Research progress on design strategies, synthesis and performance of LiMn2O4-based cathodes

Fangxin Maoa, Wei Guo*b and Jianmin Ma*ac
aKey Laboratory for Micro-/Nano-Optoelectronic Devices of the Ministry of Education, School of Physics and Electronics, Hunan University, Changsha 410082, P. R. China. E-mail: nanoelechem@hnu.edu.cn
bCollege of Chemistry and Chemical Engineering, Anyang Normal University, Anyang 455000, China. E-mail: wgnk@mail.nankai.edu.cn
cInstitute for Superconducting and Electronic Materials, University of Wollongong, Wollongong 2500, Australia

Received 19th October 2015 , Accepted 21st November 2015

First published on 24th November 2015


Abstract

Spinel LiMn2O4 (LMO)-based composites, due to their combination of low toxicity, abundant natural resources, and excellent electrochemical performance, are regarded as promising candidate cathode materials for lithium ion batteries. Current energy storage demands are not being met with existing materials, however, because of their defects, such as fast capacity fading, low rate capability, and low specific capacity in practical applications. Manganese dissolution during electrochemical processes bears the major responsibility for capacity loss, apart from the electrolyte factor. Low electrical conductivity, low ionic diffusion efficiency, and large structural variation have adverse effects on the electrochemical performance of materials. With respect to these drawbacks, significant progress has been made recently on optimizing the performance of LMO-based cathode materials. In this work, we review recent progress in 1 structural design, designing composites with graphene/carbon nanotubes, crystalline doping, and coatings for improving the electrochemical performance of these cathode materials.


1 Introduction

As the supply of fossil fuels becomes deficient and aggravates global distribution, sustainable and clean energy is becoming more and more critical for supporting electric vehicles and other electronic devices. Energy storage is needed to take full advantage of renewable electricity generation (wind, wave and solar).1 For energy storage, the lithium-ion secondary battery is the most promising device because of its high energy and power density, long cycle life, and high reliability, which are important considerations for the performance of electric vehicles and hybrid electric vehicles. In addition, its high gravimetric and volumetric energy densities provide advantages for digital electrical products, such as laptops, cell phones, and even microelectronic devices.2,3

Typically, a basic Li-ion battery consists of a cathode (positive electrode) and an anode (negative electrode), which are immersed in a specific electrolyte and isolated from each other with a separator. The lithium ions shuttle between the cathode and anode through the electrolyte and electrons move through the external circuit during the charging and discharging process, which is shown in Fig. 1 in a sketch of a classical commercial lithium ion cell from the 1970s, using LiCoO2 and graphite as the cathode and anode, respectively. During the discharging process, the cathode accepts the electrons from the external load and lithium ions in the electrolyte. Conduction happens in two ways in the cycle circuit, via electrons and ion transfer. As it is known that the ionic mobility in an electrolyte is much smaller than the electrical conductivity in a metal, a cell needs a large contact surface between the electrodes and electrolyte. And a reversible loss of capacity occurs due to the limit of ionic diffusion while charging/discharging at a high rate is what leads to a loss of Li inserted into an electrode particle. However, in an electrochemical process, changes in electrode volume, electrode–electrolyte reaction, and/or electrode decomposition can cause an irreversible loss of capacity. The chemical reaction between an electrode and electrolyte results in the irreversible formation of a passivating solid-electrolyte interphase (SEI) layer when the initial charge of a cell is made in a discharge state.4


image file: c5ra21777f-f1.tif
Fig. 1 Illustration to show the basic components and operation principle of a Li-ion cell. Reproduced with permission.6

Although there has been research on Li-ion batteries for decades, there are many challenges involved to improve practical Li-ion secondary batteries, for example, improving their poor capacity and drastic capacity fading, which are keeping these batteries from meeting the ever-increasing demands for energy storage and limit their further spread to new applications. The properties of electrode active materials in the anode and cathode determine Li-ion battery performance to a large extent. Lithium manganese oxide-based materials have been attracting much attention as cathode materials because they have particular advantages, such as low cost, high potential platform, and high rate capability.5

The electrochemical reaction (intercalation/deintercalation of Li ions) that occurs in the LiMn2O4 cathode can be expressed as follows:

 
LiMn2O4 ↔ Li1−δMn2O4 + δLi+ + δe (1)
For this material, electrochemical intercalation/deintercalation of Li-ions occurs in the potential range of 3.0–4.5 V vs. lithium electrode, while reaction (1) occurs in two stages in the composite range of 0 ≤ δ ≤ 0.5 and 0.5 ≤ δ ≤ 1, and it has a theoretical specific capacity of 148 mA h g−1.7

Lithium manganese oxide has a cubic spinel crystal structure, as shown in Fig. 2, in which oxygen ions occupy the 32e tetrahedral positions and exhibit dense cubic packing, manganese ions occupy the 16d octahedral positions, and the 8a tetrahedral positions are occupied by lithium ions. A site occupied by a lithium ion is separated from the four neighboring ions by voids at 16c, so three-dimensional (3-D) channels (8a–16c–8a–16c) provide a passage for potential migration of lithium ions in the crystal body.7,8 In an insertion cathode, the 3-D passages provide a chance for high rate capability.9 In addition, Kalantarian et al.10 concluded that the LiMn2O4 crystal structure could be interpreted as that of an n- or p-type semiconductor during the delithiation or lithiation process, respectively, considering the density of states calculated by several density functional theory methods, suggesting that the structure could sustain high current rates. Yamada et al.11 found that a LiMn2O4 thin film electrode requires less activation energy for transferring Li-ions, with a smaller increase after cycling compared with the LiCoO2 thin film electrode. During the delithiation or lithiation process, the composite can retain the stable structure of the spinel cubic. High voltage and tailoring of the potential window could be achieved, as has been reported in many papers, using LiMn2O4-based materials synthesized by a variety of methods.12,13


image file: c5ra21777f-f2.tif
Fig. 2 Supercell model for spinel-type LiMn2O4 after structural optimization. Large gray spheres are Li, medium blue (yellow) spheres are Mn3+ (Mn4+), and small red spheres are O.5 Copyright 2014, Royal Society of Chemistry.

Despite the above-mentioned advantages, fast capacity fading during cycling at high temperature, which is due to such possible factors as manganese dissolution, and Jahn–Teller distortion, etc., seriously limits such materials’ practical life span.14–17 As a semiconductor, LiMn2O4 has a band gap of 1.3 eV and low electrical conductivity (∼10−6 S cm−1), which leads to an inferior rate capability.18 Some research has been done on the electrochemical behavior and degradation mechanism in Li-ion batteries. Tang et al.19 investigated the electrochemical behavior and surface structure of LiMn2O4 when charged to high voltage, and found that drastic changes in the atomic-level structure occurred on the surface, in conjunction with faster manganese dissolution, which led to a degraded electrochemical performance. Lee et al.20 utilized in situ transmission electron microscopy (TEM) technology to separately research the local phase transformations during the lithium-ion de-intercalation process, and concluded that a cubic–tetragonal transition takes place during discharging, but not during the charging process, which is explained by the different diffusion rates of lithium ions between the surface and the bulk. The stability of the crystal phase determines the electrochemical performance of LiMn2O4 cathode materials to some extent. It is accepted by the majority of researchers that decreasing the Mn dissolution could alleviate the degradation of these cathode materials independently from the electrolyte content.21 Elevated temperature will exacerbate the performance degeneration due to the enhancement of one or more circumstances of manganese dissolution, irreversible transition of the crystal phase and oxygen deficiency.22 The capacity fade model for the spinel LMO cathode built by Dai et al.23 showed that 16% of the capacity degenerated after 50 cycles at C/3 and 55 °C between 3.5 and 4.5 V, which was approximate to the experimental result, and even more serious attenuation occurred at high rate cycles. Many factors lead to the fading of capacity, for example, the Li ion diffusion coefficient decreased from 3.5 × 10−15 m2 s−1 to less than 2 × 10−15 m2 s−1 ranging from the second cycle to the 50th cycle at 55 °C. Thus, increasing the resistance of Li ion migration in the surface films also accounted for the fading. It has become a brutal challenge to overcome the series of problems caused by high temperature.

