Next Article in Journal
Effect of Annealing Temperature on Microstructure and Resistivity of TiC Thin Films
Next Article in Special Issue
Facile Preparation of Fe3O4 Nanoparticles/Reduced Graphene Oxide Composite as an Efficient Anode Material for Lithium-Ion Batteries
Previous Article in Journal
Microstructure, Mechanical Properties and Tribological Behavior of Magnetron-Sputtered MoS2 Solid Lubricant Coatings Deposited under Industrial Conditions
Previous Article in Special Issue
Electrodeposition and Characterization of Lanthanide Elements on Carbon Sheets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enhancing Lithium Manganese Oxide Electrochemical Behavior by Doping and Surface Modifications

by
Alexandru-Horaţiu Marincaş
1 and
Petru Ilea
1,2,*
1
Faculty of Chemistry and Chemical Engineering, Babeş Bolyai University, 11 Arany Janos Street, 400028 Cluj-Napoca, Romania
2
Interdisciplinary Research Institute on Bio-Nano-Sciences, Babeş Bolyai University, 42 Treboniu Laurian Street, 400271 Cluj-Napoca, Romania
*
Author to whom correspondence should be addressed.
Coatings 2021, 11(4), 456; https://doi.org/10.3390/coatings11040456
Submission received: 13 March 2021 / Revised: 4 April 2021 / Accepted: 12 April 2021 / Published: 15 April 2021
(This article belongs to the Special Issue Electrodeposition of Thin Films for Energy Applications)

Abstract

:
Lithium manganese oxide is regarded as a capable cathode material for lithium-ion batteries, but it suffers from relative low conductivity, manganese dissolution in electrolyte and structural distortion from cubic to tetragonal during elevated temperature tests. This review covers a comprehensive study about the main directions taken into consideration to supress the drawbacks of lithium manganese oxide: structure doping and surface modification by coating. Regarding the doping of LiMn2O4, several perspectives are studied, which include doping with single or multiple cations, only anions and combined doping with cations and anions. Surface modification approach consists in coating with different materials like carbonaceous compounds, oxides, phosphates and solid electrolyte solutions. The modified lithium manganese oxide performs better than pristine samples, showing improved cyclability, better behaviour at high discharge c-rates and elevated temperate and improves lithium ions diffusion coefficient.

1. Introduction

Since their introduction to market by Sony, in 1990, lithium-ion batteries (LIBs) became the most popular power source for consumer electronics (smartphones, laptops, portable medical devices) and electric vehicles too [1,2,3]. Lithium-ion batteries are preferred for current electrically powered devices because of their advantages over the classic energy storage devices: better gravimetric density, higher working potential, large number of charge/discharge cycles and small self-discharge rate [4,5].
The most studied cathode materials are lithium cobalt oxide (LiCoO2) [6,7,8], lithium iron phosphate (LiFePO4) [9,10,11] and lithium manganese oxide (LiMn2O4) [5,12,13]. The anode consists of carbonaceous materials [12,14], silicon [15,16] or transition metal oxides [17] and the electrolyte mainly contains LiPF6 dissolved in a mixture of organic solvents (ethylene carbonate and dimethyl carbonate) [18].
Lithium manganese oxide (LMO) is considered an appropriate cathode material for lithium-ion batteries because of plenty raw materials, environmental friendliness, low cost, relatively facile manufacturing process and higher potential (4.1 vs. Li/Li+) than LiCoO2 (4 V vs. Li/Li+) and LiFePO4 (3.45 V vs. Li/Li+) [4,19]. The theoretical capacity of LMO is 148 mAh g−1, lower than LiCoO2 (274 mAh g−1) and LiFePO4 (170 mAh·g−1). However, lithium cobalt oxide fails to deliver more than half of the theoretical capacity because of structural deformation [2,8]. Pristine LMO managed to deliver more than 95% of its theoretical capacity [20].
LiMn2O4 has a close cubic package (ccp) spinel structure within the Fd3m space group. Manganese ions are located in the 16 d octahedral positions, oxygen ions in 32 e sites, while the lithium ions from the 8a tetrahedral positions can diffusive through the interstitial space of the 3D diamond shaped framework formed by [MnO6] octahedrals [21,22,23,24].
It is well known that lithium manganese oxide’s performance in organic electrolyte solution at elevate temperatures, suffers from manganese dissolution into electrolyte [25,26,27], Jahn-Teller distortion [28] and electrolyte oxidation on the surface of LMO particles [29]. Another problem related to Li-ion battery functioning is the electrolyte stability. Normally, LIBs are assembled in chambers with controlled atmosphere and humidity. If any water traces remain in the cell’s assembly, it can react with lithium hexafluorophosphate (LiPF6) to form hydrofluoric acid (HF), which has a harmful effect on LMO particles, thus the cycling ability of the cathode material is drastically affected [30].
Two main strategies have been employed to diminish LMO’s disadvantages: structure doping (cations [31,32], anions [33,34] or multiple ions [29,35,36]) and surface modification of the particles [37,38,39]).
Cation doping stabilizes the LMO spinel structure, by reducing the amount of electrochemically active Mn3+ [40], which is responsible for manganese disproportionate reaction into electrolyte [41]. However, in the case of doped LiMn2O4 a longer cycle life has been noticed at both room and elevated temperature as well, due to lower capacity loss during charge-discharge cycles, because of the more stable unit cell [42,43]. In addition, the anions inserted into LMO’s lattice do not reduce the content of Mn3+, therefore the good discharge capacities of anion-doped LiMn2O4 spinels were recorded [44,45].
Jahn Teller distortion (JT) occurs on the particle surface and it is a phase transition from cubic (c/a = 1) to tetragonal (c/a > 1), which reduce the structure stability due to a change in volume [46]. JT occurs especially by the end of the discharge process, when the average Mn oxidation state decreases from 3.5 to 3.0 [47,48]. Surface modification of the LMO particles proved to be an efficient way to hinder the JT’s effect on cathode performance, by shielding the active material particles from the electrolyte solution, where Mn3+ can easily disproportionate into Mn2+ and Mn4+ [49].
While in a previous review [50], the most common synthesis methods for obtaining pristine LMO have been discussed, in the current article, an intensive study about the dopant ions and coating materials influence upon the stability improvement and electrochemical behaviour enhancement of lithium manganese ions have been written. This review article covers not only the most common dopants and surface coatings, but also the influence of precursors and synthesis methods upon the structure and electrochemical performance of modified LiMn2O4. As a means to ease the results visualisation, the electrochemical parameters (initial discharge capacity, capacity retention, number of cycles, coulombic efficiency and charge-transfer resistance and lithium diffusion coefficient) of the modified LiMn2O4 have been tabulated.
For a better and easier visualization of electrochemical results for doped and coated lithium manganese oxide, Table 1 contains the discharge rates used during the electrochemical tests. The values are arranged from the lowest to highest discharge rate, thus, some letters may not appear in alphabetical order in the following tables. Typically, the theoretical discharge capacity of LiMn2O4 is considered to be 148 mAh·g−1, thus 1 C equals 148 mAh·g−1, but there are references which consider 1 C = 296 mAh·g−1 [51,52] and 1 C = 120 mAh·g−1 [53], respectively.

2. LiMn2O4 Heteroatoms Doping

Lithium manganese oxide’s major drawbacks like structural instability during long time cycling and manganese dissolution can be suppressed by doping the pristine spinel with alkali (Na+ [54]), alkali-earth (Mg2+ [55]), transition (including lanthanides) (Ni [56,57,58,59], Cu2+ [60,61], Zn2+ [62,63], La3+ [64], Dy4+ [65,66], Ce4+ [67]) or triels metals ions (Al3+ [31,68]).
Among the studied transitional metals, nickel offered an interesting doping solution, due to high potential oxidation (4.7 V) of Ni2+ to Ni4+ [69]. However, the molar ratio between Ni and Mn must be carefully chosen, because, during cycling, two different regions can appear: one with increased Ni content and the other one with reduce nickel. Those two regions may determine the presence of Jahn-Teller distortion [59]. It was established that the optimal ratio between Li:Ni:Mn ions is 1:0.5:1.5 [69,70]. More aspects related to LiNi0.5Mn1.5O4 will be discussed in other sections of the review (see Future prospects). Strategies employed for enhancing structural stability of the lithium manganese oxide spinel included cation multi doping [71,72,73,74,75,76].
Heteroatoms used to stabilize MnO6 decrease the discharge capacity, in comparison with undoped LMO, by reducing the content of electrochemically active Mn3+, thus, several studies aimed to introduce not only cations, but also monovalent anions, like Cl or F [77,78], into the structure of LMO spinel. A study published by Mao et al. describes the doping effects of different cations and anions upon the electrochemical performance of LiMn2O4 [79].

2.1. Simple Ions Doping

2.1.1. Cations

Aluminum-Doped LiMn2O4

Aluminum is one of the most studied cations for doping lithium manganese oxide [80,81,82]. In the Table 2, some studies related to aluminum ion employment as a stabilizing cation are presented. The reasons why aluminum was widely used as a dopant for LMO are related to its costs, stability and plentiful raw materials. As it is described in Ref [43,83], Al3+ occupies the 16 d octahedral sites, its bond with oxygen is much stronger than the Mn–O bond, and thus the aluminum ions improve spinel’s structure stability during charge-discharge at room temperature and elevated temperature, respectively. LiAl0.08Mn1.92O4 synthesized by a polymer pyrolysis method exhibited amazing results at both room and elevated temperature. Delivering an initial discharge capacity of 114 mAh·g−1, LiAl0.08Mn1.92O4 managed to offer the same capacity after 50 cycles (discharge rate: 50 mA·g−1), while at elevated temperature, a capacity loss of only 0.014% per cycle was recorded. The electrochemical performances of LiAl0.08Mn1.92O4 were strongly connected to its improved crystallinity and uniform particle size. It was considered that an amount of 0.10 wt.% of Al can suppress the structural distortion during cell’s employment. LiAl0.1Mn1.9O4 obtained by a phase inversion technology was long-time cycled (420 cycles) and it still delivered 96% of its initial capacity of 117 mAh·g−1 [84]. One of the highest initial discharge capacities furnished by an Al-doped spinel [85], comparable to the ones showed by undoped spinels [47,86,87], was the result of a hydrothermal synthesis method. Owing 129.8 mAh·g−1 in the first cycle, it displayed a capacity loss of 0.475 mAh·g−1 per cycle, having a capacity retention of 81.7% after 50 cycles.
The amount of aluminum dopant quantity should be carefully controlled because an increased amount of Al3+ ions into the LMO framework alter the insertion/deintercalation of Li+, by occupying the lithium ions 8a positions inside LiO4 tetrahedral [43]. Moreover, an increased Al content reduces the lattice parameter, thus an initial lower discharge capacity will be provided by modified spinels, like in the case of LiAl0.5Mn1.5O4 material, when the initial discharge capacity was below 100 mAh g−1 (95.2 mAh g−1). Although, the majority of the studies were conducted using moderate aluminum amount, the LiAl0.375Mn1.625O4 synthesized by solution combustion method performed above average and it granted 96.1 mAh g−1 after 50 cycles (0.2 C) [68]. Other high-content Al-doped spinels were obtained employing a solution combustion technique, with and without the adding of oleic acid [99]. Oleic acid favors the formation of a micelle structure template that influence the lattice parameter and morphology, thus LiAl0.4Mn1.6O4 nanospheres were synthesized in the presence of oleic acid, while in its absence nanorods were formed [99]. Even though a similar initial discharge capacity was noticed for both morphologies, rod-like LMO performed worse and delivered only 87.4 mAh g−1 (78% of its initial capacity), while LMO nanospheres exhibited 104 mAh g−1 (92% of the original capacity).
Al-doped spinel was tested in aqueous 5 M LiNO3 electrolyte solution. The tests implied a long-time cycling (4580 cycles) and LiAl0.1Mn1.9O4 performed impressively, maintaining 70% of its initial discharge capacity of 105.6 mAh g−1 [93]. The current achievement overcame the cycling ability recorded for other LiAl0.1Mn1.9O4 reported in Ref [88] and [96].

Rare earth-Doped Lithium Manganese Oxide

Another class of single doped lithium manganese oxides is represented by the rare earth elements doped LMO. The main studies reported in the literature are indexed in Table 3. Lanthanum and cerium are among the most investigated lanthanoids [19,100,101]. The larger La3+ ion has been inserted into spinel’s structure and due to stronger bonds between La–O than Mn–O, the lattice parameter and crystals size of doped samples are smaller than for pristine LMO [100].
Preliminary studies related to Dy compatibility as a dopant were realized [65,102]. It was noticed that an extremely low amount of Dy (0.02 wt.%) could not stabilize and improve the LMO performance, while an increased quantity led to enhanced results, offering an initial discharge capacity similar to undoped spinels, but a severe capacity fade of 31% was observed after only 10 cycles. Thus, further studies need to be done, in order to take Dy in consideration as a dopant for LiMn2O4.

Other Metal Cations—Doped LiMn2O4

Throughout the years, many other metal ions were tested as possible dopants for LiMn2O4 (Table 4).
Iron doped lithium manganese spinel was successfully synthesized by a sol-gel process [52]. Iron ions improved the electrochemical performances by partially replacing the Mn3+ from 16 octahedral sites (fact which suppressed the Jahn-Teller distortion and stabilized the spinel’s structure), by reducing the amount of Mn3+, which were responsible for blocking the lithium ions movement channels, in 8a tetrahedral sites and preventing the formation of manganese with lower valance on the particles surface. The mixture of chelating agents used during the sol-gel process, citric acid and glucose, helped the targeting of iron ions to the mentioned sites, by affecting the kinetics of the synthesis reactions. The altered LiFe0.1Mn1.9O4 with an initial discharge capacity of 134.7 mAh·g−1 demonstrated long cycle life at high discharge rate (100 C), mainly because the structural modifications occurred after Fe3+ doping. The discharge capacities exhibited by iron doped spines after 1400 cycles at RT and 1500 at 60°, respectively, were by the far better than in the case of dopants like Mg [106], Gd [102], Cr [107] or Ga [108].
A silicon modified LMO with high initial discharge capacity (146 mAh·g−1, 0.1 C) has been obtained by a freeze-drying method [108]. In comparison with divalent or trivalent ions, by doping with tetravalent ion (Si4+), the quantity of electrochemical active Mn3+ was not reduced. It has been demonstrated during the charge-discharge cycles that the entire trivalent manganese oxidized to Mn4+. After 300 cycles, LiSi0.05Mn1.95O4 exhibited a lower capacity retention (75.9%) than the gallium-doped spinel synthesized by a similar technique, because of the larger Mn3+ content.
Chromium influence upon structure, stability and electrochemical properties of LMO spinels was examined [114,121,126]. Because Cr6+ is toxic, the reagent salts must be carefully picked. A relatively low content of Cr3+ stabilized LMO’s structure by replacing Mn in 16 d sites, reduced the charge-discharge resistance, thus favoring the lithium ions diffusion during cycles [126]. Unfortunately, the calculated lithium-ion diffusion coefficient of 9.61·10−12 cm2·s−1 [126] was similar to titanium surface-doped LMO [115] and much smaller than for LMO with incorporated PO43− [128].
High content chromium doped lithium manganese oxide (LiCr0.2Mn1.8O4) exhibited an initial low discharge capacity (105.4 mAh·g−1) in an Aqueous Rechargeable Li-ion Battery (ARLB), but it proved a very long cycle ability and after 10,000 cycles, the capacity fade was only 27%. This outstanding result was attributed to stronger stabilizing Cr–O bonds, especially in low lithiated state [121].
Among other ions, pentavalent ions such as Nb5+ were inserted into LMO spinel’s lattice by different synthesis methods, solid state reactions [111] and ion implanted method [127]. Niobium has been selected as a dopant ion due to its similar ionic radius to manganese ion, which can increase the electrical conductivity of the material due to extra electron and its metal-oxygen bond has higher energy [111]. In the case of Li1.02Nb0.01Mn1.99O4 obtained by SSR, LiNbO3 impurities have been noticed, but these impurities favored lithium movement throughout the cycling tests. By increasing the Nb5+ amount, LiNbO3 hindered the lithium diffusion by blocking its diffusion paths.
A preliminary study related to wolfram doping of LMO has been published [88]. Due to wolfram oxidation to a higher state (W6+), its ionic radius (0.62 Å) became shorter than Mn3+ ionic radius (0.64 Å), thus the lattice parameter of W doped spinel was smaller than pristine sample. Poor capacity of LiW0.025Mn1.975O4 was mainly connected to the presence of electrochemical inert wolfram oxides and high-valance wolfram state.
Rod-like Sc-doped lithium manganese oxides were synthesized by a solid state reaction route by Bhuvaneswari et al. [129]. Rod length varied between 100–600 mm depending on the weight ratio of scandium ions into the structure and increased Sc content determined longer rods. Among the tested percentage of Sc, the best results were recorded for LiSc0.06Mn1.94O4. A slight decrease in the cell parameter “a” was noticed, but this aspect was attributed to the stronger Sc–O bond. Moreover, by introducing Sc3+ ions into the structure in the 16 d sites, the content of Mn3+ decreased, thus the resulted crystal structure is more stable, because by taking into consideration that Mn3+–O bond is longer than Mn4+-O bond, the MnO6 octahedrons have a more compact arrangement and less volume change during cell’s operating. The Sc-doped LMO delivered a slightly lower discharge capacity (114 mAh g−1) than pristine LMO (117 mAh·g−1) but succeeded to maintain ~96% of the initial capacity after a long time cycling of 500 cycles (discharge rate = 1 C), while pristine sample failed more than 75% of the first cycle capacity. Li+ ions’ diffusion coefficient was one magnitude order higher for the Sc-doped (1 × 10−12 cm2 s−1) sample than for undoped LMO (1 × 10−13 cm2 s−1). However, the calculated DLi+ for LiSc0.06Mn1.94O4 coefficient was not as large as in other cases like La-Sr-Mn coated LiMn2O4 [130], LaF3 coated LMO [131] or LiAl0.04Mn1.96O3.96F0.04 [95].

2.1.2. Anions

Anions were also considered as potential doping agents for stabilizing the structure of lithium manganese oxide (Table 5).
Fluorine (F) was regarded as a promising dopant, because, due to its lower ionic radius and higher electronegativity, it partially replaces divalent O2− and decreases the Mn4+ content [34,95]. By reducing Mn4+ to Mn3+, the amount of electrochemically active Mn3+ increased, thus, the initial capacity of the fluorine doped spinel was higher than of aluminum doped LMO, synthesized under similar conditions [95].
Sulfur doped spinels were synthesized by plasma assisted solid state method and sol-gel [33,44]. In both cases, high discharge capacities (above 125 mAh·g−1) were registered in the first cycle and the capacity retention was higher than 92% after 60 cycles or more. This remarkable behavior was attributed to stabilizing character of sulfur, by efficiently preventing the Jahn-Teller distortion at low voltage and by reducing the amount of dissolved manganese into the electrolyte solution, even at elevated temperature. Another important aspect of the improved electrochemical performances of LiMn2O4−xSx can be attributed to a more powerful manganese-sulfur bond, which stabilized the structure.
The phosphate group used as a dopant was assimilated into the LMO lattice and it weakened the manganese-oxygen bond, fact proven by FTIR spectra [132]. As a result of no manganese content change into the modified samples, the initial discharge capacity of the Li1.017Mn1.97O4(PO4)0.015 was almost similar to the one for pristine sample (116 mAh·g−1). An excellent benefit of the (PO4) doping was related to the lithium diffusion coefficient, which increased by three magnitude orders, from 3.49·10−11 cm2·s−1 for LiMn2O4 to 5.38·10−8 cm2·s−1 for Li1.017Mn1.97O4(PO4)0.015. Such improvement was connected to the PO43−’s structure, its ionic radius being larger than for O2−, favored the formation of a 3D framework, with large channels which facilitated the lithium ions movement during the electrochemical tests.

