Next Article in Journal
Influence of Reduction Temperature on the Structure and Naphthalene Hydrogenation Saturation Performance of Ni2P/Al2O3 Catalysts
Next Article in Special Issue
Synthesis of Needle-like Nanostructure Composite Electrode of Co3O4/rGO/NF for High-Performance Symmetric Supercapacitor
Previous Article in Journal
High-Efficiency Plasmonic Lens Based on Archimedes-Spiral with Cross Section of an Asymmetric Slot
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Spinel LiMn2O4 Cathode Materials in Wide Voltage Window: Single-Crystalline versus Polycrystalline

1
School of Chemistry and Materials Science, Nanjing University of Information Science and Technology, Nanjing 210044, China
2
Center for Advancing Electronics Dresden (cfaed) & Faculty of Chemistry and Food Chemistry, Technische Universität Dresden, Mommsenstrasse 4, 01069 Dresden, Germany
3
State Key Laboratory of Materials Processing and Die & Mould Technology, School of Materials Science and Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
*
Authors to whom correspondence should be addressed.
Submission received: 27 December 2021 / Revised: 19 February 2022 / Accepted: 23 February 2022 / Published: 24 February 2022
(This article belongs to the Special Issue Supercapacitors with High Energy Density)

Abstract

:
Single-crystal (SC) layered oxides as cathodes for Li-ion batteries have demonstrated better cycle stability than their polycrystalline (PC) counterparts due to the restrained intergranular cracking formation. However, there are rare reports on comparisons between single-crystal LiMn2O4 (SC-LMO) and polycrystalline LiMn2O4 (PC-LMO) spinel cathodes for Li-ion storage. In this work, the Li-ion storage properties of spinel LiMn2O4 single-crystalline and polycrystalline with similar particle sizes were investigated in a wide voltage window of 2–4.8 V vs. Li/Li+. The SC-LMO cathode exhibited a specific discharge capacity of 178 mA·h·g−1, which was a bit larger than that of the PC-LMO cathode. This is mainly because the SC-LMO cathode showed much higher specific capacity in the 3 V region (Li-ion storage at octahedral sites with cubic to tetragonal phase transition) than the PC-LMO cathode. However, unlike layered-oxide cathodes, the PC-LMO cathode displayed better cycle stability than the SC-LMO cathode. Our studies for the first time demonstrate that the phase transition-induced Mn(II) ion dissolution in the 3 V region rather than cracking formation is the limiting factor for the cycle performance of spinel LiMn2O4 in the wide voltage window.

