
Cite this article as: | Jinpeng Qu, Yushen Zhao, Yurui Ji, Yanrong Zhu, and Tingfeng Yi, Approaching high-performance lithium storage materials by constructing Li2ZnTi3O8@LiAlO2 composites, Int. J. Miner. Metall. Mater., 30(2023), No. 4, pp.611-620. https://dx.doi.org/10.1007/s12613-022-2532-2 |
Li-ion batteries (LIBs) have been used quite extensively in electric vehicles (EVs) and portable electronic devices due to their excellent performance, such as high energy density, high operating voltage, and excellent cycling stability [1–5]. The increasing demand for high-performance LIBs stimulates the exploration and design of novel electrode materials [6–8]. At present, graphite-based materials are commonly used as the negative electrode materials of LIBs. Nevertheless, the low voltage plateau of graphite leads to the generation of lithium dendrites, which causes a series of security questions of LIBs [9–12]. Hence, it is urgent to develop more appropriate anode materials to replace graphite [13–16]. Titanium-based compounds become a kind of potential negative electrode material for LIBs due to their good safety and splendid thermal stability [16]. Among them, Li2ZnTi3O8 (LZTO) with cubic spinel structure is intensively studied due to the low cost, high theoretical capacity (227 mAh·g−1), and excellent circulation performance, which makes it a promising alternative material for graphite-based anodes of LIBs [17]. Nevertheless, the electronic structure of LZTO is characterized by the 3d state of titanium with a band gap energy of 2.0–3.0 eV, which makes the material have poor electrical conductivity [18]. The Li+ and Zn2+ ions are situated at tetrahedral sites, and affect ions migration to a certain extent [19], which lead to the poor conductivity and low ionic diffusion coefficient of LZTO. So far, the electrochemical performance of LZTO can be improved by the following three strategies [20]: grain size reduction [21], doping [22], and coating conductive material [23], such as carbon materials [24] or oxides [14]. Surface modification can separate the active materials from the electrolyte to avoid direct contact, which restrains the occurrence of side reaction, the surface phase transition, and the dissolution of electrode material [25]. Meanwhile, ion doping can change the valence states of both oxygen and transition metal (TM) via charge compensation mechanisms and change the distance between the Li and the TM layers so that the electronic conductivity and Li-ion conductivity of the host materials are improved [26]. Thus, the selection of suitable modified material is the key to the performance optimization.
Recently, plenty of materials have been reported as modified materials, such as Li2MoO4 [27], Li2ZrO3 [19], and LiAlO2 [28]. Among them, LiAlO2 effectively avoids contact between the electrode surface and organic electrolyte, thus decreasing electrolyte decomposition [29]. In addition, LiAlO2 is an admirable surface modifier with low cost, good stability, and electrical conductivity. According to many reports [1,30], LiAlO2 has been widely used as a coating material to enhance the electrochemical property of electrode materials. For example, Ding et al. [29] observed that the electrochemical performance of LiAlO2-coated LiNi0.8Co0.1Mn0.1O2 was obviously improved at room and elevated temperatures. Wu et al. [30] fabricated a novel LiAlO2 modified graphite material, and the composite showed outstanding cycle performance at 0.1 C. Hence, it can be expected that Li2ZnTi3O8@LiAlO2 (LZTAO) composites show an enhanced electrochemical performance. It has been reported that the LZTO can be synthesized by many methods, such as hydrothermal method [31], sol–gel method [25], solid-state method [23], and so on. However, compared with the solid-state method, the other synthesis routes are more complex and expensive. Based on practical applications, the simple solid-state synthesis of LZTO is more likely to be commercialized because of low synthesis cost and high yields. Herein, LZTAO composites were synthesized by a facile solid-state route. As expected, the LZTAO composites show an obviously improved electrochemical performance.
The pure LZTO was synthesized by a high-temperature solid-state method. The mixture of (CH3COO)2Zn·2H2O, TiO2, and Li2CO3 with stoichiometric ratio was ball-milled for 12 h. A slight excess of Li2CO3 was added to compensate for the loss of lithium at high temperatures. The mixture was dried at 60°C for 24 h in vacuum drying oven. Then, the obtained precursor was pre-sintered at 700°C for 1 h in air and ground again after cooling. Finally, the powder was calcined in a muffle furnace at 800°C for 3 h to get the final product LZTO.
The prepared LZTO, Li2CO3, and Al2O3 were dispersed into anhydrous ethanol at a molar ratio of Li : Al = 1:1 and ball-milled for 6 h in a planetary ball mill. The obtained powder was dried at 80°C for 12 h in a vacuum drying oven. Then, the powder was calcined at 650°C for 10 h in a muffle furnace. The contents of LiAlO2 were 3wt%, 5wt%, 8wt%, and 10wt%, which are named as LZTAO-1, LZTAO-2, LZTAO-3, and LZTAO-4, respectively.
The crystal structure of all samples was observed by X-ray diffraction (XRD, Rigaku, Japan) using Cu Kα1 radiation. X-ray photoelectron spectroscopy (XPS) measurements (PHI-5000 versaprobe II, Japan) were used to identify the valence state and surface species. The morphology and lattice structure of the samples were tested by a JSM-7800F Prime field emission scanning electron microscope (SEM, Japan) and high-resolution transmission electron microscope (HRTEM, JEOL, JEM-2010, Japan).