2.2. Multiple Ions Doping

2.2.1. Cations Multi-Doping

Several attempts for multi-doping of LiMn2O4 have been investigated (Table 6). Except the entries marked with (g) in Table 5, all the tests have been done at room temperature or 25 °C.
An equal percentage of zinc and praseodymium led do a higher initial discharge capacity (130 mAh g−1), not only in the case of other single- [115] or multi-doped spinels [133], but also for pristine LiMn2O4 [134]. Despite of the low initial coulombic efficiency (only 76.5%), it managed to increase several percentages after a short number of cycles (10 cycles) and it is expected to slightly increase, if the cycling tests will last longer. Almost no capacity fade was recorded for the investigated number of cycles.
Even small amounts of bismuth and lanthanum had a beneficial effect on LiMn2O4 rate capability and an extremely small capacity loss of only 0.08 mAh g−1 per cycle was recorded [21].
Multi cations (Al, Si, Ti, Co) doped LiMn2O4 was suggested by Dokan et al. [135]. In comparison with the pristine spinel synthesized via the same glycine-nitrate combustion process, the doped sample exhibited an initial higher discharge capacity and the dopants stabilized the spinel structure, thus a capacity fade of only 10% was recorded after 30 cycles (1 C).
Mesoporous Al and Ni doped LiMn2O4 spherical particles, with high initial discharge capacity (140.5 mAh·g−1), related to stabilizing function of Ni, were synthesized by solid state reactions, using as-synthesized MnCO3 microspheres, LiOH·H2O and Ni(NO3)2·6H2O [94]. However, in spite the advantages offered by porous structure and aluminum and nickel ion doping, LiAl0.1Ni0.1Mn1.8O4 was also affected by the structural distortion from cubic to tetragonal during cycles, and severe capacity loss (52.6%) was noticed after only 75 cycles.
Impressive electrochemical performance was recorded for a LMO doped with low content three ions with different valences of Cu2+, Al3+ and Ti4+ through a lactic acid aided Pecchini process [149]. According to the authors, the homogenous distribution of the dopants played an important role in achieving such effectiveness. In the first cycle, 90.5% of LMO’s theoretical capacity (148 mAh g−1) was delivered and after a large number of cycles (400 cycles, RT), the capacity fade was just 3%. High capacity retention of 90.1% (119.2 mAh g−1) has been measured for LiCu0.02Al0.02Ti0.02Mn1.94O4 sample at elevated temperature (55 °C) too. The results obtained for the current doped spinel are similar to magnesium and fluoride multi-doping of LiMn2O4 [78].
LMO spinel stability has been enhanced by an equimolar structure doping with magnesium and silicon ions [140]. The co-doping with Mg2+ and Si4+ favored two aspects: first of all, by doping with Mg2+, the average valence of manganese increased to (IV), which hindered the Jahn-Teller structural modification and resulted in a more stable compound; secondly, the Si4+ determined the formation of a more lasting [MnO6] octahedral, which permitted the lithium ions diffusion much easier. By increasing the content of dopants, the discharge capacity diminished (from 139 mAh·g−1 for pristine sample down to 115.6 mAh·g−1 for LiMg0.07Si0.07Mn1.86O4), but the cycling life has been improved, from 87.1% (bare LMO) to 99.3% (LiMg0.07Si0.07Mn1.86O4), after 30 cycles. LiMg0.05Si0.05Mn1.86O4 has been chosen as the optimum doped sample, with an initial discharge capacity of 126.9 mAh·g−1 and a retention rate of 97.3% after 100 cycles. In addition, it showed enriched capacity retention at elevated temperature (92.7%, 0.5 C, 55 °C) and low charge-transfer resistance after 100 cycles (102.9 Ω), which suggested better lithium intercalation/deintercalation, because side compounds did not block Li+ paths.
Preliminary studies of double doping of LiMn2O4 with Gd and Ni have been attempted in Ref. [142]. The obtained material was tested in aqueous electrolyte (saturated Li2SO4 electrolyte) and it delivered an initial fairly poor discharge capacity of only 67.1 mAh·g−1. The doped spinel exhibited a similar behavior like in non-aqueous electrolytes and two pairs of peaks were noticed. Moreover, the doping reduced the charge transfer resistance for almost ten times, from 800 Ω (pristine sample) to 85 Ω (Li(GdNi)0.01Mn1.98O4), this fact suggested improved kinetics for the double-doped spinel.
Nickel and copper multiple metal doping of lithium manganese oxide by a citric acid aided sol-gel process has been realized by Iqbal et al. [152]. In the case of samples with low amount of dopants, Ni-Cu ions tend to occupy the tetrahedral positions 8a, while, by increasing the amount of doping ions, Ni-Cu will reside to 16d octahedral sites. The differences in lattice constant and particles size between pristine, low and high doped LMO resemble in the fact that tetrahedrally coordinated Cu2+(0.57 Å) own a slightly lower ionic radius than Li+(0.59 Å), while the Mn3+ ion, which is normally replaced by dopants, has a radius relatively lower (0.645 Å) than octahedrally coordinated Cu2+(0.73 Å) [152]. Regarding the electrochemical performances, all the samples showed two pairs of redox peaks (at 4.1 and 3.9 V, respectively), which are characteristic for bare lithium manganese oxide. While the pristine sample delivered a specific capacity of 122 mAh·g−1 (82.4% of theoretical discharge capacity) but faced drastically capacity fade over a number of 100 cycles (28%). Among all the samples, the best results were recorded in the case of low doped LMO, LiCu0.01Ni0.01Mn1.98O4. It provided a discharge capacity of 113 mAh·g−1 (c = 0.3 C) and maintained 88% of its initial capacity after 100 cycles. The very low amount of dopants managed to enhance the lithium kinetics during charging by reducing the charge-transfer resistance by more than two times, from 212 to 97 Ω.

2.2.2. Anions and Cations Co-Doping of LMO

Many synthesizes investigated the effect of anion and cation co-doping upon the structure, morphology and electrochemical performances of lithium manganese oxide (Table 7).
A multi doped LMO with Mg and F exhibited a relatively good discharge capacity in the first cycle (121.1 mAh·g−1). Moreover, it had an excellent rate capability at room temperature and elevated temperature (55 °C), delivering 116.3 and 108 mAh·g−1, respectively, after 400 cycles at 1 C [78].
In the case of simultaneous cation and anion doping, the strength of newly formed bonds, like Mg–O and Mn–Cl are better than Mn–O, thus the spinel structure was stabilized and the Mn3+ dissolution into electrolyte diminished [77].
Raja et al. produced [153] a nickel and sulfur co-doped LMO by an alanine-assisted low temperature combustion [153]. The nickel ion partially substitutes Mn3+ and sulfur, due to its improved catalytic activity, favors the oxidation of Mn3+ to Mn4+, resulting in a more stable [MnO6]2− and preventing the Jahn-Teller distortion. By increasing the amount of dopants, the LMO’s morphology changes from spherical shaped particles, in the case of pristine sample, to octahedron grain for the sample, which contained a high amount of nickel (LiNi0.5Mn1.5O4−δSδ). The multifaceted LiNi0.4Mn1.6O4−δSδ exhibited an interesting electrochemical behavior, because no capacity fade was observed in the first 25 cycles, but, moreover, the discharge capacity increased from 126 to 135.5 mAh·g−1. These phenomena might be related to sulfur catalytic activity and stabilizing nature at high voltage (4.5 V). Solid-state reaction method has been employed to obtain multiple spinel with magnesium and fluorine [78]. According to TGA analysis, prior to spinel formation exothermic process (708 °C for pristine and 687 °C for modified spinels, respectively), characteristic endothermic processes for water elimination (between 50–100 °C) and lithium carbonate discomposure occurred. Low doping level of lithium manganese oxide samples with Mg2+ and F- (LiMg0.05Mn1.95O3.9F0.1 and LiMg0.1Mn1.9O3.8F0.2) exhibited better Mn3+/Mn4+ ratio, hence the initial discharge capacities (119.1 and 121.4 mAh·g−1, respectively) were improved in comparison with pristine sample (117.2 mAh·g−1). In addition, 0.2 wt.% of magnesium led to a lower value of initial discharge capacity (~110 mAh·g−1). LiMg0.1Mn1.9O3.8F0.2 showed good stability and cycle life (90.6% of initial capacity after 400 cycles) and performed well at high c-rate (20 C), delivering 76.1 mAh·g−1, with 40% better than the pristine LMO (45.9 mAh·g−1). Magnesium and fluorine co-doped spinels demonstrate enhanced lithium ions diffusion kinetics and the charge-transfer resistance was drastically minimized by dopants.

3. Improving LiMn2O4 by Coating

Surface modification of LiMn2O4 particles proved to be an efficient way to diminish the harmful activity of electrolyte solution during cell cycling. As mentioned before, one important drawback of LMO spinel is related to manganese dissolution. The coatings not only improve the material stability by reducing its contact area with the electrolyte solution, but also enhance the electrochemical characteristics by reducing the charge transfer resistance and providing a better samples-cycling stability, for instance [162,163]. Throughout the years, many different coatings, which include organic and inorganic compounds like carbon, oxides, perovskites, solid solutions, have been studied [164,165,166,167].

3.1. Carbon

Lithium ions diffusion and rate capability life of lithium manganese oxide can be enhanced by using high conductive carbon-based materials [168,169]. Different carbonaceous coatings like hard carbon [12], activated carbon [170,171], carbon nanotubes [172,173,174,175,176,177], graphene [178,179], amorphous carbon [177,178,179,180,181,182], carbon nanospheres [183] were studied in the last years.
Reduced graphene oxide (rGO) nanosheets were employed as support for LMO uniform deposition by a low temperature microwave assisted hydrothermal reactions method [184]. The rGO presence prevented LMO agglomeration and low range particle distribution (10–40 nm). High crystalline LMO samples with stable 3D framework behaved remarkably at low and high c-rates, as well. Delivering an initial discharge capacity of 137 mAh·g−1 at 1 C, LMO/rGO hybrid material exhibited 85.4% (117 mAh·g−1) and 73.7 (101 mAh·g−1) of its initial discharge capacity, at 50 and 100 C, respectively. The improvement in material’s performance was strongly related to the properties of rGO: high surface area, protecting barrier against the electrolyte solution and the channels presence for permitting faster lithium ions diffusion [14].
LiMn2O4/graphene composite electrodes were proposed as suitable cathode materials for Li-ion batteries [168,169,179,185,186,187,188,189,190]. The unique and amazing graphene properties, such as high electrical conductivity, thermal and mechanical stability and high surface area, provided an efficient method for improving lithium manganese oxide’s electrochemical properties [14,167,168,189,191]. Brief characteristics about the electrochemical behavior of lithium manganese oxide/graphene composite materials are presented in Table 8, while a more advanced description about graphene influence on LIB’s cathode materials has been described elsewhere [192]. The challenges related to graphene-based materials usage in energy storage devices have been discussed in [14].
LiMn2O4 oxide microspheres were synthesized by a cyclohexanone hydrothermal reaction method, followed by a coating process with amorphous carbon, resulted after calcination of a LMO and glucose mixture [181]. The thin 1.5 nm coating layer enhanced the initial discharge capacity of the surface modified sample, from 127.9 to 138.5 mAh·g−1. In addition, amorphous carbon reduced the contact area between active electrode material and electrolyte, thus, a superior capacity retention (97.6%) was calculated for LMO/C sample after 100 cycles, 0.1 C.
Zhu et al. [191] published an ample review about novel carbon-based coatings (graphene, carbon nanotubes) for enhancing the behavior of LiMn2O4 in Rechargeable Hybrid Aqueous Batteries (ReHAB).

3.2. Oxides

Many researchers drove their attention to the usage of metal oxides as surface coatings for lithium manganese oxide [195].
Several research groups focused their studies on the influence of different lanthanide oxides on LMO particles. Since 2007, several studies were concentrated on CeO2 behavior as coating for lithium manganese oxides [164,165,196,197].
In order to modify the pristine samples, disparate approaches were made. Arumugam et al. [197] employed a polymeric process followed by a high temperature heat treatment (850 °C for 6 h), in order to obtain the modified samples. As expected, the initial discharge capacity of the pristine sample (135 mAh·g−1) was slightly higher than the one of 1 wt.% CeO2 (126 mAh·g−1), but the latter sample exhibited improved stability during cell’s cycling in both room and elevated temperature (60 °C) tests, providing 120 and 117 mAh·g−1, respectively, after 100 cycles at 0.5 C. While the pristine samples failed to offer more than 60% capacity retention at 30 °C and 50% at 60 °C, the coated material provided astonishing capacity retention of at least 93%, in both cases. Not only the cycle performance is important, but also the material ability to work at higher current rates is important. The 1 wt.% CeO2-LiMn2O4 was cycled for 100 times at higher current rate (5, 10, 15, 20 C, respectively) and its performance was compelling in all the case, because the initial discharge capacities were relatively high (102 mAh g−1 at 20 C) and small capacity fades were recorded (between 12%–17%). For a better understanding of coating layer importance, the electrochemical impedance spectra of LMO-based electrodes were studied. It was noticed that the charge transfer resistance was greatly decreased by coating, while the side reactions between the electrode material and electrolyte solution were minimized. Despite the fact that the initial Rct of the bare sample was slightly smaller than the one of coated material (21 Ω vs. 26 Ω), after 100 cycles it increased for more than four times, up to 98 Ω, while a Rct of only 42Ω was recorded for the coated LMO.
The precipitant influences, namely (NH4)2CO3 and NH4OH, in a precipitation synthesis of CeO2 modified LiMn2O4 particles were studied by Cho et al. [196]. Small particles size and a uniform coating layer were obtained when NH4OH was involved in the synthesis process, while the usage of (NH4)2CO3 favored particle agglomeration, when the coating amount increased. The precipitant used during the coated materials synthesis enhanced its capacity retention, from 50% for the pristine up to 88% for 5 wt.% CeO2 upon cycling for 200 times, at 60 °C and 1 C. Even though the lowest capacity fade was recorded for the 5 wt.% CeO2 (about 12%), its discharge capacity is still too low to be used in practical batteries (~70 mAh·g−1. Due to uniform coating layer formed when NH4OH was used, the electrode material was less exposed to electrolyte solution and thus the sample electrochemical characteristics were improved, with decreased recorded capacity fade (12–24%). In contrast, higher capacity fades were observed for the samples with (NH4)2CO3 as precipitant (32–40% of initial discharge capacity).
Michalska et al. [164] employed another investigation on CeO2 coating layer. The modified samples were obtained through a sol-gel synthesis followed by a low-temperature calcination (350 °C). The resultant particles presented relatively clear surface, with wide size interval (between 100 and 600 nm), which agglomerated into larger clusters. Compared to the previous studies, the initial discharge capacity of the modified sample was higher (115 mAh·g−1) than the pure LiMn2O4 (mAh·g−1), thus the lithium intercalation-deintercalation processes were not negatively influenced. In addition, the capacity loss of LMO/CeO2 sample was just 2% after 100 cycles (1 C), in comparison with 10% for bare LMO. Another benefit of the cerium oxide coating can be correlated with material performance at high current rates (Table 9).
Among the lanthanide oxides, La2O3 was another oxide studied by Amurugam et al. [198,199]. Uniform grains with average diameters of 200–300 nm and clear surface were obtained. During the electrochemical tests at 30 and 60 °C, respectively, the coated LMO samples exhibited higher capacity retentions than the pristine sample and among all the samples the performance of LMO/2 wt.% La2O3 had been highlighted. 2 wt.% La2O3 coated LMO not only showed good discharge capacity, but also possessed a high coulombic efficiency at high c-rates (2 and 5 C), which increased with the number of cycles. In the case of discharge rate of 2 C, a coulombic efficiency of 100% was obtained after 80 cycles and it remained unchanged even after 100 cycles. In the other situation, at 5 C, the highest coulombic efficiency (95%) was noted after 50 cycles and it did not modify until the experiment’s end.
Lithium manganese oxide modified with 5% La2O3 was synthesized by a classical solid-state reaction method by Feng et al. [200]. The SEM analysis of the sample showed rough surface for the coated spinel, which tended to form agglomerations. Similar discharge capacity with Ref. [199] (116.4 mAh·g−1) was recorded when the cell was cycled at a temperature of 25 °C and a c-rate of 1 C. In contrast with the pure LiMn2O4, there was no sudden capacity loss, and the material was able to provide 90.1% of its initial capacity after 200 cycles. Since LiMn2O4 suffers from elevate temperature instability, 5 wt.% La2O3 demonstrated that it could be an efficient spinel coating when the cell was tested at 55 °C. Like in similar studies, the capacity retention of the current modified sample was at least 82% after 200 cycles.
Lanthanum oxide synthesized by a combustion method, was coated on LMO particles in neopentyl glycol [201]. In contrast with the previous reported article related to La2O3 modified lithium manganese oxide, 3 wt.% La2O3 improved the cycling ability of LMO and a capacity loss of 8.3% was seen after 100 cycles at elevated temperature (60 °C).
Due to its interesting properties like thermal and structural stability and fast kinetics, titanium dioxide was employed as coating for lithium manganese oxide [202,203,204,205]. In one of the first studies about the TiO2 effect upon the improvement of electrochemical properties of LiMn2O4, it was noticed that it did not modify the spinel structure, but a spinel LiTixMn2−xO4 formed on the pristine particles [205]. In contrast, when TiO2-B particles with nano-belts morphology were coated, no intermediary Li-Ti-Mn-O layer was obtained. The TiO2 (B) porous coating hindered manganese dissolution and enhanced the lithium ions transport, throughout their network, fact proven by smaller charge-transfer and solid electrolyte interface (SEI) resistances, in comparison with bare sample [202].
Different mass weight percentages for possible coatings were studied, in order to determine de optimal amount of TiO2 coating for LiMn2O4. Zhang et al. [203] synthesized modified spinel with 0.5, 1, 2 and 3 wt.% TiO2, respectively. The charge-discharge curves and cycling performance emphasized that the optimal amount was 2 wt.% TiO2, because of its highest initial discharge capacity (129 mAh·g−1) and low-capacity loss after 50 cycles (5.5%). Using the same percentage of TiO2, Shang et al. [202] studied long time cycling of LMO/TiO2 materials at elevated temperature (55 °C), in order to determine the improvement done by the oxide coating. After 300 cycles, the modified sample furnished a capacity of 77.4 mAh·g−1 (74% of initial discharge capacity), while the LMO offered only 63% of the initial capacity (66.1 mAh·g−1). Walz et al. [206] used ZrO2 and TiO2 as coatings for bare LiMn2O4 and proposed a possible mechanism ZrO2 and TiO2 protection of LMO during cycling at elevated temperature (55 °C). The HF, which has such a harmful effect on lithium manganese oxide, can appear as a reaction’s product between LiPF6 and water traces during cell’s assembly [207]. TiO2 may react with HF and form a compound with the formula of TiO2·2HF and it prevents the manganese dissolution, improving spinel’s stability over repeated cycling. Having an average diameter of only 4.8 nm and pores size around 0.7 nm, TiO2 coating reduced to half the internal resistance of the cell (81 Ω), which remained lower than for the pristine LiMn2O4 cell (168 Ω). The discharge capacity of the modified LMO decreased with about 10% (from 120 to 103 mAh·g−1), while a serious severe capacity fading of ~75% was recorded for pristine material.
For the TiO2 (B)-nano belts LiMn2O4, two synthesis approaches were made. One of them [203] consisted of facile phase liquid mixing, while the other one was a two-steps synthesis, which employed a preliminary hydrothermal method stage, followed by electrostatic attraction [202]. In the case of the latter suggested synthesis method, the heat treatment for the obtaining of modified sample was also studied and it was clearly noticed that a heat treatment of 2 h would be enough to synthesis a porous coating layer on the pristine LMO. When a longer heat period (4 h) was tested, it was observed that a dense compact layer formed and it did not improve the electrochemical performance of the sample.
LMO/V2O5 particles were easily obtained by a solid-state method, followed by a calcination at 500 °C, for 2 h in air atmosphere [208]. The XRD patterns of the synthesized samples showed that the vanadium oxide did not modify LMO spinel structure and no remarkable differences between pristine and modified samples XRD peaks were noticed. The main reason of applying vanadium oxide coating on the pristine sample was, like in other case, the improvement of LMO spinel’s stability. The coating layer was not dense, so the modified sample solid electrolyte interphase resistance (RSEI) had a slightly higher value (8.52 Ω) than the pristine LMO (8.35 Ω) and the LMO charge-transfer resistance was greatly decreased from 69.72 to 26.51 Ω, demonstrating the development brought by V2O5 coating. During cycling at elevated temperature (>55 °C) and since the electrode material is in contact with the electrolyte on wide area, the possible presence of HF in the electrolyte tends to encourage the disproportionate reaction of manganese ions (Mn3+). Inductively coupled plasma atomic emission spectroscopy (ICP-AES) studies showed that the manganese concentration in electrolyte solution, in certain conditions (30 days’ samples immersion at 55 °C), was five times smaller (88 ppm) for the coated LMO, than for bare LMO, proving the coating efficiency again manganese dissolution. The best results were for 2.5 wt.% V2O5 when discharge capacities of 107.3 and 92 mAh·g−1, respectively, were provided after the materials were cycled for 200 times at room and elevated temperatures. In spite the fact that it was not such a large difference between LMO and 2.5 wt.% V2O5/LMO (~1%), the low discharge capacity exhibited by the pristine sample synthesized makes it unsuitable for practical devices.
In one of the latest studies on oxide coating for LiMn2O4, Tao et al. [209] coated MoO3 through a wet chemical method. Throughout this synthesis process, a uniform coating layer with a thickness of about three nm was deposited. The cyclic voltammetry tests revealed the two peak pairs, specific for LiMn2O4 spinel, but in the case of LMO/MoO3, the peaks were slightly shifted to the right, with about 0.05 V. Both pairs of peaks of the modified sample were sharper than the ones of bare LMO. The main characteristic of the modified samples was the ability to provide almost 70% of its initial discharge capacity (101.6 mAh·g−1) after long time cycling (900 times) at a current density of 2 C. All the results mentioned above were recorded for the sample modified with three-weight percentage of molybdenum oxide. A higher oxide amount did not lead to improved electrochemical results, except cycling performance. From EIS spectra, it was confirmed that the lowest slope of LMO/5 wt.% MoO3 provided the lowest lithium ions diffusion, lower even than pristine sample, owing the highest value of charge-transfer resistance (265.1 Ω).
Lithium manganese oxide thin films coated by manganese oxide exhibited an outstanding life cycle and only 20% capacity fade was calculated after 3000 cycles [210]. The bixbyite-type Mn2O3 coating layer was deposited on the LMO/Pt substrate by a spin-coating technique, followed by an annealing treatment at 750 °C for 6 h. The relatively large thickness of Mn2O3 porous layer (~600 nm) did not block the lithium ions during tests, and a slightly improved discharge capacity (~3 mAh·g−1) was recorded for coated sample in comparison pristine LMO. In order to demonstrate the efficiency of Mn2O3 layer against the electrolyte attack over the LMO modified electrode, pristine and coated LMO were immersed in 1 M LiPF6 electrolyte solution for seven days. The Mn2O3/LiMn2O4 electrode barely showed dissolution proofs, while the pristine LMO completely dissolved. While the Mn3+ ions inside LiMn2O4 spinel structure are exposed for reduction to Mn2+, fact that hinders material’s performance, it was determined that Mn3+ ions from bixbyite structure are more stable.
Recently, Yao et al. [211] suggested that tert-butanol addition during the synthesis of Al2O3 coated LMO enhanced the stability and electrochemical properties of the sample. Due to the fact that Al2O3 is inactive in the range between 3.0–4 V, the LMO-Al2O3 composite delivered a lower specific discharge capacity in the first cycle (105.46 mAh·g−1, c = 0.2 C), in comparison with the bare lithium manganese spinel (114.69 mAh·g−1), but the elevated temperature (55 °C) test proved a better capacity retention (90.78%). This cyclability improvement is related to the protective coating layer, which is strongly attached to the LMO particles. Al3+ has the role of an anchor ion between polymeric functional groups formed by tert-butanol’s hydroxyl groups and the hydrogen bonds from the aluminum salt precursor (aluminum sulfate octadecahydrate). In addition, the ageing tests showed less manganese content in LMO-Al2O3 sample (62.3 ppm) compared to bare LMO (218.5 ppm). For the ageing tests, the samples were immersed in electrolyte for 20 days at a temperature of 55 °C.
Al2O3-coated LiMn2O4 have been synthesized by a co-precipitation method [212]. In contrast with the previous reported LMO-Al2O3 composite, the current modified sample showed a higher capacity in the first cycle (133.5 mAh g−1, c = 0.5 C) and the recorded capacity faded with only 0.014% per cycle. Al2O3 uniform nano-coating can be considered efficient to prevent the apparition of Jahn-Teller structural distortion and also to hinder the electrolyte attack over the active LMO particles, because of the lower specific area (2.3 m2 g−1), four times lower than for pristine LMO (13.6 m2·g−1). In addition, Al2O3 strengthened the high discharge rate performance, at 5 C (55 °C), thus, the 100th cycle capacity of LMO-Al2O3 remained ~92% of the initial one (~108 mAh·g−1), much better than the bare spinel (73.5% capacity retention and ~90 mAh·g−1 in the first cycle).
In another research article [213], Al2O3 was coated on pre-sintered and calcined LiMn2O4, through a wet process, followed by heat treatment at 750 and 350 °C for 10 h, respectively. During the electrochemical test, it was noticed an improved cycle life for Al2O3 pre-coated LMO compared to other samples. The Al2O3 layer acted like an inhibitor layer and prevented large growth of LiMn2O4 particles, thus the smaller sized material owned better kinetics during tests. Moreover, during ageing tests, 0.05 mg of manganese from the Al2O3 pre-coated LMO dissolved into in electrolyte, which was four times lower than for bare spinel and with 25% less than Al2O3-coated LMO.
Guo et al. [214] have reported lithium manganese oxide coated by a wet process implying electrostatic attraction forces method. By adjusting the pH of a mixture of water and ethanol to 8, a Zeta potential difference of 44 mV has been recorded. After that, LMO spinel powder was added, while nano-silica sol was poured drop wise. The resulted sol was heated at 90 °C for 10 h and then calcined at 700 °C for 6 h. Different weight percentages of SiO2 were tested, but the best results were measured for 2 wt.% SiO2-LMO. The XRD patterns of coated samples were slightly shifted to the left, which can be attributed to a possible replacement of Mn4+ from 16d sites by Si4+ ions. During the cell’s cycling at room temperature (25 °C), 2 wt.% SiO2-LMO retained 97.6% of the initial capacity (101.2 mAh·g−1) after 100 cycles. Even though a low-capacity loss of only 2.4% was calculated, the achieved discharge capacity was too low for a possible practical application (only 68.73% of LMO’s theoretical capacity), whereas higher discharge capacities were recorded for LMO coated with CeO2 [197], Li2ZrO3 [215] or AlP [30]. SiO2 had benefic effect over the LiMn2O4 performance at elevated temperature (55 °C), which is considered a critical parameter for spinel. Moreover, silicon oxide coating reduced the amount of dissolved manganese into the electrolyte by 36% and increased the lithium diffusion coefficient by one magnitude order, which determined faster lithium ions migration throughout the electrochemical tests. Another improvement brought by the SiO2 coating was related to reducing the Rct increasing rate after 100 cycles, thus, in the case of surface modified sample, Rct increased by ~8 times (from 125.9 to 970.1 Ω), while bare sample’s Rct value dramatically raised by almost 20 times, reaching an enormous value of 11,800 Ω.