1. Introduction

Lithium-ion batteries (LIBs) have been the most successful power sources for portable electronics over the last four decades. Recently, LIBs are also emerging as promising energy storage device for electric vehicles (EVs). Nevertheless, the requirement of higher specific energy and improved safety for automotive LIBs is far from over due to the range and safety concerns of EVs [1,2,3]. A natural pathway to improve the energy densities of automotive LIBs is to increase the reversible capacities of the cathode materials. Compared to currently prevailing Co- and Ni-based layered-oxide cathode materials (such as LiCoO2 and Ni-rich LiNi1−x−yCoxMnyO2), the Mn-based spinel LiMn2O4 cathode possesses unique characteristics of low toxicity, abundant source, superior safety, and environmental friendliness [4,5]. The 4 V electrochemistry of the LiMn2O4 cathode has been popularly explored in LIBs [4,5,6,7,8,9,10,11,12]. Unfortunately, practical specific capacities of spinel LiMn2O4 cathodes in the 4 V region are limited to ~80–120 mA·h·g−1, which are lower than layered-oxide cathode materials. The low specific capacity of the spinel LiMn2O4 cathode originates from the lithiation/delithiation in spinel frameworks in the 4 V region occurring at Li-ion occupied tetrahedral (8a) sites only involving one-electron redox reaction. Interestingly, extending the discharge voltage of spinel LiMn2O4 cathode to 2 V vs. Li/Li+ can double its specific capacity. There is another voltage plateau at around 2.9 V vs. Li/Li+ (labeled as 3 V region) with extra Li-ion intercalation into the unoccupied octahedral sites (16c) sites of the spinel structure. However, upon discharging to the 3 V region, severe phase transition from cubic phase to tetragonal phase may appear because the reduction in average Mn valence state close to trivalence easily induces serious Jahn–Teller distortion and Mn(III) disproportionation reactions [13]. Both Jahn–Teller distortion and Mn(III) disproportionation reactions would lead to the severe capacity fading of spinel LiMn2O4 cathodes. Thus, most previous efforts in spinel LiMn2O4 cathodes were mainly focused on electrochemical studies in the 4 V region [4,5,6,7,8,9,10,11,12].
To suppress the capacity fading of cathodes, one effective strategy is to design a single-crystal morphology. Currently, single-crystal layered-oxide cathodes have demonstrated better structural integrity than their polycrystalline counterparts due to the reduced phase boundaries and limited surface reactivity [14,15,16,17,18,19,20,21,22,23]. Typically, the polycrystalline cathodes are composed of smaller primary particles agglomerated into large secondary particles. During cycles, the expansion/contraction of the small primary particles in nonuniform directions usually leads to cracking propagation in large secondary particles. On the one hand, the cracking propagation causes a severe loss of electrical contact between the active cathode material and the current collector. On the other hand, the formed cracking exacerbates the parasitic reactions between the electrolyte and freshly exposed particle surfaces [16]. Thus, cracking formation is regarded as one of the main reasons for the capacity loss and structural collapse in layered-oxide cathodes. It has been revealed that single-crystal morphology can avoid the formation of particle cracking because of the uniform lattice expansion/contraction and fewer phase boundaries. For example, Dahn et al. [22] demonstrated that LiNi0.5Mn0.3Co0.5O2 cathodes in pouch cells exhibited long-term stability up to 4700 cycles. Zhou et al. [23] reported that the serious oxygen release was effectively restrained within a single-crystal Li-rich Li1.2Ni0.2Mn0.6O2 electrode, whereas the polycrystalline sample displayed severe irreversible capacity during the initial cycle and capacity fading. In general, the research on single-crystalline and polycrystalline cathodes is still mainly focused on Ni-rich and Li-rich layered-oxide cathodes [14,17,18,19,23]. There are rare reports on comparisons between single-crystalline and polycrystalline LiMn2O4 spinel cathodes.
In this work, we synthesized single-crystalline LiMn2O4 (SC-LMO) and polycrystalline LiMn2O4 (PC-LMO) cathode materials with similar particle sizes. The differences in electrochemical properties of SC-LMO and PC-LMO cathodes were investigated in a wide voltage window of 2–4.8 V vs. Li/Li+. The SC-LMO cathode exhibited higher Li-ion storage capacity than the PC cathode due to the large contribution at the 3 V voltage plateau. In contrast to what is reported on Ni-rich and Li-rich layered-oxide cathodes, PC-LMO showed superior cycling stability to SC-LMO.

2. Materials and Methods

2.1. Synthesis of SC-LMO and PC-LMO

The PC-LMO sample was synthesized using Mn2O3 microspheres as the template [24]. First, the MnCO3 microspheres were prepared via a precipitation method [25]. NaHCO3 (8.4 g) was dissolved in 700 mL of distilled water. Then, ethanol (70 mL) and the NaHCO3 solution were added to the MnSO4·H2O solution (1.6 g in 700 mL distilled water) in sequence under stirring. The mixture was kept under stirring for 3 h at room temperature. After centrifugation, filtration, and drying, MnCO3 powders were obtained. Then, the as-obtained MnCO3 powders were calcined at 700 °C for 10 h in the air to synthesize the Mn2O3 sample. Finally, the Mn2O3 sample was thoroughly mixed with LiOH·H2O in a molar ratio of Mn2O3:LiOH = 1:1.05 and heat-treated at 650 °C for 10 h in a muffle furnace.
SC-LMO was synthesized using a molten salt method. Stoichiometric amounts of LiOH·H2O and commercial Mn2O3 (molar ratio 1.05:2) were thoroughly mixed as lithium and nickel precursors. Then, the mixed precursors were ground with a large excess of CsCl salt (melting point: 645 °C). The molar ratio between the CsCl flux and precursors (referred to as the R ratio) was 4. The mixtures of CsCl salt and precursors were put in alumina crucibles and annealed at 850 °C for 10 h. Afterward, the product was washed with deionized water to remove the residual CsCl and then filtered, before finally drying at 100 °C in an oven for 12 h.

