An overview of AB2O4- and A2BO4-structured negative electrodes for advanced Li-ion batteries

Subramanian Yuvaraja, Ramakrishnan Kalai Selvan*a and Yun Sung Lee*b
aSolid State Ionics and Energy Devices Laboratory, Department of Physics, Bharathiar University, Coimbatore 641 046, Tamil Nadu, India. E-mail: selvankram@buc.edu.in
bFaculty of Applied Chemical Engineering, Chonnam National University, Gwangju 500-757, Korea. E-mail: leeys@chonnam.ac.kr

Received 7th November 2015 , Accepted 14th January 2016

First published on 18th January 2016


Abstract

Energy-storage devices are state-of-the-art devices with many potential technical and domestic applications. Conventionally used batteries do not meet the requirements of electric or plug-in hybrid-electric vehicles due to their insufficient energy and power densities. Graphite is used for the conventional anodes of Li-ion batteries. However, the specific capacity (372 mA h g−1) of a graphite electrode is not sufficient for high-power applications. Therefore, Co-, Ni-, Mn-, and Zn-based simple oxides have been investigated as anode components due to their high specific capacities (500–1000 mA h g−1). Among these, Co-based anodes have demonstrated the best electrochemical performances; however, Co's high cost and toxicity limit its use as an ideal anode component. Recently, mixed-metal oxides with AB2O4 (A = Cu, Co, Ni, Mn, and Zn; B = Co, Mn, and Fe) and A2BO4 (A = Co, Mn, and Fe; B = Sn, Si, Ti, and Ge) structures have received much interest, and their electrochemical performances have been extensively studied. This type of mixed-metal oxide affords the following main advantages: they store charge through conversion as well as alloying–dealloying processes, and they exhibit higher electronic conductivities than that obtained with simple metal oxides. The above points indicate the importance of AB2O4- and A2BO4-structured materials. The present review emphasizes the recent literature on the electrochemical performance of AB2O4- and A2BO4-structured materials and their composites and feasible ways to implement these materials in Li-ion batteries in the near future.


1. Introduction

In recent years, the increasing usage of fossil fuels in transport vehicles and the refinery industry has produced more toxic gases such as CO, CO2, and hydrocarbons, which cause severe health problems in human beings, including pneumonia and the blocking of oxygen from the brain, heart, and other vital organs by CO. Importantly, the continued release of CO2 increases global warming, leading to severe climate change.1–3 In order to reduce the emissions of greenhouse gases, the exploitation of green-energy sources, including solar and wind energy, instead of fossil fuels is necessary in the present scenario.

Hence, to reduce air pollution, people are focusing on green-energy sources such as solar and wind energy; they can be considered as important alternative energy sources for sustainable economic growth. However, solar and wind energy as well as electric cars require highly efficient energy-storage devices. In recent decades, batteries have been viewed as the most promising energy-storage devices, as they can store electrical energy as electrochemical energy. Batteries are classified into two types: primary and secondary batteries. A primary battery is used in its charged state once it converts its chemical energy into electrical energy; it is then discarded as it is not rechargeable. A secondary battery is electrically rechargeable after discharge. To date, the most investigated secondary batteries have been lead–acid, Ni–Cd, Ni–metal hydride, and Li-ion batteries (LIBs). Among these batteries (Table 1), LIBs have exhibited the best electrochemical performances in terms of long cycle life, low self-discharge, high cell voltage, and no memory effects, and their energy densities are two times and their power densities five times greater than those of current Pb–acid and Ni–Cd batteries.4 Hence, LIBs are widely used in different electronic devices, including laptops, mobile phones, and other such electronic devices.5–7

Table 1 Comparison of lead–acid, Ni–Cd, Ni–metal hydride, and Li-ion batteries
Specifications Lead–acid Ni–Cd Ni–MH Li-ion batteries
LiCoO2 LiMn2O4 LiFePO4
Cell voltage 2.1 V 1.2 V 1.2 V 3.7 V 3.7 V 3.3 V
Energy density (W h kg−1) 30–50 45–80 60–120 150–190 100–135 90–120
Power density (W kg−1) 180 150 250–1000 1800 1200–1400 1400
Cycle life (cycles) 400 500 500 >500 >500 >2000
Energy efficiency (%) 60 70 90 75 90 95
Self-discharge (%) per month 20 30 35 10 10 8
Memory effect No Yes Little No No No


2. Construction and working principle of LIBs

The LIB is an energy-storage device that converts chemical energy into electrical energy through a controlled, thermodynamically favorable chemical reaction. Generally, such a battery consists of a number of interconnected cells. Each individual cell has a cathode, anode, and electrolyte. Both the cathode and anode are electrically isolated by the electrolyte and a separator in order to avoid a short circuit. The electrolyte is an ionically conductive medium, such as a polymeric/organic electrolyte, that facilitates Li ion transfer between the anode and cathode. Currently, such batteries consist of a layered structure of LiCoO2 and graphite, which are used as the cathode and anode, respectively.

In rechargeable batteries, the storage mechanism involves a reversible insertion–extraction of Li ions into and out of the electrode material during the charge–discharge process, based on the rocking-chair concept. While charging (which involves the loss of electrons and Li ions), Li ions are deintercalated from the cathode and intercalated into the anode. In the discharge process, Li ions are deintercalated from the anode and intercalated into the cathode, delivering the energy to an external circuit. This discharge process continues until the potential difference between the two electrodes becomes too small at which point the cell is fully discharged. Fig. 1 depicts a schematic diagram of their function, and the reversible-reaction mechanism of LIBs is described as follows,7–9

 
Anode: 6C + xLi+ + xe ↔ LixC6, (1)
 
Cathode: LiCoO2 ↔ Li1−xCoO2 + xLi+ + xe. (2)


image file: c5ra23503k-f1.tif
Fig. 1 Charge–discharge process of a Li-ion battery (LIB).

Currently, commercially used cathode materials include layer structured LiCoO2 (140–160 mA h g−1); olivine-type LiFePO4 (140–160 mA h g−1); and spinel-type LiMn2O4 (100–120 mA h g−1), and their practical energy densities versus a graphite anode are 584, 398, and 424 W h kg−1, respectively. On the other hand, while using combinations of Li4Ti5O12 and LiMn2O4 the obtained energy density is 200 W h kg−1, which is half of the reported value when using graphite as the anode, since the specific capacity of spinel Li4Ti5O12 is 175 mA h g−1.10 The currently used graphite anode has several fascinating properties such as a low working potential versus Li, long cycle life, and low cost. However, its primary drawback is its low reversible capacity, as the diffusion rate of Li is in between 10−9 and 10−6 cm2 s−1, resulting in the low power density of this type of battery.11 Therefore, to further improve the power density of this component, it is mandatory to adopt new strategies and to identify novel materials to replace the graphite anode. Another well-known anode material is Li metal, which possesses a high specific capacity of 3860 mA h g−1 and low working potential. However, the working potential of Li metal exists outside the window of the electrolyte, so the transfer of Li ions into the electrode material is achieved through a solid-electrolyte-interphase (SEI) passivation layer that is formed by way of an electrolyte-decomposition process. However, for a fast charge–discharge process, a larger number of Li ions build up on the surface of the Li metal; moreover, changes in the electrode volume break the SEI layer, forming Li plating that leads to dendrite formation. On subsequent cycles, Li dendrites grow across the electrolyte, causing these batteries to short circuit.8,12 Therefore, based on safety concerns, Li metal is highly restricted to the application of anodes in LIBs. The most important issue is that commercially used LIBs do not have sufficient energy and power densities for implementation in hybrid-electric and plug-in hybrid-electric vehicles. Generally, the energy and power densities are directly proportional to the cell voltage and specific capacity of a cell (energy density = VIt (W h kg−1)). Hence, to achieve high-power- and high-energy-density LIBs, appropriate anode and cathode materials with high specific capacities, cell voltages, and Li-diffusion coefficients are needed. After the pioneering work of Poizot et al., transition-metal-oxide (MO, M = Co, Ni, Cu, Fe) nanoparticles have been explored as potential anodes for LIBs. They demonstrated an excellent electrochemical property of 700 mA h g−1 with 100% capacity retention after 100 cycles at high charging rates and have suggested that the high electrochemical reactivities of transition-metal oxides might lead to the improved performances of such batteries.13 The electrochemical reactions of transition-metal oxides are

 
MO + 2Li+ + 2e ↔ Li2O + M (M = Co, Ni, Fe, and Mn) (3)

Here, the redox reactions between metal oxides and Li ions are thermodynamically more favorable and involve multiple electron transfers per metal atom, leading to high specific capacities in the range of 500 to 1000 mA h g−1.14,15 Even though simple metal oxides have shown high reversible capacities and energy densities, they suffer from low coulombic efficiencies in the first cycle, unstable SEI-layer formation, large potential hysteresis, and poor capacity retention.16

Hence, these drawbacks can be overcome by adopting mixed-metal oxides such as ZnFe2O4, NiCo2O4, Co2GeO4, and MnCo2O4 as possible anodes for LIBs due to their high electronic conductivities, fast ionic-transport rates, and easy methods of synthesis. Their different architectures and crystallinity controls may prove these mixed-metal oxides to be good electrochemical-energy-storage materials for LIBs. Recently, Wei et al. demonstrated that complex binary- and ternary-metal oxides possess better electronic conductivities compared to simple metal oxides, and they demonstrated that NiCo2O4 has a larger electrochemical activity and an electronic conductivity two orders of magnitude greater than those of simple Ni and Co oxides.17 These kind of mixed metal oxides has special features like complex chemical composition, interfacial and synergistic effects leads to enhanced electrochemical performance and electronic conductivity.18,19 So in order to further emphasize the importance of this type of binary spinel compounds, an attempt is made here to review the electrochemical performance of the mixed metal oxides especially AB2O4 and A2BO4 structures on the application of Li-ion anode materials.

3. AB2O4 structure

The general formula of the spinel structure (AB2O4) is [A2+][B23+][O42−], where A is a divalent cation and B is a trivalent cation. This structure consists of a closely packed array of 32 oxygen ions that form 64 tetrahedral cations and 32 octahedral cations in a single unit cell. The spinel structure can be classified into three types: normal, inverse, and mixed spinel based on the cation occupancy.20,21 In a normal spinel, the A cations preferentially occupy tetrahedral sites, and the B cations occupy octahedral sites. Examples of such normal spinels include MgAl2O4, ZnFe2O4, and CdFe2O4. In the inverse-spinel structure, the octahedral sites are occupied by the A cations and half of the B cations, while the other half of the B cations occupy the tetrahedral sites. It is denoted as [B3+][A2+B3+]O4. CoFe2O4 and NiFe2O4 are examples of the inverse-spinel structure.22,23 In a mixed spinel, both A and B cations occupy both octahedral and tetrahedral sites, and the molecular formula is [Mx2+Fe1−x3+][M1−x2+Fe1+x3+]O4. Examples of mixed spinels include NiCo2O4, MnCo2O4, and CoMn2O4.

There are three different possible charge-storage mechanisms involved in anode materials: intercalation–deintercalation, alloying–de-alloying, and conversion reactions.24–26

(1) Normally, intercalation–deintercalation mechanisms occur in carbon-based materials such as carbon nanotubes (CNTs), graphene, porous carbon, TiO2, and Li4Ti5O12.

 
MOz + xLi+ + xe ↔ LixMOz (M = Ti and W) (4)

(2) Similarly, the alloying–de-alloying mechanism occurs with Si, Ge, Sn, Bi, SnO2, SiO2, GeO2, and so on.

 
MOx + 2xLi+ + 2xe → M + xLi2O (5)
 
M + xLi+ + xe ↔ LixM (M = Si, Sn, Sb, and Ge) (6)

In the case of pure Zn, Ge, Sn and Si electrodes, the charges are stored through alloying/de-alloying mechanism by forming LixZn, LixGe, LixSn and LixSi, respectively, according to eqn (6). On the other hand, the metal oxides including ZnO, GeO2 and SnO2 based electrodes involving both conversion and alloying/de-alloying mechanism, according to eqn ((5) and (6)). This different charge storage mechanism mainly arises from the oxygen molecules.

(3) The conversion-reaction mechanism is applicable to transition-metal oxides such as MnO2, Mn2O3, NiO, Fe2O3, Fe3O4, Co3O4, CoO, and CuO.