3.3. Other Surface Changes

Coatings that are more complex were studied in the last years (Table 10). Metal phosphates [216,217,218] and fluorides [163] were among the most investigated coatings. Among other possible coatings, layered material LiCoO2 [219], spinel type material (LiNi0.5Mn1.5O4) [1] or pervoskite-type oxide [220] were also tested.
Yang et al. [217] studied iron phosphate as a coating candidate for enhancing lithium manganese oxide’s performances. It was observed that, during the tests at room temperature, a low amount of FePO4 (1%) was able to increase the capacity retention of the pristine sample by almost 20%, from 75% to 93.5%. Unfortunately, at elevated temperature (60 °C), 1 wt.% FePO4 was not sufficient to diminish the structural modifications of the material, but by adding a double amount of iron phosphate, the modified sample was more stable and lesser capacity loss was calculated. The quantity of 1 wt.% FePO4 reduced the charge-transfer resistance, there, lithium ions movement was greatly increased.
Amorphous iron phosphate proved to be a promising coating material for LiMn2O4 due to its 3D channels that enhance the lithium ions diffusion during cycles and by stabilizing the LMO structure by a small amount of iron ions incorporation in the spinel’s lattice [229]. The heat treatment during the coating process was the key parameter for obtaining the right FePO4 porous structure. The samples were calcined at different temperatures, in the range of 300–700 °C. The lowest calcination temperatures favored the formation of porous structure, while when the temperature increased, the FePO4 become crystalline and less defects were noticed. In terms of electrochemical performances, the highest lithium diffusion coefficient (9.85·10−10 cm2·s−1) was calculated for the 3 wt.% FePO4-LMO sintered at 400 °C and DLi decreased as the calcination temperature increased.
Han et al. [1] studied the composite material resulted from coating spinel LiNi0.5Mn1.5O4 (LNMO) on LiMn2O4 using a sol-gel process. Although, a coating with a similar structure as the substrate material is not very common, LNMO influenced the capacity retention of the pristine sample, especially at high temperature. While the bare LMO exhibit 82.8 mAh·g−1 (66.7% of its initial discharge capacity) after 40 cycles, the LNMO/LMO composite had 98.6 mAh·g−1 (~90% of the first cycle capacity). Changes in the structure of LMO cycled at 60 °C were noticed and peaks corresponding to tetragonal Li2M2O4 were identified after analyzing the sample by XRD. No similar peaks were observed for the modified sample, fact that proved the efficiency of 10% LNMO to suppress the Jahn-Teller distortion. The LiNi0.5Mn1.5O4 structure, which greatly favors the Li+ diffusion, was an important factor in reducing the charge-transfer resistance of the composite.
LNMO/LMO composite material has been synthesized trough a citric acid sol-gel route [227]. The core-shell material presented an initial lower discharge capacity (101 mAh·g−1) and only after five cycles, the structure has been stabilized along with the formation of LNMO’s SEI layer. Charge/discharge tests at different rates (from 1 to 65 C), demonstrated the high-rate ability of the modified sample, due to its discharge capacity of approximately 25 mAh·g−1 at 65 C. In comparison, the pristine sample exhibited a similar capacity, but for a lower discharge rate (45 C).
LiCoO2 established as a good cathode material since the appearance of the Li-ion batteries. In the last years, it was suggested as a potential coating material for spinel type LMO, in order to improve the high-temperature behavior of the latter [219]. Commercial LMO sample was modified with LiCoO2 throughout a sol-gel process, employing a two-stage heat treatment, at 350 and 650 °C, respectively. The initial discharge capacities of LiCoO2-LMO sample were higher than the ones for unmodified sample, both at room and elevated temperature. The reason for this aspect is related to the electrochemical activity of the lithium cobalt oxide in the tests’ voltage area (3.4–4.2 V). LiCoO2 improved the capacity retention of LMO by 6%, from 87.5% to 93.6%, after 100 cycles at 55 °C. In addition, the LiCoO2 coating layer acted as a barrier and halved the manganese dissolution into the electrolyte; the dissolved quality of manganese ions (determined by ICP-AES analysis) from LiCoO2-LiMn2O4 electrode was 10.17μg/cm2, in comparison to 22.54 μg/cm2 for pristine LiMn2O4 electrode.
Among the newly studied core-shell material, LMO coated with LiMnPO4 (LMP) indicated suitable performance for log lasting heavy-duty devices [230]. During the test at room temperature, the LMO coated with 3% LMP showed good capacity retention of ~96% after 100 cycles. 3 wt.% LMP was not enough to suppress the structural modifications, which occurred at elevated temperature (55 °C), thus high-capacity fade (30%) was recorded after 500 cycles. The LMP layer notably improved LiMn2O4′s cyclic performance, especially at high discharge rates of 10 and 20 C, respectively. After long time cycling (1000 cycles) at 20 C, the discharge capacity for 3 wt.% LMP (76.3 mAh·g−1) was 1.6 times higher than for the pristine sample (45.8 mAh·g−1). As a drawback, due to low electronic conductivity of LiMnPO4 material, the modified samples presented a lower DLi (5.32 × 10−12 cm2·s−1) than in other reported studies [228,229]. LMP layer did not suppress the disproportionate reaction of Mn3+, but it successfully blocked the dissolution of Mn2+ into the electrolyte.
Feng et al. [30] developed an aluminum phosphide (AlP) coating. An interesting aspect related to aluminum phosphide is its ability to react with the harmful HF, which promote Mn3+ disproportionate reaction, and the resulted product, AlF3, was also successfully tested as coating for LiMn2O4 [95,221].
AlF3 was deposited on already modified LMO by an Al3+ and F co-doping sample via a co-precipitation method [95]. The aluminum fluoride coating protected the LiAl0.04Mn1.96O3.96F0.04 particles from electrolyte action, resulting in a more stable structure during charge-discharge cycles. Moreover, it reduced the charge-transfer resistance growth rate calculated, which was calculated to be from 67 to 197 Ω for pristine LiAl0.04Mn1.96O3.96F0.04, to only 6 Ω (from 87 to 93 Ω) after the same number of cycles (100 cycles). However, a minor improvement related to DLi+ was noticed for AlF3-coated LMO (1.4·10−8 cm2 s−1) in comparison with LiAl0.04Mn1.96O3.96F0.04 (1.2·10−8 cm2 s−1), but the lithium diffusion coefficients of the latter samples was ten times higher than for pristine LiMn2O4 (1.1·10−9 cm2 s−1).
Modified LMO samples were tested with good results in ARLB cells. For example, 2 wt.% LaF3-LiMn2O4 exhibited an insignificant capacity fade after 50 cycles [131]. Due to LaF3′s structure, its large channels enhance the lithium migration to and from the electrode material, while an optimum amount of lanthanum fluoride coating may prevent the side reaction that can occur between LMO and electrolyte. The calculated lithium diffusion coefficient (4.67·10−8 cm2·s−1) was two magnitude higher than the one obtained for iron phosphate modified LiMn2O4 in organic electrolyte [229] and similar to DLi calculated for La-Sr-Mn-O coating [130]. The protective LaF3 layer acted like a barrier and protected the electrode surface, by reducing the contact area between LMO and aqueous electrolyte; therefore, the manganese dissolution into LiNO3 was barely noticed.
Solid electrolytes, like Li3BO3 [223] or La-Sr-Mn-O (LSM) [130], with good ionic conductivity demonstrated to be adequate coating materials for enhancing the cycle ability of LiMn2O4. The higher ionic conductivity of LSM thin uniform layer played an essential role in the higher initial discharge capacity (129.9 mAh·g−1, 0.1 C) and promoted a capacity retention (90.6%) with up to 50% higher than the pristine sample. LSM coating decreased the charge-transfer resistance, enhanced the lithium-ion diffusion and the calculated diffusion coefficient for lithium ions had the same magnitude order as for LaF3-coated LiMn2O4 tested in an ARLB [132]. Another improvement brought by the LSM was related to electrical conductivity (8.42·10−4 S/cm), which was larger than for pristine sample (5.31·10−5 S/cm).
An isostructured Li2CuO2-Li2NiO2 solid solution has been considered as a coating for lithium manganese oxide [166]. During the electrochemical investigations at room temperature, it was demonstrated that a low amount of solid solution (0.5 wt.%) was enough to increase the rate capability (93.2% capacity retention after 300 cycles) and to maintain a reasonable discharge capacity (113 mAh·g−1 in the first cycle). However, taking in consideration that the elevated temperature performance is a key demand, it was noticed that 0.5 wt.% was not enough to stabilize the LiMn2O4 at 55 °C, thus the capacity retention was only 81.2% after 200 cycles. In comparison, 2% w/w Li2CuO2-Li2NiO2 solid solution exhibited an improved electrochemical performance. By having higher amount of coating, the contact area between the electrolyte and LiMn2O4 was reduced, so the manganese dissolution was hindered.
Multiple doped LiCo0.025Cr0.025Ni0.025Fe0.025Mn1.90O4 and pristine LiMn2O4 were coated by lithium borosilicate (LBS) through a facile solution method [143,232]. LBS-based glass was proposed as a coating due to its improved ionic conductivity. Surface modified samples exhibited lower initial discharge capacity (109.7 mAh·g−1) than pristine LMO, because the coating layer slightly decreased the lithium diffusion. However, LBS reduced the contact surface between the electrode active material (LMO) with the electrolyte (1 M LiPF6 solution), and superior capacity retention (93.3%) was observed after 70 cycles (25 °C). At elevated temperature (55 °C), LBS-coated LiMn2O4 performed better than bare LMO, showing a higher initial discharge capacity (120 mAh·g−1 vs. 104.2 mAh·g−1). In addition, the capacity fade per cycle was slightly lower for coated LMO spinel (0.708 mAh·g−1 cycle−1 vs. 0.728 mAg·g−1 cycle−1).
Perovskite-type LaMnO3 was successfully deposited on lithium manganese oxide, without modifying the spinel structure [37]. The improved electronic conductivity of LaMnO3 determined a higher initial discharge capacity of the modified sample (114 mAh g−1) and better coulombic efficiency (95%) than for pristine lithium manganese oxide (106 mAh g−1 and 89.1%).
Titanium carbide (Ti3C2Tx) nanosheets, a MXene 2D layered structured material, has been recently used as a protective coating layer for LMO particles by their encapsulation through facile electrostatic self-assembly method. Similar synthesizes methods can be using for other cathode materials like LiCoO2 or LiFePO4 [233]. MXene layered material, with the general formula of Mn+1AXn (n = 1, 2, 3), where M is a transition metal (Sc, Ti, Zr, Nb, Mo), A corresponds to elements from the 13th or 14th groups (Al, Si, for example) and X represents either C or N atoms. Namely, in the case of titanium carbide, Tx is attributed to functional group –OH and atoms like –O and –F [233,234]. Ti3C2Tx is obtained by selective removal of Al from Ti3AlC2 in dilute HF solution media. Details about the synthesis procedure of titanium carbide along with other aspects related to MXene materials can be found in Ref [235]. Three samples were tested: pristine LMO, Ti3C2Tx-LMO (obtained using CTAB as decorator for charging the surface of LMO) and Ti3C2Tx-LiMn2O4 (synthesized in similar conditions but replacing CTAB with deionized H2O). During the electrochemical tests, at both room temperature and 55 °C, Ti3C2Tx–coated LMO (with CTAB) performed better than the other two tested samples and delivered higher specific capacity of 114.1 mAh·g−1 (c = 1 C, RT) and 115.9 mAh g−1 (c = 1 C, 55 °C), respectively. In addition, the capacity fade was not as broad as in other case, and at 55 °C, Ti3C2Tx–coated LMO (with CTAB) showed a discharge capacity with more than 18% than the pristine sample (63%), after 200 cycles. Titanium carbide layered hampered the manganese dissolution into the electrolyte by reducing the contact area between the working electrode and electrolyte. By coating the pristine LMO with Ti3C2Tx, which possesses low Li+ diffusion barrier (0.007 eV), good electronic conductivity and high Li+ diffusion coefficient, determined lower charge-transfer resistance (133.4 Ω) than 370.7 Ω for Ti3C2Tx-LMO (deionized water) and 552.6 Ω for bare spinel (at room temperature).
Li4Ti5O12(LTO) is a characteristic material which can be used as anode in Li-ion batteries [236,237], but, lately, it was successfully employed as a coating layer for enhancing the performance of LiMn2O4 [237]. LTO–LMO nanorods were manufactured by a three-step synthesis method: in the first stage, β-MnO2 were obtained by a hydrothermal process, followed a solid stage reactions method, when as-synthesized β-MnO2 were mixed with lithium hydroxide and calcined to result lithium manganese oxide. Afterwards, the composite material resulted through a sol-gel process, where a key parameter was to control the hydrolysis and condensation of titanium precursor (titanium tetrabutoxide). LTO-coated LMO was electrochemically tested for a long number of cycles (500 cycles, c = 1 C), at room (25 °C) and elevated temperature (55°), respectively. In the first cycles, the tested working electrode provided 119.7 mAh·g−1 and a capacity of 17.8% was recorded after 500 cycles, while an initial discharge of 119.2 mAh·g−1 and a 74.8% capacity retention was calculated after an identical number of cycles, at 55 °C. The results obtained by Li4Ti5O12 overcome other coatings like Li3PO4 [222], LiMnPO4 [230] or Li2CuO2-Li2NiO2 [166], but performed slightly worse than Li2MnO3 [226] or La-Sr-Mn-O [130]. Despite the fact that lithium titanium oxide improved the cycling ability of the LMO by preventing manganese solution, no significant enhancement was observed in the case of lithium diffusion coefficient (5.33 × 10−12 m2 s−1), which was lower than in other cases: La0.7Sr0.3Mn0.7Co0.3O3 [220], FePO4 [229], La-Sr-Mn-O [130] or LaF3 [131]. However, electrochemical impedance spectra demonstrated that LTO helped lithium diffusion and composite’s Rct did not increased so much (from 6.58 to 54.2 Ω) after 100 cycles, as in the case of pristine sample (from 8.45 to 156.27 Ω).
Uniform thin layer of hexafluorotitanic lithium (Li2TiF6) with high lithium conductivity has been deposited on commercial LiMn2O4 by a co-precipitation method, using precursors like lithium carbonate (Li2CO3) and hexafluorotitanic acid (H2TiF6) [238]. Low amount of hexafluorotitanic lithium was enough to stabilize and improve the lifetime of LMO. It was showed that high weight amount (10 wt.% Li2TiF6) of coating determined the formation of a thick coating layer. It managed to constrain capacity fade, although the delivered discharge capacity was not sufficient. Usually, the coatings do not have excellent lithium conductivity and this drawback affects the performance of the working electrode at high discharge rates, because the adequate lithium diffusion paths are available. However, Li2TiF6 improved the high discharge rate performance of bare spinel; thus, the specific capacity for composite material was 99.4 mAh·g−1 in the 200th cycle, much greater than 65.7 mAh·g−1 for pristine LMO. The results were superior to the ones obtained by La0.7Sr0.3Mn0.7Co0.3O3-coated lithium manganese spinel [220].