Some important quantities should be endowed to an excellent cathode, such as high operating potential, fast electrochemical reaction kinetics with Li-ions and electrons, stable structure for fast delithiation or lithiation, short diffusion distance for electrons and Li-ions, high ionic diffusivity and electrical conductivity.24 In addition, the thermostability of the materials should be noted to cope with various work circumstances caused by system thermogenesis. Through understanding the importance of the state of the crystal and surface, many reasonable strategies can be designed to improve the performance of the lithium manganese oxide cathode. In this paper, we review recent research progress on lithium manganese oxide-based cathode materials, with the focus on improving the cathode capacity and cycling performance through structural design, the use of composites with graphene/carbon nanotubes, crystalline doping, and coating methods.

2 Structural design

To enhance the battery performance, developing nanostructured electrode materials represents one of the most attractive strategies.26 Islam and Fisher27 have examined the fundamental features important to cathode performance, including voltage trends, ion diffusion paths and dimensions, intrinsic defect chemistry, and surface properties of nanostructures using computational techniques. And tuning the structure and properties of nanostructured cathode materials has been reported.28 Shukla and Kumar29 have summarized the use of nanoarchitectures, which could lead to improvements in terms of electrical and ionic conductivities, diffusion and mass transport, and electron transfer in electrochemical energy storage. It is generally recognized that a smaller particle size of the active material results in a shorter diffusion distance for lithium ions during electrochemical processes, which may be beneficial for battery performance. Smaller particles have greater structural integrity than bulk ones because of fewer pre-existing defects, as concluded by Raghavan et al.,30 which is correlated with changes in the electrode particle surface area caused by particle fracture, and is related to solid electrolyte interphase (SEI) formation, isolation of the active material, and reduction in electrical conductivity. The structure or microstructure of the material grain can directly influence the properties of the electrode containing it.

Ragavendran et al.31 demonstrated that nanoparticles derive their virtue of high rate capability not only from size effects, but also from the shape effect. Han et al.32 synthesized single-phase, high-purity LiMn2O4 nanosized crystal powders as cathode materials for Li-ion batteries, and the results showed that a capacity of 116.6 mA h g−1 was retained after 80 cycles. Cai et al.33 reported LiMn2O4 octahedral nanoparticles with excellent cycling stability and rate capability (initial discharge capacity of 118.5 and 78.3 mA h g−1, and about 72.49% and 94.6% of its initial discharge capacity could be retained, even after 1600 cycles at 10C and 3000 cycles at 20C, respectively). Octahedral and truncated octahedral spinel-type LiMn2O4 has been reported for cathode materials with excellent electrochemical performances.33–36 One-dimensional nanofiber,37–41 nanorod,42–46 and nanowire47,48 structures of spinel LiMn2O4 were studied in several papers. Aravindan et al.,37 Kalluri et al.38 and Kumar et al.39 reported that electrospun nanofibers of LiMn2O4 had outstanding properties for lithium batteries. Recently, Zhou and co-workers41 prepared an ultra-long spinel lithium manganese oxide nanofiber cathode via an electrospinning method, and the cathode materials showed a porous “network-like” morphology with nanosize diameters (∼170 nm), microsize lengths, (∼20 μm) and a pure spinel structure. Their discharge capacity was 146 mA h g−1 at 0.1C; more importantly, the discharge capacities were 112 mA h g−1, 103 mA h g−1, and 92 mA h g−1 at high discharge rates of 10C, 20C, and 30C, respectively. Xie and co-workers46 applied a templating method to synthesize a single-crystalline spinel LiMn2O4 nanorod structure that exhibited a superior long cycle life, retaining 95.6% of the initial capacity after 1000 cycles at a 3C rate, with no deterioration of the morphology. Compared with the one-dimensional structure, spherical or anomalous nano/microscale pieces49–56 are more prone to degradation and have bulk-like electrochemical performance due to their crystalline structure and size effectiveness. It should be noted that a porous cube structure as shown in Fig. 325 composed of single crystalline nanoparticles showed superior long-term cycling stability and high rate capability, delivering a reversible discharge capacity of 108 mA h g−1 at a 30C rate and yielding a capacity retention of over 81% at a rate of 10C after 4000 cycles. Hierarchical porosity and structures are involved in newly developed materials for energy storage, which include the advantages of nanoparticles and eliminate their drawbacks to some degree.57 Hierarchical LiMn2O4 microspheres,58–60 nanofibers,39 and doughnut-shaped61 structures are common. A hierarchical LiMn2O4 phase with a layered nanostructure was also studied by Lee et al., and it exhibited a high discharge capacity and excellent cycling stability at elevated temperature (60 °C).62 Loading LiMn2O4 onto carbon fiber paper to create a binder-free positive electrode63 and aligning multi-walled carbon nanotubes to form elastic and wearable wire-shaped lithium-ion batteries64 are both creative examples of structures designed to improve the material’s properties. Last, but not least, in a novel strategy, laser printing and femtosecond-laser structuring were applied in lithium-ion microbatteries, which optimized the cycling stability and capacity retention, and might enlighten researchers in the future.65 As suggested above, the structure of materials plays an important role in electrochemical processes (Fig. 4).


image file: c5ra21777f-f3.tif
Fig. 3 Architecture of porous LiMn2O4 cubes: (a) SEM image, (b) TEM image, and the cycling performance of LiMn2O4 charged/discharged at rates of 3C and 5C at 25 °C (c) and at 5C at 55 °C (d) (1C = 148 mA g−1).25 Copyright 2014, Royal Society of Chemistry.

image file: c5ra21777f-f4.tif
Fig. 4 (a) Schematic illustration of the synthesis procedure for doughnut-shaped LiMn2O4. TEM images of (b) PS spheres, (c) MnO2@PS, and (d) DS-LiMn2O4.61 Copyright 2015, Royal Society of Chemistry.

3 Graphene/carbon nanotube composites

Graphene and carbon nanotubes (CNTs) are new types of promising materials and have become prime research topics. Here, we review their recent application in lithium manganese oxide cathodes. Graphene has been found to significantly improve cathode electrochemical performance in current studies, although it was previously applied as an electron-conducting additive for lithium ion cathode materials.67 Sreelakshmi et al.68 prepared the active materials in a graphene matrix for use in electrodes and obtained an enhanced energy and power density. Composites of nanocrystalline LiMn2O4 immobilized on the surface of graphene were developed using solvothermal69 and hydrothermal70 methods, and the nanocomposites showed a greatly improved electrochemical performance in terms of the specific capacity, cycling performance, and rate capability, which was attributable both to the improvement of the surface ion transport of nanocrystalline lithium manganate and to the increase in electrical conductivity. Ragavendran et al.71 explained that graphene controlled crystal synthesis through its thermodynamic properties and this led to an orientation in the highly stable (400) direction, which offers superior electrochemical properties in general and much better rate capability in particular. Apart from graphene, one-dimensional carbon nanotubes (CNTs) have been used in cathode composites due to their unique properties. CNTs not only provide a conductive matrix, facilitating fast electron transport, but also effectively reduce agglomeration of the LiMn2O4 nanoparticles.72 Guler et al.73 investigated the effects of multiwall CNT (MWCNT) reinforcement on the electrochemical performance of LiMn2O4/MWCNT nanocomposite cathodes. The results showed that along with increasing MWCNT mass, from 0 wt% to 5.0 wt%, 10.0 wt%, and 15.0 wt%, there was capacity retention of 86.1, 129.4, 134.7, and 136.5 mA h g−1, respectively, after 50 cycles. Higher electrical conductivity, higher structural stability of the composites, and rapid Li+ diffusion, resulting from the open lattice channels and unique one-dimensional structure of the MWCNTs, could be major reasons for these changes. A high specific capacity of 145.4 mA h g−1, close to the theoretical capacity of LiMn2O4 was achieved by a LiMn2O4/MWCNT nanocomposite, which also featured superior rate capability and cyclability.74 A multiwall carbon nanotube network composite with LiNi0.5Mn1.5O4 not only delivered 80% of the 1C capacity at 20C, but also featured a high working potential and very good cycling stability.75 The same improvement effects were achieved by other researchers using a two-step hydrothermal approach (Fig. 5).66,76
image file: c5ra21777f-f5.tif
Fig. 5 TEM images of carbon nanotubes/LiMn2O4 after (a) hydrothermal processing and (b) heat-treatment. (c) Rate performance of CNT/LMO heat treated for different lengths of time.66 Copyright 2013, Elsevier.