 
MzOx + 2xLi+ + 2xezM + xLi2O (M = Co, Ni and Fe) (7)

The pure Co, Fe, Mn and Ni elements are electrochemically inactive. But, it involved in the conversion reaction, while in the form of metal oxides. So the main contribution arises from the decomposition and reformation of Li2O amorphous matrix according to eqn (7).

Generally, the mixed metal oxides store the lithium ions through both conversion reaction as well as alloying/de-alloying mechanism.27–29 In M2M′O4 (M = Co, Fe, Mn and Ni; M′ = Sn, Ge and Si) system, during first discharging process the active material disintegrated into its constituent elements (M and M′) followed by the formation of Li2O amorphous matrix and electrolyte decomposition at 0.6 to 0.7 V vs. Li/Li+.30 Subsequently, the M′ reacts with lithium to form LixM′ (Li-alloying formation) at 0.1 to 0.3 V vs. Li/Li+.30 During charging process, de-alloying reaction takes place at 0.5 to 0.8 V vs. Li/Li+, and yielded individual metals (M and M′). Subsequently, this metal particles act as catalyst to decompose the Li2O amorphous matrix and formed a corresponding oxides (MOx and M′Ox), according to conversion reaction mechanism.13 In the case of MM2′O4 (M = Ni, Mn, Co and Fe; M′ = Ni, Mn, Co and Fe), the elements M and M′ are involved in conversion reaction mechanism. During the first discharging process, the crystal structure destructed into individual metal particles accompanying with the formation of Li2O matrix. As produced metal particles facilitate the electrochemical activity by means of formation/decomposition of Li2O that give the way for conversion reaction mechanism.31

In the following sections, the main significance and electrochemical performances of these different types of structured spinel materials such as ferrites (AFe2O4), cobaltites (ACo2O4), manganites (AMn2O4) (where A = Mg, Zn, Cu, Mn, Ni, and Co), and A2BO4 (where A = Co, Zn, Mn, and B = Sn, Ge, and Si)-structured materials are discussed.

3.1. AFe2O4 (A = Mg, Zn, Cu, Mn, Ni, and Co)

Generally, AFe2O4-based spinel ferrites possess good electrical and magneto-optical properties, according to the A- and B-site cation distribution, and they have been widely investigated for applications in different fields, including the electronics industry, magnetic recording, magnetic-resonance imaging, and ferrofluids.32–34 Ferrite-based materials are abundant, cheap, and environmentally friendly. The spinel-ferrite MFe2O4 (M = Mg, Zn, Cu, Mn, Ni, and Co)-based materials exhibit significant discharge capacities of 1000 mA h g−1, which is two to three times greater than that of commercially used graphite anodes.35
3.1.1. MgFe2O4. MgFe2O4 has a mixed-spinel structure, and its formula is (Mg1−xFex)[MgxFe2−x]O4, where x is the inversion parameter and denotes the fraction of B cations in the tetrahedral sites. It consists of a face-centered close-packed oxygen sublattice in which the tetrahedral sites are filled by Mg2+ ions and a fraction of the Fe3+ ions and the octahedral sites are filled by the Fe3+ ions and a fraction of the Mg2+ ions.36 Pan et al. prepared MgFe2O4 nanoparticles via a sol–gel method and reported a discharge capacity of 474 mA h g−1 at 90 mA g−1 over 50 cycles. Sadly, it delivered only 293 mA h g−1 after 50 cycles when the current density was increased to 900 mA g−1.37 On the other hand, Gong et al. prepared carbon-coated MgFe2O4 nanoparticles (Fig. 2(a and b)) for which the thickness of each carbon layer was 4 nm.37 According to the electrochemical reaction (eqn (8) and (9)), MgO does not react with Li, preventing the aggregation of Fe oxides during the charge–discharge process and also acting as a buffer matrix during the insertion–deinsertion process,37,38
 
Li0.56MgFe2O4 + 5.44Li+ + 5.44e → MgO + 2Fe + 3Li2O, (8)
 
2Fe + 3Li2O → Fe2O3 + 6Li+ + 6e. (9)

image file: c5ra23503k-f2.tif
Fig. 2 (a–c) Carbon-coated MgFe2O4 (MFO/C-600) nanoparticles and their rate-capability curves compared with Fe3O4 (FO/C-600) (reprinted with permission from ref. 38. Copyright 2013 Elsevier).

The carbon-coated MgFe2O4 electrode demonstrated a high specific capacity of 466 mA h g−1 at a high current density of 1600 mA g−1 and a greater rate capability than carbon-coated Fe3O4 (Fig. 2(c)). This high rate capability as well as its good cycling stability was achieved through its smaller grain size, MgO matrix, and carbon coating. These factors facilitated a short diffusion-path length for the Li+ ions, preventing particle agglomeration and enhancing the electrical conductivity. The carbon coating not only enhanced the electrical conductivity of the material but also accommodated the volume expansion during the charge–discharge process. It is well known that utilizing graphene composites is an important strategy to improve the electrical conductivity and to prevent volume expansion during the insertion–deinsertion process, and it can provide a large number of accessible active sites for Li+ ion insertion and a short diffusion-path length for both Li+ ions and electrons due to the synergistic effects between the active material and graphene. The prepared MgFe2O4–graphene-nanocomposite electrode showed a high reversible discharge capacity of 764.4 mA h g−1 at 0.04C over 60 cycles and retained a capacity of 219.9 mA h g−1 at a high current rate of 4.2C.39 The carbon-coated and graphene–composite MgFe2O4 particles demonstrated better electrochemical performances than did the regular MgFe2O4 nanoparticles.

3.1.2. ZnFe2O4. ZnFe2O4 has a normal-spinel structure in which the Zn2+ cations occupy the tetrahedral sites and the Fe3+ cations occupy the octahedral sites.40,41 ZnFe2O4 is considered to be a good candidate for anode materials, given its high specific capacity around 1000 mA h g−1. The charge-storage mechanism of ZnFe2O4 results primarily from the alloying–de-alloying- and conversion-reaction mechanisms. Sharma and coworkers synthesized ZnFe2O4 using a combustion method, examined the charge-storage properties, and explained the charge-storage mechanism through TEM images and selected area electron diffraction (SAED) patterns, given the following reactions,42
 
ZnFe2O4 + 0.2Li+ + 0.2e → Li0.2ZnFe2O4 (10)
 
Li0.2ZnFe2O4 + 7.8Li+ + 7.8e → Zn + 2Fe + 4Li2O (11)
 
Zn + Li+ + e ↔ LiZn (12)
 
Zn + Li2O ↔ ZnO + 2Li+ + 2e (13)
 
2Fe + 2Li2O ↔ 2FeO + 4Li+ + 4e (14)

During the first discharge process, i.e., from OCV to 0.8 V, 0.2 mol of Li ions intercalated into the crystal structure, according to eqn (10). Below the potential of 0.8 V, intercalation of Li ions led to the deterioration of the crystal structure of ZnFe2O4 into LiZn, ZnO, FeO, and Li2O matrices in which Li2O was involved in the reversible reaction. Generally, the first cycle was considered to be an irreversible reaction, and it possessed a high irreversible capacity due to the formation of the SEI layer, the side reaction of the electrolyte with the active material layer, and the amorphous-Li2O formation. In the second cycle, the ZnFe2O4 nanoparticles possessed a discharge capacity of 800 mA h g−1 at 60 mA g−1, which decreased to 615 mA h g−1 after 50 cycles due to the volume expansion of the active material, which in turn was due to the alloying–de-alloying- and conversion-reaction mechanisms. To overcome this problem, Teh and coworkers43 prepared one-dimensional (1D) ZnFe2O4 nanofibers that exhibited a high specific capacity of 925 mA h g−1 in the first cycle at 60 mA g−1 and 733 mA h g−1 after 30 cycles. Consequently, they had a capacity of 400 mA h g−1 at 800 mA g−1, a value that was high compared with that of the commercially used graphite anode. It has been reported that this enhanced capacity arises mainly from this morphology of ZnFe2O4 as the ZnFe2O4 nanofibers are well connected with nanoparticles and they act as electronic wiring during the charge–discharge process.43 Hence, tuning the morphology and size of an electrode material effectively enhances its specific capacity and rate capability. Various researchers have proposed further increasing the rate capability of the ZnFe2O4 material [Table 2] with different carbonaceous-material composites, morphological features, and binder effects.44–57 Mesoporous, ZnFe2O4 carbon-composite microspheres demonstrated good electrochemical performance as well as a high reversible capacity of 1100 mA h g−1 at a current density of 0.05 A g−1 over 100 cycles.52 The carbon composite not only enhanced their specific capacity but also facilitated a superior rate capability that provided a discharge capacity of 600 mA h g−1 at a high current density of 1 A g−1. Recently, Yao and coworkers prepared a mesoporous ZnFe2O4–graphene composite, which exhibited excellent electrochemical performance due to the synergistic effects between the large surface areas of the ZnFe2O4 nanoparticles and the excellently flexible and highly electronically conductive graphene.54 The mesoporous ZnFe2O4–graphene composite had a high specific capacity of 870 mA h g−1 at 1 A g−1 after 100 cycles and demonstrated a good rate capability. Nonetheless, it retained a specific capacity of 713 mA h g−1 at 2 A g−1, indicating the ultrafast charging capability and good cycling stability of this mesoporous ZnFe2O4–graphene composite.54 Table 2 lists the overall electrochemical performances of ZnFe2O4 and its composites, synthesized by various techniques.42–57

Table 2 A summary of recent electrochemical studies on Zn2SnO4 based anodes for LIBs
Sample Experimental methods Specific capacity Rate capability Ref.
ZnFe2O4 Combustion synthesis 615 mA h g−1 at 60 mA g−1, 50 cycles 42
ZnFe2O4 nanofibers Electrospinning technique 733 mA h g−1 at 60 mA g−1, 30 cycles 400 mA h g−1 at 800 mA g−1 43
ZnFe2O4/graphene hybrid Solvothermal route 600 mA h g−1 at 200 mA g−1, 90 cycles 400 mA h g−1 at 800 mA g−1 44
MWCNT–ZnFe2O4 nanocomposites Solvothermal route 1152 mA h g−1 at 60 mA g−1, 50 cycles 580 mA h g−1 at 600 mA g−1 45
ZnFe2O4-nanorod Template assisted method at 700 °C 888 mA h g−1 at 150 mA g−1, 80 cycles 625 mA h g−1 at 1000 mA g−1 46
ZnFe2O4/C hollow sphere Solvothermal route 625 mA h g−1 at 500 mA g−1, 30 cycles 450 mA h g−1 at 700 mA g−1 47
ZnFe2O4 nano-octahedrons Hydrothermal synthesis 910 mA h g−1 at 60 mA g−1, 80 cycles 730 mA h g−1 at 1000 mA g−1 48
ZnFe2O4–carbon conductive network Carbonothermal method using sucrose 1100 mA h g−1 at 50 mA g−1, 60 cycles 49
Yolk–shell ZnFe2O4 Spray-drying process 862 mA h g−1 at 500 mA g−1, 200 cycles 50
ZnFe2O4/graphene nanoparticles Hydrothermal method 956 mA h g−1 at 100 mA g−1, 50 cycles 600 mA h g−1 at 1000 mA g−1 51
Mesoporous ZnFe2O4/C composite microspheres Hydrothermal method using sucrose 1150 mA h g−1 at 50 mA g−1, 100 cycles 600 mA h g−1 at 1100 mA g−1 52
Carbon coated ZnFe2O4 nanoparticles Carbonothermal method using sucrose 1300 mA h g−1 at 40 mA g−1, 100 cycles 525 mA h g−1 at 3890 mA g−1 53
Mesoporous ZnFe2O4/graphene composite Ambient pressure method 1259 mA h g−1 at 100 mA g−1, 30 cycles 726 mA h g−1 at 2000 mA g−1 54
ZnFe2O4@C/graphene nanocomposite Hydrothermal method followed by carbon coating and graphene composite process 705 mA h g−1 at 232 mA g−1, 180 cycles 403 mA h g−1 at 5C 55
ZnFe2O4/graphene hybrid films One-step electrophoretic deposition subsequent thermal annealing 881 mA h g−1 at 200 mA g−1, 200 cycles 510 mA h g−1 at 3200 mA g−1 56
ZnFe2O4–C nanocomposite Sol–gel followed by ball milling technique 681 mA h g−1 at 0.1C, 100 cycles 469 mA h g−1 at 4C 57


Another method to improve specific capacity is by forming mesoporous microspheres, since they form from well-interconnected nanoparticles. Therefore, they increase the interfacial surface reactions with Li and shorten the diffusion path for Li ions and electrons, leading to high specific capacities. The interconnected nanoparticles create these mesopores in between neighboring particles, thereby accommodating the volume expansion during the charge–discharge process.52 Furthermore, the interconnected nanoparticles embedded in the carbon matrix also enhanced the intrinsic electrical conductivity and thus speeded the transport of Li+ ions and electrons. Therefore, this type of mesoporous-carbon-composite structure facilitates high reversible specific capacities, good cycling stabilities, and good rate capabilities of electrodes.