The homogeneous slurry containing active material, acetylene black, polyvinylidene fluoride (PVDF) (80:10:10 in weight), and N-methyl-pyrrolidone (NMP) was coated on a copper foil and desiccated at 105°C for 12 h under vacuum to form the working electrode. The half-cell consisted of working electrode and lithium foil, which was separated by a porous polypropylene film, and 1 M LiPF6 in ethylene carbonate (EC) : dimethyl carbonate (DMC) (1:1 in volume) as the electrolyte. All electrochemical evaluations were performed using the CR 2025-type coin cells. The cyclic voltammetry (CV) measurements were conducted on CHI 1000C electrochemical workstation (Chenhua, Shanghai, China) between 0 and 3 V vs. Li+/Li. The galvanostatic charge–discharge (GCD) tests were conducted on the LAND CT2001A (Lanbo, Wuhan, China) systems between 0 and 3.0 V. Electrochemical impedance spectroscopy (EIS) studies were carried out on a Princeton P4000 (USA) electrochemical workstation with a frequency range of 100 kHz to 0.01 Hz.
The XRD patterns of pristine LZTO and LZTAO composites are indicated in Fig. 1(b). The diffraction peaks of all the samples are well-indexed to cubic Li2ZnTi3O8 (PDF# 86-1512) with a P4332 space group [32]. The results indicate that the LiAlO2 does not alter the crystal structure of LZTO. The main peaks at 15.0°, 18.3°, 23.7°, 26.1°, 30.2°, 35.5°, 43.2°, 53.6°, 57.1°, and 62.7° are ascribed to the crystal planes of (110), (111), (210), (211), (220), (311), (400), (422), (511), and (440), respectively. There are no impurity peaks in the XRD patterns of all samples, which indicates a successful synthesis of LZTO and LZTAO composites. At the same time, the characteristic peaks LiAlO2 are not detected due to the low amount of LiAlO2 and small X-ray scattering factor of Al. Based on the model in Fig. 2(a), the Rietveld refinements for LZTO and LZTAO compounds are performed, and the results of all samples are given in Fig. 2(b)–(f). Based on the Rietveld refinement, the lattice parameters of LZTO and LZTAO (3wt%, 5wt%, 8wt%, and 10wt%) are 0.83843, 0.83762, 0.83737, 0.83733, and 0.83733 nm, respectively. The lattice shrinkage of LZTAO composites can be observed from the Rietveld refinement results because the ionic radius of Al3+ ion (0.0535 nm) is smaller than those of Ti4+ (0.0605 nm) and Zn2+ (0.074 nm) ions.
To identify the surface chemical valence states of various elements in LZTAO composites, the XPS spectrum of LZTAO-3 is shown in Fig. 3(a)–(e). In the survey spectra, the chemical elements of Li, Zn, Ti, Al, and O can be detected (Fig. 3(a)). In Fig. 3(b), the spectrum of Zn 2p contains two main peaks located at 1021.32 and 1044.36 eV, representing Zn 2p3/2 and Zn 2p1/2, respectively, which are attributed to Zn2+ ions [31]. Two doublet peaks located at 458.63 and 464.33 eV can be found in Ti 2p spectra, representing Ti 2p3/2 and Ti 2p1/2, which proves the existence of Ti4+ state [33]. The spectrum of O 1s is suitably into two major peaks at 530.03 and 531.41 eV, due to the lattice oxygen of A–O (A = Ti, Zn, or Al) bonds and adsorptive or defective oxygen species, respectively [5,34].
As shown in Fig. 4(a)–(e), the morphology of the LZTO and LZTAO composites are examined by SEM. All samples show similar morphology with a particle size of 300–500 nm, proving that the LiAlO2 does not affect the original morphology and particle size. The transmission electron microscopy (TEM) and corresponding energy dispersive spectroscopy (EDS) mapping images of LZTAO-3 indicate the existence of Zn, Ti, Al, and O with well-proportioned distribution (Fig. 4(f)–(j)), which also confirms the successful synthesis of LZTAO composites.
The TEM and HRTEM tests were recorded to further analyze the microstructure of LZTAO. The particle size of LZTAO-3 is about 300 nm (Fig. 5(a) and (b)), and the interplanar spacings of 0.405 and 0.5934 nm represent the (101) plane of LiAlO2 and the (110) plane of LZTO (Fig. 5(c)), respectively [6]. The thicknesses of LiAlO2 layers in the LZTAO-3 composite is about 5–10 nm (Fig. 5(c)). The selected area electron diffraction (SAED) patterns of LZTAO-3 are given in Fig. 5(d). The alternant dark and white rings around LZTAO crystals represent the (210), (221), and (110) planes of LZTO crystal [35], respectively. Based on these results, it can be confirmed that LiAlO2 is successfully covered around the surfaces of LZTO particles, which offers a fine conductive connection between the LZTO particles and LiAlO2 layer.