2.2. Material Characterizations

Powder X-ray diffraction (XRD) patterns were collected using a BrukerD4 X-ray diffractometer (Bruker, Dresden, Germany) with λ (Cu Kα) = 0.154 nm (40 kV, 30 mA). All the obtained XRD patterns are shown with background (without any smoothing). The morphologies of the prepared materials were observed by scanning electron microscopy (Philip XL30, Dresden, Germany) at 15 kV. Elemental analysis was carried out on an inductively coupled plasma optical emission spectrometry (ICP-OES). Dissolution of Mn ions in the electrolyte was quantified using a two-electrode glass device with 10 mL of electrolyte after 10 cycles. The electrolyte was treated under vacuum-drying and then under a small amount of concentrated HNO3, before finally diluting to a 50 mL solution using deionized water for ICP-OES tests.

2.3. Electrochemical Measurements

The working electrode was prepared by mixing the as-prepared active material (SC-LMO or PC-LMO), conductive agent (acetylene black), and binder (polyvinylidene fluoride, PVDF) with the help of N-methyl-2-pyrrolidone (NMP). The mass ratios of the LMO cathodes, acetylene black, and PVDF were 80 wt.%, 10 wt.%, and 10 wt.%, respectively. The slurry was coated onto an aluminum foil current collector and dried at 120 °C in a vacuum oven overnight. Then, the dried foil was cut into circular electrode with an area of 1.0 cm2. The areal mass loading of cathode is in the range of 1.2–1.5 mg·cm−2. Half-cell assembly was carried out in a glovebox with moisture and oxygen concentrations of 0.1 ppm. A solution of 1.0 M LiPF6 in ethylene carbonate/ethyl methyl carbonate (EC/EMC, 3:7) solvents was used as electrolyte. Li metal was used as both the counter and reference electrodes for half-cell tests, and all voltages are reported versus the Li/Li+ redox couple. Cyclic voltammetry tests were performed using a CHI660 electrochemical workstation between 2.0 V and 5.0 V at 0.1 mV·s1. The galvanostatic charge/discharge measurement was carried out by a Land tester (CT2001A) between 2.0 and 4.8 V vs. Li/Li+. All electrochemical tests were performed at room temperature. The specific capacities were obtained as a function of the mass of active cathode materials not including the mass of electrolyte and the Li metal anode.