4 Crystalline doping

Introducing a guest element into a LMO crystalline cell could affect the valence of Mn, the operating potential plot and the lattice of the crystal, all of which adjust the performance of the LMO cathode. In pure LiMn2O4, manganese ions can be partially replaced by other metal ions, and no damage to the spinel structure occurs, which was revealed by computational methods, but the electrical properties of the material are modified.21 Single or multi-metal doping with metal ions, including Al, Ni, Co, Fe, Mg, Cu, V, Sm and Zn, was extensively investigated for optimizing the electrochemical performance of Li-ion batteries, among which, Ni-doped LiMn2O4 has become one of the current research interests for a high potential cathode.77–87 LiNi0.5Mn1.5O4 is considered to be promising as a cathode material because of its excellent electrochemical performance with an operating voltage of 4.7 V and capacity of 135 mA h g−1.88–95 Hugues et al.96 studied the relationship between the Mn3+ content, structural ordering, phase transformation, and kinetic properties in LiNixMn2−xO4, and revealed that increasing the Mn3+ content triggered the transition from ordered to disordered spinel, which led to increased solid solution behavior, reduced two-phase transformation domains, and improved transport properties during Li extraction and insertion. Further increasing the Mn3+ content, however, in an already disordered structure extends the solid solution domain and eliminates the presence of phase II, so that it yields only a limited effect on rate capability. Shimoda et al.97 used in situ and ex situ Li nuclear magnetic resonance (NMR) spectroscopy to study the delithiation/lithiation behavior of LiNi0.5Mn1.5O4 and clarify the phase reactions during electrochemical processes. Wang et al.98 reported the effect of oxygen deficiency on defect chemistry in delithiated spinel LiNi0.5Mn1.5O4 cathodes, where the results revealed the progress of the atomic-level structural changes: oxygen deficiency promoted the migration of Ni and Mn ions, and the migration assisted the formation of oxygen vacancies. Luo et al.99 reported high-voltage LiNi0.5Mn1.5O4 hollow microspheres with high reversible capacities of 135.5, 147.5, and 132.1 mA h g−1 at 0.1, 0.5, and 2C, respectively. Even at a high rate of 5C, the electrode retained 93.4% of the initial capacity at 0.1C. After investigating the improvement after Al doping, it was concluded by Guo et al.100 that Al-doping facilitated rate and cycling capabilities at room and high temperature: the LiAl0.1Mn1.9O4 sample was cycled at a rate of 5C, and the capacity retention ratio of the electrode after 100 cycles was about 95% at 25 °C and about 90% at 55 °C. Many researchers have shown that LMO-based cathode materials can be endowed with improved cycling stability and cycling performance at elevated temperatures through Al-doping methods.101–106 LMO cathode materials doped with nanoparticles of the bimetallic metal alloys Au–Fe107 and Pt–Au108 were applied in lithium-ion batteries, which showed enhanced conduction and improved cyclability.

In addition to doping with transition metal cations, the semiconductor elements Si,109,110 and Sb,111 and non-metal anions, including F,112,113 Cl,114 B3−,115 and PO43−,116,117 were embedded into LMO crystals to change their attributes. Zhao et al.109,110 researched low-level Si, Mg single and co-doping to improve the electrochemical performance of LMO cathode materials. Results showed that the cubic spinel structure of LiMn2O4 was preserved throughout the introduction of alien atoms. Equimolar Mg2+ and Si4+ ions could completely occupy the octahedral (16d) sites, replacing Mn3+ and Mn4+ ions, respectively, which significantly improved the structural stability and suppressed the Jahn–Teller distortion, so that a far higher reversible capacity was obtained. Cui et al. embedded Sb ions in a LMO crystal, and even though the impure phase of LiSbO3 appeared, the composite featured higher cycling and rate capacities than the pure LMO material. In some ways, the introduction of Si and Sb led to positive effects. The halogen elements fluorine and chlorine could bond with manganese, forming electrovalent bonds that were stronger than Mn–O covalent bonds, so that substitution of stronger bonds for weaker ones led to a more stable structure.114 The introduction of the trivalent anions B3− and PO43− led to increases in the a lattice parameter in both cases, without changing the spinel crystal structure, because their ionic radii were favorable for intercalation/deintercalation of Li-ions.

5 Surface coating

Two major activation processes, desolvation and lattice incorporation, were suggested for the Li-ion exchange reaction at the electrode/electrolyte interface.118 Reactions and charge transfer at cathode–electrolyte interfaces affect the performance and the stability of Li-ion cells.119 As it is well known that Mn dissolution is the major reason for capacity fading of the LMO-based cathode materials, formation of a stable solid electrolyte interface (SEI) film and protection of the integrity of the crystalline structure are effective in improving the electrochemical performance of electrode materials. Surface coating on active materials or the electrode could achieve this goal to some extent, because the outer layer forms a stable solid-electrolyte interphase that could protect Mn from dissolving into the electrolyte, protect the active materials from contact with HF acid, and improve the rate of lithium ion insertion and de-insertion. Wang et al.120 summarized the application of surface and interface engineering in Li-ion batteries, and analyzed surface modifications from active surface and surface functionalization perspectives. Carbon (including organic compounds), lithium-metal–oxygen, metal oxides (Al2O3, MnO2, ZrO2, ZnO, TiO2, etc.), fluorides (AlF3, FeF3, etc.), and some inorganic compounds were proposed and successfully applied to modify the surfaces of electrode materials. Considering that inorganic compounds are poor conductors, the noble metals Au121 and Ag122 were also used to modify the surfaces of active materials by some researchers, and they not only enhanced the conductivity, but also improved the cycle life. Meng et al.123 reported that several of the materials mentioned above had been coated previously onto electrode materials to optimize lithium-ion batteries using the atomic layer deposition (ALD) method.