4. Future Prospects

Nickel-doped (LiNi0.5Mn1.5O4) lithium manganese oxide is considered as a suitable cathode material for electronic devices and electrical cars, which demand high operating voltage Li-ion batteries [56,239]. Depending on the chosen annealing temperature, LiNi0.5Mn1.5O4 exhibits two type of structures: face-centered spinel (Fd3m) and primitive simple cubic crystal (P4332) [59,240]. Spinel structure is more stable and shows better electrochemical characteristics, which recommends to carefully selecting the synthesis route for obtaining LNMO spinel-type materials [240].
The working mechanism during charge/discharge cycles for a pristine LNMO includes three voltage plateaus, one at 4–4.1 V, which corresponds to Mn3+/Mn4+ redox couple and the other two around 4.75 V, which correspond to Ni2+/Ni3+ and Ni3+/Ni4+ couples. The low voltage plateau is less visible when the content of Mn4+ and Ni2+ is higher [59,240].
The improvement of LNMO’s performance during battery functioning includes doping, surface modification or the usage of a proper electrolyte. Among the doping ions, the insertion of chromium slightly increased the voltage, up to 4.9 V, because of the Cr3+/Cr4+ redox couple [59,70]. The different ways to overcome LNMO’s drawbacks are described in other studies [59,240].

5. Conclusions

Lithium manganese oxide exhibits two main drawbacks, which hinder its performance during LIB cycling: Jahn-Teller distortion [241,242,243] and manganese dissolution into electrolyte solution [164,244]. In order to hinder these two disadvantages and stabilize the structure of LMO, two main strategies have been adopted: structure doping and particle coating.
The dopants can stabilize LMO structure, by decreasing the amount of Mn3+ content, which is highly responsible for disproportion during charge/discharge cycles. However, by reducing the amount of electrochemical active Mn3+ species, the discharge capacity decreased with several percentages, but the performance of doped material has been improved, by retaining a higher capacity after a certain number of cycles.
Among the studied cations dopants, aluminum is one of the most studied. Aluminum-doped LMO spinel performs well both in aqueous and organic based electrolyte. In aqueous electrolyte, the calculated capacity retention of LiAl0.1Mn1.9O4 was 70% after an extremely long number of cycles (more than 4000 cycles) [93]. In comparison, Al-doped LMO managed to retain more than 80% of its initial discharge capacity in a cell with LiPF6 based electrolyte after 1000 charge-discharge cycles [92].
Multiple doping with low amounts of cations and anions, like Mg and F [78] or Al and S [154], led to materials capable to provide adequate discharge capacity for long cycle life.
The anions did not reduce the quantity of Mn3+, therefore the anion-doped LMO exhibited higher capacity, even than pristine sample [45].
Coating materials lessen the contact between the electrolyte and active material, thus the cycle life ability has been longer [163]. For example, TiO2 coated on lithium manganese oxide may react with HF, resulted when LiPF6 accidentally reacts with moisture, to form a protective layer, which prevents the manganese dissolution [205]. Aluminum sulfate octadecahydrate used as a precursor for aluminum oxide coating layer for LMO helped Al2O3 to be strongly bonded to LMO by hydrogen bonds, having Al as an anchor atom [211].
Another improvement provided by coatings is related to enhanced lithium diffusion, by reducing the charge-transfer resistance due to better electronic conductivity. In addition, by using a carbon-based support material, one can achieve even a greater discharge capacity than the theoretical capacity of LMO, because of high concentration of lithium ions into carbon nanotubes network [168].