In addition, other than carbon composites and mesoporous structures, a binder can also effectively alleviate the volume-expansion problem. Bresser et al. used sodium carboxymethyl cellulose (NaCMC) as a binder for carbon-coated ZnFe2O4 nanoparticles, and Fig. 3 depicts its corresponding cycling-stability curve.53 This binder promoted superior electrochemical performance as a result of the stable reversible capacity (1300 mA h g−1) at 0.02 A g−1 over 100 cycles and 525 mA h g−1 at a high current density of 3.89 A g−1. The main advantages of the CMC binder are that it is nontoxic, eco-friendly, and water soluble; therefore, it provides better contact between the active material and the current collector and modifies the SEI formation due to the continuous reaction of CMC with organic electrolytes, enhancing the cycling stability and high rate capability.


image file: c5ra23503k-f3.tif
Fig. 3 Rate capabilities of carbon-coated ZnFe2O4 nanoparticles using different binders (circles: ZnFe2O4–C, triangles: ZnFe2O4–CMC, and stars: ZnFe2O4–PVDF-HFP) (reprinted with permission from ref. 53. Copyright 2013, Wiley-VCH Verlag GmbH & Co. KGaA).
3.1.3. CuFe2O4. CuFe2O4 possesses an inverse-spinel structure, and it has been widely investigated as an anode material due to its low cost, high abundance, environmental friendliness, and high theoretical specific capacity of 895 mA h g−1. A number of factors limit the implementation of CuFe2O4 as an anode material, including its low inherent electronic conductivity and volume expansion during lithium insertion and extraction process. To overcome this problem down the particle size from bulk to nanolevel and making composite with the carbonaceous material is required.58 The spinel-CuFe2O4 structure undergoes a conversion reaction during the charge–discharge process as follows,59–62
 
CuFe2O4 + 8Li+ + 8e → 4Li2O + Cu + 2Fe (15)
 
Cu + 2Fe + 4Li2O ↔ CuO + Fe2O3 + 8Li+ + 8e (16)

Ding et al. prepared the CuFe2O4 nanoparticles which possess a high specific capacity of 869 mA h g−1 in the second cycle; further increasing the number of cycles decreases the capacity to 551 mA h g−1.58 Normally, the nanoparticles have high surface energies, which often result in self-aggregation and lead to less surface area overall as well as the easy diffusion of the electrolyte ions, in turn leading to severe capacity fading. Significantly, Jin et al. reported the preparation of hollow CuFe2O4@C nanospheres and CuFe2O4@C nanoparticles via a polymer-templated hydrothermal method, followed by calcination to create the carbon shells.60 Among these compounds, hollow, carbon-coated CuFe2O4 nanospheres demonstrated a good cyclic performance of 550 mA h g−1 at 100 mA g−1 over 70 cycles, whereas hollow CuFe2O4 only showed 120 mA h g−1 and CuFe2O4@C nanoparticles only showed 300 mA h g−1. The carbon layer and hollow space influenced the electrochemical performance of hollow CuFe2O4@C, effectively accommodating the volumetric changes during the cycling process, mitigating the agglomeration between the particles, and facilitating the electron transport that enhanced the specific capacity and good rate performance of the electrode material.60

3.1.4. MnFe2O4. The spinel manganese ferrite is a promising candidate for several applications, including semiconductors, biosensors, photocatalysts, supercapacitors, and medical applications.63–67 However, only a sparse number of reports are available concerning MnFe2O4 with regards to its application to anode materials in LIBs.68,69 MnFe2O4 stores Li ions through a conversion reaction (MnFe2O4 + 8Li+ + 8e ↔ Mn + 2Fe + 4Li2O), and it exhibits a high theoretical specific capacity of 928 mA h g−1. Zhang et al. synthesized mesoporous manganese ferrite microspheres using a solvothermal method and examined their electrochemical performance versus Li metal. These microspheres showed a high reversible discharge capacity of 712 mA h g−1 at a rate of 0.2C over 50 cycles. Only feeble capacity fading was observed over each cycle, and they retained a capacity of 552 mA h g−1 even at a high current density at a rate of 0.8C.68 This reported capacity fading was an acceptable limit and was achieved without fabricating any sort of composite with carbon, CNTs, or graphene. As explained above, the porous nature of these microspheres could accommodate the volume changes during the conversion reaction. The appropriate grain size with crystallinity might improve the surface area of the active material, and the electronic conductivity of the electrode material thus could lead to both an improved specific capacity and cycling stability.69–71 Furthermore, Xiao et al. prepared MnFe2O4–graphene nanocomposites via an ultrasonication process to improve their electrochemical properties.72 A graphene composite has several advantages such as its enhanced inherent conductivity and high surface area (2600 m2 g−1), and it stores Li+ ions through an insertion–de-insertion process, thereby demonstrating its high theoretical specific capacity of 744 mA h g−1 at the same time that it can accommodate volume expansion.73 The MnFe2O4–graphene composite possessed a high specific capacity of 1017 mA h g−1 at 100 mA g−1 over 90 cycles and 767 mA h g−1 at 1000 mA g−1. It also showed an excellent rate capability.72
3.1.5. NiFe2O4. NiFe2O4 is also considered a promising anode material for LIBs, since it can accommodate 8 mol of Li per formula unit; therefore, it has a high theoretical specific capacity of 914 mA h g−1. It stores Li+ ions through an electrochemical-conversion reaction as follows,74–77
 
NiFe2O4 + 8Li+ + 8e → Ni + 2Fe + 4Li2O (17)
 
Ni + Li2O ↔ NiO + 2Li+ + 2e (18)
 
2Fe + 3Li2O ↔ Fe2O3 + 6Li+ + 6e (19)

However, there are numerous difficulties in implementing NiFe2O4 as an anode material because of the low diffusion of Li ions in the bulk material, low conductivity, and large volume expansion during the charge–discharge process.74–76 Several reports have proposed solutions to overcome this issue so as to synthesize nanosized particles with porous structures and conducting carbon networks to improve the electrochemical performance of NiFe2O4.77–79 Macroporous NiFe2O4 particles were prepared via a sol–gel method using a citrate precursor. The sample, calcined at 800 °C, showed better cycling stability and a better rate capability due to its smaller particle sizes with larger pore sizes, as compared to a sample calcined at 1000 °C, and it maintained a capacity of 600 mA h g−1 after 80 cycles at a rate of 1C.76 Morphological features also play an important role in enhancing the electrochemical performance. Recently, NiFe2O4 nanofibers were synthesized via an electrospinning technique, and their electrochemical performance was compared with that of NiFe2O4 nanoparticles.77 The cycling-stability curve indicated that the NiFe2O4 nanofibers exhibited a superior electrochemical performance as compared to the nanoparticles. The NiFe2O4 nanofibers showed only feeble capacity fading in the initial 15 cycles due to the structural rearrangement of the active material, as they were well connected with conductive carbon and had stable SEI formation. After 15 cycles, the capacity stabilized at 870 mA h g−1 at 100 mA g−1 for another 40 cycles. In the subsequent cycles, the capacity increased steeply to 1000 mA h g−1 over 100 cycles. This reported specific capacity was greater than the theoretical specific capacity; it was attained through a kinetically activated electrolyte degradation.80,81 Imperatively, binders also effectively contribute to the electrochemical performance. Many authors have reported using different binders such as polyvinylidene fluoride (PVDF), polyacrylic acid (PAA), CMC, and sodium alginate for electrode preparation.82–85 Ramesh and coworkers reported that they synthesized nanocrystalline NiFe2O4 via a sol–gel-combustion route and examined its electrochemical performances using PVDF and sodium alginate as binders. The cycling-stability curve clearly illustrated the severe capacity fading of the NiFe2O4 nanoparticles at a rate of 0.1C, which indicated the inadequate interfacial stability of the PVDF binder.85 However, in the case of the NiFe2O4 nanoparticles with sodium alginate, the nanoparticles showed a stable discharge capacity of 880 mA h g−1 over 30 cycles at a current density of 92 mA g−1 without any degradation of the discharge capacity. Further studies of the rate performance of this electrode material were carried out using a sodium alginate binder at different current densities. The discharge-capacity values decreased to 740, 620, 504, 420, and 380 mA h g−1 when the current density was increased to 1C, 2C, 5C, 10C, and 20C, respectively. The NiFe2O4 nanoparticles with sodium alginate had a discharge capacity of 380 mA h g−1 at a high current density of 18[thin space (1/6-em)]300 mA g−1 at a rate of 20C, as shown in Fig. 4, and they demonstrated a good rate capability as compared with the ternary composite.82 The sodium alginate binder effectively improved the cycle life and exhibits good rate capability, compared to commercially used PVDF, as the PVDF could not accommodate any volume strain during the cycling process. On the other hand, sodium alginate contains carboxylic groups evenly distributed in the polymer chain that improve the transport of Li ions and increase the stablility of the SEI-layer formation. It also possesses polar molecules that improve the interfacial interaction of the polymer binder and facilitate the self-healing process during the charge–discharge process. Furthermore, it provided strong adhesion between the electrode layer and the Cu substrate.84,86,87 The NiFe2O4 nanoparticles with sodium alginate binder showed superior electrochemical performance, as compared to previously reported NiFe2O4–graphene heteroarchitectures.88


image file: c5ra23503k-f4.tif
Fig. 4 Rate capabilities of the NiFe2O4 nanoparticles with sodium alginate binder at different current densities (reproduced from ref. 85. Copyright 2013, with permission from the Royal Society of Chemistry).
3.1.6. CoFe2O4. Cobalt ferrite is a thoroughly investigated anode material used for LIBs due its low cost, environmental friendliness, and high chemical stability. It also stores Li ions through a conversation reaction similar to that of NiFe2O4, and it possesses a high theoretical specific capacity of 914 mA h g−1.89–91 However, it has severe drawbacks such as a large volume change that induces the pulverization and aggregation of the active material and low conductivity that leads to poor cycling stability and a low rate capability of the CoFe2O4 particles.92–94 To achieve better performance, CoFe2O4 nanoparticles have been synthesized with different morphologies, including hollow nanospheres, mesoporous nanospheres, and nanorods, and as composites with different conducting matrices.46,95–98 In particular, Xiong et al. prepared different morphologies of CoFe2O4, such as irregular particles, microspheres, and flower-like microspheres (Fig. 5(a)), via a hydrothermal method using various concentrations of ascorbic acid. Among these morphologies, the flower-like microspheres had the largest surface areas of 51.0 m2 g−1 compared to the regular microsphere and the irregular particles. Similarly, the flower-like microspheres (Fig. 5(b)) had a high reversible specific capacity of 790 mA h g−1 at a current density of 200 mA g−1, and even when the current density increased to 1000 mA g−1, they still maintained their high discharge capacity of 744 mA h g−1.99 The high reversible capacity and rate capability attained through good electronic conductive wetting of the active materials thus facilitated a rapid charge-transfer process during the insertion–deinsertion process.100 Recently, Zhang et al. reported the superior electrochemical performance of mesoporous CoFe2O4 nanospheres with cross-linked CNTs.96 The CNT-weight ratios varied between 10, 20, and 30 wt%. Among these values, the CoFe2O4 with 20 wt% CNTs is shown in Fig. 5(c) which possessed the highest specific capacity of 1137 mA h g−1 after 10 charge–discharge cycles at a current density of 200 mA g−1 and increase the current density to 1200 mA g−1 which has stable specific capacity of 648 mA h g−1 with a good cycling stability and rate capability (Fig. 5(d)), as compared to the CoFe2O4–graphene composite.96,98 The SEM results of after charge–discharge cycles indicated that the formation of a small amount of SEI film on the active surfaces led to a fast conversion reaction and that the CNTs accommodated the volume changes during the charge–discharge process.101 On the other hand, Kumar et al. reported that porous CoFe2O4–rGO with an alginate binder showed an excellent rate performance (Fig. 5(e)) compared to both CNTs and graphene composite.91 Therefore, a novel binder and a conductive matrix both play essential roles in improving the cycling stability as well as the rate capability of an electrode material. As such, it can be concluded that MgFe2O4, CoFe2O4, and NiFe2O4 are not suitable anode materials. After the first lithiation process, MgFe2O4 converts into MgO, Fe, and a Li2O matrix in which MgO does not participate in the electrochemical reaction, leading to a low specific capacity. So too, while CoFe2O4 has been considered as an important anode material, it is costly, and its high working potential (2.1 V vs. Li/Li+) reduces the cell voltage. Finally, NiFe2O4 is both highly expensive and toxic. Therefore, these inherent characteristics restrict the use of MgFe2O4, CoFe2O4, and NiFe2O4 in potential applications requiring high-energy-density LIBs. ZnFe2O4 and MnFe2O4 are environmentally friendly and have low working potentials, and their high specific capacities could pave the way for the construction of efficient high-energy-density batteries.
image file: c5ra23503k-f5.tif
Fig. 5 (a and b) Schematic diagram of the reaction producing various morphologies of CoFe2O4 particles and rate capability curve, (reprinted with permission from ref. 99. Copyright 2014 Elsevier). (c and d) TEM image of CoFe2O4 with 20 wt% CNT and its rate capability curve, (reproduced from ref. 96. Copyright 2013, with permission from the Royal Society of Chemistry) and (e) rate capability curve of CoFe2O4 + 20 wt% graphene with a sodium alginate binder (reproduced from ref. 91. Copyright 2014, with permission from the Centre National de la Recherche Scientifique (CNRS) and the Royal Society of Chemistry).

