni 2 + , Ni 3 + , Co 3 + , and Mn 4 + ) [29,30]. Among these TM ions, Ni 2 + has the smallest electrostatic repulsion with migrating Li ions. According to the research by Kang et al., the activation barrier is 210 meV, which is the lowest compared with   Ni 3 + , Co 3 + , and Mn 4 + when the nearby TM cation is Ni 2 + [30]. Therefore, regardless of the dominant diffusion path, when the migrating Li ion is close to the Ni 2 + in the TM layer, the activation barrier is very low. This suggests that a high Ni 2 + concentration in the TM layer is conducive to the diffusion of Li ions. However, Ni 2 + from the TM layer exchanges with Li + from the Li layer to form the Li/Ni anti-site defect because the ionic radii of Ni 2 + (0.69 Å) and Li + (0.76 Å) are similar. The anti-site Ni 2 + in the Li layer impedes the diffusion of Li ions. The Li/Ni exchange ratios of NMC materials with different TM ratios are calculated using the MD simulation, as shown in Figure 2 (the detailed values are listed in Table S2). It should be noted that the ratios obtained from our simulations are in good agreement with the experimental result [42,43,57,58,59,60]. It is clear that the Li/Ni exchange ratio increases with increasing Ni 2 + content (Figure 2). In addition, the Li/Ni exchange ratio also increases with increasing total Ni content when the Ni 2 + content remains the same. For example, when x = 0.1 , the Li/Ni exchange ratios of 514, 613, 712, and 811 are 1.8%, 2.8%, 3.96%, and 5%, respectively (Table S2). According to the research of Zheng et al. [61], there are mixed valence states of Ni 2 + and Ni 3 + in Ni-rich NMC materials. A small amount of Ni 3 + transforms into the Ni 2 + state using an additional electron from the nearest Co 3 + in the TM layer. The Co 3 + changes to Co 4 + owing to the charge compensation mechanism. The higher the total Ni content, the more Ni 3 + Ni 2 + transformations occur. Thus, a higher total Ni content leads to more Li/Ni anti-sites in Ni-rich materials. Essentially, it is still increasing with the amount of Ni 2 + . In brief, the Li/Ni anti-sites are positively correlated with Ni 2 + content, which indicates that the structural disorder increases with increasing Ni 2 + content.
The same NMC materials mentioned in the study by Wei et al. [30], including 442, 532, 622, and 71515, were simulated to compare the results obtained using our method. Figure 3 shows the relationship between the Li-ion diffusion coefficient (SOC = 1/3) and the Ni content. The variation tendency of our simulation is consistent with the experiment and the ab initio molecular dynamics (AIMD) simulation, which proves the reliability of our simulation. Although there is a discrepancy of nearly two orders of magnitude between the calculated and experimental results, this difference is unavoidable due to the errors in the experimental method and conditions [14,16,17,30]. In our work, we concentrate on the effect of Ni 2 + content on the Li-ion diffusion coefficient at SOC = 0.
Based on our previous discussions, there are primarily two mechanisms for tuning the diffusion of Li ions. The more Ni2+ in the TM layer, the more conducive to the diffusion of Li ions. However, the more Ni2+ in the TM layer would cause more Li/Ni exchange to form (Figure 2). The anti-site Ni2+ in the Li layer hinders the transportation of Li ions. In a word, the Li-ion diffusion coefficient is tuned by the Ni2+ of TM layer and the anti-site Ni2+ of Li layer. The relationship between the Li-ion diffusion coefficient ( D Li ) and Ni 2 + content is obtained from the MD simulation, as shown in Figure 4. For LiNi 0.8 Mn x Co 0.2 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2) and LiNi 0.7 Mn x Co 0.3 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3) (Figure 4a,b). The D Li   increases from 0 cm 2 s 1 for x = 0 to 1.5 ×   10 8 cm 2 s 1 for x = 0.15 and from 0 cm 2 s 1 for x = 0 to 1.6 ×   10 8   cm 2 s 1 for x = 0.15. It indicates that the positive effect of Ni2+ in the TM layer on Li ion diffusion is greater than the negative effect of anti-site Ni2+ in the Li layer when the content of Ni2+ is low. Subsequently, D Li decreases from 1.5 ×   10 8   cm 2 s 1 for x = 0.15 to 1.4 ×   10 8   cm 2 s 1 for x = 0.2 and from 1.6 ×   10 8   cm 2 s 1 for x = 0.15 to 1.2 ×   10 8   cm 2 s 1 for x = 0.25. Because the Ni2+ content continues increasing, the Li/Ni anti-site increases. The negative effect of Ni2+ in the Li layer on Li ion diffusion increases and plays a dominant role. Moreover, there is a maximum value of D Li at x = 0.15 (1.5 ×   10 8   cm 2 s 1 for LiNi 0.8 Mn x Co 0.2 x O 2 and 1.6 ×   10 8   cm 2 s 1 for LiNi 0.7 Mn x Co 0.3 x O 2 ). This indicates that the Ni 2 + content at x = 0.15 is the most favorable for Li-ion diffusion in these two series. In addition, the effect of Ni 2 + content on Li-ion diffusion is also slightly different, owing to the difference in the total Ni content. In Figure 4c,d, for the LiNi 0.6 Mn x Co 0.4 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35) and LiNi 0.5 Mn x Co 0.5 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4), the increase range is from x = 0 to x = 0.1. The decrease areas are from x = 0.1 to x = 0.2 and from x = 0.1 to x = 0.15 in Figure 4c,d, respectively. The maximum value of D Li occurs at x = 0.1 (1.5 ×   10 8   cm 2 s 1 for LiNi 0.6 Mn x Co 0.4 x O 2 and 1.43 ×   10 8 cm 2 s 1 for LiNi 0.5 Mn x Co 0.5 x O 2 ). Furthermore, when the Ni 2 + content is high (x   0.2), D Li shows a tendency to first increase and then decrease.