5.1 Carbon coating

Lee et al.125 reported carbon-coated single-crystal Li2Mn2O4 nanoparticle clusters with high gravimetric and volumetric energy and power density, which delivered 63% of the initial capacity after 2000 cycles at a charge/discharge rate of 20C. Coating with carbon may enhance the specific capacity apart from stabilizing the cycling performance. Noh et al. discovered that LMO with a thin graphitic layer doubled its capacity at a cut-off voltage of 2.5 V, which could be explained by the facile electron-transfer highways provided by the graphitic layer, stabilizing the structural distortion due to the reaction in the 3 V region. A composite of LMO with a graphene-like carbon membrane coating synthesized by an in situ method reached 131.1 mA h g−1 at room temperature, and up to 96% capacity was retained after 50 cycles at 0.1C.126 Conductive polypyrrole (PPy)-coated LiNi0.5Mn1.5O4 (LNMO)127 showed remarkable stability, even at elevated temperature, and good voltage properties. Polypyrrole can increase the electrical conductivity and work as an effective protective layer to suppress the electrolyte decomposition arising from undesirable reactions between the cathode electrode and the electrolyte on the surface of the active material at elevated temperature. Sun et al.124 reported that uniform carbon-coated spinel LiMn2O4 nanowires displayed an ultra-high rate and cycling performance, retaining 83% of the initial capacity after 1500 cycles, even at a rate of 30C. In addition to coating carbon materials directly onto active material particles, an electrode coating can also improve performance in Li-ion batteries. Sputtering a graphitic coating128 and loading an ion exchange polymer coating129 on the surfaces of cathodes have been studied, and enhanced performances were achieved (Fig. 6).
image file: c5ra21777f-f6.tif
Fig. 6 HRTEM image (a) and dark-field STEM image (b) of a carbon-coated LMO nanowire.124 Copyright 2015, Royal Society of Chemistry.

5.2 Metal oxide coating

Loïc et al.131 studied the surface chemistry of metal-oxides (ZnO, Al2O3, and ZrO2) coated on LNMO thin film cathodes via X-ray photoelectron spectroscopy (XPS) and found that ZnO decomposed during the first charge, whereas Al2O3 and ZrO2 were stable for more than 100 cycles, which explains the poor cycle life of the ZnO coated cathode. Al2O3 and ZrO2 coatings demonstrated good performance in other research.132,133 Coating metal oxides with stable electrochemical properties onto the surface of active materials may yield a better improvement. Coatings of the manganese oxides Mn2O3,134 MnO2,135,136 and MnO137 were demonstrated as suitable ways to promote cycle life, due to decreased Mn dissolution after the coating process. Y2O3,138 V2O5,139 and TiO2 (ref. 140) coatings all resulted in appreciable improvement of capacity and cycling stability at elevated temperature, and in addition, excellent rate capability was also found in the case of the former two. Composites with the multi-metal oxides In–Sn–O,141 La–Sr–Mn–O,142 and La0.7Sr0.3Mn0.7Co0.3O3 (ref. 143) coated on LMO composites were also studied, among which, the last composite was also obviously enhanced, especially at high rates (10C, 20C, and 30C), in addition to having an excellent cycling stability. Michalska and Monika144 modified the surface of LMO grains with CeO2 by a low temperature method and obtained a super-stable cycling performance, with only 2% loss after 100 cycles at 1C. RuO2 was applied to modify the surface of LNMO cathode materials, and a good cycling performance was the result: a specific capacity as high as 113.8 mA h g−1 was preserved beyond 1000 cycles (Fig. 7).130
image file: c5ra21777f-f7.tif
Fig. 7 SEM images of RuO2-modified LiNi0.5Mn1.5O4 (a and b); Al2O3 modified LiNi0.5Mn1.5O4 (c and d).130 Copyright 2015, Royal Society of Chemistry.

5.3 Lithium-metal–oxygen coating

Lu et al.146 reported a surface-doping strategy, which combined the bulk-doping strategy and the surface-coating strategy, to yield a TiO2-surface-doped LiMn2O4 material with a LiMn2−xTixO4 layer on the outside of the particles. The same phenomenon occurs for the Al–Ga coating LMO spinel: the LiMn2−xyAlxGayO4 phase was formed on the outside, and an excellent electrochemical performance was obtained from −30 to 55 °C.147 LMO spinel cathode materials coated with lithium-metal–oxygen have a more stable structure. This coating can enhance both the cycling stability, and the specific capacity. Pure spinel LMO with an epitaxial coating of a highly doped spinel has a hierarchical atomic structure consisting of a cubic spinel, tetragonal spinel, and layered structures, and can retain 90% of its capacity after 800 cycles at 60 °C.148 Coating stoichiometric LiMn2O4 with a cobalt-substituted spinel results in good elevated temperature stability and rate capability.149 The high voltage cathode material LiNi0.5Mn1.5O4 was used to modify the surfaces of Li2Mn2O4 cores, and good electrochemical performance was achieved.150–153 Li2TiO3, Li0.17La0.61TiO3, Li2ZrO3, LiNbO3, and LiBO, etc. were researched with the aim of suppressing Mn dissolution and promoting good cell performance (Fig. 8).145,154–159
image file: c5ra21777f-f8.tif
Fig. 8 (a and b) TEM images of LiMn2O4 with a Li0.34La0.51TiO3 coating layer; (c) CV curves of the composites at a scan rate of 0.1 mV s−1 after different cycles at a 1C rate at room temperature; (d) schematic illustration of the surface coating of a cathode.145 Copyright 2015, Royal Society of Chemistry.

5.4 Fluoride/phosphate coatings

As identified by a number of researchers, HF acid attacks the active electrode materials and dissolves the transition metal cations in cells with a LiPF6-based electrolyte, leading to solid deposits on the anode surface through the reduction reaction of the dissolved cations. Metal oxides on the cathode surface turn into metal fluorides, which are expected to be resistant to HF. Coating with fluorides inactive to HF provides an available method to protect active materials and decrease the interfacial charge transfer resistance.120 Wu et al.161 coated MgF2 onto the LiNi0.5Mn1.5O4 surface via a wet coating strategy and obtained an improved electrochemical performance. The Zhao research group studied FeF3 (ref. 162) and LaF3 (ref. 163) coated spinel LiMn2O4 and found that the cycling stabilities were enhanced at both room temperature (25 °C) and elevated temperature (55 °C) for both coatings. Phosphate (MPO4) has been applied as a cathode for lithium-ion batteries and could provide continuous lithium insertion channels and considerable electrical conductivity. Xiao et al.160 coated LiNi0.5Mn1.5O4 with active FePO4 by atomic layer deposition, and the results showed that the cycling stability and the conductivity increased with the coating mass, but there was a lower specific capacity. Li3PO4,164 YPO4 (ref. 165) and SrHPO4 (ref. 166) coatings were found to improve the cycling stability and thermal stability in a LMO cathode (Fig. 9).
image file: c5ra21777f-f9.tif
Fig. 9 (a) SEM and (b) TEM images of LiNi0.5Mn1.5O4 with an active FePO4 coating by atomic layer deposition, with the inset to (b) showing the corresponding electron diffraction pattern; (c) first charge/discharge curves and (d) cycling stability at 0.5C of samples with different coating times. Reproduced with permission.160

6 Conclusions

Spinel lithium-manganese-oxides are one of the most promising candidate cathode materials for Li-ion batteries with a high working potential and high rate performance. Many current challenges, related to cycling stability, thermal stability, rate capability, and specific capacity, have impeded further application in energy storage. Much work has been done and some achievements obtained towards the improvement of the electrochemical performance of LMO-based cathode materials. Several strategies, including structural design, the use of composites, and crystalline and surface engineering, have proven to be available ways to optimize their performance for Li-ion batteries.

To further develop LiMn2O4-based cathode materials, in addition to the strategies mentioned above, real reasons and mechanisms for material degradation during electrochemical processes should be explored clearly, and then better solutions can be studied. In addition: (i) new advanced materials, other than graphene and carbon nanotubes, applied to the LiMn2O4-based cathode should be an available breakthrough, (ii) new strategies for synthesizing LiMn2O4-based cathodes should be explored and (iii) LiMn2O4-based batteries can be improved through exploring novel electrolyte formulations.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 51302079). We also thank Dr Tania Silver from the Institute for Superconducting and Electronic Materials (University of Wollongong) for revising our manuscript.