3. Results

The crystallographic structure and crystallinity of the SC-LMO and PC-LMO were characterized by XRD patterns. The diffraction peaks (Figure 1a) of these two samples could be well indexed to the cubic structure in a space group Fd 3 ¯ m (standard card, PDF#53-1237) [13], where Li, O and Mn ions occupy the 8a tetrahedral, 32e octahedral, and 16d octahedral sites (Figure 1b). The Li-ion can deintercalate from the 8a tetrahedral sites, and the 8a tetrahedral sites together with 16c octahedral sites form a 3D pathway for Li-ion diffusion. The spinel structure has a three-dimensional network of channels for fast Li-ion conduction. Moreover, the Li-ion can intercalate into the empty 16c octahedral sites in an over-lithiation state, which may cause the transition of crystal structure from cubic LiMn2O4 into tetragonal Li2Mn2O4 [26]. Notably, the SC-LMO showed relatively stronger diffraction peaks than PC-LMO, indicating a higher degree of crystallinity in SC-LMO than PC-LMO.
The morphology of the as-synthesized LiMn2O4 samples was examined by SEM. As exhibited in Figure 2a, the morphology of PC-LiMn2O4 was microspherical with a rough surface. The diameter of the microsphere was in the range of 2.0–2.5 μm. The enlarged SEM image indicated that the spherical particle was densely composed of numerous small primary particles with sizes in the range of 100–200 nm (Figure 2b). In contrast to the spherical-shaped polycrystalline particles consisting of small grains, SC-LMO had a relatively regular polyhedral shape with a smooth surface and clear edges. SC-LMO had an average particle size of ~2.0 μm.
The electrochemical properties of SC-LMO and PC-LMO samples were evaluated on the basis of cyclic voltammograms (CV) and galvanostatic charge–discharge (GCD) measurements. As shown in the CV curves of both SC-LMO and PC-LMO (Figure 3a), there were two pairs of redox peaks in the 4 V region, corresponding to a two-stage Li-ion intercalation/deintercalation process at the 8a tetrahedral sites with LiMn2O4/Li0.5Mn2O4 and Li0.5Mn2O4/Mn2O4 transition [27,28]. Another pair of redox peaks in the 3 V region resulted from the further lithiation at the octahedral vacancies, forming over-lithiated Li2Mn2O4 with the phase transformation of cubic to tetragonal phases (LiMn2O4/Li2Mn2O4) [13,29]. Interestingly, in the CV curve of the SC-LMO cathode, the current of the cathodic peak associated with the lithiation of LiMn2O4 to Li2MnO4 in the 3 V region was slightly larger than that associated with the lithiation of Mn2O4 to Li0.5Mn2 O4 and Li0.5Mn2O4 to LiMn2O4 in the 4 V region. On the contrary, PC-LMO presented a smaller current of the cathodic peak in the 4 V region than in the 3 V region.
Figure 3b,c show the GCD profiles of SC-LMO and PC-LMO cathodes at 50 mA·g−1 in a wide voltage range of 2.0–4.8 V vs. Li/Li+. During the first charging process, both SC-LMO and PC-LMO cathodes exhibited two distinct voltage plateaus at ~4.1/4.2 V vs. Li/Li+. The first discharge capacities of SC-LMO and PC-LMO cathodes at a current density of 50 mA·g−1 were 178 and 172 mA·h·g−1, respectively. Clearly, the SC-LMO had a much higher specific capacity (88 mA·h·g−1) than PC-LMO (56 mA·h·g−1) in the 3 V region, which is consistent with the CV tests. This indicates that extra Li-ion storage at the 16c octahedral vacancies in spinel LiMn2O4 [13] was more easily activated in single-crystalline particles than in polycrystalline secondary particles.
In the voltage window of 2.0 and 4.8 V vs. Li/Li+, the SC-LMO cathode delivered a discharge capacity of 114 mA·h·g−1 with ~67% of its original capacity after 50 cycles, which was much better than the PC-LMO cathode with a capacity retention of ~48% after 50 cycles. Better cycling performance was achieved in the polycrystalline cathode than in the single-crystalline cathode. This is quite different from previous Ni-rich NMC and Li-rich cathodes [14,18,19,21].