The ratio of anti-site Ni 2 + ( Ni anti 2 + ) to total Ni 2 + ( Ni total 2 + ) is calculated for further analysis, and we named it η, which is illustrated in Figure 4. The corresponding curves exhibit the same tendency. The maximum values of the corresponding curves are not the same (Figure 4). However, they are within the range of low Ni 2 + content (x 0.15 ). Within this range, there are a few Li/Ni anti-sites. Ni 2 + mainly exists in the TM layer, which facilitates the diffusion of Li ions. Taking the case of LiNi 0.7 Mn x Co 0.3 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3) as an example, in Figure 4b, D Li reaches its maximum value at x = 0.15 , while η is maximized at x = 0.1 . The D Li of 71515 is greater than that of 712, which can be explained by their corresponding η. Although both the Ni anti 2 + and Ni total 2 + of 71515 are higher than that of 712, the value of η of 71515 is smaller than that of 712 (Figure 4b). Thus, there is more Ni 2 + left in the TM layer of 71515 than of 712. Ni 2 + in the TM layer is beneficial for the diffusion of Li ions. Thus, the D Li of 71515 is greater than that of 712 (Figure 4b), which is consistent with the previous discussion. The facilitating effect caused by Ni 2 + in the TM layer plays a leading role when the Ni 2 + content is low (x 0.15). Thus, the D Li of 71515 is greater than that of 721 (Figure 4b). As the Ni 2 + content continues increasing, η begins to decline and reaches a minimum value. The minimum values of D Li   and η correspond to each other in Figure 4. This indicates that Li-ion diffusion is more impeded by a smaller η. As discussed above, the impeding effect caused by the anti-site Ni 2 + in the Li layer plays a dominant role when the Ni 2 + content is high (x   0.2). In a word, D Li depends strongly on η, especially when the Ni 2 + concentration is high (x   0.2). D Li is positively correlated to η.

3.2. Local Coordination Structure of Li/Ni Anti-Site

The formation of the Li/Ni anti-site indicates a change in the local coordination structure, as shown in Figure 5. Before the Li/Ni anti-site, the Ni ion is surrounded by the six nearest TM ions in the TM layer and is located in the center of a hexagon composed of these six TM ions. After the Li/Ni anti-site, the anti-site Li ion is located in the center of these six TM ions. The anti-site Ni ion connects to the five nearest TM ions and one Li ion in the upside TM layer, and six nearest TM ions in the downside TM layer via intermediate oxygen atoms. In this work, 333 is chosen for further investigation as it contains the same contents of Ni, Mn, and Co. We calculated the defect energy of different local coordination environments, in which the number of Mn ions increased from zero to six. Moreover, we also considered the influence of different arrangements on the defect energy when the numbers of Ni, Mn, and Co are determined. Thus, the defect energies of all the cases were calculated. The result is illustrated in Figure 6, which is supported by the research of Lee et al. The defect energy of the Li/Ni anti-site is approximately 0.84 eV, which is consistent with our calculation [62]. As shown in Figure 6, the defect energy ( E d   ) exhibits a downward tendency with an increasing number of Mn ions. When there are few Mn ions (Figure 6a–c), E d   is typically high. This is not conducive to the stable existence of the Li/Ni anti-site. When the number of Mn ions is in the majority (Figure 6e–g), E d   is generally low. This is favorable for the formation of the Li/Ni anti-site. Additionally, we calculated the variation in the defect energy with the increasing number of Co ions. The results are presented in Figure 7. E d   exhibits no evident increase or decrease tendency with an increasing Co number from Figure 7a–g. However, E d   changes significantly at the same Co number. For instance, in Figure 7b, which contains one Co ion, E d exhibits a tendency to decline when the Mn number increases from left to right. This means that it is easy to form the Li/Ni anti-site when Mn ions are in the majority among the six TM ions which surround the Ni ion. Furthermore, in both Figure 6 and Figure 7, the defect energy of the Li/Ni anti-site is different when the Ni, Mn, and Co numbers are confirmed. This reveals that different arrangements influence the E d   of the Li/Ni anti-site.