Notes and references

  1. N.-S. Choi, Z. Chen, S. A. Freunberger, X. Ji, Y.-K. Sun, K. Amine, G. Yushin, L. F. Nazar, J. Cho and P. G. Bruce, Angew. Chem., Int. Ed., 2012, 51, 9994–10024 CrossRef CAS PubMed.
  2. Y. Wang, B. Liu, Q. Li, S. Cartmell, S. Ferrara, Z. D. Deng and J. Xiao, J. Power Sources, 2015, 286, 330–345 CrossRef CAS.
  3. E. M. Erickson, C. Ghanty and D. Aurbach, J. Phys. Chem. Lett., 2014, 5, 3313–3324 CrossRef CAS PubMed.
  4. J. B. Goodenough and K.-S. Park, J. Am. Chem. Soc., 2013, 135, 1167–1176 CrossRef CAS PubMed.
  5. L. Ghadbeigi, J. K. Harada, B. R. Lettiere and T. D. Sparks, Energy Environ. Sci., 2015, 8, 1640–1650 CAS.
  6. D. Deng, Energy Sci. Eng., 2015, 3, 385–418 CrossRef.
  7. A. V. Potapenko and S. A. Kirillov, J. Energy Chem., 2014, 23, 543–558 CrossRef.
  8. K. Hoang, J. Mater. Chem. A, 2014, 2, 18271–18280 CAS.
  9. M. Ø. Filsø, M. J. Turner, G. V. Gibbs, S. Adams, M. A. Spackman and B. B. Iversen, Chem.–Eur. J., 2013, 19, 15535–15544 CrossRef.
  10. M. M. Kalantarian, S. Asgari and P. Mustarelli, J. Mater. Chem. A, 2014, 2, 107–115 CAS.
  11. I. Yamada, K. Miyazaki, T. Fukutsuka, Y. Iriyama, T. Abe and Z. Ogumi, J. Power Sources, 2015, 294, 460–464 CrossRef CAS.
  12. A. Bhaskar, S. Krueger, V. Siozios, J. Li, S. Nowak and M. Winter, Adv. Energy Mater., 2015, 5, 1401156 Search PubMed.
  13. M. D. Levi, V. Dargel, Y. Shilina, V. Borgel, D. Aurbach and I. C. Halalay, J. Power Sources, 2015, 278, 599–607 CrossRef CAS.
  14. Y. Kim, J. Lim and S. Kang, Int. J. Quantum Chem., 2013, 113, 148–154 CrossRef CAS.
  15. L. Cai, Y. Dai, M. Nicholson, R. E. White, K. Jagannathan and G. Bhatia, J. Power Sources, 2013, 221, 191–200 CrossRef CAS.
  16. C.-H. Lu and S.-W. Lin, J. Mater. Res., 2002, 17, 1476–1481 CrossRef CAS.
  17. J. Darul, C. Lathe and P. Piszora, RSC Adv., 2014, 4, 65205–65212 RSC.
  18. J. He, Y. Chen, P. Li, F. Fu, J. Liu and Z. Wang, RSC Adv., 2015, 5, 80063–80068 RSC.
  19. D. Tang, L. Ben, Y. Sun, B. Chen, Z. Yang, L. Gu and X. Huang, J. Mater. Chem. A, 2014, 2, 14519–14527 CAS.
  20. S. Lee, Y. Oshima, E. Hosono, H. Zhou, K. Kim, H. M. Chang, R. Kanno and K. Takayanagi, J. Phys. Chem. C, 2013, 117, 24236–24241 CAS.
  21. A. Kraytsberg and Y. Ein-Eli, Adv. Energy Mater., 2012, 2, 922–939 CrossRef CAS.
  22. G. Xu, Z. Liu, C. Zhang, G. Cui and L. Chen, J. Mater. Chem. A, 2015, 3, 4092–4123 CAS.
  23. Y. Dai, L. Cai and R. E. White, J. Electrochem. Soc., 2013, 160, A182–A190 CrossRef CAS.
  24. Y. Tang, Y. Zhang, W. Li, B. Ma and X. Chen, Chem. Soc. Rev., 2015, 44, 5926–5940 RSC.
  25. H. B. Lin, J. N. Hu, H. B. Rong, Y. M. Zhang, S. W. Mai, L. D. Xing, M. Q. Xu, X. P. Li and W. S. Li, J. Mater. Chem. A, 2014, 2, 9272–9279 CAS.
  26. H. Xia, Z. Luo and J. Xie, Progress in Natural Science: Materials International, 2012, 22, 572–584 CrossRef.
  27. M. S. Islam and C. A. J. Fisher, Chem. Soc. Rev., 2014, 43, 185–204 RSC.
  28. G.-L. Xu, Q. Wang, J.-C. Fang, Y.-F. Xu, J.-T. Li, L. Huang and S.-G. Sun, J. Mater. Chem. A, 2014, 2, 19941–19962 CAS.
  29. A. K. Shukla and T. Prem Kumar, Wiley Interdiscip. Rev.: Energy Environ., 2013, 2, 14–30 CrossRef CAS.
  30. R. S. Raghavan, A. M. Kumar and R. Narayanrao, Z. Angew. Math. Mech., 2014, 1–14 Search PubMed.
  31. K. R. Ragavendran, H. Xia, G. Yang, D. Vasudevan, B. Emmanuel, D. Sherwood and A. K. Arof, Phys. Chem. Chem. Phys., 2014, 16, 2553–2560 RSC.
  32. C.-G. Han, C. Zhu, G. Saito and T. Akiyama, Adv. Powder Technol., 2015, 26, 665–671 CrossRef CAS.
  33. Y. Cai, Y. Huang, X. Wang, D. Jia, W. Pang, Z. Guo, Y. Du and X. Tang, J. Power Sources, 2015, 278, 574–581 CrossRef CAS.
  34. B.-M. Hwang, S.-J. Kim, Y.-W. Lee, H.-C. Park, D.-M. Kim and K.-W. Park, Mater. Chem. Phys., 2015, 158, 138–143 CrossRef CAS.
  35. S. Huang, H. Wu, P. Chen, Y. Guo, B. Nie, B. Chen, H. Liu and Y. Zhang, J. Mater. Chem. A, 2015, 3, 3633–3640 CAS.
  36. G. Jin, H. Qiao, H. Xie, H. Wang, K. He, P. Liu, J. Chen, Y. Tang, S. Liu and C. Huang, Electrochim. Acta, 2014, 150, 1–7 CrossRef CAS.
  37. V. Aravindan, J. Sundaramurthy, P. S. Kumar, N. Shubha, W. C. Ling, S. Ramakrishna and S. Madhavi, Nanoscale, 2013, 5, 10636–10645 RSC.
  38. S. Kalluri, K. H. Seng, Z. Guo, H. K. Liu and S. X. Dou, RSC Adv., 2013, 3, 25576–25601 RSC.
  39. P. S. Kumar, J. Sundaramurthy, S. Sundarrajan, V. J. Babu, G. Singh, S. I. Allakhverdiev and S. Ramakrishna, Energy Environ. Sci., 2014, 7, 3192–3222 CAS.
  40. M. Qian, J. Huang, S. Han and X. Cai, Electrochim. Acta, 2014, 120, 16–22 CrossRef CAS.
  41. H. Zhou, X. Ding, G. Liu, Y. Jiang, Z. Yin and X. Wang, Electrochim. Acta, 2015, 152, 274–279 CrossRef CAS.
  42. Z. Li, L. Wang, K. Li and D. Xue, J. Alloys Compd., 2013, 580, 592–597 CrossRef CAS.
  43. D. Zhan, Q. Zhang, X. Hu, G. Zhu and T. Peng, Solid State Ionics, 2013, 239, 8–14 CrossRef CAS.