To further understand the relationship between the cycle stability and crystallinity in the LiMn2O4 cathode, we carried out post-cycling structural characterizations of SC-LMO and PC-LMO electrodes using SEM, XRD, and ICP analysis. Figure 4a,b compare the morphology evolution of PC-LMO and SC-LMO cathodes after 10 cycles. The PC-LMO particles presented micro-cracks (intergranular fracture), whereas the SC-LMO particles showed no sign of grain boundaries after cycles. Regarding the cycle performance of polycrystalline LiMn2O4 cathodes, there was a slight gradual increase in discharge capacity during the 5th–20th cycles. In contrast, the increase in discharge capacity was not observed for the single-crystalline LiMn2O4 cathode (free of cracks after 10 cycles). Recent work from Jürgen Janek et al. [30] reported that the cracking of secondary polycrystalline particles leads to liquid electrolyte infiltration in the cathode active materials, lowering the charge-transfer resistance and increasing the Li-ion diffusion coefficient by more than one order of magnitude. This was also the case in our polycrystalline LiMn2O4 cathode. The slight increase in discharge capacities of PC-LMO might have resulted from the gradual electrolyte wetting due to the cracking formation.
Furthermore, post-XRD characterization (Figure 4c) suggested that more impurity phases (tetragonal Li2Mn2O4 phase and orthorhombic LiMnO2 phase) [31] appeared in the single-crystalline LiMn2O4 electrode than in the polycrystalline LiMn2O4 electrode after 10 cycles. Results from elemental analysis by ICP show that the dissolved Mn-ion amount in the electrolyte from the SC-LMO cathode was 1.5 times higher than that from PC-LMO (Figure 4d). The detected Mn ions were dissoluble Mn(II) ions, resulting from the disproportion reaction of Mn(III) before dissolving into the electrolyte. The loss of Mn at the LiMn2O4 cathode not only caused a gradual loss of active cathode, but also deteriorated the stable cathode–electrolyte interface, thereby leading to severe capacity decay during cycles [32]. It is well known that Mn(II) dissolution is also associated with the atomic arrangement of Mn-ions on the crystal facets of LiMn2O4. A large portion of Mn(II) ions exist at the {110} facet of LiMn2O4, leading to the {110} facet being more vulnerable to the Mn dissolution [33]. Compared to densely packed particles in polycrystalline LiMn2O4, it is probable that {110} facets are more exposed in single-crystalline LiMn2O4. This may be another reason for the more severe Mn(II) ion dissolution and worse cycling stability of single-crystalline LiMn2O4. Accurate crystal facet/orientation analysis is required for further characterization based on transmission electron microscopy (TEM) with a focused ion beam (FIB) in the future. Moreover, the high number of Mn(III) ions produced from SC-LMO would induce Jahn–Teller distortion due to the valence electronic configuration of t2g3eg1, which may lead to the structural degradation of spinel phases [34]. Consequently, in spite of cracking formation, the PC-LMO cathode had better cycle stability than the SC-LMO cathode due to the low capacity contribution in the 3 V region with cubic/tetragonal phase transition. In general, these studies indicate that the phase transition-induced Mn(II) ion dissolution in the 3V region instead of cracking formation is the limiting factor for the cycle performance of the spinel LiMn2O4 cathode operating in a wide voltage window.