The first GCD curves of pristine LZTO and LZTAO composites in the potential window of 0–3.0 V vs. Li/Li+ at 0.5 C are given in Fig. 6(a). All discharge curves show two distinct plateaus at 1.5–1.7 V and 0.4–0.6 V. The reason for this phenomenon is that the tetrahedral sites of LZTO, [Li0.5Zn0.5]tet[Ti1.5Li0.5]octO4 (tet and oct represent tetrahedral and octahedral sites, respectively), can be taken up by Li+ at low voltage and the octahedral empty sites can be taken up by Li+ at the high operating potential [36]. The electrochemical mechanism of LZTO is described as [28]:
[Li0.5Zn0.5]tet[Ti1.5Li0.5]octO4+xe−+xLi+→[Li0.5Zn0.5]tet[Ti1.5Li0.5+x]octO4 |
(1) |
[Li0.5Zn0.5]tet[Ti1.5Li0.5+x]octO4+ye−+yLi+→[Li0.5+yZn0.5]tet[Ti1.5Li0.5+x]octO4 |
(2) |
The initial charge (discharge) capacities of LZTO, LZTAO-1, LZTAO-2, LZTAO-3, and LZTAO-4 are 124.5 (208.8), 156.3 (273.2), 149.9 (359.3), 187.6 (325.5), and 164.3 (322.1) mAh·g−1 at 0.5 C, respectively. Obviously, all LZTAO composites exhibit higher capacity than pure LZTO. This indicates that an appropriate LiAlO2 modification increases the initial charge and discharge capacity of LZTO, but only the charge capacity of the LZTAO-3 sample is significantly improved. The corresponding initial Coulombic efficiencies (CEs) are 59.9%, 57.2%, 41.7%, 57.6%, and 51.0%, respectively. The initial CE of the LZTAO composites is lower than that of pristine LZTO, mainly because there is a layer of LiAlO2 on the surface of LZTO. The low initial CE of all samples may be due to the formation of solid electrolyte interface (SEI) and irreversible electrolyte decomposition when discharged to such low potential. Good cycling performance and rate performance are important criteria to evaluate the electrochemical performance of anode material. Fig. 6(b) and (c) depicts the rate performance of all samples at continuously increased current densities from 0.5 to 5 C. It can be seen that the rate performance of the LZTAO samples is superior to the pristine sample. Especially, the reversible discharge (charge) capacities of LZTAO-3 composite at 0.5, 1, 2, 3, and 5 C are about 212.6 (203.9), 199.9 (194.8), 190.4 (187.4), 183 (180.6), and 175.6 (177.1) mAh·g−1, respectively. Moreover, the LZTAO-3 sample shows the highest charge/discharge capacities (219.9/224.5 mAh·g−1) when the current density returns to 0.5 C. However, the discharge (charge) capacities of pristine LZTO are only 148.8 (134.5), 112.6 (109.7), 90.4 (89.4), 80.3 (79.9), and 73.5 (72.9) mAh·g−1 at the corresponding rates, which is much smaller than those of LZTAO-3 sample. That means the LiAlO2 layer contributes to the ion and electron transfer and effectively enhances the charge and discharge capacity. Fig. 6(d) and (e) exhibits the cycling properties of all samples at a current density of 5 C between 0 and 3.0 V. All samples are activated at a low charge–discharge rate for 20 cycles before testing their cycling performance. The loss of capacity in the first few cycles may be caused by the generation of SEI film on the surface of electrode material, which is produced by the irreversible reaction between electrode material and electrolyte at the solid–liquid interface [37]. After 150 cycles, the charge (discharge) capacities of LZTO, LZTAO-1, LZTAO-2, LZTAO-3, and LZTAO-4 are 195.9 (198.2), 223.8 (225.6), 236.3 (232.7), 263.5 (265.8), and 269.2 (270.9) mAh·g−1, respectively. Compared with the pure sample, it is obvious that the specific capacity of the LZTAO-3 composite is more excellent. Moreover, the trend of all curves remains rising during cycling, which is mainly caused by the continuous activation process, such as the optimization of the lithiation-induced structure, reconstruction of steady SEI layer, the reduced grain size of electrolytic material (Fig. 7(a) and (b)), and the formation of organic polymeric/gelatinous layer attributed to the electrolyte decomposition at low voltage [16,38]. The XRD patterns the LZTO and LZTAO-3 anodes discharged to 0 V after 150 cycles are shown in Fig. 7(c), and an overlapping between LZTO (311), (421), and (533) and Cu (111), (200), and (220) peaks can be found. No additional phases generate upon the electrochemical Li storage reactions, revealing that the crystal structures of LZTO and LZTAO-3 are not destroyed during cycling. These results reveal the good structure stability of LZTO and LZTAO-3 during cycling. Compared with other reports of LZTO-based materials as shown in Fig. 7(d), the electrochemical performance of LZTAO-3 is superior to some previously reported works, such as Li2ZnTi3O8@Na2WO4 (NWLZTO) [6], LZTO prepared by solid state method [21], Li2ZnTi2.9Mo0.1O8 [23], Li2MoO4 modified Li2ZnTi3O8 (LMO-LZT) [28], Li2ZnTi3O8-carbon microspheres (LZTO-M) [39], and nitrogen-doped carbon on Li2ZnTi3O8/TiO2 (N–C@LZTO/TO) [40].