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

This research did not receive any specific grant from funding public, governmental or non-profit agencies, nor from private/commercial sectors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Han, Z.; Jia, X.; Zhan, H.; Zhou, Y. LiMn2O4/LiNi0.5Mn1.5O4 composite with improved electrochemical property. Electrochim. Acta 2013, 114, 772–778. [Google Scholar] [CrossRef]
  2. Li, C.; Zhang, H.; Fu, L.; Liu, H.; Wu, Y.; Rahm, E.; Holze, R.; Wu, H. Cathode materials modified by surface coating for lithium ion batteries. Electrochim. Acta 2006, 51, 3872–3883. [Google Scholar] [CrossRef]
  3. Zhao, Q.; Wu, Y.; Ma, X.; Wang, R.; Xu, X.; Cao, C. Mn oxidation state controllable spinel manganese-based intergrown cathode for excellent reversible lithium storage. J. Power Source 2017, 359, 295–302. [Google Scholar] [CrossRef]
  4. Li, X.; Shao, Z.; Liu, K.; Liu, G.; Xu, B. Synthesis and electrochemical characterizations of LiMn2O4 prepared by high temperature ball milling combustion method with citric acid as fuel. J. Electroanal. Chem. 2018, 818, 204–209. [Google Scholar] [CrossRef]
  5. Wang, Y.; Liu, B.; Li, Q.; Cartmell, S.; Ferrara, S.; Deng, Z.D.; Xiao, J. Lithium and lithium ion batteries for applications in microelectronic devices: A review. J. Power Source 2015, 286, 330–345. [Google Scholar] [CrossRef] [Green Version]
  6. Dai, X.; Zhou, A.; Xu, J.; Lu, Y.; Wang, L.; Fan, C.; Li, J. Extending the high-voltage capacity of LiCoO2 cathode by direct coating of the composite electrode with Li2CO3 via magnetron sputtering. J. Phys. Chem. C 2015, 120, 422–430. [Google Scholar] [CrossRef]
  7. Wang, G.; Qu, Q.; Wang, B.; Shi, Y.; Tian, S.; Wu, Y.; Holze, R. Electrochemical behavior of LiCoO2 in a saturated aqueous Li2SO4 solution. Electrochim. Acta 2009, 54, 1199–1203. [Google Scholar] [CrossRef]
  8. Chen, J.-M.; Cho, Y.-D.; Hsiao, C.-L.; Fey, G.T.-K. Electrochemical studies on LiCoO2 surface coated with Y3Al5O12 for lithium-ion cells. J. Power Source 2009, 189, 279–287. [Google Scholar] [CrossRef]
  9. Satyavani, T.; Kumar, A.S.; Rao, P.S. Methods of synthesis and performance improvement of lithium iron phosphate for high rate Li-ion batteries: A review. Eng. Sci. Technol. Int. J. 2016, 19, 178–188. [Google Scholar] [CrossRef] [Green Version]
  10. Hou, Y.; Wang, X.; Zhu, Y.; Hu, C.; Chang, Z.; Wu, Y.; Holze, R. Macroporous LiFePO4 as a cathode for an aqueous rechargeable lithium battery of high energy density. J. Mater. Chem. A 2013, 1, 14713. [Google Scholar] [CrossRef]
  11. Li, Z.; Zhang, D.; Yang, F. Developments of lithium-ion batteries and challenges of LiFePO4 as one promising cathode material. J. Mater. Sci. 2009, 44, 2435–2443. [Google Scholar] [CrossRef]
  12. Yu, H.; Dong, X.; Pang, Y.; Wang, Y.; Xia, Y. High power lithium-ion battery based on spinel cathode and hard carbon anode. Electrochim. Acta 2017, 228, 251–258. [Google Scholar] [CrossRef]
  13. Liu, Q.; Wang, S.; Tan, H.; Yang, Z.; Zeng, J. Preparation and doping mode of doped LiMn2O4 for Li-ion batteries. Energies 2013, 6, 1718–1730. [Google Scholar] [CrossRef]
  14. Lv, W.; Li, Z.; Deng, Y.; Yang, Q.-H.; Kang, F. Graphene-based materials for electrochemical energy storage devices: Opportunities and challenges. Energy Storage Mater. 2016, 2, 107–138. [Google Scholar] [CrossRef]
  15. Ozanam, F.; Rosso, M. Silicon as anode material for Li-ion batteries. Mater. Sci. Eng. B 2016, 213, 2–11. [Google Scholar] [CrossRef]
  16. Dash, R.; Pannala, S. The potential of silicon anode based lithium ion batteries. Mater. Today 2016, 19, 483–484. [Google Scholar] [CrossRef]
  17. Wang, X.F.; Liu, B.; Hou, X.J.; Wang, Q.F.; Li, W.W.; Chen, D.; Shen, G.Z. Ultralong-life and high-rate web-like Li4Ti5O12 anode for high-performance flexible lithium-ion batteries. Nano Res. 2014, 7, 1073–1082. [Google Scholar] [CrossRef]
  18. Li, Q.; Chen, J.; Fan, L.; Kong, X.; Lu, Y. Progress in electrolytes for rechargeable Li-based batteries and beyond. Green Energy Environ. 2016, 1, 18–42. [Google Scholar] [CrossRef] [Green Version]
  19. Michalska, M.; Ziolkowska, D.A.; Jasinski, J.B.; Lee, P.H.; Lawniczak, P.; Andrzejewski, B.; Ostrowski, A.; Bednarski, W.; Wu, S.H.; Lin, J.Y. Improved electrochemical performance of LiMn2O4 cathode material by Ce doping. Electrochim. Acta 2018, 276, 37–46. [Google Scholar] [CrossRef]
  20. Kiani, M.A.; Mousavi, M.F.; Rahmanifar, M.S. Synthesis of nano- and micro-particles of LiMn2O4: Electrochemical investigation and assessment as a cathode in Li battery. Int. J. Electrochem. Sci. 2011, 6, 2581–2595. [Google Scholar]
  21. Han, C.-G.; Zhu, C.; Saito, G.; Akiyama, T. Improved electrochemical properties of LiMn2O4with the Bi and La co-doping for lithium-ion batteries. RSC Adv. 2015, 5, 73315–73322. [Google Scholar] [CrossRef]
  22. Wang, J.-G.; Jin, D.; Liu, H.; Zhang, C.; Zhou, R.; Shen, C.; Xie, K.; Wei, B. All-manganese-based Li-ion batteries with high rate capability and ultralong cycle life. Nano Energy 2016, 22, 524–532. [Google Scholar] [CrossRef]
  23. Thackeray, M.M.; Picciotto, L.A.; de Kock, A.; Johnson, P.J.; Nicholas, V.A.; Adendorff, K.T. Spinel electrodes for Lithium batteries—A review. J. Power Source 1987, 21, 1–8. [Google Scholar] [CrossRef]
  24. Thackeray, M.M.; David, W.I.F.; Bruce, P.G.; Goodenough, J.B. Lithium insertion into manganese spinels. Mater. Res. Bull. 1983, 18, 461–472. [Google Scholar] [CrossRef]
  25. Zuo, D.; Tian, G.; Li, X.; Chen, D.; Shu, K. Recent progress in surface coating of cathode materials for lithium ion secondary batteries. J. Alloys Compd. 2017, 706, 24–40. [Google Scholar] [CrossRef]
  26. Peng, Z.; Li, Y.; Du, K.; Cao, Y.; Hu, G. Improved elevated temperature performance of spinel LiMn2O4 via surface-modified by Li-rich Li1.2Ni0.2Mn0.6O2 for lithium-ion batteries. J. Alloys Compd. 2017, 728, 1209–1216. [Google Scholar] [CrossRef]
  27. He, X.; Li, J.; Cai, Y.; Wang, Y.; Ying, J.; Jiang, C.; Wan, C. Preparation of co-doped spherical spinel LiMn2O4 cathode materials for Li-ion batteries. J. Power Source 2005, 150, 216–222. [Google Scholar] [CrossRef]
  28. Chung, K.Y.; Ryu, C.-W.; Kim, K.-B. Onset mechanism of Jahn-Teller distortion in 4 V LiMn2O4 and its suppression by LiM0.05Mn1.95O4 (M = Co, Ni) coating. J. Electrochem. Soc. 2005, 152, A791. [Google Scholar] [CrossRef]
  29. Liang, X.; Zeng, S.; Liu, Y.; Shi, L.; Liu, T. Enhance cycling performance of LiMn2O4 cathode by Sr2+ and Cr3+ doping. Mater. Sci. Technol. 2014, 31, 443–447. [Google Scholar] [CrossRef]
  30. Feng, X.; Zhang, J.; Yin, L. Effect of AlP coating on electrochemical properties of LiMn2O4 cathode material for lithium ion battery. Mater. Res. Bull. 2016, 74, 421–424. [Google Scholar] [CrossRef]
  31. Thirunakaran, R.; Sivashanmugam, A.; Gopukumar, S.; Dunnill, C.W.; Gregory, D.H. Electrochemical behaviour of nano-sized spinel LiMn2O4 and LiAlxMn2−xO4 (x = Al: 0.00–0.40) synthesized via fumaric acid-assisted sol-gel synthesis for use in lithium rechargeable batteries. J. Phys. Chem. Solids 2008, 69, 2082–2090. [Google Scholar] [CrossRef]
  32. Wang, M.; Yang, M.; Zhao, X.; Ma, L.; Shen, X.; Cao, G. Spinel LiMn2−xSiO4 (x < 1) through Si4+ substitution as a potential cathode material for lithium-ion batteries. Sci. China Mater. 2016, 59, 558–566. [Google Scholar]
  33. Jiang, Q.Q.; Liu, D.D.; Zhang, H.; Wang, S.Y. Plasma-assisted sulfur doping of LiMn2O4 for high-performance lithium-ion batteries. J. Phys. Chem. C 2015, 119, 28776–28782. [Google Scholar] [CrossRef]
  34. Xiang, M.W.; Zhou, X.Y.; Zhang, Z.F.; Chen, M.M.; Bai, H.L.; Guo, J.M. LiMn2O4 prepared by liquid phase flameless combustion with F-doped for lithium-ion battery cathode materials. In Advances in Materials and Materials Processing; Jiang, Z.Y., Liu, X.H., Jiao, S.H., Han, J.T., Eds.; Trans Tech Publications Ltd.: Stafa-Zurich, Switzerland, 2013; pp. 825–830. [Google Scholar]
  35. Thirunakaran, R.; Ravikumar, R.; Vanitha, S.; Gopukumar, S.; Sivashanmugam, A. Glutamic acid-assisted sol-gel synthesis of multi-doped spinel lithium manganate as cathode materials for lithium rechargeable batteries. Electrochim. Acta 2011, 58, 348–358. [Google Scholar] [CrossRef]
  36. Bao, S.J.; Liang, Y.Y.; Zhou, W.J.; He, B.L.; Li, H.L. Enhancement of the electrochemical properties of LiMn2O4 through Al3+ and F-co-substitution. J. Colloid Interface Sci. 2005, 291, 433–437. [Google Scholar] [CrossRef]
  37. Wang, C.-M.; Jin, F.-M.; Shi, T.; Chen, L. The effect of LaMnO3 with high electronic conductivity on the high rate charge-discharge performance of LiMn2O. J. Electroanal. Chem. 2016, 775, 306–310. [Google Scholar] [CrossRef]
  38. Park, K.; Park, J.-H.; Hong, S.-G.; Yoon, J.; Park, S.; Kim, J.-H.; Yoon, D.; Kim, H.; Son, Y.-H.; Park, J.-H.; et al. Induced AlF3 segregation for the generation of reciprocal Al2O3 and LiF coating layer on self-generated LiMn2O4 surface of over-lithiated oxide based Li-ion battery. Electrochim. Acta 2016, 222, 830–837. [Google Scholar] [CrossRef]
  39. Zhang, Z.J.; Chou, S.L.; Gu, Q.F.; Liu, H.K.; Li, H.J.; Ozawa, K.; Wang, J.Z. Enhancing the high rate capability and cycling stability of LiMn2O4 by coating of solid-state electrolyte LiNbO3. ACS Appl. Mater. Interfaces 2014, 6, 22155–22165. [Google Scholar] [CrossRef] [PubMed]
  40. Fang, D.-L.; Li, J.-C.; Liu, X.; Huang, P.-F.; Xu, T.-R.; Qian, M.-C.; Zheng, C.-H. Synthesis of a Co-Ni doped LiMn2O4 spinel cathode material for high-power Li-ion batteries by a sol-gel mediated solid-state route. J. Alloys Compd. 2015, 640, 82–89. [Google Scholar] [CrossRef]
  41. Yi, Z. Rheological phase reaction synthesis of Co-doped LiMn2O4 octahedral particles. J. Mater. Sci. Mater. Electron. 2016, 27, 10347–10352. [Google Scholar] [CrossRef]
  42. Li, S.Y.; Zhu, K.L.; Du, S.L. Enhanced elevated-temperature performance of Al-doped LiMn2O4 as cathodes for lithium ion batteries. In Proceedings of the 2nd International Conference on Materials Science, Resource and Environmental Engineering, Wuhan, China, 27–29 October 2017. [Google Scholar] [CrossRef]
  43. Zhan, D.; Liang, Y.; Cui, P.; Xiao, Z.A. Al-doped LiMn2O4 single crystalline nanorods with enhanced elevated-temperature electrochemical performance via a template-engaged method as a cathode material for lithium ion batteries. RSC Adv. 2015, 5, 6372–6377. [Google Scholar] [CrossRef]
  44. Bakierska, M.; Świętosławski, M.; Chudzik, K.; Lis, M.; Molenda, M. Enhancing the lithium ion diffusivity in LiMn2O4−ySy cathode materials through potassium doping. Solid State Ion. 2018, 317, 190–193. [Google Scholar] [CrossRef]
  45. Molenda, M.; Bakierska, M.; Majda, D.; Świętosławski, M.; Dziembaj, R. Structural and electrochemical characterization of sulphur-doped lithium manganese spinel cathode materials for lithium ion batteries. Solid State Ion. 2015, 272, 127–132. [Google Scholar] [CrossRef]
  46. Sun, Y.-K.; Jeon, Y.; Leeb, H.J. Overcoming Jahn-Teller Distortion for Spinel Mn Phase. Electrochem. Solid-State Lett. 1999, 3, 7–9. [Google Scholar] [CrossRef]
  47. Nkosi, F.P.; Jafta, C.J.; Kebede, M.; le Roux, L.; Mathe, M.K.; Ozoemena, K.I. Microwave-assisted optimization of the manganese redox states for enhanced capacity and capacity retention of LiAlxMn2−xO4(x = 0 and 0.3) spinel materials. RSC Adv. 2015, 5, 32256–32262. [Google Scholar] [CrossRef] [Green Version]
  48. Liu, J.; Li, G.; Yu, Y.; Bai, H.; Shao, M.; Guo, J.; Su, C.; Liu, X.; Bai, W. Synthesis and electrochemical performance evaluations of polyhedra spinel LiAlxMn2−xO4 (x ≦ 0.20) cathode materials prepared by a solution combustion technique. J. Alloys Compd. 2017, 728, 1315–1328. [Google Scholar] [CrossRef]
  49. Waller, G.; Brooke, P.; Rainwater, B.; Lai, S.; Hu, R.; Ding, Y.; Alamgir, F.; Sandhage, K.; Liu, M. Structure and surface chemistry of Al2O3 coated LiMn2O4 nanostructured electrodes with improved lifetime. J. Power Source 2016, 306, 162–170. [Google Scholar] [CrossRef]
  50. Marincaş, A.-H.; Goga, F.; Dorneanu, S.-A.; Ilea, P. Review on synthesis methods to obtain LiMn2O4-based cathode materials for Li-ion batteries. J. Solid State Electrochem. 2020, 24, 473–497. [Google Scholar] [CrossRef]
  51. Kebede, M.A.; Phasha, M.J.; Kunjuzwa, N.; le Roux, L.J.; Mkhonto, D.; Ozoemena, K.I.; Mathe, M.K. Structural and electrochemical properties of aluminium doped LiMn2O4 cathode materials for Li battery: Experimental and ab initio calculations. Sustain. Energy Technol. Assess. 2014, 5, 44–49. [Google Scholar] [CrossRef]
  52. Liu, H.; Tian, R.; Jiang, Y.; Tan, X.; Chen, J.; Zhang, L.; Guo, Y.; Wang, H.; Sun, L.; Chu, W. On the drastically improved performance of Fe-doped LiMn2O4 nanoparticles prepared by a facile solution-gelation route. Electrochim. Acta 2015, 180, 138–146. [Google Scholar] [CrossRef]
  53. Zhang, L.-X.; Wang, Y.-Z.; Jiu, H.-F.; Wang, Y.-L.; Sun, Y.-X.; Li, Z. Controllable synthesis of Co-doped spinel LiMn2O4 nanotubes as cathodes for Li-ion batteries. Electron. Mater. Lett. 2014, 10, 439–444. [Google Scholar] [CrossRef]
  54. Guo, H.-J.; Li, X.-Q.; He, F.-Y.; Li, X.-H.; Wang, Z.-X.; Peng, W.-J. Effects of sodium substitution on properties of LiMn2O4 cathode for lithium ion batteries. Trans. Nonferrous Met. Soc. China 2010, 20, 1043–1048. [Google Scholar] [CrossRef]
  55. Singh, P.; Sil, A.; Nath, M.; Ray, S. Phase stability and charge capacity of cubic spinel in Li1+δMn2−δO4 and consequence of magnesium doping. J. Electrochem. Soc. 2010, 157, A259–A266. [Google Scholar] [CrossRef]
  56. Wu, Y.; Zhang, J.; Cao, C.; Khalid, S.; Zhao, Q.; Wang, R.; Butt, F.K. LiNi0.5Mn1.5O4 nano-submicro cubes as high-performance 5 V cathode materials for lithium-ion batteries. Electrochim. Acta 2017, 230, 293–298. [Google Scholar] [CrossRef]
  57. Feng, X.-Y.; Shen, C.; Xiang, H.-F.; Liu, H.-K.; Wu, Y.-C.; Chen, C.-H. High rate capability of 5 V LiNi0.5Mn1.5O4 cathode material synthesized via a microwave assist method. J. Alloys Compd. 2017, 695, 227–232. [Google Scholar] [CrossRef] [Green Version]
  58. Sha, O.; Wang, S.; Qiao, Z.; Yuan, W.; Tang, Z.; Xu, Q.; Su, Y. Synthesis of spinel LiNi0.5Mn1.5O4 cathode material with excellent cycle stability using urea-based sol–gel method. Mater. Lett. 2012, 89, 251–253. [Google Scholar] [CrossRef]
  59. Liu, G.Q.; Wen, L.; Liu, Y.M. Spinel LiNi0.5Mn1.5O4 and its derivatives as cathodes for high-voltage Li-ion batteries. J. Solid State Electrochem. 2010, 14, 2191–2202. [Google Scholar] [CrossRef]
  60. Murali, K.R.; Saravanan, T.; Jayachandran, M. Synthesis and characterization of copper substituted lithium manganate spinels. J. Mater. Sci. Mater. Electron. 2008, 19, 533–537. [Google Scholar] [CrossRef]
  61. Sulochana, A.; Thirunakaran, R.; Sivashanmugam, A.; Gopukumar, S.; Yamaki, J. Sol-gel synthesis of 5 V LiCuxMn2−xO4 as a cathode material for lithium rechargeable batteries. J. Electrochem. Soc. 2008, 155, A206–A210. [Google Scholar] [CrossRef]
  62. Xu, W.; Li, Q.; Guo, J.; Bai, H.; Su, C.-W.; Ruan, R.; Peng, J. Electrochemical evaluation of LiZnxMn2−xO4 (x ≤ 0.10) cathode material synthesized by solution combustion method. Ceram. Int. 2016, 42, 5693–5698. [Google Scholar] [CrossRef]
  63. Rahman, A.; Rusdi, R.; Elong, K.; Kamarulzaman, N. Synthesis and electrochemical behaviour of LiMn2−xZnxO4 Cathode Materials. In Advancement of Materials and Nanotechnology II; Kamarulzaman, N., Wui, W.T., Mohamed, N.S., Rusdi, R., Chayed, N.F., Eds.; Trans Tech Publications Ltd.: Durnten-Zurich, Switzerland, 2012; pp. 199–203. [Google Scholar]
  64. Yi, T.-F.; Zhu, Y.-R.; Hu, X.-G. Structure and electrochemical properties of LiLaxMn2–xO4 cathode material by the ultrasonic-assisted sol-gel method. Int. J. Min. Met. Mater. 2009, 16, 119–123. [Google Scholar] [CrossRef]
  65. Ram, P.; Singhal, R.; Sharma, R.K. Preliminary study of dysprosium doped LiMn2O4 spinel cathode materials. Mater. Today Proc. 2017, 4, 9365–9370. [Google Scholar] [CrossRef]
  66. Balaji, S.; Chandran, T.M.; Mutharasu, D. A study on the influence of dysprosium cation substitution on the structural, morphological, and electrochemical properties of lithium manganese oxide. Ionics 2011, 18, 549–558. [Google Scholar] [CrossRef]
  67. Wu, W.; Qin, X.; Guo, J.; Wang, J.; Yang, H.; Wang, L. Influence of cerium doping on structure and electrochemical properties of LiNi0.5Mn1.5O4 cathode materials. J. Rare Earths 2017, 35, 887–895. [Google Scholar] [CrossRef]
  68. Kebede, M.A.; Phasha, M.J.; Kunjuzwa, N.; Mathe, M.K.; Ozoemena, K.I. Solution-combustion synthesized aluminium-doped spinel (LiAlxMn2−xO4) as a high-performance lithium-ion battery cathode material. Appl. Phys. A 2015, 121, 51–57. [Google Scholar] [CrossRef]
  69. Lee, Y.; Sun, Y.; Ota, S.; Miyashita, T.; Yoshio, M. Preparation and characterization of nano-crystalline LiNi0.5Mn1.5O4 for 5 V cathode material by composite carbonate process. Electrochem. Commun. 2002, 4, 989–994. [Google Scholar] [CrossRef]
  70. Wang, S.; Li, P.; Shao, L.; Wu, K.; Lin, X.; Shui, M.; Long, N.; Wang, D.; Shu, J. Preparation of spinel LiNi0.5Mn1.5O4 and Cr-doped LiNi0.5Mn1.5O4 cathode materials by tartaric acid assisted sol-gel method. Ceram. Int. 2015, 41, 1347–1353. [Google Scholar] [CrossRef]
  71. Li, Y.; Mao, Y.; Zhan, T.; Li, C.; Xiao, S. Effects of aluminium and sodium co-doping on the structural and electrochemical performances of the spinel LiMn2O4 cathode materials. Micro Nano Lett. 2016, 11, 430–432. [Google Scholar] [CrossRef]
  72. Thirunakaran, R.; Kim, T.; Yoon, W.-S. Effects of multi-dopants (Zn and Ho) in stabilizing spinel structure for cathode materials in lithium rechargeable batteries—Novel chelated sol–gel synthesis. Particuology 2016, 24, 87–95. [Google Scholar] [CrossRef]
  73. Thirunakaran, R.; Lew, G.H.; Yoon, W.-S. Synthesis and electrochemical properties of dual doped spinels LiNixAlyMn2−xyO4 via facile novel chelated sol-gel method as possible cathode material for lithium rechargeable batteries. J. Energy Chem. 2017, 26, 101–114. [Google Scholar] [CrossRef]
  74. Thirunakaran, R.; Lew, G.H.; Yoon, W.-S. Cerotic acid assisted sol-gel synthesis and electrochemical performance of double doped spinels (LiCrxMgyMn2−xyO4) as cathode materials for lithium rechargeable batteries. Powder Technol. 2016, 301, 197–210. [Google Scholar] [CrossRef]