4. Conclusions

In this article, we performed MD simulations to investigate the influence of Ni 2 + on Li-ion diffusion in layered LiNi 1 x y Mn x Co y O 2 with Li/Ni anti-sites. Firstly, the Li-ion diffusion coefficients of NMC materials with different TM ratios were calculated. We found a peak value for the Li-ion diffusion coefficient with increasing Ni 2 + content in each series of NMC materials. Then, we calculated the ratio of the anti-site Ni 2 + to total Ni 2 + . The variation tendency of this ratio is consistent with that of the Li-ion diffusion coefficient. This indicates that the Li-ion diffusion coefficient strongly depends on the ratio of anti-site Ni 2 + to total Ni 2 + . Besides, by calculating the defect energy of the Li/Ni anti-site for different local coordination environments, we found that it is easy to form the Li/Ni anti-site when Mn ions are in the majority among the six TM ions which surround the Ni ion. According to our research, an appropriate Ni 2 + can be selected to maximize the diffusion of Li ions, which is of great significance for the design and optimization of layered NMC materials.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cryst11050465/s1, Table S1: LiNi 1 x y Mn x Co y O 2 materials with different TM ratios in this work. LiNi 0.8 Mn x Co 0.2 x O 2 (x = 0, 0.05, 0.1 0.15, 0.2), LiNi 0.7 Mn x Co 0.3 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3), LiNi 0.6 Mn x Co 0.4 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35), and LiNi 0.5 Mn x Co 0.5 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4). Table S2: The value of the Li/Ni exchange ratio, for LiNi 0.8 Mn x Co 0.2 x O 2 (x = 0, 0.05, 0.1 0.15, 0.2), LiNi 0.7 Mn x Co 0.3 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3), LiNi 0.6 Mn x Co 0.4 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35), and LiNi 0.5 Mn x Co 0.5 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4). Figure S1: The comparison of lattice parameters between MD simulations and experimental values.

Author Contributions

Formal analysis and writing—original draft, Y.Z.; methodology, Y.H.; software, R.D. and M.T.; writing—review and editing, J.Z.; resources, B.W.; funding acquisition and supervision, J.Z.; All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (Grant No.11675195), the National Key R&D Program of China (2017YFA0403700), and Key Research Program of Frontier Sciences, CAS, Grant No. 292016YQYKXJ00135.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, B.; Qian, D.; Wang, Z.; Meng, Y.S. Recent progress in cathode materials research for advanced lithium ion batteries. Mater. Sci. Eng. R 2012, 73, 51–65. [Google Scholar] [CrossRef]
  2. Lu, L.; Han, X.; Li, J.; Hua, J.; Ouyang, M. A review on the key issues for lithium-ion battery management in electric vehicles. J. Power Source 2013, 226, 272–288. [Google Scholar] [CrossRef]
  3. Zhao, E.; He, L.; Wang, B.; Li, X.; Zhang, J.; Wu, Y.; Chen, J.; Zhang, S.; Liang, T.; Chen, Y.; et al. Structural and mechanistic revelations on high capacity cation-disordered Li-rich oxides for rechargeable Li-ion batteries. Energy Storage Mater. 2019, 16, 354–363. [Google Scholar] [CrossRef]
  4. Huang, Y.-R.; Chen, C.-L.; Pu, N.-W.; Wu, C.-H.; Liu, Y.-M.; Chen, Y.-H.; Youh, M.-J.; Ger, M.-D. Experimental and Modeling Analysis of Holey Graphene Electrodes for High-Power-Density Li-Ion Batteries. Crystals 2020, 10, 1063. [Google Scholar] [CrossRef]
  5. 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]