The CV profiles of LZTO and LZTAO composites within the voltage window of 0–3.0 V are depicted in Fig. 8. There is a pair of redox peaks between 1 and 2 V, which is attributed to the redox of Ti4+/Ti3+ couples [41]. In addition, the extra peaks between 0.2 and 0.5 V are relate to the multiple conversions of the Ti4+ ions for all samples, certifying that the reaction of LZTO is a multi-step continuous conversion process [39,41–42]. The peak positions of pure samples and LZTAO compounds are similar, indicating that the electrochemical reaction of LZTO is not changed by the LiAlO2 layer. Compared to the pristine LZTO, the LZTAO compounds reveal higher redox peaks at around 1.5 V and smaller potential differences between reduction and oxidation peaks. That means the LZTAO materials have less polarization behavior and superior dynamic performance, which can illustrate that LiAlO2 enhances the reversibility of Li ions insertion and extraction. In the 2nd and 3th cycles, all redox peaks at ~1.3/1.6 V are attributed to the redox of Ti4+/Ti3+ couples, revealing that the insertion/extraction of Li ions within the electrodes is highly reversible. In addition, the reduction peak at about 0.51 V moves to 0.57 V and oxidation peak at about 1.1 V shifts to 1.25 V after the first cycle, which is related to the multiple restoration of Ti4+ because of order/disorder transitions [31,42].
To further investigate the lithium-ion diffusion mechanism of the pristine LZTO and LZTAO composites, EIS test was carried out (Fig. 9). The Nyquist plots are given in Fig. 9(a) and the equivalent circuit is given in Fig. 9(d). All curves are constituted by a semicircle in the high-frequency area and an oblique line in the low-frequency area. The high frequency intercept of the semicircle always corresponds to the charge transfer resistance (Rct), reflecting the electrochemical reaction activity at the electrode interface [43]. The oblique line corresponds to the Warburg impedance (W), which is related to the diffusion process of Li+ [44]. Besides, the intercept of the real axis represents the internal resistance of the solution (Rs) [45]. Obviously, the Rct of LZTAO samples is lower than the pristine LZTO (Fig. 9(c)), revealing that the introduction of LiAlO2 can reduce the charge transfer resistance, thus facilitating the electron transfer and improving the electronic conductivity. In addition, LZTAO composites also show much smaller solution resistance than pure LZTO, revealing that LiAlO2 modification can efficiently reduce the resistance between anodes and electrolytes. In particular, the solution resistance of the LZTAO-3 electrode material is the smallest among all samples, meaning the best electrochemical activity.
The Li-ion diffusion coefficients (
Z′=Rs+Rct+σω−1/2 |
(3) |
DLi+=R2T22A2n4F4C2Li+σ2 |
(4) |
where ω is the angular frequency; R is the gas constant; T is the absolute temperature; A is the surface area; n is the number of electrons involved in the redox reaction; F is the Faraday constant;
According to HRTEM image (Fig. 5(c)), it is reasonable to infer that an interface between the LZTO and LiAlO2 can exist. The interface model between LZTO and LiAlO2 is plotted in Fig. 10. From the interface model and HRTEM image, it can be observed that the LiAlO2 can be generated in situ on the surface of LZTO, and then form many LZTO–LiAlO2 stable interfaces. These interfaces strengthen the bonding of the composite and improve the structural stability of the material. Moreover, the formed phase interface can provide more sites for electrolyte storage, thus accelerating the rate of lithiation and delithiation. The suitable amount of LiAlO2 modification can not only reduces the transfer resistances and electrochemical polarization of LZTO, but also improves the diffusion coefficient of lithium-ion. Therefore, in situ modification of LiAlO2 is an available strategy to enhance the electrochemical property of LZTO.
In summary, pristine Li2ZnTi3O8 and Li2ZnTi3O8@LiAlO2 composites were successfully fabricated using a facile solid-state method. All samples exhibit similar morphology with a particle size of 300–500 nm, proving that the LiAlO2 does not affect the original morphology and particle size. LiAlO2 modification can stabilize the structure, enhance ionic conductivity, reduce interfacial charge transfer impedance, and improve the transfer rate of Li ions of Li2ZnTi3O8, leading to a good rate performance and cycling stability. Among all composites, LZTAO-3 possesses the highest capacities at each rate and the best cycling stability. The LZATO-3 can provide charge capacities of 203.9, 194.8, 187.4, 180.6, and 177.1 mAh·g−1 at 0.5, 1, 2, 3, and 5 C, respectively. Nevertheless, pure LZTO only shows charge capacities of 134.5, 109.7, 89.4, 79.9, and 72.9 mAh·g−1 at the corresponding rates. Consequently, Li2ZnTi3O8@LiAlO2 (8wt%) composite as anode material shows a potential application prospect for LIBs with high performance.
This work was supported by the National Natural Science Foundation of China (No. U1960107), the “333” Talent Project of Hebei Province, China (No. A202005018), the Fundamental Research Funds for the Central Universities (No. N2123001), and the Performance Subsidy Fund for Key Laboratory of Dielectric and Electrolyte Functional Material Hebei Province, China (No. 22567627H).