4. Conclusions

In summary, we investigated the electrochemical properties of single-crystal LiMn2O4 particles and polycrystalline LiMn2O4 particles with similar sizes as cathodes for Li-ion storage. Our studies suggest that, while single-crystalline LiMn2O4 cathodes might have the advantage of enhancing specific capacity, polycrystalline LiMn2O4 cathodes have less cubic–tetragonal phase transition in the 3 V region, thus presenting better cycle stability. Unlike layered Ni-rich cathodes, the single-crystalline design does not work well for stable spinel LiMn2O4 cathodes. Post-cycling structural characterizations confirmed that the limiting factor for the cycle performance of the spinel LiMn2O4 cathode in the wide voltage window was Mn(II) ion dissolution rather than intergranular cracking formation.

Author Contributions

Conceptualization, F.W. and P.Z.; methodology, H.L.; validation, C.G. and Y.W.; data analysis, W.B. and J.L.; original draft preparation, F.Y.; writing—review and editing, F.Y., F.W. and P.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by National Natural Science Foundation of China (No.21601089), the Jiangsu Specially Appointed Professor Program, and the Startup Foundation for Introducing Talent of NUIST.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goodenough, J.B.; Park, K.S. The Li-Ion Rechargeable Battery: A Perspective. J. Am. Chem. Soc. 2013, 135, 1167–1176. [Google Scholar] [CrossRef] [PubMed]
  2. Whittingham, M.S. Ultimate Limits to Intercalation Reactions for Lithium Batteries. Chem. Rev. 2014, 114, 11414–11443. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, F.; Zhang, P.; Wang, G.; Nia, A.S.; Yu, M.; Feng, X. Functional Electrolytes: Game Changers for Smart Electrochemical Energy Storage Devices. Small Sci. 2021, 2, 2100080. [Google Scholar] [CrossRef]
  4. Thackeray, M.M.; Amine, K. LiMn2O4 spinel and substituted cathodes. Nat. Energy 2021, 6, 566. [Google Scholar] [CrossRef]
  5. Lee, S.; Cho, Y.; Song, H.; Lee, K.T.; Cho, J. Carbon-Coated single-crystal LiMn2O4 nanoparticle clusters as cathode material for high-energy and high-power lithium-ion batteries. Angew. Chem. Int. Ed. 2012, 51, 8748–8752. [Google Scholar] [CrossRef] [PubMed]
  6. Jiao, F.; Bao, J.; Hill, A.H.; Bruce, P.G. Synthesis of Ordered Mesoporous Li–Mn–O Spinel as a Positive Electrode for Rechargeable Lithium Batteries. Angew. Chem. Int. Ed. 2008, 47, 9711–9716. [Google Scholar] [CrossRef]
  7. Abou-Rjeily, J.; Bezza, I.; Laziz, N.A.; Autret-Lambert, C.; Sougrati, M.T.; Ghamouss, F. High-rate cyclability and stability of LiMn2O4 cathode materials for lithium-ion batteries from low-cost natural β−MnO2. Energy Storage Mater. 2020, 26, 423–432. [Google Scholar] [CrossRef]
  8. Ha, Y.; Harvey, S.P.; Teeter, G.; Colclasure, A.M.; Trask, S.E.; Jansen, A.N.; Burrell, A.; Park, K. Long-term cyclability of Li4Ti5O12/LiMn2O4 cells using carbonate-based electrolytes for behind-the-meter storage applications. Energy Storage Mater. 2021, 38, 581–589. [Google Scholar] [CrossRef]
  9. Zhang, S.; Deng, W.; Momen, R.; Yin, S.; Chen, J.; Massoudi, A.; Zou, C.; Hou, H.; Deng, W.; Ji, X. Element substitution of a spinel LiMn2O4 cathode. J. Mater. Chem. A 2021, 9, 21532–21550. [Google Scholar] [CrossRef]
  10. Lee, J.C.; Abbas, M.A.; Lee, M.D.; Lee, J.; Lee, J.; Bang, J.H. Lithiation Mechanism Change Driven by Thermally Induced Grain Fining and Its Impact on the Performance of LiMn2O4 in Lithium-Ion Batteries. Small 2020, 16, 2002292. [Google Scholar] [CrossRef]
  11. Labyedh, N.; Mattelaer, F.; Detavernier, C.; Vereecken, P.M. 3D LiMn2O4 thin-film electrodes for high rate all solid-state lithium and Li-ion microbatteries. J. Mater. Chem. A 2019, 7, 18996–19007. [Google Scholar] [CrossRef]
  12. Liang, J.; Zhang, R.; Li, H.; Wang, L.; Cai, J.; Yan, H.; Cao, W. The Electrochemical Mechanism of Preparing Mn from LiMn2O4 in Waste Batteries in Molten Salt. Crystals 2021, 11, 1066. [Google Scholar] [CrossRef]