  6. Schipper, F.; Erickson, E.M.; Erk, C.; Shin, J.-Y.; Chesneau, F.F.; Aurbach, D. Review—Recent Advances and Remaining Challenges for Lithium Ion Battery Cathodes I. Nickel-Rich, LiNixCoyMnzO2. J. Electrochem. Soc. 2017, 164, A6220–A6228. [Google Scholar] [CrossRef]
  7. Tuccillo, M.; Palumbo, O.; Pavone, M.; Muñoz-García, A.B.; Paolone, A.; Brutti, S. Analysis of the Phase Stability of LiMO2 Layered Oxides (M = Co, Mn, Ni). Crystals 2020, 10, 526. [Google Scholar] [CrossRef]
  8. Wang, J.-H.; Wang, Y.; Guo, Y.-Z.; Liu, C.-W.; Dan, L.-L. Electrochemical characterization of AlPO4 coated LiNi1/3Co1/3Mn1/3O2 cathode materials for high temperature lithium battery application. Rare Met. 2014, 40, 78–83. [Google Scholar] [CrossRef]
  9. Satyanarayana, M.; James, J.; Varadaraju, U.V. Electrochemical performance of LiNi0.4Co0.2Mn0.4O2 prepared by different molten salt flux: LiNO3-LiCl and LiNO3-KNO3. Appl. Surf. Sci. 2017, 418, 72–78. [Google Scholar]
  10. Zhao, X.; Liu, B.; Yang, J.; Hou, J.; Wang, Y.; Zhu, Y.; Zhao, X.; Liu, B.; Yang, J.; Hou, J.; et al. Synthesizing LiNi0.5 Co0.2Mn0.3O2 with microsized peanut-like structure for enhanced electrochemical properties of lithium ion batteries. J. Alloys Compd. 2020, 832, 154464. [Google Scholar] [CrossRef]
  11. Chen, Y.; Tang, S.; Deng, S.; Lei, T.; Li, Y.; Li, W.; Cao, G.; Zhu, J.; Zhang, J. Chemical coupling constructs amorphous silica modified LiNi0.6Co0.2Mn0.2O2 cathode materials and its electrochemical performances. J. Power Source 2019, 431, 8–16. [Google Scholar] [CrossRef]
  12. Li, W.; Li, Y.; Yang, L.; Chen, Y.; Guo, J.; Zhu, J.; Cao, G.L. Enhancing high-voltage electrochemical performance of LiNi0.7Mn0.15Co0.15O2 cathode materials with SiO2 coatings via electrostatic attraction forces method. Ionics 2020, 26, 5393–5403. [Google Scholar] [CrossRef]
  13. Liang, J.; Lu, Y.; Wang, J.; Liu, X.; Chen, K.; Ji, W.; Zhu, Y.; Wang, D. Well-ordered layered LiNi0.8Co0.1Mn0.1O2 submicron sphere with fast electrochemical kinetics for cathodic lithium storage. J. Energy Chem. 2020, 47, 188–195. [Google Scholar] [CrossRef]
  14. Park, M.; Zhang, X.; Chung, M.; Less, G.B.; Sastry, A.M. A review of conduction phenomena in Li-ion batteries. J. Power Source 2010, 195, 7904–7929. [Google Scholar] [CrossRef]
  15. Yang, S.; Wang, X.; Yang, X.; Bai, Y.; Liu, Z.; Shu, H.; Wei, Q. Determination of the chemical diffusion coefficient of lithium ions in spherical Li[Ni0.5Mn0.3Co0.2]O2. Electrochim. Acta 2012, 66, 88–93. [Google Scholar] [CrossRef]
  16. Noh, H.J.; Youn, S.; Yoon, C.S.; Sun, Y.K. 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 Source 2013, 233, 121–130. [Google Scholar] [CrossRef]
  17. Li, Z.; Ban, C.; Chernova, N.A.; Wu, Z.; Upreti, S.; Dillon, A.; Whittingham, M.S. Towards understanding the rate capability of layered transition metal oxides. J. Power Source 2014, 268, 106–112. [Google Scholar] [CrossRef]
  18. Gao, A.; Sun, Y.; Zhang, Q.; Zheng, J.; Lu, X. Evolution of Ni/Li antisites under the phase transition of a layered LiNi1/3Co1/3Mn1/3O2 cathode. J. Mater. Chem. A 2020, 8, 6337–6348. [Google Scholar] [CrossRef]
  19. Cui, S.; Wei, Y.; Liu, T.; Deng, W.; Hu, Z.; Su, Y.; Li, H.; Li, M.; Guo, H.; Duan, Y.; et al. Optimized Temperature Effect of Li-Ion Diffusion with Layer Distance in Li(NixMnyCoz)O2 Cathode Materials for High Performance Li-Ion Battery. Adv. Energy Mater. 2016, 6, 1501309. [Google Scholar] [CrossRef]
  20. Yamada, A.; Tanaka, M.; Tanaka, K.; Sekai, K. Jahn–Teller instability in spinel Li–Mn–O. J. Power Source 1999, 81–82, 73–78. [Google Scholar] [CrossRef]