The authors declare no conflict of interests.
[1] |
Y.Q. Cao, X.B. Meng, and A.D. Li, Enhanced electrochemical performances, Energy Environ. Mater., 4(2021), No. 3, p. 363. DOI: 10.1002/eem2.12132
|
[2] |
S.Q. Zhao, Y.J. He, Z.W. Wang, et al., Advancing performance and unfolding mechanism of lithium and sodium storage in SnO2 via precision synthesis of monodisperse PEG-ligated nanoparticles, Adv. Energy Mater., 12(2022), No. 26, art. No. 2201015. DOI: 10.1002/aenm.202201015
|
[3] |
Z.Y. Feng, W.J. Peng, Z.X. Wang, et al., Review of silicon-based alloys for lithium-ion battery anodes, Int. J. Miner. Metall. Mater., 28(2021), No. 10, p. 1549. DOI: 10.1007/s12613-021-2335-x
|
[4] |
A.M. Huang, Y.C. Ma, J. Peng, et al., Tailoring the structure of silicon-based materials for lithium-ion batteries via electrospinning technology, eScience, 1(2021), No. 2, p. 141. DOI: 10.1016/j.esci.2021.11.006
|
[5] |
M.R. Wu, M.Y. Gao, S.Y. Zhang, et al., High-performance lithium-sulfur battery based on porous N-rich g-C3N4 nanotubes via a self-template method, Int. J. Miner. Metall. Mater., 28(2021), No. 10, p. 1656. DOI: 10.1007/s12613-021-2319-x
|
[6] |
L.Y. Qiu, X.Q. Lai, F.F. Wang, et al., Promoting the Li storage performances of Li2ZnTi3O8@Na2WO4 composite anode for Li-ion battery, Ceram. Int., 47(2021), No. 14, p. 19455. DOI: 10.1016/j.ceramint.2021.03.282
|
[7] |
H.M. Qian and X.F. Li, Progress in functional solid electrolyte interphases for boosting Li metal anode, Acta Phys. Chim. Sin., 37(2021), No. 2, art. No. 2008092.
|
[8] |
C. Sun, X. Li, X.Z. Wu, et al., Improved the lithium storage capability of Na2Li2Ti6O14 by barium doping, J. Electroanal. Chem., 802(2017), p. 100. DOI: 10.1016/j.jelechem.2017.09.007
|
[9] |
Y.K. Ye, Z.X. Hu, J.H. Liu, et al., Research progress of theoretical studies on polarons in cathode materials of lithium-ion batteries, Acta Phys. Chim. Sin., 37(2021), No. 11, art. No. 2011003.
|
[10] |
H. Chang, Y.R. Wu, X. Han, and T.F. Yi, Recent developments in advanced anode materials for lithium-ion batteries, Energy Mater., 1(2021), art. No. 100003.
|
[11] |
Y. Xiao, R. Xu, L. Xu, J.F. Ding, and J.Q. Huang, Recent advances in anion-derived SEIs for fast-charging and stable lithium batteries, Energy Mater., 1(2021), art. No. 100013.
|
[12] |
X. Qiao, X.B. Yang, N. Zhang, et al., One-step in situ encapsulation of Ge nanoparticles into porous carbon network with enhanced electron/ion conductivity for lithium storage, Rare Met., 40(2021), No. 9, p. 2432. DOI: 10.1007/s12598-020-01662-4
|
[13] |
S.Q. Zhao, C.D. Sewell, R.P. Liu, et al., SnO2 as advanced anode of alkali-ion batteries: Inhibiting Sn coarsening by crafting robust physical barriers, void boundaries, and heterophase interfaces for superior electrochemical reaction reversibility, Adv. Energy Mater., 10(2020), No. 6, art. No. 1902657. DOI: 10.1002/aenm.201902657
|
[14] |
Y. Li, T.F. Yi, X.Z. Li, et al., Li2ZnTi3O8@α-Fe2O3 composite anode material for Li-ion batteries, Ceram. Int., 47(2021), No. 13, p. 18732. DOI: 10.1016/j.ceramint.2021.03.208
|
[15] |
J.B. Ye, T. Chen, Q.N. Chen, W.X. Chen, Z.T. Yu, and S.R. Xu, Facile hydrothermal synthesis of SnCoS4/graphene composites with excellent electrochemical performance for reversible lithium ion storage, J. Mater. Chem. A, 4(2016), No. 34, p. 13194. DOI: 10.1039/C6TA04196E
|
[16] |
L.F. Wang, M.M. Geng, X.N. Ding, et al., Research progress of the electrochemical impedance technique applied to the high-capacity lithium-ion battery, Int. J. Miner. Metall. Mater., 28(2021), No. 4, p. 538. DOI: 10.1007/s12613-020-2218-6
|
[17] |
H. Yang, H.L. Zhu, Y.X. Qi, N. Lun, and Y.J. Bai, Optimizing the cycling life and high-rate performance of Li2ZnTi3O8 by forming thin uniform carbon coating derived from citric acid, J. Mater. Sci., 55(2020), No. 32, p. 15538. DOI: 10.1007/s10853-020-04980-1
|
[18] |
C. Chen, C.C. Ai, and X.Y. Liu, Ti(Ⅲ) self-doped Li2ZnTi3O8 as a superior anode material for Li-ion batteries, Electrochim. Acta, 265(2018), p. 448. DOI: 10.1016/j.electacta.2018.01.159