3.2. AMn2O4 (A = Co, Ni, and Zn)

It is well known that Mn is environmentally benign, highly safe, low in cost, and abundant in nature compared to Co. With respect to anode materials, Mn-based oxides have lower operating voltages (∼1.5 V), reduce the source of Co in mixed-transition-metal oxides, and correspondingly increase the energy density as well as the output cell voltage of such anodes. Therefore, investigating the peculiar properties of manganite-based mixed-transition-metal oxides [AMn2O4 (A = Co, Ni, Zn)] for applications to LIBs is very necessary.102–124 CoMn2O4 spinels have been synthesized with different morphologies, including hierarchical microspheres, double-shelled hollow microcubes, CNF@CoMn2O4 nanocables, nanorods, and nanofibers, in attempts to achieve stable capacities and good rate capabilities for electrode materials.105–110 Significantly, Hu et al. prepared hierarchical CoMn2O4 microspheres, assembled with porous nanosheets, which possessed large pores between the neighbouring nanosheets [Fig. 6(a)], via a solvothermal method.105 These microspheres exhibited a slight decrease of their specific capacity only a few cycles after that capacity reached 894 mA h g−1, a value that was close to their theoretical specific capacity of 920 mA h g−1, over 60 cycles with excellent cycling stability [Fig. 6(b)]. It has been reported that generally at low current densities, the electrolyte ions easily penetrated into the inner parts of the active material and utilized all the active sites, and over subsequent cycles, the observed increase in capacity resulted from the full conversion reaction of the active material. These hierarchical microspheres also had a high specific capacity of 559 mA h g−1 at a rate of 4C [Fig. 6(c)], ensuring a good rate capability of the active material. It has been concluded that porous CoMn2O4 hierarchical microspheres provided a short diffusion length for both Li ions and electron transport within the electrolyte–electrode interface and accommodated the volume expansion and easy diffusion of the electrolyte into open pores, while the well-interconnected nanosheets enhanced the energy and power densities. Zhang et al. recently synthesized carbon-nanofiber–metal-oxide coaxial nanocables using two steps, a polyol method and subsequent thermal annealing.107 The morphology of the CNF@CoMn2O4 nanocables and their cycling stability are shown in Fig. 6(d)–(f). The cycling-stability curve of these nanocables provided a high specific capacity of 870 mA h g−1 at a current density of 200 mA g−1 over 150 cycles, and they maintained a capacity of 650 mA h g−1 even at a high current density of 1000 mA g−1. These results confirmed the superior electrochemical performance of CNF@CoMn2O4, as the outer shells of the CoMn2O4 cables were strongly anchored to the carbon nanofibers and the inner cores of the CNFs could accommodate the volume change during the Li insertion–extraction process, providing good electrical conductivity and thus leading to a high capacity, long cycle life, and good rate performance.107
image file: c5ra23503k-f6.tif
Fig. 6 (a)–(c) Hierarchical microsphere of CoMn2O4 and its cycling stability and rate-capability curve, respectively (reprinted with permission from ref. 105. Copyright 2012, Nature Publishing Group). (d)–(f) HRSEM images of CNF@CoMn2O4 nanocables and their cycling-stability curve, respectively (reproduced from ref. 107. Copyright 2014, with permission from the Royal Society of Chemistry).

NiMn2O4 has also been considered as an attractive anode material for LIBs. Lee's group recently synthesized hierarchical tremella-like NiMn2O4/C nanostructures through a simple solvothermal method followed by a calcination process.111 The tremella-like nanostructures demonstrated a discharge capacity of 2130 mA h g−1 after 350 cycles at a current density of 1000 mA g−1, a value that is greater than the theoretical specific capacity of NiMn2O4 (∼922 mA h g−1), and they retained a high discharge capacity of 1773 mA h g−1 at 4000 mA g−1 after 1500 cycles. This high discharge capacity arose due to an interfacial-storage mechanism, which originated from the reversible formation–dissolution of an organic polymeric gel-like layer via electrolyte decomposition that induces the extra capacity in the electrode material by way of pseudocapacitive behavior. The small amount of carbon facilitated this high rate capability, prevented the destruction of the hierarchical structure, and provided for continuous electron transport.111,112

Compared with CoMn2O4 and NiMn2O4, ZnMn2O4 has several attractive features,113–124 including the fact that it is environmentally friendly; Zn and Mn are also more abundant and lower in cost compared to Ni and Co. In terms of applications, Zn (1.2 V) and Mn (1.5 V) exhibit lower oxidation potentials than does Co (2.2–2.4 V). The low oxidation potentials of Zn and Mn eventually increase the cell voltage and energy density of a battery. ZnMn2O4 has a high theoretical capacity of 1008 mA h g−1 in the first lithiation process due to the conversion and alloying-reaction mechanisms,113–116

 
ZnMn2O4 + 9Li+ + 9e → ZnLi + 2Mn + 4Li2O (20)
 
ZnLi + 2Mn + 3Li2O ↔ ZnO + 2MnO + 7Li+ + 7e (21)

However, in the second reversible reaction, only 7 mol of Li is contributed to the energy storage, and therefore, this compound has a specific capacity of 784 mA h g−1. As such, its capacity is twice as large as that of a conventional graphite anode. Yang et al. synthesized nanocrystalline ZnMn2O4 (Fig. 7(a)) using a polymer-pyrolysis route and reported the first investigation of this compound for use as an anode material for LIBs.113 In the initial ten cycles, the discharge capacity decreased up to 20% due to the formation of an SEI film on the surface of the active material is shown in Fig. 7(b). However, after ten cycles, only 0.20% capacity fading was observed, and a stable capacity of 569 mA h g−1 was maintained after 50 cycles. To implement ZnMn2O4 in LIB applications, its electrochemical performance was enhanced with different morphologies and sizes, using various synthetic techniques that are listed in Table 3. Courtel et al. synthesized ZnMn2O4 particles via a simple coprecipitation route and optimized their electrochemical performance by varying the different reaction conditions such as the sintering temperature, binder, and electrolyte.102 Among the sintered (400, 600, 800, and 1000 °C) samples, the one sintered at 800 °C showed a stable capacity of 620 mA h g−1 over 70 cycles at a rate of 0.1C. The optimized sample was used as an anode material with different binding agents such as polyvinylidene fluoride (PVDF), sodium carboxymethyl cellulose (NaCMC), lithium carboxymethyl cellulose (LiCMC), Xanthan Gum (XG), and poly-3,4-ethlenedioxythiophene-poly-styrenesulfonate (PEDOT/PSS or Baytron). Among these agents, CMC-based salts and Baytron delivered the most stable capacities. In contrast, the use of PVDF and XG resulted in severe capacity fading [Fig. 7(c)]. However, the Baytron binder, while highly conductive (10–500 mS cm−1), is also expensive. The PVDF and XG binders could not accommodate the volume strain during the insertion–extraction process, so the electrical connections between the active material and current collector broke, leading to capacity fading. CMC-based salts alleviated the volume expansion during the cycling process, thus facilitating the improvement of the cycling stability. Li-based CMC resulted in a capacity of 690 mA h g−1 over 70 cycles at a rate of 0.1C, which is greater than the capacity that resulted from the use of Na-based CMC as it only provided a specific capacity of 615 mA h g−1 over 70 cycles. Furthermore, to analyze the effects of the electrolyte composition, a sample sintered at 800 °C and a LiCMC binder were used with different compositions of the electrolyte propylene carbonate (PC)–1 M lithium hexafluorophosphate (LiPF6), ethylene carbonate (EC)[thin space (1/6-em)]:[thin space (1/6-em)]dimethyl carbonate (DMC) (1[thin space (1/6-em)]:[thin space (1/6-em)]1)–1 M LiPF6, and ethylene carbonate (EC)[thin space (1/6-em)]:[thin space (1/6-em)]diethylene carbonate (DEC) (3[thin space (1/6-em)]:[thin space (1/6-em)]7)–1 M LiPF6]. The specific capacity was increased for both EC[thin space (1/6-em)]:[thin space (1/6-em)]DMC (1[thin space (1/6-em)]:[thin space (1/6-em)]1) and EC[thin space (1/6-em)]:[thin space (1/6-em)]DEC (3[thin space (1/6-em)]:[thin space (1/6-em)]7) when the cycle number increased, but such a change did not occur with the PC-based electrolyte because PC does not contribute to a stable SEI layer and is also inactive for Li storage.117 The EC[thin space (1/6-em)]:[thin space (1/6-em)]DEC electrolyte-composition capacity increased faster, as compared to that of the EC[thin space (1/6-em)]:[thin space (1/6-em)]DMC, as the former compound retained a specific capacity of 800 mA h g−1 after 100 cycles. This observation indicated that EC[thin space (1/6-em)]:[thin space (1/6-em)]DEC electrolyte promoted the faster formation of an SEI layer on the surface of the active material. Finally, the optimized sintered sample, LiCMC binder, and EC[thin space (1/6-em)]:[thin space (1/6-em)]DEC (3[thin space (1/6-em)]:[thin space (1/6-em)]7)–1 M LiPF6 (electrolyte) were used to fully fabricate a cell.