  44. D. Zhan, F. Yang, Q. Zhang, X. Hu and T. Peng, Electrochim. Acta, 2014, 129, 364–372 CrossRef CAS.
  45. H. Zhao, F. Li, X. Liu, W. Xiong, B. Chen, H. Shao, D. Que, Z. Zhang and Y. Wu, Electrochim. Acta, 2015, 166, 124–133 CrossRef CAS.
  46. X. Xie, D. Su, B. Sun, J. Zhang, C. Wang and G. Wang, Chem.–Eur. J., 2014, 20, 17125–17131 CrossRef CAS PubMed.
  47. L. Mai, X. Tian, X. Xu, L. Chang and L. Xu, Chem. Rev., 2014, 114, 11828–11862 CrossRef CAS PubMed.
  48. Y. Wang, Y. Wang, D. Jia, Z. Peng, Y. Xia and G. Zheng, Nano Lett., 2014, 14, 1080–1084 CrossRef CAS PubMed.
  49. D. Guo, Z. Chang, B. Li, H. Tang, X.-Z. Yuan and H. Wang, Solid State Ionics, 2013, 237, 34–39 CrossRef CAS.
  50. J. Jiang, K. Du, Y. Cao, Z. Peng, G. Hu and J. Duan, J. Alloys Compd., 2013, 577, 138–142 CrossRef CAS.
  51. Q. Jiang, X. Wang, C. Miao and Z. Tang, RSC Adv., 2013, 3, 12088–12090 RSC.
  52. Y. Wang, X. Shao, H. Xu, M. Xie, S. Deng, H. Wang, J. Liu and H. Yan, J. Power Sources, 2013, 226, 140–148 CrossRef CAS.
  53. Z. Yang, Y. Jiang, H.-H. Xu and Y.-H. Huang, Electrochim. Acta, 2013, 106, 63–68 CrossRef CAS.
  54. D. Guo, Z. Chang, H. Tang, B. Li, X. Xu, X.-Z. Yuan and H. Wang, Electrochim. Acta, 2014, 123, 254–259 CrossRef CAS.
  55. Y. Mizuno, N. Zettsu, H. Inagaki, S. Komine, K. Kami, K. Yubuta, H. Wagata, S. Oishi and K. Teshima, CrystEngComm, 2014, 16, 1157–1162 RSC.
  56. D. Guo, X. Wei, Z. Chang, H. Tang, B. Li, E. Shangguan, K. Chang, X.-Z. Yuan and H. Wang, J. Alloys Compd., 2015, 632, 222–228 CrossRef CAS.
  57. Y. Li, Z.-Y. Fu and B.-L. Su, Adv. Funct. Mater., 2012, 22, 4634–4667 CrossRef CAS.
  58. Y. Gu, Z. Tang, Y. Deng and L. Wang, Electrochim. Acta, 2013, 94, 165–171 CrossRef CAS.
  59. H. Zhang, Y. Xu and D. Liu, RSC Adv., 2015, 5, 11091–11095 RSC.
  60. L. Xiao, Y. Guo, D. Qu, B. Deng, H. Liu and D. Tang, J. Power Sources, 2013, 225, 286–292 CrossRef CAS.
  61. W. Sun, H. Liu, T. Peng, Y. Liu, G. Bai, S. Kong, S. Guo, M. Li and X.-Z. Zhao, J. Mater. Chem. A, 2015, 3, 8165–8170 CAS.
  62. M.-J. Lee, S. Lee, P. Oh, Y. Kim and J. Cho, Nano Lett., 2014, 14, 993–999 CrossRef CAS PubMed.
  63. G. H. Waller, S. Y. Lai, B. H. Rainwater and M. Liu, J. Power Sources, 2014, 251, 411–416 CrossRef CAS.
  64. J. Ren, Y. Zhang, W. Bai, X. Chen, Z. Zhang, X. Fang, W. Weng, Y. Wang and H. Peng, Angew. Chem., 2014, 126, 7998–8003 CrossRef.
  65. J. Pröll, H. Kim, A. Piqué, H. J. Seifert and W. Pfleging, J. Power Sources, 2014, 255, 116–124 CrossRef.
  66. M. Tang, A. Yuan, H. Zhao and J. Xu, J. Power Sources, 2013, 235, 5–13 CrossRef CAS.
  67. G. Kucinskis, G. Bajars and J. Kleperis, J. Power Sources, 2013, 240, 66–79 CrossRef CAS.
  68. K. V. Sreelakshmi, S. Sasi, A. Balakrishnan, N. Sivakumar, A. S. Nair, S. V. Nair and K. R. V. Subramanian, Energy Technol., 2014, 2, 257–262 CrossRef CAS.
  69. K.-Y. Jo, S.-Y. Han, J. M. Lee, I. Y. Kim, S. Nahm, J.-W. Choi and S.-J. Hwang, Electrochim. Acta, 2013, 92, 188–196 CrossRef CAS.
  70. B. Lin, Q. Yin, H. Hu, F. Lu and H. Xia, J. Solid State Chem., 2014, 209, 23–28 CrossRef CAS.
  71. K. Ragavendran, X. Hui, X. Gu, D. Sherwood, B. Emmanuel and A. K. Arof, RSC Adv., 2014, 4, 60106–60111 RSC.
  72. H. Xia, K. R. Ragavendran, J. Xie and L. Lu, J. Power Sources, 2012, 212, 28–34 CrossRef CAS.
  73. M. O. Guler, A. Akbulut, T. Cetinkaya and H. Akbulut, Int. J. Energy Res., 2014, 38, 509–517 CrossRef CAS.
  74. M. Tang, A. Yuan and J. Xu, Electrochim. Acta, 2015, 166, 244–252 CrossRef CAS.
  75. X. Fang, C. Shen, M. Ge, J. Rong, Y. Liu, A. Zhang, F. Wei and C. Zhou, Nano Energy, 2015, 12, 43–51 CrossRef CAS.
  76. B.-K. Zou, X.-H. Ma, Z.-F. Tang, C.-X. Ding, Z.-Y. Wen and C.-H. Chen, J. Power Sources, 2014, 268, 491–497 CrossRef CAS.
  77. J. Liu, W. Liu, S. Ji, Y. Zhou, P. Hodgson and Y. Li, ChemPlusChem, 2013, 78, 636–641 CrossRef CAS.
  78. M. Prabu, M. V. Reddy, S. Selvasekarapandian, G. V. Subba Rao and B. V. R. Chowdari, Electrochim. Acta, 2013, 88, 745–755 CrossRef CAS.
  79. K. Dai, J. Mao, Z. Li, Y. Zhai, Z. Wang, X. Song, V. Battaglia and G. Liu, J. Power Sources, 2014, 248, 22–27 CrossRef CAS.
  80. B. Ebin, S. Gürmen and G. Lindbergh, Ceram. Int., 2014, 40, 1019–1027 CrossRef CAS.
  81. M. Xiang, C.-W. Su, L. Feng, M. Yuan and J. Guo, Electrochim. Acta, 2014, 125, 524–529 CrossRef CAS.
  82. M. Yavuz, N. Kiziltas-Yavuz, A. Bhaskar, M. Scheuermann, S. Indris, F. Fauth, M. Knapp and H. Ehrenberg, Z. Anorg. Allg. Chem., 2014, 640, 3118–3126 CrossRef CAS.
  83. D.-L. Fang, J.-C. Li, X. Liu, P.-F. Huang, T.-R. Xu, M.-C. Qian and C.-H. Zheng, J. Alloys Compd., 2015, 640, 82–89 CrossRef CAS.
  84. Y. Shi, S. Zhu, C. Zhu, Y. Li, Z. Chen and D. Zhang, Electrochim. Acta, 2015, 154, 17–23 CrossRef CAS.
  85. R. Thirunakaran, R. Ravikumar, S. Gopukumar and A. Sivashanmugam, J. Alloys Compd., 2013, 556, 266–273 CrossRef CAS.