|
[19] |
H. Yang, J. Park, C.S. Kim, et al., Boosted electrochemical performance of Li2ZnTi3O8 enabled by ion-conductive Li2ZrO3 concomitant with superficial Zr-doping, J. Power Sources, 379(2018), p. 270. DOI: 10.1016/j.jpowsour.2018.01.064
|
[20] |
A.I. Inamdar, A.T.A. Ahmed, H.S. Chavan, et al., Influence of operating temperature on Li2ZnTi3O8 anode performance and high-rate charging activity of Li-ion battery, Ceram. Int., 44(2018), No. 15, p. 18625. DOI: 10.1016/j.ceramint.2018.07.087
|
[21] |
L. Wang, L.J. Wu, Z.H. Li, G.T. Lei, Q.Z. Xiao, and P. Zhang, Synthesis and electrochemical properties of Li2ZnTi3O8 fibers as an anode material for lithium-ion batteries, Electrochim. Acta, 56(2011), No. 15, p. 5343. DOI: 10.1016/j.electacta.2011.03.122
|
[22] |
S. Qi, J. Pan, L.N. Shi, Y.R. Zhu, T.F. Yi, and Y. Xie, Achieving high-performance Li2ZnTi3O8 anode for advanced Li-ion batteries by molybdenum doping, Mater. Today Chem., 21(2021), art. No. 100523. DOI: 10.1016/j.mtchem.2021.100523
|
[23] |
S. Wang, L.J. Wang, Z.H. Meng, and R. Xun, Design of a three-dimensional-network Li2ZnTi3O8 co-modified with graphene nanosheets and carbon nanotubes as a high performance anode material for lithium-ion batteries, J. Alloys Compd., 774(2019), p. 581. DOI: 10.1016/j.jallcom.2018.10.035
|
[24] |
Y.X. Xu, Z.S. Hong, L.C. Xia, J. Yang, and M.D. Wei, One step sol–gel synthesis of Li2ZnTi3O8/C nanocomposite with enhanced lithium-ion storage properties, Electrochim. Acta, 88(2013), p. 74. DOI: 10.1016/j.electacta.2012.10.044
|
[25] |
H.Q. Tang, Y.K. Zhou, L.X. Zan, N.Q. Zhao, and Z.Y. Tang, Long cycle life of carbon coated lithium zinc titanate using copper as conductive additive for lithium ion batteries, Electrochim. Acta, 191(2016), p. 887. DOI: 10.1016/j.electacta.2016.01.141
|
[26] |
C.Y. Xu, J.L. Li, J. Sun, W.Z. Zhang, and B.M. Ji, Li-rich layered oxide single crystal with Na doping as a high-performance cathode for Li ion batteries, J. Alloys Compd., 895(2022), art. No. 162613. DOI: 10.1016/j.jallcom.2021.162613
|
[27] |
Z.F. Li, H. Li, Y.H. Cui, et al., Li2MoO4 modified Li2ZnTi3O8 as a high property anode material for lithium ion battery, J. Alloys Compd., 692(2017), p. 131. DOI: 10.1016/j.jallcom.2016.09.042
|
[28] |
Z.K. Fang, Y.R. Zhu, T.F. Yi, and Y. Xie, Li4Ti5O12–LiAlO2 composite as high performance anode material for lithium-ion battery, ACS Sustainable Chem. Eng., 4(2016), No. 4, p. 1994. DOI: 10.1021/acssuschemeng.5b01271
|
[29] |
G.Y. Ding, F.Q. Yan, Z. Zhu, et al., Mussel-inspired polydopamine-assisted uniform coating of Li+ conductive LiAlO2 on nickel-rich LiNi0.8Co0.1Mn0.1O2 for high-performance Li-ion batteries, Ceram. Int., 48(2022), No. 4, p. 5714. DOI: 10.1016/j.ceramint.2021.11.118
|
[30] |
Y. Wu, Y.F. Li, L.Y. Wang, et al., Enhancing the Li-ion storage performance of graphite anode material modified by LiAlO2, Electrochim. Acta, 235(2017), p. 463. DOI: 10.1016/j.electacta.2017.03.129
|
[31] |
H. Yang, N. Lun, Y.X. Qi, et al., Li2ZnTi3O8 coated with uniform lithium magnesium silicate layer revealing enhanced rate capability as anode material for Li-ion battery, Electrochim. Acta, 315(2019), p. 24. DOI: 10.1016/j.electacta.2019.05.087
|
[32] |
T.B. Lan, L. Chen, Y.B. Liu, W.F. Zhang, and M.D. Wei, Nanocomposite Li2ZnTi3O8/C with enhanced electrochemical performances for lithium-ion batteries, J. Electroanal. Chem., 794(2017), p. 120. DOI: 10.1016/j.jelechem.2017.04.013
|
[33] |
H.Q. Tang, L.X. Zan, and Z.Y. Tang, Predominant electronic conductivity of Li2ZnTi3O8 anode material prepared in nitrogen for rechargeable lithium-ion batteries, J. Electroanal. Chem., 823(2018), p. 269. DOI: 10.1016/j.jelechem.2018.06.025
|
[34] |
S.Q. Zhao, Z.W. Wang, Y.J. He, et al., A robust route to Co2(OH)2CO3 ultrathin nanosheets with superior lithium storage capability templated by aspartic acid-functionalized graphene oxide, Adv. Energy Mater., 9(2019), No. 26, art. No. 1901093.