  75. Thirunakaran, R.; Lew, G.H.; Yoon, W.-S. Sol-Gel synthesis using novel chelating agent and electrochemical characterization of binary doped LiMn2O4 spinel as cathode material for lithium rechargeable batteries. World J. Nano Sci. Eng. 2016, 6, 1–19. [Google Scholar] [CrossRef] [Green Version]
  76. Thirunakaran, R.; Ravikumar, R.; Gopukumar, S.; Sivashanmugam, A. Electrochemical evaluation of dual-doped LiMn2O4 spinels synthesized via co-precipitation method as cathode material for lithium rechargeable batteries. J. Alloys Compd. 2013, 556, 266–273. [Google Scholar] [CrossRef]
  77. Zeng, X.; Wu, J.; Hu, Q. Effects of magnesium and chlorine co-doping on the structural and electrochemical performance of the spinel LiMn2O4 cathode materials. Micro Nano Lett. 2016, 11, 789–791. [Google Scholar] [CrossRef]
  78. Wen, W.; Ju, B.; Wang, X.; Wu, C.; Shu, H.; Yang, X. Effects of magnesium and fluorine co-doping on the structural and electrochemical performance of the spinel LiMn2O4 cathode materials. Electrochim. Acta 2014, 147, 271–278. [Google Scholar] [CrossRef]
  79. Mao, F.; Guo, W.; Ma, J. Research progress on design strategies, synthesis and performance of LiMn2O4-based cathodes. RSC Adv. 2015, 5, 105248–105258. [Google Scholar] [CrossRef] [Green Version]
  80. Sun, X.Y.; He, G.; Hong, J.H.; He, M.Z.; Li, H.F. Synthesis of Al-doped lithium manganese oxide from Al2O3-Li2CO3-MnO2 ternary system. In Advanced Materials; Bu, J.L., Jiang, Z.Y., Jiao, S., Eds.; Trans Tech Publications Ltd.: Stafa-Zurich, Switzerland, 2012; pp. 1721–1724. [Google Scholar]
  81. Gao, T.H.; Liu, H.Y.; Zhang, P.; Wu, S.Q.; Yang, Y.; Zhu, Z.Z. Structural and electronic properties of Al-doped spinel LiMn2O4. Acta Phys. Sin. 2012, 61, 9. [Google Scholar]
  82. Cai, Z.; Ma, Y.; Huang, X.; Yan, X.; Yu, Z.; Zhang, S.; Song, G.; Xu, Y.; Wen, C.; Yang, W. High electrochemical stability Al-doped spinel LiMn2O4 cathode material for Li-ion batteries. J. Energy Storage 2020, 27, 101036. [Google Scholar] [CrossRef]
  83. Xiao, L.; Zhao, Y.; Yang, Y.; Cao, Y.; Ai, X.; Yang, H. Enhanced electrochemical stability of Al-doped LiMn2O4 synthesized by a polymer-pyrolysis method. Electrochim. Acta 2008, 54, 545–550. [Google Scholar] [CrossRef]
  84. Wang, J.; Li, Z.; Yang, J.; Tang, J.; Yu, J.; Nie, W.; Lei, G.; Xiao, Q. Effect of Al-doping on the electrochemical properties of a three-dimensionally porous lithium manganese oxide for lithium-ion batteries. Electrochim. Acta 2012, 75, 115–122. [Google Scholar] [CrossRef]
  85. Ryu, W.-H.; Eom, J.-Y.; Yin, R.-Z.; Han, D.-W.; Kim, W.-K.; Kwon, H.-S. Synergistic effects of various morphologies and Al doping of spinel LiMn2O4 nanostructures on the electrochemical performance of lithium-rechargeable batteries. J. Mater. Chem. 2011, 21, 15337–15342. [Google Scholar] [CrossRef]
  86. Han, S.C.; Singh, S.P.; Hwang, Y.-H.; Bae, E.G.; Park, B.K.; Sohn, K.-S.; Pyo, M. Gadolinium-doped LiMn2O4 cathodes in Li ion batteries: Understanding the stabilized structure and enhanced electrochemical kinetics. J. Electrochem. Soc. 2012, 159, A1867–A1873. [Google Scholar] [CrossRef]
  87. Aziz, S.; Zhao, J.; Cain, C.; Wang, Y. Nanoarchitectured LiMn2O4/graphene/ZnO composites as electrodes for lithium ion batteries. J. Mater. Sci. Technol. 2014, 30, 427–433. [Google Scholar] [CrossRef]
  88. Lee, D.K.; Prabakar, S.J.R.; Pyo, M. Influence of W-doping on electrochemical performance of spinel LiMn2O4. J. Nanosci. Nanotechnol. 2013, 13, 5517–5521. [Google Scholar] [CrossRef] [PubMed]
  89. Mao, L.; Du, S.; Li, S.; Ren, Z. Excellent stability of Al-doped LiMn2O4 prepared by a sol-gel method. In Proceedings of the 5th International Conference on Environment, Materials, Chemistry and Power Electronics, Zhengzhou, China, 11–12 April 2016; pp. 338–341. [Google Scholar]
  90. Fu, Y.; Jiang, H.; Hu, Y.; Dai, Y.; Zhang, L.; Li, C. Synergistic enhancement effect of Al doping and highly active facets of LiMn2O4 cathode materials for lithium-ion batteries. Ind. Eng. Chem. Res. 2015, 54, 3800–3805. [Google Scholar] [CrossRef]
  91. Nguyen, C.C.; Bae, Y.-S.; Lee, K.-H.; Song, J.-W.; Min, J.-H.; Kim, J.-S.; Ko, H.-S.; Paik, Y.; Song, S.-W. Roles of fluorine-doping in enhancing initial cycle efficiency and SEI formation of Li-, Al-cosubstituted spinel battery cathodes. Bull. Korean Chem. Soc. 2013, 34, 384–388. [Google Scholar] [CrossRef] [Green Version]
  92. Zhang, Y.; Zhang, Y.; Zhang, M.; Xu, M.; Li, X.; Yu, X.; Dong, P. Synthesis of spherical Al-doping LiMn2O4 via a high-pressure spray-drying method as cathode materials for lithium-ion batteries. JOM 2018, 71, 608–612. [Google Scholar] [CrossRef]
  93. Yuan, A.; Tian, L.; Xu, W.; Wang, Y. Al-doped spinel LiAl0.1Mn1.9O4 with improved high-rate cyclability in aqueous electrolyte. J. Power Source 2010, 195, 5032–5038. [Google Scholar] [CrossRef]
  94. Mukherjee, S.; Schuppert, N.; Bates, A.; Lee, S.C.; Park, S. Novel mesoporous microspheres of Al and Ni doped LMO spinels and their performance as cathodes in secondary lithium ion batteries. Int. J. Green Energy 2017, 14, 656–664. [Google Scholar] [CrossRef]
  95. Wang, M.-S.; Wang, J.; Zhang, J.; Fan, L.-Z. Improving electrochemical performance of spherical LiMn2O4 cathode materials for lithium ion batteries by Al-F codoping and AlF3 surface coating. Ionics 2015, 21, 27–35. [Google Scholar] [CrossRef]
  96. Ding, Y.L.; Xie, J.; Cao, G.S.; Zhu, T.J.; Yu, H.M.; Zhao, X.B. Enhanced elevated-temperature performance of ai-doped single-crystalline LiMn2O4 nanotubes as cathodes for lithium ion batteries. J. Phys. Chem. C 2011, 115, 9821–9825. [Google Scholar] [CrossRef]
  97. Liu, G.Y.; Guo, J.M.; Wang, B.S.; He, Y. Flameless solution combustion synthesis of Al3+ doped LiMn2O4. In New Trends and Applications of Computer-Aided Material and Engineering; Tian, W.Y., Xu, L.L., Eds.; Trans Tech Publications Ltd.: Durnten-Zurich, Switzerland, 2011; pp. 7–10. [Google Scholar]
  98. Liu, G.Y.; Guo, J.M.; Zhang, L.L.; Wang, B.S.; He, Y. Al doped LiMn2O4 prepared by a solution combustion synthesis using acetate salts as raw materials and acetic acid as fuel. In Intelligent Materials, Applied Mechanics and Design Science; Zhang, H., Jin, D., Eds.; Trans Tech Publications Ltd.: Durnten-Zurich, Switzerland, 2012; pp. 209–212. [Google Scholar]
  99. Akhoon, S.A.; Rubab, S.; Shah, M.A. Enhanced cycling properties and better rate capabilities of Al-doped LiMn2O4 nanorods and nanospheres. Mater. Res. Express 2017, 4, 105016. [Google Scholar] [CrossRef] [Green Version]
  100. Chen, M.; Li, S.J.; Yang, C. Structure and electrochemical properties of La, F dual-dopedLiLa0.01Mn1.99O3.99F0.01cathode materials. J. Univ. Sci. Technol. Beijing 2008, 15, 468–473. [Google Scholar] [CrossRef]
  101. Arumugam, D.; Kalaignan, G.P. Synthesis and electrochemical characterizations of nano size Ce doped LiMn2O4 cathode materials for rechargeable lithium batteries. J. Electroanal. Chem. 2010, 648, 54–59. [Google Scholar] [CrossRef]
  102. Ram, P.; Gören, A.; Ferdov, S.; Silva, M.M.; Singhal, R.; Costa, C.M.; Sharma, R.K.; Lanceros-Méndez, S. Improved performance of rare earth doped LiMn2O4 cathodes for lithium-ion battery applications. New J. Chem. 2016, 40, 6244–6252. [Google Scholar] [CrossRef]
  103. Du, R.B.; Liu, T.; Jiang, X.J. Synthesis and electrochemical properties of Pr3+-doped LiPrxMn2−xO4 spinel. Rare Met. Mater. Eng. 2010, 39, 932–935. [Google Scholar]
  104. Zhang, H.L.; Ren, R.; An, J. Improvement of the electrochemical properties of LiMn2O4 by doping rare earth element Ce. In Advanced Structural Materials; Han, Y.F., Pan, F., Tang, J.M., Zhou, C.G., Eds.; Trans Tech Publications Ltd.: Stafa-Zurich, Switzerland, 2011; pp. 716–719. [Google Scholar]
  105. Sun, H.; Chen, Y.; Xu, C.; Zhu, D.; Huang, L. Electrochemical performance of rare-earth doped LiMn2O4 spinel cathode materials for Li-ion rechargeable battery. J. Solid State Electrochem. 2012, 16, 1247–1254. [Google Scholar] [CrossRef]
  106. Singh, P.; Sil, A.; Nath, M.; Ray, S. Preparation and characterization of lithium manganese oxide cubic spinel Li1.03Mn1.97O4 doped with Mg and Fe. Phys. B Condens. Matter 2010, 405, 649–654. [Google Scholar] [CrossRef]
  107. Zhang, Z.F.; Chen, M.M.; Xiang, M.W.; Feng, L.L.; Zhang, Y.J.; Guo, J.M. Effect of Cr doping on electrochemical performance of LiMn2O4. In Advances in Materials and Materials Processing; Jiang, Z.Y., Liu, X.H., Jiao, S.H., Han, J.T., Eds.; Trans Tech Publications Ltd.: Stafa-Zurich, Switzerland, 2013; pp. 848–852. [Google Scholar]
  108. Iturrondobeitia, A.; Goñi, A.; Palomares, V.; Gil de Muro, I.; Lezama, L.; Rojo, T. Effect of doping LiMn2O4 spinel with a tetravalent species such as Si(IV) versus with a trivalent species such as Ga(III). Electrochemical, magnetic and ESR study. J. Power Source 2012, 216, 482–488. [Google Scholar] [CrossRef]
  109. Reddy, M.V.; Manoharan, S.S.; John, J.; Singh, B.; Rao, G.V.S.; Chowdari, B.V.R. Synthesis, characterization, and electrochemical cycling behavior of the Ru-doped spinel, LiMn2−xRuxO4 (x = 0, 0.1, and 0.25). J. Electrochem. Soc. 2009, 156, A652–A660. [Google Scholar] [CrossRef]
  110. Wang, C.; Lu, S.; Kan, S.; Pang, J.; Jin, W.; Zhang, X. Enhanced capacity retention of Co and Li doubly doped LiMn2O4. J. Power Source 2009, 189, 607–610. [Google Scholar] [CrossRef]
  111. Li, J.; Tian, Y.; Xu, C. Influence of Nb5+ doping on structure and electrochemical properties of spinel Li1.02Mn2O4. J. Mater. Sci. Technol. 2012, 28, 817–822. [Google Scholar] [CrossRef]
  112. Jiang, J.; Li, W.; Deng, H.; Gong, G.; Li, N. Research on improving the electrochemical performance of LiMn2O4 via Cr-doping. J. Nanosci. Nanotechnol. 2019, 19, 125–129. [Google Scholar] [CrossRef] [PubMed]
  113. Rao, M.; Satyanarayana, T. Spectroscopic and electrochemical investigations on LiCrxMn2−xO4 cathodes for rechargeable battery application. Mater. Res. Express 2019, 6, 085514. [Google Scholar] [CrossRef]
  114. Singh, G.; Panwar, A.; Sil, A.; Ghosh, S. Synthesis and characterization of citric acid assisted Cr doped lithium manganese oxide spinel. Ceram.-Silik 2009, 53, 260–267. [Google Scholar]
  115. Lu, J.; Zhan, C.; Wu, T.; Wen, J.; Lei, Y.; Kropf, A.J.; Wu, H.; Miller, D.J.; Elam, J.W.; Sun, Y.-K.; et al. Effectively suppressing dissolution of manganese from spinel lithium manganate via a nanoscale surface-doping approach. Nat. Commun. 2014, 5, 5693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Guo, S.; Zhang, S.; He, X.; Pu, W.; Jiang, C.; Wan, C. Synthesis and characterization of Sn-doped LiMn2O4 cathode materials for rechargeable Li-ion batteries. J. Electrochem. Soc. 2008, 155, A760. [Google Scholar] [CrossRef]
  117. Wang, J.-F.; Zhang, H.-L. Improvement on the electrochemical properties of LiMn2O4 by codoping. Funct. Mater. Lett. 2010, 3, 189–191. [Google Scholar] [CrossRef]
  118. Şahan, H.; Goktepe, H.; Patat, Ş. Synthesis and cycling performance of double metal doped LiMn2O4 cathode materials for rechargeable lithium ion batteries. Inorg. Mater. 2008, 44, 420–425. [Google Scholar] [CrossRef]
  119. Arumugam, D.; Kalaignan, G.P.; Vediappan, K.; Lee, C.W. Synthesis and electrochemical characterizations of nano-scaled Zn doped LiMn2O4 cathode materials for rechargeable lithium batteries. Electrochim. Acta 2010, 55, 8439–8444. [Google Scholar] [CrossRef]
  120. Xu, H.; Li, G.; Han, E.S.; Wang, S.; Zhang, X.B. Influences of calcination temperature on the crystal structure and electrochemical performance of co-doped LiMn2O4. Synth. React. Inorg. Met. Org. Nano-Met. Chem. 2011, 41, 1338–1341. [Google Scholar] [CrossRef]
  121. Xu, W.M.; Yuan, A.B.; Tian, L.; Wang, Y.Q. Improved high-rate cyclability of sol-gel derived Cr-doped spinel LiCryMn2−yO4 in an aqueous electrolyte. J. Appl. Electrochem. 2011, 41, 453–460. [Google Scholar] [CrossRef]
  122. Sahan, H.; Goktepe, H.; Patat, S. Cycling behaviour of barium doped LiMn2O4 cathode materials for Li ion secondary batteries. Mater. Sci. Pol. 2010, 28, 773–780. [Google Scholar]
  123. Xiang, M.; Li, B.; Zhang, Z.; Peng, C.; Bai, H.; Su, C.; Guo, J. The effect of calcination time on the electrochemical performance of doped LiMg0.02Mn1.98O4 cathode material prepared by solid-state combustion synthesis. Int. J. Electrochem. Sci. 2014, 9, 901–908. [Google Scholar]
  124. Kebede, M.A.; Kunjuzwa, N.; Ozoemena, K.; Mathe, M. Synthesis and electrochemical properties of Ni doped spinel LiNixMn2−xO4 (0 ≤ x ≤ 0.5) cathode materials for Li-ion battery. ESC Trans. 2013, 50, 1–14. [Google Scholar] [CrossRef] [Green Version]
  125. Kokubu, T.; Oaki, Y.; Uchiyama, H.; Hosono, E.; Zhou, H.S.; Imai, H. Biomimetic synthesis of metal ion-doped hierarchical crystals using a gel matrix: Formation of cobalt-doped LiMn2O4 with improved electrochemical properties through a cobalt-doped MnCO3 precursor. Chem. Asian J. 2010, 5, 792–798. [Google Scholar] [CrossRef] [PubMed]
  126. Li, X.L.; Zhou, Q.; Wang, H.L.; Liu, S. Ionothermal synthesis and enhanced electrochemical performance of nanostructure Cr-doped LiMn2O4 for lithium-ion batteries. Ionics 2015, 21, 1517–1523. [Google Scholar] [CrossRef]
  127. Li, W.; Siqin, G.-W.; Zhu, Z.; Qi, L.; Tian, W.-H. Electrochemical properties of niobium and phosphate doped spherical Li-rich spinel LiMn2O4 synthesized by ion implantation method. Chin. Chem. Lett. 2017, 28, 1438–1446. [Google Scholar] [CrossRef]
  128. Rodriguez, R.A.; Laffita, Y.M.; Cappe, E.P.; Frutis, M.A.A.; Salazar, J.S.; Alves, O.L. A new strategy toward enhancing the phosphate doping in LixMn2O4 cathode materials. Ceram. Int. 2014, 40, 12413–12422. [Google Scholar] [CrossRef] [Green Version]
  129. Bhuvaneswari, S.; Varadaraju, U.V.; Gopalan, R.; Prakash, R. Structural stability and superior electrochemical performance of Sc-doped LiMn2O4 spinel as cathode for lithium ion batteries. Electrochim. Acta 2019, 301, 342–351. [Google Scholar] [CrossRef]
  130. Wang, H.-Q.; Lai, F.-Y.; Li, Y.; Zhang, X.-H.; Huang, Y.-G.; Hu, S.-J.; Li, Q.-Y. Excellent stability of spinel LiMn2O4-based cathode materials for lithium-ion batteries. Electrochim. Acta 2015, 177, 290–297. [Google Scholar] [CrossRef]
  131. Zhu, Q.; Zheng, S.; Lu, X.; Wan, Y.; Chen, Q.; Yang, J.; Zhang, L.-Z.; Lu, Z. Improved cycle performance of LiMn2O4 cathode material for aqueous rechargeable lithium battery by LaF3 coating. J. Alloys Compd. 2016, 654, 384–391. [Google Scholar] [CrossRef]
  132. Ye, S.; Bo, J.; Li, C.; Cao, J.; Sun, Q.; Wang, Y. Improvement of the high-rate discharge capability of phosphate-doped spinel LiMn2O4 by a hydrothermal method. Electrochim. Acta 2010, 55, 2972–2977. [Google Scholar] [CrossRef]
  133. Iqbal, A.; Iqbal, Y.; Khan, A.M.; Ahmed, S. Low content Ni and Cr co-doped LiMn2O4 with enhanced capacity retention. Ionics 2017, 23, 1995–2003. [Google Scholar] [CrossRef]
  134. Lee, H.-W.; Muralidharan, P.; Ruffo, R.; Mari, C.M.; Cui, Y.; Kim, D.K. Ultrathin spinel LiMn2O4 nanowires as high power cathode materials for Li-ion batteries. Nano Lett. 2010, 10, 3852–3856. [Google Scholar] [CrossRef] [PubMed]
  135. Dokan, F.K.; Sahan, H.; Veziroglu, S.; Tas, N.A.; Aydin, A.; Patat, S. The effect of multiple cation (Al, Si, Ti, Co) doping on electrochemical performance of LiMn2O4 cathode active material. Acta Phys. Pol. A 2013, 123, 365–367. [Google Scholar] [CrossRef]
  136. Thirunakaran, R. Synthesis and electrochemical characterization of duo doped spinels (zinc and praseodymium) for use in lithium rechargeable batteries. J. Sol-Gel Sci. Technol. 2013, 69, 397–406. [Google Scholar] [CrossRef]
  137. Thirunakaran, R.; Kim, T.; Yoon, W.-S. Zinc and aluminium: Glutamic acid assisted dual-doped LiMn2O4 spinels via sol–gel method as cathode material for use in lithium rechargeable batteries. J. Sol-Gel Sci. Technol. 2014, 73, 62–71. [Google Scholar] [CrossRef]
  138. Thirunakaran, R.; Lew, G.H.; Yoon, W.-S. Novel chelating agent assisted dual doped spinel via sol–gel method for lithium rechargeable batteries. J. Electroanal. Chem. 2016, 767, 141–152. [Google Scholar] [CrossRef]
  139. Thirunakaran, R.; Sivashanmugam, A.; Gopukumar, S.; Rajalakshmi, R. Cerium and zinc: Dual-doped LiMn2O4 spinels as cathode material for use in lithium rechargeable batteries. J. Power Source 2009, 187, 565–574. [Google Scholar] [CrossRef]
  140. Zhao, H.; Li, F.; Liu, X.; Cheng, C.; Zhang, Z.; Wu, Y.; Xiong, W.; Chen, B. Effects of equimolar Mg (II) and Si (IV) co-doping on the electrochemical properties of spinel LiMn2−2xMgxSixO4 prepared by citric acid assisted sol–gel method. Electrochim. Acta 2015, 151, 263–269. [Google Scholar] [CrossRef]
  141. Jian-Kun, T.; Fu-Cheng, W.; Battaglia, V.S.; Hai-Lang, Z. Synthesis and electrochemical performance of nanosized multiple-doped LiMn2O4 prepared at low temperature for Li-ion battery. Int. J. Electrochem. Sci. 2014, 9, 931–942. [Google Scholar]
  142. Rao, A.V.; Kumar, B.R. Structural, electrical and electrochemical studies on doubly doped LiMn2−x(GdNi)xO4 cathode materials for Li-ion batteries. Mater. Lett. 2018, 227, 250–253. [Google Scholar]
  143. Sahan, H.; Ates, M.N.; Dokan, F.K.; Ulgen, A.; Patat, S. Synergetic action of doping and coating on electrochemical performance of lithium manganese spinel as an electrode material for lithium-ion batteries. Bull. Mater. Sci. 2015, 38, 141–149. [Google Scholar] [CrossRef]
  144. Kim, S.W.; Kumar, V.G.; Seo, D.H.; Park, Y.U.; Kim, J.; Kim, H.; Kim, J.; Hong, J.; Kang, K. Invited paper: Preparation and electrochemical characterization of doped spinel LiMn1.88Ge0.1Li0.02O4 cathode material. Electron. Mater. Lett. 2011, 7, 105–108. [Google Scholar] [CrossRef]
  145. Zhang, Z.F.; Chen, Z.L.; Wang, G.J.; Ren, H.; Pan, M.; Xiao, L.L.; Wu, K.C.; Zhao, L.T.; Yang, J.Q.; Wu, Q.G.; et al. Dual-doping to suppress cracking in spinel LiMn2O4: A joint theoretical and experimental study. Phys. Chem. Chem. Phys. 2016, 18, 6893–6900. [Google Scholar] [CrossRef]
  146. Zhao, H.Y.; Liu, S.S.; Tan, M.; Wang, Z.W.; Cai, Y.; Liu, X.Q. Enhanced Cycling Stability of Multi-Cations Doped Spinel Lithium Manganese Oxide for Rechargeable Lithium Batteries; Walter De Gruyter Gmbh: Berlin, Germany, 2017. [Google Scholar]
  147. Chen, M.F.; Chen, P.; Yang, F.; Song, H.Y.; Liao, S.J. Ni, Mo Co-doped lithium manganate with significantly enhanced discharge capacity and cycling stability. Electrochim. Acta 2016, 206, 356–365. [Google Scholar] [CrossRef]