image file: c5ra23503k-f7.tif
Fig. 7 (a)–(b) ZnMn2O4 nanoparticles synthesized via polymer pyrolysis route and its cycling stability, (reprinted with permission from ref. 113. Copyright 2008 Elsevier). (c) Cycling stability of ZnMn2O4 calcined at 800 °C, and its cycling-stability curves with different binders, respectively (reproduced from ref. 102. Copyright 2011, with permission from the Royal Society of Chemistry). (d)–(f) FESEM image of ZnMn2O4 microspheres, their cycling-stability curve, and different current densities, respectively (reprinted with permission from ref. 119. Copyright 2014 Elsevier). (g)–(i) TEM image of twin microspheres of ZnMn2O4, their cycling stability, and their rate-capability curves, respectively (reproduced from ref. 120. Copyright 2014, with permission from the Royal Society of Chemistry). (j)–(l) TEM image of ZnMn2O4–graphene, its cycling-stability curves, and different current densities shows the comparison of 2D ZMO-G, ZMO-SD (physical mixing of Super P carbon) and control ZMO-G (physical mixing of rGO) respectively (reprinted with permission from ref. 122. Copyright 2014, American Chemical Society).
Table 3 A summary on recent electrochemical performance of MMn2O4 (M = Co, Zn) anodes for LIBs
Sample Experimental methods Reversible capacity Rate performance Ref.
CoMn2O4 hierarchical microspheres Solvothermal method 894 mA h g−1 at 100 mA g−1, 65 cycles 344 mA h g−1 at 7200 mA g−1 105
Double-shelled CoMn2O4 hollow microcubes Co-precipitation and annealing process 624 mA h g−1 at 200 mA g−1, 50 cycles 406 mA h g−1 at 800 mA g−1 106
CNF@CoMn2O4 nanocables Polyol method and subsequent annealing treatment 870 mA h g−1 at 200 mA g−1, 150 cycles 390 mA h g−1 at 2400 mA g−1 107
CoMn2O4 nanorods Hydrothermal and subsequent annealing process 520 mA h g−1 at 200 mA g−1, 50 cycles 450 mA h g−1 at 800 mA g−1 108
CoMn2O4 nanofibers Electrospinning technique 526 mA h g−1 at 400 mA g−1, 90 cycles 463 mA h g−1 at 800 mA g−1 109
CoMn2O4 quasi-hollow spheres Solvothermal route 706 mA h g−1 at 200 mA g−1, 25 cycles 450 mA h g−1 at 400 mA g−1 110
Nanocrystalline-ZnMn2O4 Polymer-pyrolysis route 569 mA h g−1 at 100 mA g−1, 50 cycles 113
Flower-like ZnMn2O4 Solvothermal process 626 mA h g−1 at 100 mA g−1, 50 cycles 114
Nano-ZnMn2O4 Single precursor route 640 mA h g−1 at 100 mA g−1, 50 cycles 405 mA h g−1 at 600 mA g−1 115
ZnMn2O4 nanoparticles Hydrothermal method 430 mA h g−1 at 78 mA g−1, 100 cycles 120 mA h g−1 at 784 mA g−1 116
ZnMn2O4-tubular arrays Template route 784 mA h g−1 at 100 mA g−1, 100 cycles 363 mA h g−1 at 1600 mA g−1 118
Porous ZnMn2O4 microsphere Solvothermal method 800 mA h g−1 at 500 mA g−1, 300 cycles 395 mA h g−1 at 2000 mA g−1 119
Quasi-mesocrystal ZnMn2O4 twin microsphere Solvothermal method 860 mA h g−1 at 500 mA g−1, 130 cycles 329 mA h g−1 at 5000 mA g−1 120
Loaf-like ZnMn2O4 nanorods Solid state reaction 517 mA h g−1 at 500 mA g−1, 100 cycles 457 mA h g−1 at 1000 mA g−1 121
2-D hybrid ZnMn2O4–graphene nanosheet Reflux method 800 mA h g−1 at 500 mA g−1, 100 cycles 650 mA h g−1 at 2000 mA g−1 122
ZnMn2O4 hollow microsphere Co-precipitation and annealing method 607 mA h g−1 at 400 mA g−1, 100 cycles 361 mA h g−1 at 1600 mA g−1 123
MWCNT/ZnMn2O4 Polyol method subsequent thermal annealing process 847 mA h g−1 at 400 mA g−1, 100 cycles 527 mA h g−1 at 1600 mA g−1 124


Furthermore, to attain a stable capacity and high rate capability, ZnMn2O4 has been synthesized in various morphologies such as mesoscale tubular arrays, porous microspheres, quasimesocrystal twin microspheres, loaf-like nanorods, and hybrid ZnMn2O4–graphene nanosheets.118–122 Wang and coworkers reported porous ZnMn2O4 microspheres prepared through the calcination of metal carbonates via a solvothermal reaction [Fig. 7(d)].119 BET analysis indicated that the surface area and pore volume of these porous ZnMn2O4 microspheres were 17.7 m2 g−1 and 0.27 cm3 g−1, respectively, indicating that the pores were in the range of 45 to 90 nm. Fig. 7(e) shows the cycling performance of the ZnMn2O4 microspheres, analyzed in a potential window of 0.01–3 V at 500 mA g−1. Over the first 50 cycles, the specific capacity decreased to 700 mA h g−1 after it had gradually increased, and it reached a value of 800 mA h g−1 on the 100th cycle, where it was maintained over 300 cycles. Fig. 7(f) indicates that at a higher current density of 2000 mA g−1, the capacity was maintained at 395 mA h g−1. Overall, this excellent cycling stability and rate performance were achieved through the porous structures of the ZnMn2O4 microspheres, which provided high accessibility of the electrolyte over the electrode surface, high surface area that facilitated both charge transfer and a short diffusion distance for Li ions, and the alleviation of the pulverization problems during the insertion–extraction of Li+ ions.

Compared to porous microspheres, solvothermally prepared quasimesocrystal ZnMn2O4 twin microspheres [Fig. 7(g)] demonstrated better electrochemical performance [Fig. 7(h) and (i)]. They demonstrated a high specific capacity of 484 mA h g−1 at a current density of 2 A g−1,120 and current density return to 0.2 mA g−1 they exhibited a capacity of 1084 mA h g−1. This excellent electrochemical performance was mainly attributed to the structural assembly, i.e., the twin microspheres that consisted of a 3D, porous, hierarchical structure composed of smaller primary nanoparticles that increased the number of active sites available for Li ions. The pores on the surfaces and within the interiors alleviated strain during the insertion–extraction process and provided a short diffusion path for Li ions, so an oriented mesocrystal could facilitate more electronic conductivity and in turn excellent specific capacity, cycling stability, and high rate capability. Xiong et al. reported the high rate capability of a ZnMn2O4 anode material using a 2D graphene nanoarchitecture prepared via a reflux method.122 The ZnMn2O4 nanoparticles uniformly distributed on the 2D nanosheet, forging a synergistic effect between the nanoparticles and graphene sheets that enhanced the conductivity and large active surface area, which maintained the structural integrity during the cycling process [Fig. 7(j)]. Fig. 7(k) and (l) show the astonishing properties of this 2D ZnMn2O4–graphene that had a stable capacity of 806 mA h g−1 at a low current density of 200 mA g−1, but it still maintained a high specific capacity of 568 mA h g−1 even when the current density was increased to 3200 mA g−1, indicating the high rate capability of these ZnMn2O4 nanoparticles with a graphene nanoarchitecture.122 For simple comparison, the electrochemical performances of CoMn2O4 and ZnMn2O4 with various morphologies, synthesized via different techniques, are provided in Table 3.

Among the manganites, ZnMn2O4 exhibits a lower oxidation potential and higher cell-output voltage than NiMn2O4 and CoMn2O4. However, ZnMn2O4 exhibits a higher cell-output voltage than ZnFe2O4, since Fe has a higher oxidation potential (∼1.61 V). Therefore, the synthesis of a large surface area with a proper composite is necessary to enhance the electrochemical performance of ZnMn2O4, and it may be expected to become a valuable source for Li ion anode materials for applications to LIBs.

3.3. ACo2O4 (A = Mg, Cu, Fe, Zn, Mn, and Ni)

Spinel cobaltites are an important group of spinel-type materials and have been widely investigated for use in applications in different fields such as sensors, bifunctional electrocatalysts, supercapacitors, and LIBs.125–128 ACo2O4 has demonstrated excellent performances in LIBs, including such characteristics as high specific capacities, long cycle lives, and good rate capabilities, and it is slightly less toxic than Co3O4. Table 4 lists the electrochemical performances of spinel cobaltites.128–165 For the most part, spinel cobaltites store Li ions through a conversion-reaction mechanism except for ZnCo2O4, since it follows both the conversion and alloying–de-alloying mechanisms. Sharma et al. first investigated the electrochemical performance of CuCo2O4, MgCo2O4, and FeCo2O4, synthesized via a low-temperature urea-combustion method and an oxalate-decomposition method, for applications in LIBs.128,129 The electrochemical reaction was analyzed using cyclic-voltammetry analysis in the potential range of 0.005–3.0 V, and their reaction mechanisms are written as follows,128–130
 
MCo2O4 + 8Li+ + 8e → M + 2Co + 4Li2O (M = Cu, Fe Ni, Mn) (22)
 
MgCo2O4 + 6Li+ + 6e → MgO + 2Co + 3Li2O (23)
 
M + Li2O ↔ MO + 2Li+ + 2e (24)
 
2Co + 2Li2O ↔ 2CoO + 4Li+ + 4e (25)
 
image file: c5ra23503k-t1.tif(26)
Table 4 A summary of recent studies of spinel cobaltites (MCo2O4; M = Zn, Mn, Ni and Fe) for LIB anode materials
Sample Experimental method Reversible capacity Rate performance Ref.
Nanophase ZnCo2O4 Combustion method 894 mA h g−1 at 60 mA g−1, 60 cycles 400 mA h g−1 at 0.7C 131
Porous ZnCo2O4 nanotubes Electrospinning technique 1454 mA h g−1 at 100 mA g−1, 30 cycles 794 mA h g−1 at 2000 mA g−1 132
ZnCo2O4 nanowires/carbon cloth Hydrothermal method 1278 mA h g−1 at 200 mA g−1, 100 cycles 605 mA h g−1 at 5C 133
Double-shelled ZnCo2O4 hollow microspheres Citrate route followed by calcination 1019 mA h g−1 at 90 mA g−1, 120 cycles 570 mA h g−1 at 5C 134
Hierarchical ZnCo2O4/Ni current collector Hydrothermal process 1050 mA h g−1 at 100 mA g−1, 60 cycles 240 mA h g−1 at 2778 mA g−1 135
Mesoporous ZnCo2O4 microspheres Solvothermal route 721 mA h g−1 at 100 mA g−1, 80 cycles 382 mA h g−1 at 5000 mA g−1 136
Yolk–shelled ZnCo2O4 microsphere Refluxing route 949 mA h g−1 at 200 mA g−1, 100 cycles 331 mA h g−1 at 1000 mA g−1 137
Hierarchical ZnCo2O4 nanostructure on Ni foam Hydrothermal process 1122 mA h g−1 at 0.5 A g−1, 50 cycles 542 mA h g−1 at 2000 mA g−1 138
Mesoporous rose-like ZnCo2O4 Microwave assisted synthesis 1000 mA h g−1 at 50 mA g−1, 50 cycles 800 mA h g−1 at 800 mA h g−1 139
ZnCo2O4/graphene nanosheets Auto-combustion method 755 mA h g−1 at 0.1C, 70 cycles 302 mA h g−1 at 4.5C 140
Fiber bundle structure ZnCo2O4 Co-precipitation method 1026 mA h g−1 at 100 mA g−1, 100 cycles 606 mA h g−1 at 2000 mA g−1 141
Flake-by-flake ZnCo2O4 Co-precipitation 1275 mA h g−1 at 100 mA g−1, 50 cycles 730 mA h g−1 at 3000 mA g−1 142
Porous ZnCo2O4 nanowires Sacrificial template route 1197 mA h g−1 at 100 mA g−1, 20 cycles 143
Porous ZnxCo3−xO4 hollow polyhedra ZIFs templating approach 990 mA h g−1 at 100 mA g−1, 50 cycles 575 mA h g−1 at 10C 144
Yolk–shell ZnCo2O4 Gas-phase reaction 753 mA h g−1 at 3000 mA g−1, 200 cycles 145
Flower-like ZnCo2O4 nanowires Hydrothermal method 900 mA h g−1 at 200 mA g−1, 50 cycles 347 mA h g−1 at 800 mA g−1 146
1D-MnCo2O4 nanowires Hydrothermal 895 mA h g−1 at 100 mA g−1, 50 cycles 345 mA h g−1 at 1000 mA g−1 147
Hierarchical MnCo2O4 nanowires Hydrothermal/calcination approach 1038 mA h g−1 at 200 mA g−1, 45 cycles 388 mA h g−1 at 1600 mA g−1 148
Hierarchical MnCo2O4 nanosheets/carbon cloth Hydrothermal method 3 mA h cm−2 at 800 μA cm−2, 60 cycles 2 mA h cm−2 at 1600 μA cm−2 149
MnCo2O4 microsphere Solvothermal route 722 mA h g−1 at 200 mA g−1, 25 cycles 320 mA h g−1 at 900 mA g−1 150
Mn1.5Co1.5O4 core–shell microsphere Urea assisted solvothermal route 618 mA h g−1 at 400 mA g−1, 300 cycles 455 mA h g−1 at 800 mA g−1 151
Porous NiCo2O4 microflowers Solvothermal route 952 mA h g−1 at 100 mA g−1, 60 cycles 720 mA h g−1 at 500 mA g−1 153
Uniform NiCo2O4 hollow spheres Solvothermal route 885 mA h g−1 at 150 mA g−1, 50 cycles 533 mA h g−1 at 2000 mA g−1 154
NiCo2O4 mesoporous microspheres Solvothermal route 1198 mA h g−1 at 200 mA g−1, 30 cycles 705 mA h g−1 at 800 mA g−1 155
Mesoporous NiCo2O4 Co-precipitation 1000 mA h g−1 at 0.5C, 400 cycles 718 mA h g−1 at 10C 156
Mesopoprous NiCo2O4 nanosheets Microwave method 767 mA h g−1 at 100 mA g−1, 50 cycles 487 mA h g−1 at 1000 mA g−1 158
Porous NiCo2O4 nanobelts Hydrothermal technique 981 mA h g−1 at 500 mA g−1, 100 cycles 1062 mA h g−1 at 2000 mA g−1 159
Porous NixCo3−xO4 nanosheets Thermal decomposition 844 mA h g−1 at 500 mA g−1, 200 cycles 293 mA h g−1 at 1600 mA g−1 160
Hollow NiCo2O4 nanospheres Template method 1346 mA h g−1 at 100 mA g−1, 20 cycles 695 mA h g−1 at 2000 mA g−1 161
NiCo2O4 nanocubes Etching and precipitation 1160 mA h g−1 at 200 mA g−1, 200 cycles 750 mA h g−1 at 1000 mA g−1 162
Porous NiCo2O4@C nanocomposite Hydrothermal 1389 mA h g−1 at 0.55C, 180 cycles 625 mA h g−1 at 4.4C 163
Peapod-like NiCo2O4–C nanorods Hydrothermal 1183 mA h g−1 at 100 mA g−1, 200 cycles 664 mA h g−1 at 2000 mA g−1 164
FeCo2O4 nanoflakes Hydrothermal 905 mA h g−1 at 200 mA g−1, 170 cycles 1222 mA h g−1 at 800 mA g−1 165