|
[35] |
Q.Y. Li, G.C. Yang, Y.Q. Chu, et al., Enhanced electrochemical performance of Ni-rich cathode material by N-doped LiAlO2 surface modification for lithium-ion batteries, Electrochim. Acta, 372(2021), art. No. 137882. DOI: 10.1016/j.electacta.2021.137882
|
[36] |
C.Y. An, C.H. Li, H.Q. Tang, T. Liu, and Z.Y. Tang, Binder-free flexible Li2ZnTi3O8@MWCNTs stereoscopic network as lightweight and superior rate performance anode for lithium-ion batteries, J. Alloys Compd., 816(2020), art. No. 152580. DOI: 10.1016/j.jallcom.2019.152580
|
[37] |
P.P. Peng, Y.R. Wu, X.Z. Li, et al., Toward superior lithium/sodium storage performance: Design and construction of novel TiO2-based anode materials, Rare Met., 40(2021), No. 11, p. 3049. DOI: 10.1007/s12598-021-01742-z
|
[38] |
C.X. Hou, J. Wang, W. Du, et al., One-pot synthesized molybdenum dioxide–molybdenum carbide heterostructures coupled with 3D holey carbon nanosheets for highly efficient and ultrastable cycling lithium-ion storage, J. Mater. Chem. A, 7(2019), No. 22, p. 13460. DOI: 10.1039/C9TA03551F
|
[39] |
T. Liu, H.Q. Tang, J.Y. Liu, et al., Improved electrochemical performance of Li2ZnTi3O8 using carbon materials as loose and porous agent, Electrochim. Acta, 259(2018), p. 28. DOI: 10.1016/j.electacta.2017.10.139
|
[40] |
G.S. Sim, P. Santhoshkumar, J.W. Park, et al., Chitosan-derived nitrogen-doped carbon on Li2ZnTi3O8/TiO2 composite as an anode material for lithium-ion batteries, Ceram. Int., 47(2021), No. 23, p. 33554. DOI: 10.1016/j.ceramint.2021.08.264
|
[41] |
S. Wang, Y.F. Bi, L.J. Wang, Z.H. Meng, and B.M. Luo, Mo-doped Li2ZnTi3O8@graphene as a high performance anode material for lithium-ion batteries, Electrochim. Acta, 301(2019), p. 319. DOI: 10.1016/j.electacta.2019.01.168
|
[42] |
Z.S. Hong, M.D. Wei, X.K. Ding, L.L. Jiang, and K.M. Wei, Li2ZnTi3O8 nanorods: A new anode material for lithium-ion battery, Electrochem. Commun., 12(2010), No. 6, p. 720. DOI: 10.1016/j.elecom.2010.03.016
|
[43] |
H.S. Ren, H.Y. Peng, T.Y. Xie, et al., Temperature stable microwave dielectric ceramics in Li2ZnTi3O8-based composite for LTCC applications, J. Mater. Sci. Mater. Electron., 29(2018), No. 15, p. 12978. DOI: 10.1007/s10854-018-9418-0
|
[44] |
T.T. Liu, N. Peng, X.K. Zhang, et al., Controllable defect engineering enhanced bond strength for stable electrochemical energy storage, Nano Energy, 79(2021), art. No. 105460. DOI: 10.1016/j.nanoen.2020.105460
|
[45] |
P. Fu, Z.Y. Li, Y. Pan, et al., Synthesis and characterization of Sm-doped Li2ZnTi3O8 as anode material for lithium-ion batteries, Mater. Chem. Phys., 277(2022), art. No. 125449. DOI: 10.1016/j.matchemphys.2021.125449
|
[46] |
T.F. Yi, J.Z. Wu, J. Yuan, Y.R. Zhu, and P.F. Wang, Rapid lithiation and delithiation property of V-doped Li2ZnTi3O8 as anode material for lithium-ion battery, ACS Sustainable Chem. Eng., 3(2015), No. 12, p. 3062. DOI: 10.1021/acssuschemeng.5b00505
|
[47] |
L.Y. Qiu, Y.R. Ji, Z.C. Lv, et al., Enhanced lithium storage property of porous Na2Li2Ti6O14@PEDOT spheres as anodes for lithium-ion batteries, Mater. Chem. Phys., 278(2022), art. No. 125700. DOI: 10.1016/j.matchemphys.2022.125700
|
[48] |
H. Chang, Y. Li, Z.K. Fang, J.P. Qu, Y.R. Zhu, and T.F. Yi, Construction of carbon-coated LiMn0.5Fe0.5PO4@Li0.33La0.56TiO3 nanorod composites for high-performance Li-ion batteries, ACS Appl. Mater. Interfaces, 13(2021), No. 28, p. 33102. DOI: 10.1021/acsami.1c08373
|
[1] | Zhen He, Jiaming Liu, Yuqian Wei, Yunfei Song, Wuxin Yang, Aobo Yang, Yuxin Wang, Bo Li. Polypyrrole-coated triple-layer yolk–shell Fe2O3 anode materials with their superior overall performance in lithium-ion batteries [J]. International Journal of Minerals, Metallurgy and Materials, 2024, 31(12): 2737-2748. DOI: 10.1007/s12613-024-2954-0 |