  148. Meng, M.W.; Ji, X.M.; Liu, Q.Y.; Hu, C.H.; Huang, S.Y.; Huang, Y.; Zhao, Y.M.; Liao, Q.H.; Kang, C.Y. Characterization of structure and Properties of lithium manganese oxygen cathode materials doped with rare earth. Rare Met. Mater. Eng. 2010, 39, 652–655. [Google Scholar]
  149. Xiong, L.L.; Xu, Y.L.; Tao, T.; Goodenough, J.B. Synthesis and electrochemical characterization of multi-cations doped spinel LiMn2O4 used for lithium ion batteries. J. Power Source 2012, 199, 214–219. [Google Scholar] [CrossRef]
  150. Abou-Sekkina, M.M.; Saad, F.A.; El-Metwaly, F.G.; Khedr, A.M. Synthesis, characterization, DC-electrical conductivity and gamma-ray effect on Ag1+, Y3+ double doped nano lithium manganates (LiMn2−2xAgxYxO4) for rechargeable batteries. Mater. Sci-Pol. 2014, 32, 315–323. [Google Scholar] [CrossRef] [Green Version]
  151. Xu, C.R.; Li, Y.J.; Xu, H.; Li, P.L.; Kong, L.; Su, Q.Y.; Cao, X.L. Electrochemical evaluation of Co-Al dual-doped LiMn2O4 spinels synthesized via hydrothermal method. Int. J. Electrochem. Sci. 2017, 12, 5185–5198. [Google Scholar] [CrossRef]
  152. Iqbal, A.; Khan, A.M.; Wang, T.; Li, D.; Gao, Y. Effect of Ni and Cu Substitution on the crystal structure, morphology and electrochemical performance of spinel LiMn2O4. Int. J. Electrochem. Sci. 2019, 929–942. [Google Scholar] [CrossRef]
  153. Raja, M.W.; Mahanty, S.; Basu, R.N. Influence of S and Ni co-doping on structure, band gap and electrochemical properties of lithium manganese oxide synthesized by soft chemical method. J. Power Source 2009, 192, 618–626. [Google Scholar] [CrossRef]
  154. Amaral, F.A.; Bocchi, N.; Brocenschi, R.F.; Biaggio, S.R.; Rocha, R.C. Structural and electrochemical properties of the doped spinels Li1.05M0.02Mn1.98O3.98N0.02 (M = Ga3+, Al3+, or Co3+; N = S2− or F) for use as cathode material in lithium batteries. J. Power Source 2010, 195, 3293–3299. [Google Scholar] [CrossRef]
  155. Meng, M.; Liao, Q.; Jiang, E.; Liu, X.; Lu, G.; Huang, H. Effects of doping elements lanthanum and fluorine on structure and electrochemical performance of spinel LiMn2O4 cathode material. Rare Met. Mater. Eng. 2009, 38, 995–998. [Google Scholar]
  156. Fan, X.F.; Zhao, S.X.; Li, L.; Nan, C.W. Structure and electrochemical performance of modificated LiMn2O4 by S-Co codoping and nano SiO2 surface coating. In Functional Nanocomposite Materials; Yu, S., Sun, R., Du, R.X., Eds.; Trans Tech Publications Ltd.: Stafa-Zurich, Switzerland, 2012; pp. 1–9. [Google Scholar]
  157. Fu-Cheng, W.; Jian-Kun, T.; Hai-Lang, Z. Synthesis and electrochemical performance of Co-doped LiMn2O4 by Mg and F with one improved soild-state method. Int. J. Electrochem. Sci. 2014, 9, 4627–4634. [Google Scholar]
  158. Lee, H.R.; Lee, B.; Chung, K.Y.; Cho, B.W.; Lee, K.Y.; Oh, S.H. Scalable synthesis and electrochemical investigations of fluorine-doped lithium manganese spinel oxide. Electrochim. Acta 2014, 136, 396–403. [Google Scholar] [CrossRef]
  159. Zhang, H.L.; Zhang, L.; Ye, L. Synthesis and improvement of Li-Mn spinel by multiple Co-doping. Int. J. Electrochem. Sci. 2014, 9, 4635–4642. [Google Scholar]
  160. Zhang, H.L.; Zhang, L.; Yang, S.W. Synthesis and electrochemical study of Co-doped LiMn2O4 at room temperature and high temperature. Int. J. Electrochem. Sci. 2014, 9, 8182–8188. [Google Scholar]
  161. Wang, Q.Q.; Zhang, X.S.; Xu, Y.L.; Liu, D.; Dong, H.; Zhang, Y. Synthesis and electrochemical performance of Ni and F doped LiMn2O4 cathode materials. RSC Adv. 2015, 5, 75333–75340. [Google Scholar] [CrossRef]
  162. Guan, D.; Wang, Y. Ultrathin surface coatings to enhance cycling stability of LiMn2O4 cathode in lithium-ion batteries. Ionics 2012, 19, 1–8. [Google Scholar] [CrossRef]
  163. Chen, Q.; Wang, Y.; Zhang, T.; Yin, W.; Yang, J.; Wang, X. Electrochemical performance of LaF3-coated LiMn2O4 cathode materials for lithium ion batteries. Electrochim. Acta 2012, 83, 65–72. [Google Scholar] [CrossRef]
  164. Michalska, M.; Hamankiewicz, B.; Ziółkowska, D.; Krajewski, M.; Lipińska, L.; Andrzejczuk, M.; Czerwiński, A. Influence of LiMn2O4 modification with CeO2 on electrode performance. Electrochim. Acta 2014, 136, 286–291. [Google Scholar] [CrossRef]
  165. Ha, H.-W.; Yun, N.J.; Kim, K. Improvement of electrochemical stability of LiMn2O4 by CeO2 coating for lithium-ion batteries. Electrochim. Acta 2007, 52, 3236–3241. [Google Scholar] [CrossRef]
  166. Han, C.-G.; Zhu, C.; Saito, G.; Sheng, N.; Nomura, T.; Akiyama, T. Enhanced cycling performance of surface-doped LiMn2O4 modified by a Li2CuO2-Li2NiO2 solid solution for rechargeable lithium-ion batteries. Electrochim. Acta 2017, 224, 71–79. [Google Scholar] [CrossRef]
  167. Wang, X.; Wang, H.; Wen, J.; Tan, Y.; Zeng, Y. Surface modification of LiMn2O4 cathode with LaCoO3 by a molten salt method for lithium ion batteries. Ceram. Int. 2021, 47, 6434–6441. [Google Scholar] [CrossRef]
  168. Zhao, X.; Hayner, C.M.; Kung, H.H. Self-assembled lithium manganese oxide nanoparticles on carbon nanotube or graphene as high-performance cathode material for lithium-ion batteries. J. Mater. Chem. 2011, 21, 17297–17303. [Google Scholar] [CrossRef]
  169. Jaber-Ansari, L.; Puntambekar, K.P.; Kim, S.; Aykol, M.; Luo, L.L.; Wu, J.S.; Myers, B.D.; Iddir, H.; Russell, J.T.; Saldana, S.J.; et al. Suppressing manganese dissolution from lithium manganese oxide spinel cathodes with single-layer graphene. Adv. Energy Mater. 2015, 5, 10. [Google Scholar] [CrossRef]
  170. Cericola, D.; Novak, P.; Wokaun, A.; Kotz, R. Segmented bi-material electrodes of activated carbon and LiMn2O4 for electrochemical hybrid storage devices: Effect of mass ratio and C-rate on current sharing. Electrochim. Acta 2011, 56, 1288–1293. [Google Scholar] [CrossRef]
  171. Cericola, D.; Novak, P.; Wokaun, A.; Kotz, R. Mixed bi-material electrodes based on LiMn2O4 and activated carbon for hybrid electrochemical energy storage devices. Electrochim. Acta 2011, 56, 8403–8411. [Google Scholar] [CrossRef]
  172. Tang, M.; Yuan, A.; Xu, J. Synthesis of highly crystalline LiMn2O4/multiwalled carbon nanotube composite material with high performance as lithium-ion battery cathode via an improved two-step approach. Electrochim. Acta 2015, 166, 244–252. [Google Scholar] [CrossRef]
  173. Shah, A.; Ates, M.N.; Kotz, S.; Seo, J.; Abraham, K.M.; Somu, S.; Busnaina, A. A Layered carbon nanotube architecture for high power lithium ion batteries. J. Electrochem. Soc. 2014, 161, A989–A995. [Google Scholar] [CrossRef]
  174. Hong, H.P.; Kim, M.S.; Lee, Y.H.; Yu, J.S.; Lee, C.J.; Min, N.K. Spray deposition of LiMn2O4 nanoparticle-decorated multiwalled carbon nanotube films as cathode material for lithium-ion batteries. Thin Solid Film. 2013, 547, 68–71. [Google Scholar] [CrossRef]
  175. Xia, H.; Ragavendran, K.R.; Xie, J.P.; Lu, L. Ultrafine LiMn2O4/carbon nanotube nanocomposite with excellent rate capability and cycling stability for lithium-ion batteries. J. Power Source 2012, 212, 28–34. [Google Scholar] [CrossRef]
  176. Ding, Y.H.; Li, J.X.; Zhao, Y.; Guan, L.H. Direct growth of LiMn2O4 on carbon nanotubes as cathode materials for lithium ion batteries. Mater. Lett. 2012, 68, 197–200. [Google Scholar] [CrossRef]
  177. Liu, X.-M.; Huang, Z.-D.; Oh, S.; Ma, P.-C.; Chan, P.C.H.; Vedam, G.K.; Kang, K.; Kim, J.-K. Sol–gel synthesis of multiwalled carbon nanotube-LiMn2O4 nanocomposites as cathode materials for Li-ion batteries. J. Power Source 2010, 195, 4290–4296. [Google Scholar] [CrossRef]
  178. Zhuo, H.T.; Wan, S.; He, C.X.; Zhang, Q.L.; Li, C.H.; Gui, D.Y.; Zhu, C.Z.; Niu, H.B.; Liu, J.H. Improved electrochemical performance of spinel LiMn2O4 in situ coated with graphene-like membrane. J. Power Source 2014, 247, 721–728. [Google Scholar] [CrossRef]
  179. Pyun, M.H.; Park, Y.J. Graphene/LiMn2O4 nanocomposites for enhanced lithium ion batteries with high rate capability. J. Alloys Compd. 2015, 643, S90–S94. [Google Scholar] [CrossRef]
  180. Jiang, Q.Q.; Xu, L.; Ma, Z.L.; Zhang, H. Carbon coated to improve the electrochemical properties of LiMn2O4 cathode material synthesized by the novel acetone hydrothermal method. Appl. Phys. A Mater. Sci. Process. 2015, 119, 1069–1074. [Google Scholar] [CrossRef]
  181. Jiang, Q.; Wang, X.; Tang, Z. Improving the electrochemical performance of LiMn2O4 by amorphous carbon coating. Fuller. Nanotub. Carbon Nanostruct. 2014, 23, 676–679. [Google Scholar] [CrossRef]
  182. Lee, S.; Cho, Y.; Song, H.K.; 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. Engl. 2012, 51, 8748–8752. [Google Scholar] [CrossRef]
  183. Wutthiprom, J.; Phattharasupakun, N.; Sawangphruk, M. Turning carbon black to hollow carbon nanospheres for enhancing charge storage capacities of LiMn2O4, LiCoO2, LiNiMnCoO2, and LiFePO4 lithium-ion batteries. ACS Omega 2017, 2, 3730–3738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Bak, S.-M.; Nam, K.-W.; Lee, C.-W.; Kim, K.-H.; Jung, H.-C.; Yang, X.-Q.; Kim, K.-B. Spinel LiMn2O4/reduced graphene oxide hybrid for high rate lithium ion batteries. J. Mater. Chem. 2011, 21, 17309. [Google Scholar] [CrossRef]
  185. Li, J.; Zhang, X.; Peng, R.F.; Huang, Y.J.; Guo, L.; Qi, Y.C. LiMn2O4/graphene composites as cathodes with enhanced electrochemical performance for lithium-ion capacitors. RSC Adv. 2016, 6, 54866–54873. [Google Scholar] [CrossRef]
  186. Ge, Q.S.; Wang, D.F.; Li, F.L.; Chen, D.; Ping, G.X.; Fan, M.Q.; Qin, L.S.; Bai, L.Q.; Tian, G.L.; Lv, C.J.; et al. Enhanced cycling stability of spinel LiMn2O4 cathode by incorporating graphene sheets. Russ. J. Electrochem. 2015, 51, 125–133. [Google Scholar] [CrossRef]
  187. Sreelakshmi, K.V.; Sasi, S.; Balakrishnan, A.; Sivakumar, N.; Nair, A.S.; Nair, S.V.; Subramanian, K.R.V. Hybrid composites of LiMn2O4-graphene as rechargeable electrodes in energy storage devices. Energy Technol. 2014, 2, 257–262. [Google Scholar] [CrossRef]
  188. Lin, B.H.; Yin, Q.; Hu, H.R.; Lu, F.J.; Xia, H. LiMn2O4 nanoparticles anchored on graphene nanosheets as high-performance cathode material for lithium-ion batteries. J. Solid State Chem. 2014, 209, 23–28. [Google Scholar] [CrossRef]
  189. Jasinski, J.B.; Ziolkowska, D.; Michalska, M.; Lipinska, L.; Korona, K.P.; Kaminska, M. Novel graphene oxide/manganese oxide nanocomposites. RAS Adv. 2013, 3, 22857–22862. [Google Scholar] [CrossRef]
  190. Cui, Y.L.; Xu, K.; Yuan, Z.; Xie, R.J.; Zhu, G.L.; Zhuang, Q.C.; Qiang, Y.H. Synthesis and electrochemical performance of graphene modified nano-spinel LiMn2O4 cathode materials. Chin. J. Inorg. Chem. 2013, 29, 50–56. [Google Scholar]
  191. Zhu, X.; Hoang, T.K.A.; Chen, P. Novel carbon materials in the cathode formulation for high rate rechargeable hybrid aqueous batteries. Energies 2017, 10, 17. [Google Scholar] [CrossRef] [Green Version]
  192. Zhu, J.P.; Duan, R.; Zhang, S.; Jiang, N.; Zhang, Y.Y.; Zhu, J. The application of graphene in lithium ion battery electrode materials. Springerplus 2014, 3, 585. [Google Scholar] [CrossRef] [Green Version]
  193. Rangappa, D.; Hari Mohan, E.; Siddhartha, V.; Gopalan, R.; Narasinga Rao, T. Preparation of LiMn2O4 graphene hybrid nanostructure by combustion synthesis and their electrochemical properties. AIMS Mater. Sci. 2014, 1, 174–183. [Google Scholar] [CrossRef]
  194. Ragavendran, K.; Hui, X.; Gu, X.; Sherwood, D.; Emmanuel, B.; Arof, A.K. On the graphene incorporated LiMn2O4 nanostructures: Possibilities for tuning the preferred orientations and high rate capabilities. RSC Adv. 2014, 4, 60106–60111. [Google Scholar] [CrossRef]
  195. Liu, D.; He, Z.; Liu, X. Increased cycling stability of AlPO4-coated LiMn2O4 for lithium ion batteries. Mater. Lett. 2007, 61, 4703–4706. [Google Scholar] [CrossRef]
  196. Cho, M.-Y.; Roh, K.-C.; Park, S.-M.; Lee, J.-W. Effects of CeO2 coating uniformity on high temperature cycle life performance of LiMn2O4. Mater. Lett. 2011, 65, 2011–2014. [Google Scholar] [CrossRef]
  197. Arumugam, D.; Kalaignan, G.P. Synthesis and electrochemical characterization of nano-CeO2-coated nanostructure LiMn2O4 cathode materials for rechargeable lithium batteries. Electrochim. Acta 2010, 55, 8709–8716. [Google Scholar] [CrossRef]
  198. Arumugam, D.; Kalaignan, G.P. Electrochemical characterizations of surface modified LiMn2O4 cathode materials for high temperature lithium battery applications. Thin Solid Film. 2011, 520, 338–343. [Google Scholar] [CrossRef]
  199. Arumugam, D.; Paruthimal Kalaignan, G. Synthesis and electrochemical characterizations of nano-La2O3-coated nanostructure LiMn2O4 cathode materials for rechargeable lithium batteries. Mater. Res. Bull. 2010, 45, 1825–1831. [Google Scholar] [CrossRef]
  200. Feng, L.; Wang, S.; Han, L.; Qin, X.; Wei, H.; Yang, Y. Enhanced electrochemical properties of LiMn2O4 cathode material coated by 5 wt.% of nano-La2O3. Mater. Lett. 2012, 78, 116–119. [Google Scholar] [CrossRef]
  201. Zhang, Y.N.; Dong, P.; Zhang, M.Y.; Sun, X.L.; Yu, X.H.; Song, J.J.; Meng, Q.; Li, X.; Zhang, Y.J. Combustion combined with ball milling to produce nanoscale La2O3 coated on LiMn2O4 for optimized Li-ion storage performance at high temperature. J. Appl. Electrochem. 2018, 48, 135–145. [Google Scholar] [CrossRef]
  202. Shang, Y.; Lin, X.; Lu, X.; Huang, T.; Yu, A. Nano-TiO2(B) coated LiMn2O4 as cathode materials for lithium-ion batteries at elevated temperatures. Electrochim. Acta 2015, 156, 121–126. [Google Scholar] [CrossRef]
  203. Zhang, J.; Shen, J.; Wang, T.; Wei, C.; Ma, Y.; Zhu, C.; Yue, Y. Improvement of capacity and cycling performance of spinel LiMn2O4 cathode materials with TiO2-B nanobelts. Electrochim. Acta 2013, 111, 691–697. [Google Scholar] [CrossRef]
  204. Lai, C.; Ye, W.; Liu, H.; Wang, W. Preparation of TiO2-coated LiMn2O4 by carrier transfer method. Ionics 2008, 15, 389–392. [Google Scholar] [CrossRef]
  205. Yu, L.; Qiu, X.; Xi, J.; Zhu, W.; Chen, L. Enhanced high-potential and elevated-temperature cycling stability of LiMn2O4 cathode by TiO2 modification for Li-ion battery. Electrochim. Acta 2006, 51, 6406–6411. [Google Scholar] [CrossRef]
  206. Walz, K.A.; Johnson, C.S.; Genthe, J.; Stoiber, L.C.; Zeltner, W.A.; Anderson, M.A.; Thackeray, M.M. Elevated temperature cycling stability and electrochemical impedance of LiMn2O4 cathodes with nanoporous ZrO2 and TiO2 coatings. J. Power Source 2010, 195, 4943–4951. [Google Scholar] [CrossRef]
  207. Guler, M.O.; Akbulut, A.; Cetinkaya, T.; Uysal, M.; Akbulut, H. Improvement of electrochemical and structural properties of LiMn2O4 spinel based electrode materials for Li-ion batteries. Int. J. Hydrogen Energy 2014, 39, 21447–21460. [Google Scholar] [CrossRef]
  208. Ming, H.; Yan, Y.; Ming, J.; Adkins, J.; Li, X.; Zhou, Q.; Zheng, J. Gradient V2O5 surface-coated LiMn2O4 cathode towards enhanced performance in Li-ion battery applications. Electrochim. Acta 2014, 120, 390–397. [Google Scholar] [CrossRef]
  209. Tao, S.; Zhao, H.; Wu, C.; Xie, H.; Cui, P.; Xiang, T.; Chen, S.; Zhang, L.; Fang, Y.; Wang, Z.; et al. Enhanced electrochemical performance of MoO3-coated LiMn2O4 cathode for rechargeable lithium-ion batteries. Mater. Chem. Phys. 2017, 199, 203–208. [Google Scholar] [CrossRef]
  210. Lee, J.H.; Kim, K.J. Superior electrochemical properties of porous Mn2O3-coated LiMn2O4 thin-film cathodes for Li-ion microbatteries. Electrochim. Acta 2013, 102, 196–201. [Google Scholar] [CrossRef]
  211. Yao, Y.; Wang, Z.; Yu, X.; Zhang, Y.; Duan, J.; Zhu, C.; Hu, Z.; Shen, Z.; Wang, Q.; Zhan, Z.; et al. Interface control strategy of synthesis LiMn2O4@Al2O3 assisted by tert-butanol. Int. J. Electrochem. Sci. 2019, 14, 6478–6487. [Google Scholar] [CrossRef]
  212. Li, S.; Zhu, K.; Zhao, D.; Zhao, Q.; Zhang, N. Porous LiMn2O4 with Al2O3 coating as high-performance positive materials. Ionics 2018, 25, 1991–1998. [Google Scholar] [CrossRef]
  213. Zhou, H.-M.; Zhu, Y.-H.; Li, J.; Sun, W.-J.; Liu, Z.-Z. Electrochemical performance of Al2O3 pre-coated spinel LiMn2O4. Rare Met. 2015, 38, 128–135. [Google Scholar] [CrossRef] [Green Version]
  214. Guo, J.; Chen, Y.; Xu, C.; Li, Y.; Deng, S.; Xu, H.; Su, Q. Enhanced electrochemical performance of LiMn2O4 by SiO2 modifying via electrostatic attraction forces method. Ionics 2019, 25, 2977–2985. [Google Scholar] [CrossRef]
  215. Yi, X.; Wang, X.; Ju, B.; Shu, H.; Wen, W.; Yu, R.; Wang, D.; Yang, X. Effective enhancement of electrochemical performance for spherical spinel LiMn2O4 via Li ion conductive Li2ZrO3 coating. Electrochim. Acta 2014, 134, 143–149. [Google Scholar] [CrossRef]
  216. Qing, C.; Bai, Y.; Yang, J.; Zhang, W. Enhanced cycling stability of LiMn2O4 cathode by amorphous FePO4 coating. Electrochim. Acta 2011, 56, 6612–6618. [Google Scholar] [CrossRef]
  217. Yang, Z.; Li, S.; Xia, S.-A.; Jiang, Y.; Zhang, W.-X.; Huang, Y.-H. Significant improved electrochemical performance of spinel LiMn2O4 promoted by FePO4 incorporation. Electrochem. Solid-State Lett. 2011, 14, A109. [Google Scholar] [CrossRef]
  218. Zhao, S.; Bai, Y.; Ding, L.; Wang, B.; Zhang, W. Enhanced cycling stability and thermal stability of YPO4-coated LiMn2O4 cathode materials for lithium ion batteries. Solid State Ion. 2013, 247, 22–29. [Google Scholar] [CrossRef]
  219. Yan, J.; Liu, H.; Wang, Y.; Zhao, X.; Mi, Y.; Xia, B. Enhanced high-temperature cycling stability of LiMn2O4 by LiCoO2 coating as cathode material for lithium ion batteries. J. Mater. Sci. Chem. Eng. 2014, 02, 12–18. [Google Scholar]
  220. Shi, T.; Dong, Y.; Wang, C.-M.; Tao, F.; Chen, L. Enhanced cycle stability at high rate and excellent high rate capability of La0.7Sr0.3Mn0.7Co0.3O3-coated LiMn2O4. J. Power Source 2015, 273, 959–965. [Google Scholar] [CrossRef]
  221. Tron, A.; Park, Y.D.; Mun, J. AlF3-coated LiMn2O4 as cathode material for aqueous rechargeable lithium battery with improved cycling stability. J. Power Source 2016, 325, 360–364. [Google Scholar] [CrossRef]
  222. Li, X.; Yang, R.; Cheng, B.; Hao, Q.; Xu, H.; Yang, J.; Qian, Y. Enhanced electrochemical properties of nano-Li3PO4 coated on the LiMn2O4 cathode material for lithium ion battery at 55 °C. Mater. Lett. 2012, 66, 168–171. [Google Scholar] [CrossRef]
  223. Liu, J.; Wu, X.; Chen, S.; Liu, J.; He, Z. Enhanced high temperature performance of LiMn2O4 coated with Li3BO3 solid electrolyte. Bull. Mater. Sci. 2013, 36, 687–691. [Google Scholar]