  86. M. C. Kim, K.-W. Nam, E. Hu, X.-Q. Yang, H. Kim, K. Kang, V. Aravindan, W.-S. Kim and Y.-S. Lee, ChemSusChem, 2014, 7, 829–834 CrossRef CAS.
  87. A. M. Khedr, M. M. Abou-Sekkina and F. G. El-Metwaly, J. Electron. Mater., 2013, 42, 1275–1281 CrossRef CAS.
  88. Z. Lu, X. Rui, H. Tan, W. Zhang, H. H. Hng and Q. Yan, ChemPlusChem, 2013, 78, 218–221 CrossRef CAS.
  89. Z. Chen, R. Zhao, P. Du, H. Hu, T. Wang, L. Zhu and H. Chen, J. Mater. Chem. A, 2014, 2, 12835–12848 CAS.
  90. J.-H. Kim, N. P. W. Pieczonka and L. Yang, ChemPhysChem, 2014, 15, 1940–1954 CrossRef CAS PubMed.
  91. D. Liu, W. Zhu, J. Trottier, C. Gagnon, F. Barray, A. Guerfi, A. Mauger, H. Groult, C. M. Julien, J. B. Goodenough and K. Zaghib, RSC Adv., 2014, 4, 154–167 RSC.
  92. A. Manthiram, K. Chemelewski and E.-S. Lee, Energy Environ. Sci., 2014, 7, 1339–1350 CAS.
  93. H. Liu, G. Zhu, L. Zhang, Q. Qu, M. Shen and H. Zheng, J. Power Sources, 2015, 274, 1180–1187 CrossRef CAS.
  94. X. Feng, Y. Tian, J. Zhang and L. Yin, Powder Technol., 2014, 253, 35–40 CrossRef CAS.
  95. S. Yang, J. Chen, Y. Liu and B. Yi, J. Mater. Chem. A, 2014, 2, 9322–9330 CAS.
  96. H. Duncan, B. Hai, M. Leskes, C. P. Grey and G. Chen, Chem. Mater., 2014, 26, 5374–5382 CrossRef CAS.
  97. K. Shimoda, M. Murakami, H. Komatsu, H. Arai, Y. Uchimoto and Z. Ogumi, J. Phys. Chem. C, 2015, 119, 13472–13480 CAS.
  98. Z. Wang, Q. Su, H. Deng and Y. Fu, ChemElectroChem, 2015, 2, 1182–1186 CrossRef CAS.
  99. H. Luo, P. Nie, L. Shen, H. Li, H. Deng, Y. Zhu and X. Zhang, ChemElectroChem, 2015, 2, 127–133 CrossRef CAS.
  100. D. Guo, B. Li, Z. Chang, H. Tang, X. Xu, K. Chang, E. Shangguan, X.-Z. Yuan and H. Wang, Electrochim. Acta, 2014, 134, 338–346 CrossRef CAS.
  101. M. A. Kebede, M. J. Phasha, N. Kunjuzwa, L. J. le Roux, D. Mkhonto, K. I. Ozoemena and M. K. Mathe, Sustainable Energy Technologies and Assessments, 2014, 5, 44–49 CrossRef.
  102. X. Yi, X. Wang, B. Ju, Q. Wei, X. Yang, G. Zou, H. Shu and L. Hu, J. Alloys Compd., 2014, 604, 50–56 CrossRef CAS.
  103. F.-D. Yu, Z.-B. Wang, F. Chen, J. Wu, X.-G. Zhang and D.-M. Gu, J. Power Sources, 2014, 262, 104–111 CrossRef CAS.
  104. X. Ding, H. Zhou, G. Liu, Z. Yin, Y. Jiang and X. Wang, J. Alloys Compd., 2015, 632, 147–151 CrossRef CAS.
  105. Y. Fu, H. Jiang, Y. Hu, Y. Dai, L. Zhang and C. Li, Ind. Eng. Chem. Res., 2015, 54, 3800–3805 CrossRef CAS.
  106. F. P. Nkosi, C. J. Jafta, M. Kebede, L. le Roux, M. K. Mathe and K. I. Ozoemena, RSC Adv., 2015, 5, 32256–32262 RSC.
  107. N. West, K. I. Ozoemena, C. O. Ikpo, P. G. L. Baker and E. I. Iwuoha, Electrochim. Acta, 2013, 101, 86–92 CrossRef CAS.
  108. N. Ross, E. I. Iwuoha, C. O. Ikpo, P. Baker, N. Njomo, S. N. Mailu, M. Masikini, N. Matinise, A. Tsegaye, N. Mayedwa, T. Waryo, K. I. Ozoemena and A. Williams, Electrochim. Acta, 2014, 128, 178–183 CrossRef CAS.
  109. H. Zhao, F. Li, X. Liu, C. Cheng, Z. Zhang, Y. Wu, W. Xiong and B. Chen, Electrochim. Acta, 2015, 151, 263–269 CrossRef CAS.
  110. H. Zhao, X. Liu, C. Cheng, Q. Li, Z. Zhang, Y. Wu, B. Chen and W. Xiong, J. Power Sources, 2015, 282, 118–128 CrossRef CAS.
  111. P. Cui, Y. Liang, D. Zhan, Y. Zhao and R. Peng, RSC Adv., 2014, 4, 43821–43827 RSC.
  112. H. R. Lee, B. Lee, K. Y. Chung, B. W. Cho, K.-Y. Lee and S. H. Oh, Electrochim. Acta, 2014, 136, 396–403 CrossRef CAS.
  113. H. Li, Y. Luo, J. Xie, Q. Zhang and L. Yan, J. Alloys Compd., 2015, 639, 346–351 CrossRef CAS.
  114. D.-W. Han, W.-H. Ryu, W.-K. Kim, J.-Y. Eom and H.-S. Kwon, J. Phys. Chem. C, 2013, 117, 4913–4919 CAS.
  115. B. Ebin, G. Lindbergh and S. Gürmen, J. Alloys Compd., 2015, 620, 399–406 CrossRef CAS.
  116. R. A. Rodríguez, Y. M. Laffita, E. P. Cappe, M. A. A. Frutis, J. S. Salazar and O. L. Alves, Ceram. Int., 2014, 40, 12413–12422 CrossRef.
  117. J. Kang, J. Song, S. Kim, J. Gim, J. Jo, V. Mathew, J. Han and J. Kim, RSC Adv., 2013, 3, 25640–25643 RSC.
  118. M. Nakayama, H. Taki, T. Nakamura, S. Tokuda, R. Jalem and T. Kasuga, J. Phys. Chem. C, 2014, 118, 27245–27251 CAS.
  119. R. Hausbrand, D. Becker and W. Jaegermann, Prog. Solid State Chem., 2014, 42, 175–183 CrossRef CAS.
  120. K. X. Wang, X. H. Li and J. S. Chen, Adv. Mater., 2015, 27, 527–545 CrossRef CAS PubMed.
  121. J. L. Esbenshade, M. D. Fox and A. A. Gewirth, J. Electrochem. Soc., 2015, 162, A26–A29 CrossRef CAS.
  122. R. Jiang, C. Cui, H. Ma, H. Ma and T. Chen, J. Electroanal. Chem., 2015, 744, 69–76 CrossRef CAS.
  123. X. Meng, X.-Q. Yang and X. Sun, Adv. Mater., 2012, 24, 3589–3615 CrossRef CAS PubMed.
  124. W. Sun, H. Liu, Y. Liu, G. Bai, W. Liu, S. Guo and X. Zhao, Nanoscale, 2015, 7, 13173–13180 RSC.
  125. S. Lee, Y. Cho, H.-K. Song, K. T. Lee and J. Cho, Angew. Chem., 2012, 124, 8878–8882 CrossRef.
  126. H. Zhuo, S. Wan, C. He, Q. Zhang, C. Li, D. Gui, C. Zhu, H. Niu and J. Liu, J. Power Sources, 2014, 247, 721–728 CrossRef CAS.