After the first discharge, MgCo2O4 converted into MgO, which was then no longer involved in the reaction mechanism, since it acted as an inactive matrix and avoided the volume expansion. Galvanostatic-cycling studies indicated that MgCo2O4 demonstrated serious capacity fading, i.e., it retained only a very small capacity of 108 mA h g−1 after 50 cycles.129 On the other hand, CuCo2O4 and FeCo2O4 possessed reasonable discharge capacities of 747 and 752 mA h g−1, respectively, after 50 cycles. These results might be due to the conversion reactions of both the A- and B-site metal cations. In contrast, only the B-site cations contributed to the charge storage in MgCo2O4.

Among these cobaltites, ZnCo2O4 and its composite has received much attention as a potential material for anodes in LIBs.131–146 In the ZnCo2O4-spinel structure, Zn2+ ions occupy the tetrahedrally coordinated sites, and Co3+ ions occupy the octahedrally coordinated sites, and this structure involves an overarching electrochemical reaction through a conversion reaction between the Co and Zn ions and an alloying reaction between the Zn and Li that provides the overall charge-storage capacity. This reaction mechanism is given as,131–135

 
ZnCo2O4 + 8Li+ + 8e → Zn + 2Co + 4Li2O (27)
 
Zn + Li+ + e ↔ LiZn (28)
 
Zn + Li2O ↔ ZnO + 2Li+ + 2e (29)
 
2Co + 2Li2O ↔ 2CoO + 4Li+ + 4e (30)
 
image file: c5ra23503k-t2.tif(31)

ZnCo2O4 has been widely investigated, and to study its electrochemical performance, researchers have reported various morphologies, including nanotubes,132 3D hierarchical nanowires,133 uniform mesoporous microspheres,136 and yolk–shelled microspheres,137 and they have also grown it on different current collectors, including Ni foam and carbon cloth.135,138 Importantly, porous ZnCo2O4 nanotubes, synthesized via an electrospinning technique with subsequent heat treatment. Fig. 8(a) and (b) show the nanotubes (∼200 nm), which consist of interconnected primary nanocrystals (∼30 nm) and nanopores (∼3 nm) in the walls which exhibited a capacity of 1454 mA h g−1 at 100 mA g−1 over 30 cycles with 100% coulombic efficiency (Fig. 8(c)).132 Even at high current densities of 500, 1000, and 2000 mA g−1, they demonstrated high reversible capacities of 1011, 841, and 794 mA h g−1, indicating that these porous nanotubes could accommodate the volume strain and that the well-interconnected nanoparticles shortened the pathway of Li ion diffusion and enhanced the electronic conductivity, and they possessed large surface areas that led to an increase in the active sites for Li ion diffusion. Liu et al. reported a binder- and current-collector-free 3D hierarchical ZnCo2O4 nanowire array grown on flexible carbon cloth using a hydrothermal method.133 Fig. 9(a) depicts its schematic diagram, while Fig. 9(b)–(e) depict the morphologies of the well-ordered woven structure of the array on carbon cloth in which the ZnCo2O4 nanowires have uniform diameters of 80–100 nm and lengths of 5 μm. Fig. 9(f) shows the cycling-stability curve at 200 mA h g−1. This array had a high irreversible capacity of 1530 mA h g−1 over the initial cycles. After 50 and 100 cycles, it possessed capacities of 1280 and 1278 mA h g−1, respectively, with 99% capacity retention, and still it retained a stable specific capacity of about 1200 mA h g−1 after 160 cycles. The reported capacity was very high when compared with the theoretical specific capacity due to the reversible growth of a polymeric/gel-like film on the surface of the active material. The current density rises upto 5C rate it maintain the stable capacity of 605 mA h g−1 (Fig. 9(g)). The superior electrochemical performance of ZnCo2O4 on carbon cloth depended primarily on the following factors. The ZnCo2O4 nanowires strongly adhered to the carbon cloth, facilitating good electronic conductivity, and the open spaces between the nanowires accommodated the volume strain and allowed electrolyte penetration into the active materials, leading to an increase in the number of active sites. The nanowires also shortened the Li+ ion diffusion path, enhancing the rate capability. It has been suggested that binder- and current-collector-free flexible ZnCo2O4 nanowires on carbon cloth are a good candidate for flexible LIBs.


image file: c5ra23503k-f8.tif
Fig. 8 (a and b) Porous ZnCo2O4 nanotubes and (c) their cycles at different current densities (reproduced from ref. 132. Copyright 2012, with permission from the Royal Society of Chemistry).

image file: c5ra23503k-f9.tif
Fig. 9 (a) Schematic diagram of a ZnCo2O4 nanowire array on carbon cloth. (b)–(e) Morphologies of these ZnCo2O4 nanowires. (f) and (g) Their cycling stability and different current densities (reprinted with permission from ref. 133. Copyright 2012, American Chemical Society).

In addition to that of ZnCo2O4, the Co-based-spinel structures of MnCo2O4 and NiCo2O4 have also been widely examined,147–151,153–163 as Mn-based cobaltite (MnCo2O4) is considered a promising candidate for LIBs due to its low cost, environmental benignity, and lower operating voltage.149 Recently, Fu et al. reported the electrochemical performance of stoichiometric MnCo2O4-spinel mesoporous microspheres prepared via a one-step low-temperature solvothermal method. The microspheres showed a discharge capacity of 722 mA h g−1 over 25 cycles at 200 mA g−1. Due to the hierarchical structure of MnCo2O4,150 they demonstrated a superior rate-capability performance. Li et al. prepared core–shell Mn1.5Co1.5O4 with a nonstoichiometric, spinel structure via a solvothermal route and reported its discharge capacity of 618 mA h g−1 at 400 mA g−1 with good cycling stability even after 300 cycles.151 Interestingly, Hou et al. fabricated the hierarchical structure of MnCo2O4 nanosheet arrays on carbon cloth [Fig. 10(a)–(c)] that demonstrated excellent electrochemical performance, which was even better than that of Co3O4 on carbon cloth149 since the carbon cloth assured good electronic conductivity between the active material and the current collector. Therefore, this type of binder- and additive-free electrode is a promising anode material for LIBs.


image file: c5ra23503k-f10.tif
Fig. 10 (a)–(c) Morphologies of MnCo2O4 nanosheet arrays on carbon cloth. (d) Their cycling stability and (e) rate-capability curve (reproduced from ref. 149. Copyright 2014, with permission from the Royal Society of Chemistry).

Similarly, NiCo2O4 has also been exploited as an efficient energy-storage material for diverse applications, including bifunctional electrocatalysts, supercapacitors, and LIBs.126,152,153–164 It exhibits a high electronic conductivity due to the mixed valences of the same cations in the metal oxides. NiCo2O4 [Co1−x2+Co3+[Co3+Nix2+Ni1−x3+]O4] exhibits two- to three-fold increases in the electrical conductivity, compared to simple metal oxides. For example, the electrical conductivity of Co3O4 is 3.1 × 10−5 S cm−1, and that of NixCo3−xO4 is 0.1–0.3 S cm−1. Hence, the high electronic conductivity of NiCo2O4 enhances the rate capabilities and cycle lives of LIBs.156,157 Park and his group recently reported the excellent electrochemical performance of urchin-like, mesoporous [Fig. 11(a)] NiCo2O4 spinel, which was synthesized via a simple, cost-effective coprecipitation method followed by a low-temperature calcination process.156 It possessed a discharge capacity of 1000 mA h g−1 after 400 cycles at a current rate of 0.1C with a high coulombic efficiency greater than 98%, and it showed an extraordinary rate capability [Fig. 11(b) and (c)]. Instead of conventional PVDF, they used PAA as a binder as it could accommodate the volume strain due to its elastic nature and could assist the efficient electronic conductivity between the active material and the current collector.


image file: c5ra23503k-f11.tif
Fig. 11 (a)–(c) Urchin-like morphology of NiCo2O4, its cycling stability curve at a 0.1C rate, and its current densities at different C rates (reproduced from ref. 156. Copyright 2014 with permission from the Royal Society of Chemistry).

From the observation of these cobaltites, urchin-like mesoporous NiCo2O4 demonstrated an excellent electrochemical performance without any composites because its good electronic conductivity and mesoporous structure led to a high rate capability and high specific capacity. To further improve the electronic and Li ion transports of the urchin-like NiCo2O4, a composite containing carbonaceous material is necessary.

4. Germanates, stannates, and silicates, A2BO4 (A = Zn, Co, Mn, and Co; B = Ge, Sn, Si, and Ti)

It is believed that Zn2GeO4 is an efficient anode material, since ZnO and GeO2 are electrochemically active for Li ion storage through the alloying–de-alloying-reaction mechanism and they exhibit high reversible capacities of 978 and 1100 mA h g−1, respectively.167,168 Ge has 400 and 10[thin space (1/6-em)]000 times greater Li diffusivity and electronic conductivity, respectively, than Si, but during the alloying–de-alloying process, it undergoes a volume change of up to 260% that leads to cracks in the electrode material.168,169 Therefore, to avoid the aggregation of Ge particles and reduce the cost of the electrode material, some active metal oxides such as BaO, CaO, and ZnO have been implemented.170,171 Among these compounds, BaO and CaO act as inactive matrices that reduce the capacity of the material, but ZnO facilitates Li storage and prevents particle aggregation. The alloying–de-alloying-reaction mechanism of Zn2GeO4 is,168,169,172–178
 
Zn2GeO4 + 8Li+ + 8e → 2Zn + Ge + 4Li2O (32)
 
Zn + xLi+ + xe ↔ LixZn (0 ≤ x ≤ 1) (33)
 
Ge + yLi+ + ye ↔ LiyGe (0 ≤ y ≤ 1) (34)
 
Zn + Li2O ↔ ZnO + 2Li+ + 2e (35)
 
Ge + 2Li2O ↔ GeO2 + 4Li+ + 4e (36)

Feng and coworkers analyzed the electrochemical performance of Zn2GeO4 nanorods, synthesized via a hydrothermal method,168 and obtained a discharge capacity of 1820 mA h g−1 in the first cycle, but this value linearly decreased to 616 mA h g−1 after 100 cycles. Therefore, to further enhance the specific capacity and cycling stability, they synthesized amorphous nanoparticles and graphene composite with the Zn2GeO4 particles in order to overcome the volume expansion and enhance the conductivity of the electrode material.169,172–174 The amorphous phase of the Zn2GeO4 nanoparticles showed a high reversible capacity of 1250 mA h g−1 at 400 mA g−1 over 500 cycles, and the capacity was maintained at 470 mA h g−1 at 6400 mA g−1. The high specific capacity and rate capability resulted from the facts that the amorphous structure did not suffer from stresses that mitigated the pulverization of the material and that the incorporation of Zn and oxygen acts as a buffering matrix, thus contributing to the reversible capacity.172