  13. Zuo, C.; Hu, Z.; Qi, R.; Liu, J.; Li, Z.; Lu, J.; Dong, C.; Yang, K.; Huang, W.; Chen, C.; et al. Double the Capacity of Manganese Spinel for Lithium-Ion Storage by Suppression of Cooperative Jahn–Teller Distortion. Adv. Energy Mater. 2020, 10, 2000363. [Google Scholar] [CrossRef]
  14. Bi, Y.; Tao, J.; Wu, Y.; Li, L.; Xu, Y.; Hu, E.; Wu, B.; Hu, J.; Wang, C.; Zhang, J.; et al. Reversible planar gliding and microcracking in a single-crystalline Ni-rich cathode. Science 2020, 370, 1313–1317. [Google Scholar] [CrossRef]
  15. Langdon, C.; Manthiram, A. A perspective on single-crystal layered oxide cathodes for lithium-ion batteries. Energy Storage Mater. 2021, 37, 143–160. [Google Scholar] [CrossRef]
  16. Xu, C.; Märker, K.; Lee, J.; Mahadevegowda, A.; Reeves, P.J.; Day, S.J.; Groh, M.F.; Emge, S.P.; Ducati, C.; Mehdi, B.L.; et al. Bulk fatigue induced by surface reconstruction in layered Ni-rich cathodes for Li-ion batteries. Nat. Mater. 2021, 20, 84–92. [Google Scholar] [CrossRef]
  17. Logan, E.R.; Hebecker, H.; Ma, X.; Quinn, J.; HyeJeong, Y.; Kumakura, S.; Paulsen, J.; Dahn, J.R. A Comparison of the Performance of Different Morphologies of LiNi0.8Mn0.1Co0.1O2 Using Isothermal Microcalorimetry, UltraHigh Precision Coulometry, and Long-Term Cycling. J. Electrochem. Soc. 2020, 167, 060530. [Google Scholar] [CrossRef]
  18. Deng, X.; Zhang, R.; Zhou, K.; Gao, Z.; He, W.; Zhang, L.; Han, C.; Kang, F.; Li, B. A Comparative Investigation of Single Crystal and Polycrystalline Ni-Rich NCMs as Cathodes for Lithium-Ion Batteries. Energy Environ. Mater. 2018, 1, 88–98. [Google Scholar] [CrossRef]
  19. Wang, F.; Xiao, S.; Li, M.; Wang, X.; Zhu, Y.; Wu, Y.; Shirakawa, A.; Peng, J. A nanocomposite of Li2MnO3 coated by FePO4 as cathode material for lithium ion batteries. J. Power Sources 2015, 287, 416–421. [Google Scholar] [CrossRef]
  20. Leng, J.; Wang, J.; Peng, W.; Tang, Z.; Xu, S.; Liu, Y.; Wang, J. Highly-Dispersed Submicrometer Single-Crystal Nickel-Rich Layered Cathode: Spray Synthesis and Accelerated Lithium-Ion Transport. Small 2021, 17, 2006869. [Google Scholar] [CrossRef]
  21. Qian, G.; Zhang, Y.; Li, L.; Zhang, R.; Xu, J.; Cheng, Z.; Xie, S.; Wang, H.; Rao, Q.; He, Y.; et al. Single-crystal nickel-rich layered-oxide battery cathode materials: Synthesis, electrochemistry, and intra-granular fracture. Energy Storage Mater. 2020, 27, 140–149. [Google Scholar] [CrossRef]
  22. Harlow, J.E.; Ma, X.; Li, J.; Logan, E.; Liu, Y.; Zhang, N.; Ma, L.; Glazier, S.L.; Cormier, M.M.E.; Genovese, M.; et al. A Wide Range of Testing Results on an Excellent Lithium-Ion Cell Chemistry to be used as Benchmarks for New Battery Technologies. J. Electrochem. Soc. 2019, 166, A3031–A3044. [Google Scholar] [CrossRef]
  23. Sun, J.; Sheng, C.; Cao, X.; Wang, P.; He, P.; Yang, H.; Chang, Z.; Yue, X.; Zhou, H. Restraining Oxygen Release and Suppressing Structure Distortion in Single-Crystal Li-Rich Layered Cathode Materials. Adv. Funct. Mater. 2021, 31, 2110295. [Google Scholar] [CrossRef]
  24. Hai, Y.; Zhang, Z.; Liu, H.; Liao, L.; Fan, P.; Wu, Y.; Lv, G.; Mei, L. Facile Controlled Synthesis of Spinel LiMn2O4 Porous Microspheres as Cathode Material for Lithium Ion Batteries. Front. Chem. 2019, 7, 437. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, F.; Chang, Z.; Wang, X.; Wang, Y.; Chen, B.; Zhu, Y.; Wu, Y. Composites of porous Co3O4 grown on Li2MnO3 microspheres as cathode materials for lithium ion batteries. J. Mater. Chem. A 2015, 3, 4840–4845. [Google Scholar] [CrossRef]
  26. Xu, G.; Liu, Z.; Zhang, C.; Cui, G.; Chen, L. Strategies for improving the cyclability and thermo-stability of LiMn2O4-based batteries at elevated temperatures. J. Mater. Chem. A 2015, 3, 4092–4123. [Google Scholar] [CrossRef]
  27. Wang, F.X.; Xiao, S.Y.; Gao, X.W.; Zhu, Y.S.; Zhang, H.P.; Wu, Y.P.; Holze, R. Nanoporous LiMn2O4 spinel prepared at low temperature as cathode material for aqueous supercapacitors. J. Power Sources 2013, 242, 560–565. [Google Scholar] [CrossRef]
  28. Wang, F.X.; Xiao, S.Y.; Shi, Y.; Liu, L.L.; Zhu, Y.S.; Wu, Y.P.; Wang, J.Z.; Holze, R. Spinel LiNixMn2−xO4 as cathode material for aqueous rechargeable lithium batteries. Electrochim. Acta 2013, 93, 301–306. [Google Scholar] [CrossRef] [Green Version]