  224. Zhao, S.; Bai, Y.; Chang, Q.; Yang, Y.; Zhang, W. Surface modification of spinel LiMn2O4 with FeF3 for lithium ion batteries. Electrochim. Acta 2013, 108, 727–735. [Google Scholar] [CrossRef]
  225. Zhao, S.; Chang, Q.; Jiang, K.; Bai, Y.; Yang, Y.; Zhang, W. Performance improvement of spinel LiMn2O4 cathode material by LaF3 surface modification. Solid State Ion. 2013, 253, 1–7. [Google Scholar] [CrossRef]
  226. Peng, Z.; Wang, G.; Cao, Y.; Zhang, Z.; Du, K.; Hu, G. Enhanced high power and long life performance of spinel LiMn2O4 with Li2MnO3 coating for lithium-ion batteries. J. Solid State Electrochem. 2016, 20, 2865–2871. [Google Scholar] [CrossRef]
  227. Potapenko, A.V.; Kirillov, S.A. Enhancing high-rate electrochemical properties of LiMn2O4 in a LiMn2O4/LiNi0.5Mn1.5O4 core/shell composite. Electrochim. Acta 2017, 259, 9–16. [Google Scholar] [CrossRef]
  228. Wen, W.; Chen, S.; Fu, Y.; Wang, X.; Shu, H. A core–shell structure spinel cathode material with a concentration-gradient shell for high performance lithium-ion batteries. J. Power Source 2015, 274, 219–228. [Google Scholar] [CrossRef]
  229. Shang, Y.; Liu, J.; Huang, T.; Yu, A. Effect of heat treatment on the structure and electrochemical performance of FePO4 coated spinel LiMn2O4. Electrochim. Acta 2013, 113, 248–255. [Google Scholar] [CrossRef]
  230. Lu, Z.; Lu, X.; Ding, J.; Zhou, T.; Ge, T.; Yang, G.; Yin, F.; Wu, M. Enhanced electrochemical performance of LiMn2O4 by constructing a stable Mn2+-rich interface. Appl. Surf. Sci. 2017, 426, 19–28. [Google Scholar] [CrossRef]
  231. Mohan, P.; Paruthimal-Kalaignan, G. Structure and electrochemical performance of surface modified LaPO4 coated LiMn2O4 cathode materials for rechargeable lithium batteries. Ceram. Int. 2014, 40, 1415–1421. [Google Scholar] [CrossRef]
  232. Şahan, H.; Göktepe, H.; Patat, Ş.; Ülgen, A. Improvement of the electrochemical performance of LiMn2O4 cathode active material by lithium borosilicate (LBS) surface coating for lithium-ion batteries. J. Alloys Compd. 2011, 509, 4235–4241. [Google Scholar] [CrossRef]
  233. Wei, C.; Fei, H.; An, Y.; Zhang, Y.; Feng, J. Crumpled Ti3C2Tx (MXene) nanosheet encapsulated LiMn2O4 for high performance lithium-ion batteries. Electrochim. Acta 2019, 309, 362–370. [Google Scholar] [CrossRef]
  234. Sinha, A.; Dhanjai; Zhao, H.; Huang, Y.; Lu, X.; Chen, J.; Jain, R. MXene: An emerging material for sensing and biosensing. Trac. Trends Anal. Chem. 2018, 105, 424–435. [Google Scholar] [CrossRef]
  235. Tang, H.; Hu, Q.; Zheng, M.; Chi, Y.; Qin, X.; Pang, H.; Xu, Q. MXene-2D layered electrode materials for energy storage. Prog. Nat. Sci. Mater. Int. 2018, 28, 133–147. [Google Scholar] [CrossRef]
  236. Yan, H.; Zhang, D.; Guo, G.; Wang, Z.; Liu, Y.; Wang, X. Hydrothermal synthesis of spherical Li4Ti5O12 material for a novel durable Li4Ti5O12/LiMn2O4 full lithium ion battery. Ceram. Int. 2016, 42, 14855–14861. [Google Scholar] [CrossRef]
  237. Su, X.L.; Liu, J.Y.; Zhang, C.C.; Huang, T.; Wang, Y.G.; Yu, A.S. High power lithium-ion battery based on a LiMn2O4 nanorod cathode and a carbon-coated Li4Ti5O12 nanowire anode. RSC Adv. 2016, 6, 107355–107363. [Google Scholar] [CrossRef]
  238. Kim, M.-K.; Kim, J.; Yu, S.-H.; Mun, J.; Sung, Y.-E. A facile process for surface modification with lithium ion conducting material of Li2TiF6 for LiMn2O4 in lithium ion batteries. J. Electrochem. Sci. Technol. 2019, 10, 223–230. [Google Scholar]
  239. Kang, H.-B.; Myung, S.-T.; Amine, K.; Lee, S.-M.; Sun, Y.-K. Improved electrochemical properties of BiOF-coated 5V spinel LiNi0.5Mn1.5O4 for rechargeable lithium batteries. J. Power Source 2010, 195, 2023–2028. [Google Scholar] [CrossRef]
  240. Xu, X.; Deng, S.; Wang, H.; Liu, J.; Yan, H. Research progress in improving the cycling stability of high-voltage LiNi0.5Mn1.5O4 cathode in lithium-ion battery. Nano-Micro Lett. 2017, 9, 22. [Google Scholar] [CrossRef] [Green Version]
  241. Ouyang, C.Y.; Shi, S.Q.; Lei, M.S. Jahn–Teller distortion and electronic structure of LiMn2O4. J. Alloys Compd. 2009, 474, 370–374. [Google Scholar] [CrossRef]
  242. Li, X.; Xu, Y.; Wang, C. Suppression of Jahn–Teller distortion of spinel LiMn2O4 cathode. J. Alloys Compd. 2009, 479, 310–313. [Google Scholar] [CrossRef]
  243. Capsoni, D.; Bini, M.; Chiodelli, G.; Mustarelli, P.; Massarotti, V.; Azzoni, C.B.; Mozzati, M.C.; Linati, L. Inhibition of Jahn-Teller cooperative distortion in LiMn2O4 spinel by Ga3+ doping. J. Phys. Chem. B 2002, 106, 7432–7438. [Google Scholar] [CrossRef]
  244. Tu, J.; Zhao, X.B.; Xie, J.; Cao, G.S.; Zhuang, D.G.; Zhu, T.J.; Tu, J.P. Enhanced low voltage cycling stability of LiMn2O4 cathode by ZnO coating for lithium ion batteries. J. Alloys Compd. 2007, 432, 313–317. [Google Scholar] [CrossRef]
Table 1. List of discharge rates abbreviations (ordered from the lowest to highest discharge rate value).
Table 1. List of discharge rates abbreviations (ordered from the lowest to highest discharge rate value).
Abbr. *Discharge Rate ValueUnit of MeasurementAbbr. *Discharge Rate ValueUnit of MeasurementAbbr.*Discharge Rate ValueUnit of Measurement
(a)0.05(C)(i)1(C)(q)100(mA·g−1)
(b)0.1(j)2(r)500
(c)0.14(k)3(s)0.1(mA·cm−2)
(d)0.2(l)5(t)0.2
(e)0.25(m)6.75(u)0.3
(f)0.3(n)100(μA)(v)0.5
(g)0.5(o)50(mA·g−1)(w)0.55
(h)0.7(p)75(x)70(mA·g−1)
* Abbreviation.
Table 2. Al-doped lithium manganese oxide and its performances during electrochemical tests.
Table 2. Al-doped lithium manganese oxide and its performances during electrochemical tests.
Nr. Crt.MaterialSynthesis MethodDischarge CapacityCoulombic EfficiencyDischarge Capacity/N CyclesCapacity RetentionRef
(mAh·g−1)(%)(mAh·g−1)(%)
1LiAl0.02Mn1.98O4Hydrothermal reaction129.8 (d)95.4106.05/5081.7[85]
2LiAl0.1Mn1.9O4Phase-inversion technology117 (d)96112.32/42096[84]
120 (i)-102/5085
3LiAl0.025Mn1.975O4Sol-gel109 (i)-88/5080.7[88]
4LiAl0.1Mn1.9O4113.2 (g)-109.9/10097.1[89]
5Al-doped LMO120.199.9115.9/20096.5[42]
119.6 (l)111.6/10093.3
6LiAl0.1Mn1.9O481.6 (j)-72/100088.7[90]
89.9 (j)-77.4/50086.1
7Li1.08 Al0.09 Mn1.83O4Spray drying102 (b)86--[91]
8LiAl0.075Mn1.925O4111 (j)-90.6/100081.6[92]
9LiAl0.08Mn1.92O4Polymer pyrolysis method114 (o)98.6114/50100[83]
116 (o)96.3115.2/5099.3
10LiAl0.1Mn1.9O4Solid state reaction105.6 (m)-73.9/458070[93]
11LiAl0.1Mn1.9O4122.5 (v)91.442.2/7534.7[94]
12LiAl0.04 Mn1.96O4107.1 (d)94.398.5/10092[95]
13Al doped LMOSelf-template100 (l)-78.4/77478.4[96]
100 (l)-72.3/20072.3
14LiAl0.1Mn1.9O4107 (g)92.8--[43]
90 (k)62/50070
15LiAl0.10Mn1.90O4Solution combustion method124.8 (d)-111.1/2089[97]
16LiAl0.1Mn1.9O4122 (d)-113.3/2592.9[98]
17LiAl0.375Mn1.625O4113.1 (d)-96.1/5085[68]
18LiAl0.4Mn1.6O4113 (h)-104/10092[99]
19LiAl0.05Mn1.95O4Combustion method159 (b) *-75.7/5047.6[51]
20LiAl0.1Mn1.9O4146 (b) *-56.5/5038.7
21LiAl0.5Mn1.5O495 (b) *-67.3/5070.8
22LiAl0.16Mn1.5O4100.7 (g)-92.8/40093.9[82]
* 1 C = 296 mA·g−1 in the voltage range 2.4–4.8V vs. Li [51].
Table 3. Rare-earth doped LMO.
Table 3. Rare-earth doped LMO.
Nr. Crt.MaterialSynthesis MethodDischarge CapacityCoulombic EfficiencyDischarge Capacity/N CyclesCapacity RetentionRef
(mAh g−1)(%)(mAh g−1)(%)
1LiLa0.01Mn1.99O4Sol-gel127.6 (d)
109.8 (i)
92.1116.8 (c)
92.5 (h)
91.6/0.2 C
84.3/1 C
[64]
2LiCe0.1Mn1.9O4101 (i)-98/15097[19]
3LiLa0.01Mn1.99O4122.8 (p)96.4113.5/5092.4[100]
4LiCe0.05Mn1.95O4126 (r)94.7117/10092[101]
5LiPr0.2Mn1.98O4118 (f)-115.6/10098[103]
6LiDy0.05Mn1.95O454.4 (g)-49.1/5093[102]
7LiDy0.2Mn1.8O4124 (b)8185.6/1069[65]
8Li Ce0.02Mn1.98O4Solid state reaction119.6 (v)97.6108.5/50 91.0[104]
9LiLa0.1Mn1.9O4100 (g)92.296.9/10096.9[105]
101.4 (g)-95.8/5094.5
10LiCe0.1Mn1.9O499.1 (g)9692.4/10093.2
101.2 (g)-93.4/5092.3
11LiNd0.1Mn1.9O499.4 (g)9195.2/10095.8
101.3 (g)-92.4/5091.2
12LiSm0.1Mn1.9O499.8 (g)99.293.6/10093.8
102.1 (g)-94.1/5092.2
Table 4. Cations doped lithium manganese oxide.
Table 4. Cations doped lithium manganese oxide.
Nr. Crt.MaterialSynthesis MethodDischarge Capacity
(mAh g−1)
Coulombic Efficiency
(%)
Discharge Capacity/N Cycles
(mAh g−1)
Capacity Retention
(%)
Ref
1LiRu0.25Mn1.75O4Solid state reaction125 (c)85113.8/2591[109]
2Li1.05Co0.10Mn1.85O481.3 (d)97.380/3098.4[110]
3Co-LiMn2O4 nanotubes126.6 (b)93.1100/5079[53]
4Li1.02Nb0.01Mn1.99O4127.9 (t) 113.2/6088.5[111]
5LiCr0.04Mn1.98O4Solid State Method119.6 (i)-109.9/35091.9[112]
LiCr0.15Mn1.85O4147.2 (i)106.6--[113]
6LiCr0.1Mn1.9O4Sol-gel123.6 (s)-123/1099.5[114]
7LiW0.025Mn1.975O477.1 (i)-42/5054.5[88]
8LiTixMn2−xO4111.2 (b)94.989.2/100 d)83.7[115]
9LiSn0.02Mn1.98O4123.1 (v)94.6111.7/10090.7[116]
114.0 (v)-92.7/5081.3
10Li1.05Cr0.1Mn1.85O4113.7 (t)-108.6/3895.5[117]
11LiCo0.15Mn1.85O494.4 (u) 90.6/3096[118]
12LiZn0.10Mn1.90O4122 (g)96.1107/10087[119]
13Li1.03Mg0.2Mn1.77O488 (n)8887/1098.9[106]
14Li1.03Fe0.2Mn1.77O4110 (n)-109/1099.1
15Li1.05Co0.05Mn1.95O4120.7 (b)95.7116.2/2096.3[120]
16LiCr0.2Mn1.8O4105.4 (r)10054.1/10000/2000 mA/g73[121]
17LiGd0.02Mn1.98O4116.0 (d)-98.6/7085[86]
116.5 (i)-92.04/7079
18LiGd0.05Mn1.95O459.4 (g)-48.4/5086[102]
19LiFe0.1Mn1.9O4134.7 (g)85.492.5/1400 100C92[52]
--75.8/1500 100C73(d)
20LiBa0.02Mn1.98O4Combustion method115.2 (i)-111.2/3096.5[122]
21LiMg0.02Mn1.98O4107.2 (d)87.495/10088.62[123]
22LiNi0.1Mn1.9O4Solution combustion method116.4 (b)-115.2/5099[124]
23LiNi0.1Mn1.9O4116.4 (h)-115.2/5099
24LiCr0.02Mn1.98O4Molten salt combustion120.7 (d)8894.7/4078.5[107]
25LiSi0.05Mn1.95O4Freeze-drying method146 (b)
139 (i)
-105.5/30075.9[108]
26LiGa0.1Mn1.9O499 (b)
95 (i)
-80.2/30085
27LiCo0.05Mn1.95O4Biomimetic synthesis106.9 (q)92.193.6/10087.6[125]
28LiCr0.12Mn1.88O4Ionothermal synthesis129.6 (g)96.6125.5/20096.8[126]
29Co-doped LiMn2O4Rheological phase reaction112.5 (d)88.4100/10088.9[41]
30Li1.1Nb0.03Mn1.87O4Ion implanted method120.5 (i)-113.8/5094.4[127]
122.1 (i)-112.5/5092.1
Table 5. Anions doped LMO.
Table 5. Anions doped LMO.
Nr. Crt.MaterialSynthesis MethodDischarge Capacity
(mAh g−1)
Coulombic Efficiency
(%)
Discharge Capacity/N Cycles
(mAh g−1)
Capacity Retention
(%)
Ref
1LiMn2O4−xSxPlasma assisted solid state method125.2 (d)-122.4/6097.8[33]
123.5 (b)-108.2/6087.6
2LiMn2O3.94F0.06Solid state reactions116 (d)98.4107.8/10093[95]
3LiMn2O3.99S0.01Sol-gel126 (i)83.2116.6/7092.1[44]
4LiMn2O3.99S0.01Sol-gel158.9 (b) *95.5--[45]
5LiMn2O3.96F0.4Combustion method101.2 (d)-74.5/4073.6[34]
6Li1.017Mn1.97O4(PO4)0.015Hydrothermal method116 (i)98.3--[132]
* in the case of LiMn2O3.99S0.01 [45] the discharge capacity is considered to be after the 5th cycle.
Table 6. Cations multi-doping of LMO.
Table 6. Cations multi-doping of LMO.
Nr. Crt.MaterialSynthesis MethodDischarge Capacity
(mAh g−1)
Coulombic Efficiency
(%)
Discharge Capacity/N Cycles
(mAh g−1)
Capacity Retention
(%)
Ref
1LiZn0.10Pr0.10Mn1.75O4Sol-gel130 (b)76.5128.8/1099.1[136]
2LiZn0.10Al0.15Mn1.75O4122 (a)71.3112/1091.8[137]
3LiZn0.10Ho0.10Mn1.75O4127 (a)91.0115/1090.6[72]
4LiCr0.5Mg0.05Mn1.45O4123 (b)81.5112/1091.1[74]
5LiCu0.5Cr0.05Mn1.45O4124 (b)79113/1091.1[75]
6LiCu0.5Al0.05Mn1.45O4120 (b)78114/1095[138]
7Li0.94Na0.06Al0.1Mn1.9O4115.9 (d)---[71]
8LiNi0.5Al0.05Mn1.45O4126 (b)86116/1092.0[73]
9LiZn0.01Ce0.01Mn1.98O4124 (b)97.6123/1099.2[139]
10LiMg0.05Si0.05Mn1.9O4126.9 (g)-123.5/10097.3[140]
126.4 (g)117.1/2092.7 (g)
11LiAl0.025W0.025Mn1.95O445.6 (i)-36.2/5079.4[88]
12Li1.01Co0.05Ni0.099Mn1.83O491.5 (u)-89.7/3098[118]
13Li1.01Co0.05Zn0.094Mn1.82O486.7 (u)-85.8/3099
14Li1.01Co0.048Cu0.099Mn1.82O485.7 (u)-84.1/3098.1
15Li1.03Mg0.1Fe0.1Mn1.77O4110 (n)-107/1097.3[106]
16LiMg0.01Sn0.04Al0.35Mn1.6O4114.6 (b)61.8113.5/1099[35]
17LiLa0.01Zn0.01Mn1.98O4118.0 (f)-99.1/10084[136]
18Li1.03Zn0.03Mg0.03Mn1.91O4107.5 (f)97.4107.4/4097.4[141]
19LiNi0.01Cr0.01Mn1.98O4113 (b)-98/10087[133]
20Li1.00 Co0.028Ni0.026Mn1.95O4116.1 (i)99.4112.6/20097[40]
115 (i)-110.4/10096
21Li(GdNi)0.01Mn1.98O467.199--[142]
22LiAl0.0125Si0.0125Ti0.0125Co0.0125O4Combustion process111.8 (i)-106.7/3090[135]
23LiCo0.025Cr0.025Ni0.025Fe0.025Mn1.9O4105.6 (i)91.885.5/7081[143]
24LiBi0.005La0.005Mn1.99O4Solution combustion method130.2 (i)97.9123.7/8095[21]
25LiGe0.1Li0.02Mn1.88O4Solid state reaction124 (i)98.463/10050.8[144]
26LiSr0.1Cr0.1Mn1.8O4128 (i)93.3116/10090.6[29]
27LiAl0.125Co0.125Mn1.75O4107.3 (i)89.990.2/35084.1[145]
108.2 (j)96.182.4/18085.7
28LiAl0.1Ni0.1Mn1.8O4140.5 (v)95.666.2/7547.4[94]
29LiAl0.03Si0.05Mg0.05Mn1.87O4118.7 (g)-115.2/5097.1[146]
30LiNi0.03Mo0.01Mn1.96O4114 (i)95.6/30091.2[147]
31LiCe0.012Nd0.012Mn1.976O4Pechini process123.5 *-113.2/3091.6[148]
32LiCu0.02Al0.02Ti0.02Mn1.94O4134 (g)-130/40097[149]
131.2 (g)-119.2/5091
33LiAg0.05Y0.05Mn1.9O4Co-precipitation method137 **99--[150]
34LiZn0.15W0.10Mn1.75O4120 (b)95.299/2582.5[76]
35Li1.088Al0.037Co0.028Mn1.847O4Hydrothermal method102.9 (i)-97.3/10094.5[151]
* Discharge rate not available. ** Discharge rate not available; the sample was also irradiated with 30 kGy (non-irradiated sample delivered 129 mAh·g−1).
Table 7. Anion and cation multi-doping of LMO.
Table 7. Anion and cation multi-doping of LMO.
Nr. Crt.MaterialSynthesis MethodDischarge CapacityCoulombic EfficiencyDischarge Capacity/N CyclesCapacity RetentionRef
(mAh g−1)(%)(mAh g−1)(%)
37LiNi0.4Mn1.6O4-δSδCombustion Method126 (t)96.9135.5/25107.1 *[153]
38Li1.05Ga0.02Mn1.98O3.98S0.02Solid state reaction120-115/30095.8[154]
39Li1.05Al0.02Mn1.98O3.98S0.02126-118.4/30094
40LiMg0.1Mn1.9O3.8F0.2121.1 (i)98.9116.3/40096.0[78]
108/40089.2 (g)
41LiMg0.05Mn1.95O3.9Cl0.1125.2 (d)-111.8/10089.3[77]
42LiLa0.02Mn1.98O3.95F0.05123.6-114.6/3092.7[155]
43Li1.02Co0.1Mn1.9O3.98S0.02122.1 (g)86.9109.8/10090[156]
113 (g)70/50 (g)61.9
44LiMn1.96Al0.04O3.94F0.06109.1 (d)93.1103.6/10095[95]
45Li1.02Mg0.05Mn1.93O3.95F0.05115.9 (f)99.7111.8/4096.5[157]
116.3 (f)-105.1/4090.4
46Li[Li0.01Al0.1Mn1.89]O3.9F0.1110 (i)97.3101/10091.8[158]
111 (i)95.9100.9/10090.9
47LiZn0.03Mg0.03Mn1.94O3.95F0.05Sol-gel120.1 (t)93.5118.3/4098.5[159]
48LiLa0.01Mn1.99O3.99F0.01126.4 (e)96.9116.3/5092[100]
49Li0.99K0.01Mn2O3.99S0.01117.8 (i)86.4108.5/7092.1[44]
50LiCo0.01Al0.05Mn1.94O3.95F0.05122.9 (f)96.7118. 9/4096.8[160]
120.8 (f)-111.7/4092.5
51LiNi0.03Mn1.97O3.95F0.05120.3 (b)-113.7/10094.5[161]
96.7/100g)80.4
52Li1.05Mn1.84Al0.11O3.99F0.01Spray drying process113 (b)98--[91]
53Li1.09Mn1.87Nb0.031O3.99(PO4)0.021Ion implanted method119.2 (i)-115.6/5097.0[127]
121 (i)111.3/5092.0
Table 8. Graphene-coated LMO and its electrochemical properties.
Table 8. Graphene-coated LMO and its electrochemical properties.
No.MaterialSynthesisDischarge Capacity
(mAh g−1)
Coulombic Efficiency
(%)
Discharge Capacity/N Cycles
(mAh g−1 N−1)
Capacity Retention
(%)
Ref
1LiMn2O4/grapheneHydrothermal method143.6 (x) *-137.4/5095.7[179]
141.5 (x) **133.4/5094.3
2Sputtering + CVD131.6 ***-119/750 ***90.4[169]
3Ball-milling method78.893.764.6/50082[186]
4Solution method131.1 (b)-125.3/5096[178]
5Electrophoretic deposition49.6 (r)---[187]
6Spray drying combustion process139 (b)92.512187[193]
7Hydrothermal method139.2 (i)98.2117.5/20084.5[194]
8Self-assembly process+Solid-state lithiation155 (d)94.9149.8/8096.6[168]
* t = 30 °C, ** t = 45 °C, *** the considered unit of measurement is mAh·cm−3.
Table 9. Comparative data of pristine and CeO2 modified LMO at high current rates.
Table 9. Comparative data of pristine and CeO2 modified LMO at high current rates.
MaterialInitial Discharge Capacity (mAh g−1)Ref.
5 C20 C15 C20 C30 C
Pristine LiMn2O48879--23[164]
CeO2 1 wt.%-LiMn2O4108100--35
125121117102-[197]
Table 10. Other surface modification of LMO.
Table 10. Other surface modification of LMO.
Nr. Crt.MaterialSynthesis MethodDischarge Capacity
(mAh g−1)
Coulombic Efficiency
(%)
Discharge Capacity/N Cycles *
(mAh·g−1)
Capacity Retetion
(%)
DLi × 1010 (cm2·s−1)Rct
(Ω)
Ref
13 wt.% FePO4-LiMn2O4Chemical deposition method122.0 (d)-83/8068--[216]
118.2 (d)78/8066-
21 wt.% FePO4-LiMn2O4124.0 (d)98%118.7/10093.5--[217]
127.0 (g)-68/10053.5-
33% LaF3-LiMn2O4118.7 (i)99.9107/10090.1--[163]
117.1 (i)99.598.6/10084.2-
43% AlP-LiMn2O4124.2 (i)-116.7/10094 66/100 cycles *[30]
-108.1/10087
52% AlF3-LiMn2O4103.4 (i)9092.9/5089.8--[221]
61% Li3PO4-LiMn2O4Ball milling method450 °C (5 h)114.2(i) 97.1/10085-293[222]
710% LiNi0.5Mn1.5O4- LiMn2O4Sol-gel112.0 (q)-107.8/5096.3--[1]
109.5 (q)-98.6/10090
80.3% Li3BO3-LiMn2O4118.2 (d)90108.7/4092 14.3/20[223]
119.3 (d)-104.5/4087.822/20
95% FeF3-LiMn2O4110.7 (d)-75.5/20068.2--[224]
108.1 (d)-66.5/10061.5
103% YPO4-LiMn2O4107.0 (d)-90.0/20084.1-151.7[218]
107.2 (d)-86.1/10080.3
113% LaF3-LiMn2O4115.9 (d) 84.3/20072.7- [225]
115.3 (d) 81.1/10070.3
12LiCoO2-LiMn2O4117.7 (j)-111.1/10094.4--[219]
117.2 (j)97.6110.1/10093.6
130.5% Li2CuO2-Li2NiO2 solid solution-LiMn2O4121.2 (i)-113/30093.2 [166]
12 1 **9698.1/20081.1-49.23
143% Li2ZrO3-LiMn2O4126.7 (i)97.5125.5/10099.0-34.2[215]
129.5 (i)99.8116.8/10090.2
15La-Sr-Mn-O-LMO129.9 (b)120.0 (i)90.7108.1/500/1 C90.62.54 × 10242.1[130]
107.3 (i)-97.4/13090.8
162% Li1.2Ni0.2Mn0.6O2-LiMn2O4116.6 (i)-111.8/20095.88-192.6/100 cycles[26]
-/20089.74
173% LaMnO3-LiMn2O4114.0 (g)106.0 (l)95.097/50091.55.12 × 10−1-[37]
183% Li2MnO3-LiMn2O4113.3 (i)-106.7/50094.17-75.2/50 cycles[226]
115.0 (i)-103.2/20089.75
19LiNi0.5Mn1.5O4-LiMn2O4101.0 (i)94.4----[227]
203% La0.7Sr0.3Mn0.7Co0.3O3-LiMn2O4 *123 (d)108.0 (x)95.8101/100 x)93.52.154.05[220]
21LiMn1.912Ni0.072Co0.016O4Co-precipitation118.0 (i)97113.4/20096.18 × 1045.31[228]
223% LaF3-LMO118.4 (i)97.5118.04/5099.74.67 × 102-[131]
233% FePO4-LiMn2O4105.3 (g)-93.3/10088.69.85171/20[229]
243% LiMnPO4-LiMn2O4Hydrothermal reaction112.4 (g)94.5104/30092.55.32 × 10−2400.6[230]
115.7 (g)94.2111.7/10096.6
107.0 (g)8375.1/50070.2
252 wt.% LaPO4-LiMn2O4Polymeric process123.9 (g)103.0 (g)95.4----[231]
94.582/10081.5
* t = 60 °C ** 1 C = 150 mAh g−1.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Marincaş, A.-H.; Ilea, P. Enhancing Lithium Manganese Oxide Electrochemical Behavior by Doping and Surface Modifications. Coatings 2021, 11, 456. https://doi.org/10.3390/coatings11040456

AMA Style

Marincaş A-H, Ilea P. Enhancing Lithium Manganese Oxide Electrochemical Behavior by Doping and Surface Modifications. Coatings. 2021; 11(4):456. https://doi.org/10.3390/coatings11040456

Chicago/Turabian Style

Marincaş, Alexandru-Horaţiu, and Petru Ilea. 2021. "Enhancing Lithium Manganese Oxide Electrochemical Behavior by Doping and Surface Modifications" Coatings 11, no. 4: 456. https://doi.org/10.3390/coatings11040456

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