The Huang group reported sandwiched Zn2GeO4–graphene-oxide nanocomposites using an ion-exchange-reaction method [Fig. 12(a)].169 Among these composites, 12.1 wt% of the Zn2GeO4–graphene-oxide nanocomposite showed a good cycling stability, had a specific capacity of 1150 mA h g−1 at 200 mA g−1 over 100 cycles, and maintained a stable capacity of 522 mA h g−1 at a high current density of 3.2 A g−1, as shown in Fig. 12(b) and (c). These Zn2GeO4 nanorods were completely wrapped in the graphene oxide, which thereby accommodated the strain from the volume expansion and preserved the integrity of the active material, thus leading to the improved mobilities of the ions and electrons. Recently, the Wang group explored Fe2GeO4 as an anode material for LIBs, as it has shown excellent electrochemical performance. Fe2GeO4 exhibits a high theoretical specific capacity of 1119 mA h g−1, according to the following electrochemical reactions,27

 
Fe2GeO4 + 8Li+ + 8e → 2Fe + Ge + 4Li2O (37)
 
2Fe + 3Li2O ↔ Fe2O3 + 6Li + 6e (38)
 
Ge + 4.4Li+ + 4.4e ↔ Li4.4Ge (39)


image file: c5ra23503k-f12.tif
Fig. 12 (a) Schematic diagram of Zn2GeO4–graphene-oxide nanocomposites. (b) Cycling-stability curve at 200 mA g−1 at different weight percentages of carbon and (c) the rate-capability curve (reproduced from ref. 169. Copyright 2014, with permission from the Royal Society of Chemistry). (d) TEM image of an Fe2GeO4–graphene composite and (e) its rate-capability curve (reprinted with permission from ref. 27. Copyright 2014 Elsevier).

Fig. 12(d) shows the TEM image of an Fe2GeO4–graphene composite that had a capacity of 980 mA h g−1 at 360 mA g−1 after 175 cycles and a high reversible capacity of 340 mA h g−1 at 4800 mA g−1 is shown in Fig. 12(e).27 This type of A2BO4 structure has some advantages, as the first lithiation evenly distributed between the Li2O matrix and the metal particles. Here, Li2O acts as a buffering matrix, as it can effectively accommodate the volume strain, and the secondary metal nanoparticles enhance the electrical conductivity of the electrode material. These factors lead to electrode materials with both good cyclabilities and good rate performances.179 Zhang et al. first reported a Co2GeO4@reduced-graphene composite as a Li ion anode material that demonstrated a specific capacity of 1085 mA h g−1 over 100 cycles with cycling stability superior to that of Co2GeO4. Reduced-graphene oxide (rGO) served as a good electronic conductor between the particles and an excellent buffering matrix to accommodate the volume strain during the charge–discharge process. Co2GeO4 combined with rGO specifically demonstrated a high specific capacity, good rate capability, and cycling stability.30 Among the germanates, Zn2GeO4 is considered an ideal anode material for LIBs because of its low working potential and high energy density, and the reduced usage of Ge is also beneficial in reducing the cost of the material (27% of Ge in Zn2GeO4) compared to other Ge-based materials.178

A different set of mixed-metal oxides that contain Sn is another group of important prospects for improving the specific capacity of an electrode material. Metal stannates (M2SnO4; M = Mn, Co, Zn) store Li ions through both a conversion-reaction mechanism and an alloying–de-alloying mechanism as follows,180–189

 
M2SnO4 + 8Li → 2M + 4Li2O + Sn (M = Mn, Co) (40)
 
Sn + 4.4Li ↔ Li4.4Sn (41)
 
Sn + 2Li2O ↔ SnO2 + 4Li (42)
 
M + Li2O ↔ MO + 2Li (43)

Lei et al. prepared ultrafine Mn2SnO4 nanoparticles using the thermal decomposition of a MnSn(OH)6 precursor.183 The Mn2SnO4 nanoparticles showed severe capacity fading upon cycling due to volume expansion. During the alloying–de-alloying process, Sn metal expands up to 200–300% (volume expansion), causing peeling of the active material and breaking the electronic conductivity between the active material and the current collector. A similar type of capacity was observed for Co2SnO4 nanoparticles, synthesized via a facile hydrothermal method, which had a capacity of 555 mA h g−1 after 50 cycles at 30 mA g−1 with a capacity retention of 50.3%, and even after 50 cycles, they still could not attain a stable capacity.184 To overcome this drawback, a Co2SnO4@C core–shell structure, a composite with MWCNTs and Co2SnO4 wrapped with graphene, was analyzed to enhance the electrochemical performance of the Co2SnO4.185,188 Qi and coworkers reported that the core–shell structure of Co2SnO4@C, prepared through a hydrothermal process, with different thicknesses of carbon coating (10.5 and 25.2 wt%) was achieved using glucose as the carbon source. Among these materials, the one with the thick carbon coating showed a high discharge capacity of 474 mA h g−1 with a capacity retention of 60.4% after 75 cycles, a value that was greater than that of the one with the thin carbon coating (173 mA h g−1).185 The cycling stability of Co2SnO4 was attributed to the core–shell structure of Co2SnO4@C, and it had several advantages such as stable SEI-film formation, a lack of aggregation with the active material, the enhancement of the electrical conduction with the active material, and functioning as a buffering matrix during the charge–discharge process.186 CNTs are also considered to be important carbonaceous materials that enhance the electrochemical performances of metal oxides, and they have various attractive features, including excellent electrical conductivity, a large length versus diameter ratio, a large surface area, structural flexibility, and chemical stability.187,189–192 Liu and his group proposed a multiwalled-carbon-nanotube (MWCNT) composite with Co2SnO4 nanoparticles that possessed a stable capacity of 898 mA h g−1 at 50 mAg−1 over 50 cycles.187 Compared to the Co2SnO4 nanoparticles, the Co2SnO4–MWCNT composite showed good reversibility with feeble capacity fading that was due to its highly conductive matrix of MWCNTs that facilitated contact between the nanoparticles, accommodated the volume expansion, and prevented the aggregation of nanoparticles during the cycling process.191,192 Recently, Qian and his groups reported Co2SnO4 hollow cubes@rGO, synthesized via the pyrolysis-induced transformation from the hollow, cubic precursor prepared via a hydrothermal method, where the composite with graphene sheets was attained through an electrostatic-interaction mechanism.188 Fig. 13(a–f) show the schematic diagram and TEM images of hollow Co2SnO4 cubes with graphene composite, which exhibit excellent electrochemical performance, as compared to the core–shell structure and the MWCNT composite.187,188 The BET analysis indicated that the Co2SnO4 hollow cubes@rGO exhibited a high surface area of 62.88 m2 g−1 that facilitated an interfacial electrochemical reaction. Fig. 13(g) depicts the discharge capacity versus cycle number and indicates that the specific capacity of Co2SnO4 hollow cubes@rGO was 1126 mA h g−1 at 100 mA g−1 after 50 cycles. This value was greater than the theoretically calculated value (1105 mA h g−1), a discrepancy that resulted from either the interfacial-charge storage or the thick-SEI-film formation,111,193 and this material maintained its capacity of 1016 mA h g−1 after even 100 cycles. For practical applications, they varied the mass loading from 1–2 mg to 5–6 mg, but no obvious change was observed in the cycling process; the discharge capacity was only reduced to 40 mA h g−1, compared to that of the low mass loading. In Fig. 13(h), the Co2SnO4 hollow cubes@rGO had a greatly enhanced rate capability of 775 mA h g−1 at a current density of 500 mA g−1.


image file: c5ra23503k-f13.tif
Fig. 13 (a–d) and (e and f) Schematic diagram of Co2SnO4–graphene-composite formation and its corresponding TEM images respectively. (g) The cycling stability and (h) rate capability of Co2SnO4 nanocrystals, hollow cubes, and graphene composite (reproduced from ref. 188. Copyright 2014, with permission from the Royal Society of Chemistry).

The electrochemical performance was mainly attributed to the hollow cubes of Co2SnO4 and the combined effects of the graphene nanosheets. Both the hollow cubes and the graphene could accommodate the volume expansion and contraction during the charge–discharge process, thereby facilitating the extra capacity from the interfacial storage of Li ions and graphene, improving the conductivity between the particles, and thus leading to a good rate capability of the active material.

Zn2SnO4 belongs to an inverse-spinel group (space group Fd3m) with a band gap of 3.6 eV, and it possesses high electron mobility, high electron conductivity, and low visible absorption.194,195 Due to these astonishing properties, it has been widely investigated for applications in different fields such as gas sensors, solar cells, and photocatalysts.196–198 Apart from these studies, Zn2SnO4 has mainly been investigated as a Li ion anode material due to the low working potentials of ZnO (0.05–1.50 V) and SnO2 (0–1.0 V) as well as their high specific capacities of 980 and 780 mA h g−1, respectively.199,200 Both Zn and Sn are electrochemically active and are involved in the charge-storage mechanism through the alloying mechanism (Li–Zn and Li–Sn), and they could accommodate 14.4 Li ions in the first cycle, demonstrating an irreversible capacity of 1231 mA h g−1 and exhibiting a reversible capacity of 547 mA h g−1, according to the electrochemical-reaction mechanism,201–203

 
Zn2SnO4 + 4Li+ + 4e → Sn + 2Li2O + 2ZnO (44)
 
Zn2SnO4 + 8Li+ + 8e → 2Zn + Sn + 4Li2O (45)
 
Sn + xLi+ + xe ↔ LixSn, (x ≤ 4.4) (46)
 
Zn + yLi+ + ye ↔ LiyZn, (y ≤ 1) (47)

During the charge–discharge process, the Zn2SnO4 material undergoes a large volume change, but it cannot accommodate the volume strain during the cycling process, thus leading to severe capacity fading.201–203 Therefore, many efforts have been made to improve the cycling stability and specific capacity of Zn2SnO4, so particles have been synthesized in different shapes, including cubes, hollow cubes, wires, nanoplates, hollow nanospheres, composites with carbon, graphene metal–organic frameworks, and conducting polymers,195,204–213 and their electrochemical performances are listed in Table 5. Zhang et al. developed hollow Zn2SnO4 nanospheres [Fig. 14(a)] via a hydrothermal method followed by an annealing process, and these hollow nanospheres exhibited a high surface area of 42.53 m2 g−1 and pore volume of 0.24 cm3 g−1.205 Fig. 14(b) and (c) depict these hollow nanospheres with a capacity of 602 mA h g−1 at 100 mA g−1 over 60 cycles and 442 mA h g−1 even at 1000 mA g−1 after 60 cycles. Therefore, this superior electrochemical performance of the hollow nanospheres resulted from different factors, including their large surface areas, enlarging the electrolyte–electrode-contact area and porous structure that allowed the electrolyte into the electrode, increasing the active sites for Li storage and accommodating the volume strain during the charge–discharge process. Recently, hollow Zn2SnO4 boxes@carbon/graphene ternary composite was prepared via a hydrothermal method followed by a calcination process.206 Fig. 14(d) and (e) show the TEM image of the ternary composite and its electrochemical performance, and it exhibited high irreversible discharge capacities of 1863 and 1256 mA h g−1 on the first two cycles. The discharge-capacity values were large compared to the theoretical specific capacity, a result that is due to the extra Li storage on the active material through a pseudocapacitance mechanism.213