  21. Ellis, B.L.; Lee, K.T.; Nazar, L.F. Positive Electrode Materials for Li-Ion and Li-Batteries. Chem. Mater. 2010, 22, 691–714. [Google Scholar] [CrossRef]
  22. Guo, J.; Jiao, L.F.; Yuan, H.T.; Li, H.X.; Zhang, M.; Wang, Y.M. Effect of synthesis condition on the structural and electrochemical properties of Li[Ni1/3Mn1/3Co1/3]O2 prepared by the metal acetates decomposition method. Electrochim. Acta 2006, 51, 3731–3735. [Google Scholar] [CrossRef]
  23. Hwang, B.J.; Santhanam, R.; Chen, C.H. Effect of synthesis conditions on electrochemical properties of LiNi1−yCoyO2 cathode for lithium rechargeable batteries. J. Power Source 2003, 114, 244–252. [Google Scholar] [CrossRef]
  24. Ren, H.; Huang, Y.; Wang, Y.; Li, Z.; Cai, P.; Peng, Z.; Zhou, Y. Effects of different carbonate precipitators on LiNi1/3Co1/3Mn1/3O2 morphology and electrochemical performance. Mater. Chem. Phys. 2009, 117, 41–45. [Google Scholar] [CrossRef]
  25. Kang, K.; Ceder, G. Factors that affect Li mobility in layered lithium transition metal oxides. Phys. Rev. B 2006, 74, 094105. [Google Scholar] [CrossRef] [Green Version]
  26. Yu, H.; Qian, Y.; Otani, M.; Tang, D.; Guo, S.; Zhu, Y.; Zhou, H. Study of the lithium/nickel ions exchange in the layered LiNi0.42Mn0.42Co0.16O2 cathode material for lithium ion batteries: Experimental and firstprinciples calculations. Energy Environ. Sci. 2014, 7, 1068–1078. [Google Scholar] [CrossRef]
  27. Yan, P.; Nie, A.; Zheng, J.; Zhou, Y.; Lu, D.; Zhang, X.; Xu, R.; Belharouak, I.; Zu, X.; 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]
  28. Yan, P.; Zheng, J.; Lv, D.; Wei, Y.; Zheng, J.; Wang, Z.; Kuppan, S.; Yu, J.; Luo, L.; Edwards, D.; et al. Atomic-Resolution Visualization of Distinctive Chemical Mixing Behavior of Ni, Co, and Mn with Li in Layered Lithium Transition-Metal Oxide Cathode Materials. Chem. Mater. 2015, 27, 5393–5401. [Google Scholar] [CrossRef]
  29. Kang, K.; Meng, Y.S.; Breger, J.; Grey, C.P.; Ceder, G. Electrodes with High Power and High Capacity for Rechargeable Lithium Batteries. Science 2006, 311, 977–980. [Google Scholar] [CrossRef]
  30. Wei, Y.; Zheng, J.; Cui, S.; Song, X.; Su, Y.; Deng, W.; Wu, Z.; Wang, X.; Wang, W.; Rao, M.; et al. Kinetics Tuning of Li-Ion Diffusion in Layered Li(NixMnyCoz)O2. J. Am. Chem. Soc. 2015, 137, 8364–8367. [Google Scholar] [CrossRef]
  31. Lychagin, D.; Dmitriev, A.; Nikonov, A. Alfyorova, Crystallographic and Geometric Factors in the Shear Development in <001> FCC Single Crystals: Molecular Dynamics Simulation and Experimental Study. Crystals 2020, 10, 666. [Google Scholar] [CrossRef]
  32. Wang, B.-B.; Xiao, Y.; Xu, Z.-M. Variation in Properties of Pre-Nucleation Calcium Carbonate Clusters Induced by Aggregation: A Molecular Dynamics Study. Crystals 2021, 11, 102. [Google Scholar] [CrossRef]
  33. Hu, B.; Tao, G. Molecular Dynamics Simulations on Lithium Diffusion in LiFePO4: The effect of anti-site defects. J. Mater. Chem. A 2015, 3, 20399–20407. [Google Scholar] [CrossRef]
  34. Tateishi, K.; du Boulay, D.; Ishizawa, N. The effect of mixed Mn valences on Li migration in LiMn2O4 spinel: A molecular dynamics study. Appl. Phys. Lett. 2004, 84, 529–531. [Google Scholar] [CrossRef]
  35. Smith, W.; Forester, T.R.; Todorov, I.T. The DL_POLY_4 User Manual; Version 4.08; STFC Daresbury Laboratory Daresbury: Warrington, UK, 2016. [Google Scholar]
  36. CrystalMaker Software Ltd., Centre for Innovation & Enterprise, Begbroke Science Park, Woodstock Road, Begbroke, Oxfordshire, OX5 1PF, UK. Available online: http://www.crystalmaker.com (accessed on 15 April 2021).