[2] | Kangzhe Cao, Sitian Wang, Yanan He, Jiahui Ma, Ziwei Yue, Huiqiao Liu. Constructing Al@C–Sn pellet anode without passivation layer for lithium-ion battery [J]. International Journal of Minerals, Metallurgy and Materials, 2024, 31(3): 552-561. DOI: 10.1007/s12613-023-2720-8 |
[3] | Kai-lin Cheng, Dao-bin Mu, Bo-rong Wu, Lei Wang, Ying Jiang, Rui Wang. Electrochemical performance of a nickel-rich LiNi0.6Co0.2Mn0.2O2 cathode material for lithium-ion batteries under different cut-off voltages [J]. International Journal of Minerals, Metallurgy and Materials, 2017, 24(3): 342-351. DOI: 10.1007/s12613-017-1413-6 |
[4] | Fei-biao Chen, Bo-rong Wu, Yun-kui Xiong, Wei-lin Liao, Dao-Bin Mu, Feng Wu. Modified disordered carbon prepared from 3,4,9,10-perylenetetracarboxylic dianhydride as an anode material for Li-ion batteries [J]. International Journal of Minerals, Metallurgy and Materials, 2015, 22(2): 203-209. DOI: 10.1007/s12613-015-1062-6 |
[5] | Jin Yi, Yan-lin Liu, Yuan Wang, Xiao-ping Li, She-jun Hu, Wei-shan Li. Synthesis of dandelion-like TiO2 microspheres as anode materials for lithium ion batteries with enhanced rate capacity and cyclic performances [J]. International Journal of Minerals, Metallurgy and Materials, 2012, 19(11): 1058-1062. DOI: 10.1007/s12613-012-0670-7 |
[6] | Tao Li, Juan-yu Yang, Shi-gang Lu. Effect of modified elastomeric binders on the electrochemical properties of silicon anodes for lithium-ion batteries [J]. International Journal of Minerals, Metallurgy and Materials, 2012, 19(8): 752-756. DOI: 10.1007/s12613-012-0623-1 |
[7] | Xueping Zheng, Ping Li, Xuanhui Qu, I. S. Humail, Yanghuan Zhang, Guoqing Wang. Effect of different additives on the properties of lithium alanate [J]. International Journal of Minerals, Metallurgy and Materials, 2008, 15(6): 786-790. DOI: 10.1016/S1005-8850(08)60288-1 |
[8] | Chaoli Yin, Hailei Zhao, Hong Guo, Xianliang Huang, Weihua Qiu. Effect of the synthesis method of SnSb anode materials on their electrochemical properties [J]. International Journal of Minerals, Metallurgy and Materials, 2007, 14(4): 345-349. DOI: 10.1016/S1005-8850(07)60068-1 |
[9] | Chuangang Lin, Weihua Qiu, Zhimin Li, Jialing Hong, Huanyu Hu, Zhongqin Lu. Comparative study of LiNiO2, LiNi0.8Co0.2O2 and LiNi0.75Al0.25O2 for lithium-ion batteries [J]. International Journal of Minerals, Metallurgy and Materials, 2002, 9(6): 448-452. |
[10] | LIN Qin, CHEN Ning, YE Wen, LIU Renmin. Kinetics of Hydrogen Absorption in Hydrogen Storage Alloy [J]. International Journal of Minerals, Metallurgy and Materials, 1997, 4(2): 34-37. |
1. | Fei-Long Li, Meng-Cheng Han, Guo-Chen Bian, et al. Porous Zn2Ti3O8@C rods via in situ carbon coating for superior lithium storage capability. Ionics, 2025. DOI:10.1007/s11581-024-06054-7 |
2. | Guodong Wen, Shuai Yuan, Zaizheng Dong, et al. Recycling of spent lithium iron phosphate battery cathode materials: A review. Journal of Cleaner Production, 2024, 474: 143625. DOI:10.1016/j.jclepro.2024.143625 |
3. | Daixiang Wei, Wei Wang, Longjin Jiang, et al. Preferentially selective extraction of lithium from spent LiCoO2 cathodes by medium-temperature carbon reduction roasting. International Journal of Minerals, Metallurgy and Materials, 2024, 31(2): 315. DOI:10.1007/s12613-023-2698-2 |
4. | Huanhuan Liu, Xue Zhang, Haoran Xu, et al. Li2ZnTi3O8anode: design from material to electrode and devices. Inorganic Chemistry Frontiers, 2023, 10(17): 4943. DOI:10.1039/D3QI00934C |