  127. X.-W. Gao, Y.-F. Deng, D. Wexler, G.-H. Chen, S.-L. Chou, H.-K. Liu, Z.-C. Shi and J.-Z. Wang, J. Mater. Chem. A, 2015, 3, 404–411 CAS.
  128. J. Wang, Q. Zhang, X. Li, Z. Wang, H. Guo, D. Xu and K. Zhang, Phys. Chem. Chem. Phys., 2014, 16, 16021–16029 RSC.
  129. P. Xue, D. Gao, S. Chen, S. Zhao, B. Wang and L. Li, RSC Adv., 2014, 4, 52624–52628 RSC.
  130. D. Hong, Y. Guo, H. Wang, J. Zhou and H.-T. Fang, J. Mater. Chem. A, 2015, 3, 15457–15465 CAS.
  131. L. Baggetto, N. J. Dudney and G. M. Veith, Electrochim. Acta, 2013, 90, 135–147 CrossRef CAS.
  132. J. Zhao and Y. Wang, Nano Energy, 2013, 2, 882–889 CrossRef CAS.
  133. X. Fang, M. Ge, J. Rong, Y. Che, N. Aroonyadet, X. Wang, Y. Liu, A. Zhang and C. Zhou, Energy Technol., 2014, 2, 159–165 CrossRef CAS.
  134. J. H. Lee and K. J. Kim, Electrochim. Acta, 2013, 102, 196–201 CrossRef CAS.
  135. B. J. Kang, J.-B. Joo, J. K. Lee and W. Choi, J. Electroanal. Chem., 2014, 728, 34–40 CrossRef CAS.
  136. L. Li, W. Qu, F. Liu, T. Zhao, X. Zhang, R. Chen and F. Wu, Appl. Surf. Sci., 2014, 315, 59–65 CrossRef CAS.
  137. J. Zeng, M. Li, X. Li, C. Chen, D. Xiong, L. Dong, D. Li, A. Lushington and X. Sun, Appl. Surf. Sci., 2014, 317, 884–891 CrossRef CAS.
  138. B. Ju, X. Wang, C. Wu, X. Yang, H. Shu, Y. Bai, W. Wen and X. Yi, J. Alloys Compd., 2014, 584, 454–460 CrossRef CAS.
  139. J. Wang, S. Yao, W. Lin, B. Wu, X. He, J. Li and J. Zhao, J. Power Sources, 2015, 280, 114–124 CrossRef CAS.
  140. Y. Shang, X. Lin, X. Lu, T. Huang and A. Yu, Electrochim. Acta, 2015, 156, 121–126 CrossRef CAS.
  141. C.-S. Kim, S.-H. Kwon and J.-W. Yoon, J. Alloys Compd., 2014, 586, 574–580 CrossRef CAS.
  142. H.-Q. Wang, F.-Y. Lai, Y. Li, X.-H. Zhang, Y.-G. Huang, S.-J. Hu and Q.-Y. Li, Electrochim. Acta, 2015, 177, 290–297 CrossRef CAS.
  143. T. Shi, Y. Dong, C.-M. Wang, F. Tao and L. Chen, J. Power Sources, 2015, 273, 959–965 CrossRef CAS.
  144. M. Michalska, B. Hamankiewicz, D. Ziółkowska, M. Krajewski, L. Lipińska, M. Andrzejczuk and A. Czerwiński, Electrochim. Acta, 2014, 136, 286–291 CrossRef CAS.
  145. S. Hu, Y. Li, F. Lai, X. Zhang, Q. Li, Y. Huang, X. Yuan, J. Chen and H. Wang, RSC Adv., 2015, 5, 17592–17600 RSC.
  146. J. Lu, C. Zhan, T. Wu, J. Wen, Y. Lei, A. J. Kropf, H. Wu, D. J. Miller, J. W. Elam, Y.-K. Sun, X. Qiu and K. Amine, Nat. Commun., 2014, 5, 5693 CrossRef CAS PubMed.
  147. F.-Y. Hung and K.-Y. Yang, J. Power Sources, 2014, 268, 7–13 CrossRef CAS.
  148. S. Lee, G. Yoon, M. Jeong, M.-J. Lee, K. Kang and J. Cho, Angew. Chem., 2015, 127, 1169–1174 CrossRef.
  149. M. Jeong, M.-J. Lee, J. Cho and S. Lee, Adv. Energy Mater., 2015, 1500440–1500447 Search PubMed.
  150. Z. Han, X. Jia, H. Zhan and Y. Zhou, Electrochim. Acta, 2013, 114, 772–778 CrossRef CAS.
  151. Y. Chae, J. K. Lee and W. Choi, J. Electroanal. Chem., 2014, 730, 20–25 CrossRef CAS.
  152. W. Liu, J. Liu, K. Chen, S. Ji, Y. Wan, Y. Zhou, D. Xue, P. Hodgson and Y. Li, Chem.–Eur. J., 2014, 20, 824–830 CrossRef CAS PubMed.
  153. T. Qiu, J. Wang, Y. Lu and W. Yang, Electrochim. Acta, 2014, 147, 626–635 CrossRef CAS.
  154. H. Deng, P. Nie, H. Luo, Y. Zhang, J. Wang and X. Zhang, J. Mater. Chem. A, 2014, 2, 18256–18262 CAS.
  155. S. Kim, M. Hirayama, K. Suzuki and R. Kanno, Solid State Ionics, 2014, 262, 578–581 CrossRef CAS.
  156. X. Yi, X. Wang, B. Ju, H. Shu, W. Wen, R. Yu, D. Wang and X. Yang, Electrochim. Acta, 2014, 134, 143–149 CrossRef CAS.
  157. Z.-J. Zhang, S.-L. Chou, Q.-F. Gu, H.-K. Liu, H.-J. Li, K. Ozawa and J.-Z. Wang, ACS Appl. Mater. Interfaces, 2014, 6, 22155–22165 CAS.
  158. C. Du, M. Yang, J. Liu, S. Sun, Z. Tang, D. Qu and X. Zhang, RSC Adv., 2015, 5, 57293–57299 RSC.
  159. J.-H. Kim, N. P. W. Pieczonka, P. Lu, Z. Liu, R. Qiao, W. Yang, M. M. Tessema, Y.-K. Sun and B. R. Powell, ACS Appl. Mater. Interfaces, 2015, 2, 1500109–1500122 Search PubMed.
  160. B. Xiao, J. Liu, Q. Sun, B. Wang, M. N. Banis, D. Zhao, Z. Wang, R. Li, X. Cui, T.-K. Sham and X. Sun, Adv. Sci., 2015, 2, 1500022–1500028 Search PubMed.
  161. Q. Wu, X. Zhang, S. Sun, N. Wan, D. Pan, Y. Bai, H. Zhu, Y.-S. Hu and S. Dai, Nanoscale, 2015, 7, 15609–15617 RSC.
  162. S. Zhao, Y. Bai, Q. Chang, Y. Yang and W. Zhang, Electrochim. Acta, 2013, 108, 727–735 CrossRef CAS.
  163. S. Zhao, Q. Chang, K. Jiang, Y. Bai, Y. Yang and W. Zhang, Solid State Ionics, 2013, 253, 1–7 CrossRef CAS.
  164. J. Chong, S. Xun, J. Zhang, X. Song, H. Xie, V. Battaglia and R. Wang, Chem.–Eur. J., 2014, 20, 7479–7485 CrossRef CAS PubMed.
  165. S. Zhao, Y. Bai, L. Ding, B. Wang and W. Zhang, Solid State Ionics, 2013, 247–248, 22–29 CrossRef CAS.
  166. X. Zhang, Y. Xu, H. Zhang, C. Zhao and X. Qian, Electrochim. Acta, 2014, 145, 201–208 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.