  37. Ahn, W.; Lim, S.N.; Jung, K.-N.; Yeon, S.-H.; Kim, K.-B.; Song, H.S.; Shin, K.-H. Combustion-synthesized LiNi0.6Mn0.2Co0.2O2 as cathode material for lithium ion batteries. J. Alloys Compd. 2014, 609, 143–149. [Google Scholar] [CrossRef]
  38. Guo, J.; Hu, C.Y. Synthesis of nanosized LiNi0.7Mn0.2Co0.1O2 cathode material for lithium ion batteries by combination method of forced hydrolytic and hydroxide coprecipitation. Mater. Res. Innov. 2014, 19, 238–243. [Google Scholar] [CrossRef]
  39. Kosova, N.V.; Devyatkina, E.T.; Kaichev, V.V. Optimization of Ni2+/Ni3+ ratio in layered Li(Ni,Mn,Co)O2 cathodes for better electrochemistry. J. Power Source 2007, 174, 965–969. [Google Scholar] [CrossRef]
  40. Ewald, P.P. The calculation of optical and electrostatic grid potential. Ann Phys. 1921, 64, 253–287. [Google Scholar] [CrossRef] [Green Version]
  41. Verlet, L. Computer “Experiments” on Classical Fluids. I. Thermodynamical Properties of Lennard-Jones Molecules. Phys. Rev. 1967, 159, 98–103. [Google Scholar] [CrossRef]
  42. Soo, S.P.; Idris, M.S.; Osman, R.A.; Rahmat, A. The Effect of Synthesis Temperature on Interlayer Mixing in Layered Rock Salt Cathode Materials LiNi0.7Mn0.1Co0.2O2 for Li-ion Batteries Application. Mater. Sci. Forum. 2015, 819, 155–160. [Google Scholar] [CrossRef]
  43. Xiao, Y.; Liu, T.; Liu, J.; He, L.; Chen, J.; Zhang, J.; Luo, P.; Lu, H.; Wang, R.; Zhu, W.; et al. Insight into the origin of lithium/nickel ions exchange in layered Li(NixMnyCoz)O2 cathode materials. Nano Energy 2018, 49, 77–85. [Google Scholar] [CrossRef] [Green Version]
  44. Pedone, A.; Malavasi, G.; Menziani, M.C.; Cormack, A.N.; Segre, U. A New Self-Consistent Empirical Interatomic Potential Model for Oxides, Silicates, and Silica-Based Glasses. J. Phys. Chem. B 2006, 110, 11780–11795. [Google Scholar] [CrossRef] [PubMed]
  45. Huang, Y.; Liu, L.; Gao, M. Molecular dynamics study on the Li diffusion mechanism and delithiation process of Li2MnO3. Solid State Ion. 2020, 346, 115195. [Google Scholar] [CrossRef]
  46. Gale, J.D. Gulp: A Computer Program for the Symmetry-Adapted Simulation of Solids. J. Chem. Soc. Faraday Trans. 1997, 93, 629–637. [Google Scholar] [CrossRef]
  47. Asadi, A.; Aghamiri, S.F.; Talaie, M.R. Molecular dynamics simulation of a LixMn2O4 spinel cathode material in Li-ion batteries. RSC Adv. 2016, 6, 115354–115363. [Google Scholar] [CrossRef]
  48. Fisher, C.A.J.; Prieto, V.M.H.; Islam, M.S. Lithium Battery Materials LiMPO4(M = Mn, Fe, Co, and Ni): Insights into Defect Association, Transport Mechanisms, and Doping Behavior. Chem. Mater. 2008, 20, 5907–5915. [Google Scholar] [CrossRef]
  49. Islam, M.S.; Driscoll, D.J.; Fisher, C.A.J.; Slater, P.R. Atomic-Scale Investigation of Defects, Dopants, and Lithium Transport in the LiFePO4 Olivine-Type Battery Material. Chem. Mater. 2005, 17, 5085–5092. [Google Scholar] [CrossRef]
  50. Armstrong, A.R.; Kuganathan, N.; Islam, M.S.; Bruce, P.G. Structure and Lithium Transport Pathways in Li2FeSiO4, Cathodes for Lithium Batteries. J. Am. Chem. Soc. 2011, 133, 13031–13035. [Google Scholar] [CrossRef] [Green Version]
  51. Fisher, C.A.J.; Kuganathan, N.; Islam, M.S. Defect chemistry and lithium-ion migration in polymorphs of the cathode material Li2MnSiO4. J. Mater. Chem. A 2013, 1, 4207–4214. [Google Scholar] [CrossRef] [Green Version]
  52. Tripathi, R.; Gardiner, G.R.; Islam, M.S.; Nazar, L.F. Alkali-ion Conduction Paths in LiFeSO4F and NaFeSO4F Tavorite-Type Cathode Materials. Chem. Mater. 2011, 23, 2278–2284. [Google Scholar] [CrossRef]
  53. Perera, D.; Ganeshalingam, S.; Kuganathan, N.; Chroneos, A. A Computational Study of Defects, Li-Ion Migration and Dopants in Li2ZnSiO4 Polymorphs. Crystals 2019, 9, 563. [Google Scholar] [CrossRef] [Green Version]