  29. Huang, W.; Zhang, M.; Liu, T.; Zhao, W.; He, L.; Yin, L.; Tan, Z.; Lin, C.; Liu, J.; Zhao, Q.; et al. Tunning the linkage of structure units to enable stable spin el-based cathode in the wide potential window. Nano Energy 2021, 89, 106457. [Google Scholar] [CrossRef]
  30. Trevisanello, E.; Ruess, R.; Conforto, G.; Richter, F.H.; Janek, J. Polycrystalline and single crystalline NCM cathode materials—quantifying particle cracking, active surface area, and lithium diffusion. Adv. Energy Mater. 2021, 11, 2003400. [Google Scholar] [CrossRef]
  31. West, K.; Vitins, G.; Koksbang, R. Synthesis and host properties of tetragonal Li2Mn2O4 and Li2Co0.4Mn1.6O4. Electrochim. Acta 2000, 45, 3141–3149. [Google Scholar] [CrossRef]
  32. Piao, J.; Duan, S.; Lin, X.; Tao, X.; Xu, Y.; Cao, A.; Wan, L. Surface Zn doped LiMn2O4 for an improved high temperature performance. Chem. Commun. 2018, 54, 5326–5329. [Google Scholar] [CrossRef]
  33. Kim, J.S.; Kim, K.S.; Cho, W.; Shin, W.H.; Kanno, R.; Choi, J.W. A truncated manganese spinel cathode for excellent power and lifetime in lithium-ion batteries. Nano Lett. 2012, 12, 6358–6365. [Google Scholar] [CrossRef]
  34. De Taeye, L.L.; Vereecken, P.M. Detrimental MnPO4F and MnF2 formation on LiMn2O4 in the 3 V region. J. Mater. Chem. A 2021, 9, 23256–23268. [Google Scholar] [CrossRef]
Figure 1. (a) Powder XRD patterns of single-crystal LiMn2O4 and polycrystalline LiMn2O4. (b) Crystal structure of spinel LiMn2O4.
Figure 1. (a) Powder XRD patterns of single-crystal LiMn2O4 and polycrystalline LiMn2O4. (b) Crystal structure of spinel LiMn2O4.
Crystals 12 00317 g001
Figure 2. SEM images of (a,b) polycrystalline LiMn2O4 and (c) single-crystal LiMn2O4.
Figure 2. SEM images of (a,b) polycrystalline LiMn2O4 and (c) single-crystal LiMn2O4.
Crystals 12 00317 g002
Figure 3. (a) CV curves of single-crystal LiMn2O4 and polycrystalline LiMn2O4 cathodes. GCD profiles of (b) single-crystal LiMn2O4 and (c) polycrystalline LiMn2O4 cathodes at 50 mA·g−1. (d) The cycle behaviors of single-crystal LiMn2O4 and polycrystalline LiMn2O4 cathodes.
Figure 3. (a) CV curves of single-crystal LiMn2O4 and polycrystalline LiMn2O4 cathodes. GCD profiles of (b) single-crystal LiMn2O4 and (c) polycrystalline LiMn2O4 cathodes at 50 mA·g−1. (d) The cycle behaviors of single-crystal LiMn2O4 and polycrystalline LiMn2O4 cathodes.
Crystals 12 00317 g003
Figure 4. Post-cycling structural characterizations of single-crystal LiMn2O4 and polycrystalline LiMn2O4 cathodes. SEM images of (ac) polycrystalline LiMn2O4 and (d) single-crystal LiMn2O4 cathodes after 10 cycles. (e) XRD patterns and (f) detected Mn(II) ion concentrations in electrolyte of polycrystalline LiMn2O4 and single-crystal LiMn2O4 cathodes after 10 cycles.
Figure 4. Post-cycling structural characterizations of single-crystal LiMn2O4 and polycrystalline LiMn2O4 cathodes. SEM images of (ac) polycrystalline LiMn2O4 and (d) single-crystal LiMn2O4 cathodes after 10 cycles. (e) XRD patterns and (f) detected Mn(II) ion concentrations in electrolyte of polycrystalline LiMn2O4 and single-crystal LiMn2O4 cathodes after 10 cycles.
Crystals 12 00317 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yu, F.; Wang, Y.; Guo, C.; Liu, H.; Bao, W.; Li, J.; Zhang, P.; Wang, F. Spinel LiMn2O4 Cathode Materials in Wide Voltage Window: Single-Crystalline versus Polycrystalline. Crystals 2022, 12, 317. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12030317

AMA Style

Yu F, Wang Y, Guo C, Liu H, Bao W, Li J, Zhang P, Wang F. Spinel LiMn2O4 Cathode Materials in Wide Voltage Window: Single-Crystalline versus Polycrystalline. Crystals. 2022; 12(3):317. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12030317

Chicago/Turabian Style

Yu, Feng, Yi Wang, Cong Guo, He Liu, Weizhai Bao, Jingfa Li, Panpan Zhang, and Faxing Wang. 2022. "Spinel LiMn2O4 Cathode Materials in Wide Voltage Window: Single-Crystalline versus Polycrystalline" Crystals 12, no. 3: 317. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12030317

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