  54. Read, M.S.D.; Jackson, R.A. Derivation of enhanced potentials for uranium dioxide and the calculation of lattice and intrinsic defect properties. J. Nucl. Mater. 2010, 406, 293–303. [Google Scholar] [CrossRef]
  55. Lee, E.; Lee, K.-R.; Lee, B.-J. An interatomic potential for the Li-Co-O ternary system. Comput. Mater. Sci. 2018, 142, 47–58. [Google Scholar] [CrossRef]
  56. Van der Ven, A.; Ceder, G.; Asta, M.; Tepesch, P.D. First-principles theory of ionic diffusion with nondilute carriers. Phys. Rev. B 2001, 64, 184307. [Google Scholar] [CrossRef] [Green Version]
  57. Ngala, J.K.; Chernova, N.A.; Ma, M.; Mamak, M.; Zavalij, P.Y.; Whittingham, M.S. The synthesis, characterization and electrochemical behavior of the layered LiNi0.4Mn0.4Co0.2O2 compound. J. Mater. Chem. 2004, 14, 214–220. [Google Scholar] [CrossRef]
  58. Julien, C.; Mauger, A.; Zaghib, K.; Groult, H. Optimization of Layered Cathode Materials for Lithium-Ion Batteries. Materials 2016, 9, 595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Wang, D.; Xin, C.; Zhang, M.; Bai, J.; Zheng, J.; Kou, R.; Ko, J.Y.P.; Huq, A.; Zhong, G.; Sun, C.-J.; et al. Intrinsic Role of Cationic Substitution in Tuning Li/Ni Mixing in High-Ni Layered Oxides. Chem. Mater. 2019, 31, 2731–2740. [Google Scholar] [CrossRef]
  60. Zhang, N.; Li, J.; Li, H.; Liu, A.; Huang, Q.; Ma, L.; Li, Y.; Dahn, J.R. Structural, Electrochemical, and Thermal Properties of Nickel-Rich LiNixMnyCozO2 Materials. Chem. Mater. 2018, 30, 8852–8860. [Google Scholar] [CrossRef]
  61. Zheng, J.; Teng, G.; Xin, C.; Zhuo, Z.; Liu, J.; Li, Q.; Hu, Z.; Xu, M.; Yan, S.; Yang, W.; et al. Role of Superexchange Interaction on Tuning of Ni/Li Disordering in Layered Li(NixMnyCoz)O2. J. Phys. Chem. Lett. 2017, 8, 5537–5542. [Google Scholar] [CrossRef]
  62. Lee, S.; Park, S.S. Atomistic Simulation Study of Mixed-Metal Oxide (LiNi1⁄3Mn1⁄3Co1⁄3O2) Cathode Material for Lithium Ion Battery. J. Phys. Chem. C 2012, 116, 6484–6489. [Google Scholar] [CrossRef]
Figure 1. Crystal structure diagram of layered NMC materials. Ni, Mn, and Co are randomly distributed in the TM layer. Purple represents Mn, white is Ni, and dark blue is Co.
Figure 1. Crystal structure diagram of layered NMC materials. Ni, Mn, and Co are randomly distributed in the TM layer. Purple represents Mn, white is Ni, and dark blue is Co.
Crystals 11 00465 g001
Figure 2. The Li/Ni exchange ratio with the content of Ni 2 + in four series of NMC materials (a) LiNi 0.8 Mn x Co 0.2 x O 2 (x = 0, 0.05, 0.1 0.15, 0.2), (b) LiNi 0.7 Mn x Co 0.3 x O 2 (x= 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3), (c) LiNi 0.6 Mn x Co 0.4 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35) and (d) LiNi 0.5 Mn x Co 0.5 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4). Blue represents the result from our simulation, and red represents experimental values from previous literature [42,43,56,57,58,59].
Figure 2. The Li/Ni exchange ratio with the content of Ni 2 + in four series of NMC materials (a) LiNi 0.8 Mn x Co 0.2 x O 2 (x = 0, 0.05, 0.1 0.15, 0.2), (b) LiNi 0.7 Mn x Co 0.3 x O 2 (x= 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3), (c) LiNi 0.6 Mn x Co 0.4 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35) and (d) LiNi 0.5 Mn x Co 0.5 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4). Blue represents the result from our simulation, and red represents experimental values from previous literature [42,43,56,57,58,59].
Crystals 11 00465 g002
Figure 3. The diffusion coefficients of 442, 532, 622, and 71515. Black represents the experimental values measured by GITT [30]. Blue represents the values calculated obtained from AIMD simulations [30]. Red represents the calculated values by MD simulations in our work. It should be noted that all diffusion coefficients are obtained when SOC = 1/3.
Figure 3. The diffusion coefficients of 442, 532, 622, and 71515. Black represents the experimental values measured by GITT [30]. Blue represents the values calculated obtained from AIMD simulations [30]. Red represents the calculated values by MD simulations in our work. It should be noted that all diffusion coefficients are obtained when SOC = 1/3.
Crystals 11 00465 g003
Figure 4. The ratio of anti-site Ni 2 + to total Ni 2 + and D Li in four series of NMC materials. Black represents D Li and red represents η. (a) LiNi 0.8 Mn x Co 0.2 x O 2 (x = 0, 0.05, 0.1 0.15, 0.2), (b) LiNi 0.7 Mn x Co 0.3 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3), (c) LiNi 0.6 Mn x Co 0.4 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35), (d) LiNi 0.5 Mn x Co 0.5 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4).
Figure 4. The ratio of anti-site Ni 2 + to total Ni 2 + and D Li in four series of NMC materials. Black represents D Li and red represents η. (a) LiNi 0.8 Mn x Co 0.2 x O 2 (x = 0, 0.05, 0.1 0.15, 0.2), (b) LiNi 0.7 Mn x Co 0.3 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3), (c) LiNi 0.6 Mn x Co 0.4 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35), (d) LiNi 0.5 Mn x Co 0.5 x O 2 (x = 0, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.35, 0.4).
Crystals 11 00465 g004
Figure 5. The local coordination structure of Li/Ni anti-site. (a) Before Li/Ni anti-site, (b) after Li/Ni anti-site. Purple represents Mn, white represents Ni, dark blue represents Co, light blue represents Li, and red represents O.
Figure 5. The local coordination structure of Li/Ni anti-site. (a) Before Li/Ni anti-site, (b) after Li/Ni anti-site. Purple represents Mn, white represents Ni, dark blue represents Co, light blue represents Li, and red represents O.
Crystals 11 00465 g005
Figure 6. The defect energy of the Li/Ni anti-site with different local coordination environments. From (ag), the Mn number of six local TM ions increases in turn. The schematics at the bottom represent the local coordination ion environment for the anti-site Li-ion in the TM layer. For one schematic of the bottom, the different points of the longitudinal axis represent the defect energy of various arrangements under the corresponding number of Ni, Mn, and Co.
Figure 6. The defect energy of the Li/Ni anti-site with different local coordination environments. From (ag), the Mn number of six local TM ions increases in turn. The schematics at the bottom represent the local coordination ion environment for the anti-site Li-ion in the TM layer. For one schematic of the bottom, the different points of the longitudinal axis represent the defect energy of various arrangements under the corresponding number of Ni, Mn, and Co.
Crystals 11 00465 g006
Figure 7. The defect energy of Li/Ni anti-site with different local coordination environments. From (ag), the Co number of six local TM ions increases in turn. The Mn number increases from left to right under the same Co number.
Figure 7. The defect energy of Li/Ni anti-site with different local coordination environments. From (ag), the Co number of six local TM ions increases in turn. The Mn number increases from left to right under the same Co number.
Crystals 11 00465 g007
Table 1. Potential function parameters.
Table 1. Potential function parameters.
Interaction D i j   ( eV ) r 0   ( Å ) a i j   ( Å 2 )
Li + 0.6 O 1.2 0.0011142.6813603.429506
Ni + 1.2 O 1.2
Ni + 1.8 O 1.2
0.0293562.5007542.679137
O 1.2 O 1.2 0.0423953.3587011.659316
Mn + 2.4 O 1.2 0.0296582.4400003.012000
Co + 1.8 O 1.2 0.0109582.4006283.461272
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhu, Y.; Huang, Y.; Du, R.; Tang, M.; Wang, B.; Zhang, J. Effect of Ni2+ on Lithium-Ion Diffusion in Layered LiNi1−x−yMnxCoyO2 Materials. Crystals 2021, 11, 465. https://doi.org/10.3390/cryst11050465

AMA Style

Zhu Y, Huang Y, Du R, Tang M, Wang B, Zhang J. Effect of Ni2+ on Lithium-Ion Diffusion in Layered LiNi1−x−yMnxCoyO2 Materials. Crystals. 2021; 11(5):465. https://doi.org/10.3390/cryst11050465

Chicago/Turabian Style

Zhu, Yuanyuan, Yang Huang, Rong Du, Ming Tang, Baotian Wang, and Junrong Zhang. 2021. "Effect of Ni2+ on Lithium-Ion Diffusion in Layered LiNi1−x−yMnxCoyO2 Materials" Crystals 11, no. 5: 465. https://doi.org/10.3390/cryst11050465

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