Table 5 A summary on electrochemical performance of M2M′O4 (M = Zn and Co, M′ = Sn and Ge) based anodes for LIBs
Sample Experimental methods Reversible capacity Rate performance Ref.
Ultrathin 2D Co2GeO4 nanosheets Hydrothermal method 1026 mA h g−1 at 0.22 A g−1, 150 cycles 600 mA h g−1 at 1.16 A g−1 29
Mn doped Zn2GeO4 nanosheet array Hydrothermal synthesis 1301 mA h g−1 at 100 mA g−1, 100 cycles 500 mA h g−1 at 2 A g−1 175
Coaxial Zn2GeO4@C nanowires Single-step chemical vapour deposition 790 mA h g−1 at 2000 mA g−1, 100 cycles 465 mA h g−1 at 10 A g−1 176
Carbon coated Zn2GeO4 on Ni foam Chemical vapour deposition method 1100 mA h g−1 at 400 mA g−1, 1000 cycles 396 mA h g−1 at 9.6 A g−1 177
Zn2SnO4 nanowires Vapour transport method 660 mA h g−1 at 120 mA g−1, 50 cycles 28
Zn2SnO4 cubic crystals Solid state reaction 689 mA h g−1 at 50 mA g−1, 50 cycles 194
Zn2SnO4 cubes Hydrothermal method 775 mA h g−1 at 50 mA g−1, 20 cycles 195
Zn2SnO4 Supercritical water 856 mA h g−1 at 0.75 mA cm−2, 50 cycles 199
Cube shaped-Zn2SnO4 particles Hydrothermal method 580 mA h g−1 at 100 mA g−1, 50 cycles 201
Zn2SnO4/C composite Hydrothermal method 563 mA h g−1 at 60 mA g−1, 40 cycles 202
Crystalline Zn2SnO4 nanoparticles Hydrothermal method 521 mA h g−1 at 50 mA g−1, 40 cycles 203
Hollow Zn2SnO4@N–C composite Co-precipitation method followed by calcination 616 mA h g−1 at 300 mA g−1, 45 cycles 354 mA h g−1 at 1000 mA g−1 204
Zn2SnO4 hollow nanosphere Hydrothermal method 602 mA h g−1 at 100 mA g−1, 60 cycles 442 mA h g−1 at 1000 mA g−1 205
Hollow Zn2SnO4 boxes@C/graphene ternary composite Hydrothermal process followed by calcination process 726 mA h g−1 at 300 mA g−1, 50 cycles 396 mA h g−1 at 1800 mA g−1 206
Zn2SnO4/graphene composite Hydrothermal method 688 mA h g−1 at 200 mA g−1, 50 cycles 300 mA h g−1 at 1600 mA g−1 207
Hollow Zn2SnO4 boxes wrapped graphene Co-precipitation and electrostatic interaction 678 mA h g−1 at 300 mA g−1, 45 cycles 355 mA h g−1 at 2000 mA g−1 208
Flower-like Zn2SnO4 composite Green hydrothermal synthesis 501 mA h g−1 at 300 mA g−1, 50 cycles 209
PPY coated Zn2SnO4 Microemulsion polymerization 478 mA h g−1 at 60 mA g−1, 50 cycles 211
PANI coated Zn2SnO4 Emulsion polymerization method 491 mA h g−1 at 600 mA g−1, 50 cycles 212
Ex situ carbon coated Zn2SnO4 Hydrothermal synthesis 533 mA h g−1 at 700 mA g−1, 50 cycles 214



image file: c5ra23503k-f14.tif
Fig. 14 (a)–(c) Hollow Zn2SnO4 nanospheres, their cycling stability at 100 mA g−1, and different current densities, respectively (reproduced from ref. 205. Copyright 2014, with permission from the Royal Society of Chemistry). (d)–(f) Zn2SnO4 ternary composite, cycling stabilities of different morphologies of Zn2SnO4 and its composites, and their stabilities at different current densities of the ternary composite (reprinted with permission from ref. 206. Copyright 2014 Elsevier).

The ternary composite retained a capacity of 726 mA h g−1 at 300 mA g−1 over 50 cycles, and when the current density increased to 1200 mA g−1, it maintained a capacity of 438 mA h g−1 (Fig. 14(f)). The ternary composite exhibited an excellent cycling stability and rate performance, as compared to those of hollow Zn2SnO4 boxes wrapped with flexible graphene.206,208 Therefore, their excellent electrochemical performance was attributed to their hollow structure, carbon layer, and graphene composite, which act as a triple buffering structure, efficiently accommodating the volume strain during alloying–de-alloying process. Additionally, the carbon layer and graphene composite enhanced the electronic conductivity of the composite and thus improved the rate capability of the material. Overall, it can be concluded that the Zn2SnO4@graphene composite showed linear capacity fading during the cycling process but that Co2SnO4 showed a high reversible capacity. Nonetheless, the synthetic process of Co2SnO4 is difficult because of the high temperature it requires.186

Silicate-based mixed-metal oxides have also been analyzed for their potential applications to Li ion anode materials. Only a few reports are currently available that analyze silicate-based materials like Zn2SiO4 and Co2SiO4, but these materials do not exhibit high specific capacities compared to that of Co3O4.215,216 Zhang et al. reported Zn2SiO4 nanorods using a facile hydrothermal method, but they possessed a capability of only 388 mA h g−1 at 50 mA g−1 after 20 cycles with that capacity fading on subsequent cycles.215 Mueller and coworkers first reported an orthorhombic α-Co2SiO4 structure, prepared through a solid-state reaction, and examined its electrochemical performance as a Li ion anode material.216

From the cyclic voltammetry, in situ XRD analysis, and XPS results, they deduced the following reaction mechanism,

 
Co2SiO4 + 4Li+ + 4e → 2Co + Li4SiO4, (48)
 
xCo + Li4SiO4 ↔ 4Li+ + 4e + CoxSiO4. (49)

The reported reversible formation of Li4SiO4 enhances the Li ion conductivity that facilitates the high rate capability of the active material and yields a capability of 370 mA h g−1 at a high current density of 1580 mA g−1. However, exploring the detailed electrochemical performances of such silicate-based mixed-metal oxides is still necessary. For instance, Co2SiO4 and Zn2SiO4 materials yield SiO2 and Li4SiO4 during the first discharge processes, respectively, so they are not involved in the electrochemical reactions, leading to low specific capacities and difficulties attaining pure phase formations.217,218 Therefore, it is mandatory to discover alternative anode materials apart from those mentioned above.

5. Role of binder

In Li-ion batteries, binder plays a crucial role to improve the electrochemical performances. It is used as glue, which interconnected the active material and conducting agent that ensures the electronic conductivity between the active materials and the current collector.219 The commercially used PVDF has several drawbacks, i.e., low electronic conductivity, toxic and it dissolves in expensive organic solvent of NMP. Similarly, it could not accommodate volume change during charging–discharging process and adsorb more electrolyte ions that lead to high irreversible capacity loss during first cycle. Therefore, it creates an new avenue to identify an alternate binder instead of PVDF having the characteristics features including water soluble, low cost, high elasticity nature, high electronic conductivity and environmentally friendly.220,221 In this regard, recently many groups have identified different water soluble binders such as CMC (carboxy methyl cellulose), SBR (styrene butadiene rubber), PAA (polyacrylic acid) and SA (sodium alginate) for the applications of Li-ion batteries. CMC is a polymeric derivative of cellulose which consists of carboxylate anion and hydroxyl functional groups and these groups makes CMC as a water-soluble binder. Similarly, SBR is an elastomer, which has high flexibility, stronger binding force and better heat resistance than PVDF.222 On the other hand, PAA is an environmentally friendly and it not only soluble in water it also soluble in organic solvent like ethanol. It has high concentration of functional groups which can give spacing between them through copolymerization with other monomers. The copolymerization of PAA offer many properties including, elastic modulus, maximum elongation, swelling in electrolyte, surface chemistry and the strength of adhesion between the active material and current collector.83 The SA is a high modulus polysaccharide derived from brown algae which mainly consists of copolymer of 1 → 4 linked β-D-mannuronic acid (M) and α-L-guluronic acid (G) residues.223 It has several advantages including ∼6.7 times higher stiffness than dry films of PVDF, small swelling of electrolyte in its structure which restricts the undesirable reaction with the electrode/electrolyte interface, polar hydrogen bonds between the carboxyl groups in the binder that leads to strong adhesive between the active material and current collector and exhibits self healing effect during charging–discharging process.223,224 The following literatures are well depicted the importance of binders. Wei groups have reported that electrochemical performance of ZnFe2O4 using both PVDF and SBR/CMC as binders.222 ZnFe2O4–SBR/CMC delivered high capacity of 627.6 mA h g−1 at 1C rate, which was much larger than that of ZnFe2O4–PVDF (144.9 mA h g−1). Similarly, yolk–shelled microsphere ZnCo2O4, delivered the stable specific capacity of 718 mA h g−1 and 325 mA h g−1, while using CMC and PVDF, respectively as binders over 100 cycles. The result assured that SBR/CMC is a better choice for replacing PVDF, since SBR/CMC makes a unique three dimensional network between the active materials, which enhanced the stability of the electrode during cycling process.137,225 Wang group have analyzed the electrochemical performance of amorphous Zn2GeO4 nanoparticles using PAA as a binder.172 The amorphous ZnGe2O4 nanoparticles showed the high reversible specific capacity of 1250 mA h g−1 over 100 cycles and it exhibited the high discharge capacity of 610 mA h g−1 at the high current density of 3.2 A g−1. This excellent electrochemical performance was mainly attributed to the combined effect of amorphous Zn2GeO4 nanoparticles and PAA binder. Yushin group has first time introduced the sodium alginate as a binder for high capacity silicon nanopowder based lithium ion batteries.84 Mitra groups recently evaluated the electrochemical performance of spinel CoFe2O4 nanoparticles using sodium alginate binder which exhibits the stable discharge capacity of 890 mA h g−1 over 50 cycles but using PVDF it stabilized at 160 mA h g−1 over 50 cycles. On the other hand, PAA used CoFe2O4 electrode delivered 470 mA h g−1 even at high current density of 20C.226 Similarly, Wang group reported the Co2GeO4 nanosheets for anode materials using sodium alginate as a binder that could possess stable capacity of 1026 mA h g−1 over 150 cycles. So, this excellent electrochemical performance such as cyclability and rate capability mainly attributed to the peculiar properties of sodium alginate binder.29 Overall binders, the sodium alginate shows the good cycling stability and high rate capability compared to other so it is one the best binder to construct the Li-ion batteries.

6. Summary and outlook

In this overview, numerous research has been examined that focuses on the electrochemical performances of AB2O4- and A2BO4-structured materials and their corresponding pristine materials and that expounds on the large electrode polarization, high irreversible capacity loss, and unstable SEI-film formation that lead to poor cycling stability and rate capability. Hence, to achieve better electrochemical performances, researchers have focused on different strategies as summarized here. (i) The electrode materials have been prepared in different morphologies such as nanoparticles, hierarchical structures, nanowires, and nanotubes using different synthetic methods, including hydrothermal, solvothermal, thermal decomposition, and coprecipitation. (ii) Pristine materials have been composited with carbonaceous compounds such as core–shells of carbon and graphene. (iii) Binders such as CMC, PAA, and sodium alginate also play an important role in facilitating the electrochemical performance. (iv) Finally, electrolyte additives such as fluoroethylene carbonate (FEC) also contribute to cycling stability.

The first potential approach, i.e., while scaling down the electrode materials in to nanoscale, the large surface area will be obtained. This nanosize provides an access to the maximum number of reaction sites at the electrode–electrolyte interface as well as creating short diffusion-path length. Also it minimizes the volume expansion during Li-insertion when compared to the bulk-material counterparts, thus enhancing the specific capacities. However, this large surface area will leads to some unwanted side reactions with electrode–electrolyte interface which in turn the formation of a large amount of SEI layer that degrades the capacity. Mainly nanostructured electrode material has isolated often repeated cycling that inhibits the efficient flow of ions and electrons which leads to fast capacity fading on, especially at high current densities.227–230

The next approach involves hierarchical structures. Here, the secondary nanoparticles that tend to aggregate themselves in oriented manner to form hierarchical microstructures. These are potentially reducing the side reactions with the electrolyte, increases the tap density and the available nanoscale building blocks increases the kinetics of lithium ion as well as the electrical properties when compared to nanoparticles.105,231,232 However, unstable SEI-film formation causes high irreversible capacities and low coulombic efficiencies in the initial cycles.

Another approach involves the formation of composites with carbonaceous materials. The main advantages is that this prevents the side reactions with electrode/electrolyte interface and accommodate more volume strain. Among the composite, core–shell structure is the simplest strategy to enhance the electrochemical performance. Core consists of active material to involve the electrochemical reaction and shell consists of amorphous carbon which can act as protection layer to strengthen the performance of core particles. The shell has several advantages including, protecting the core from the unwanted side reactions with electrolyte, due to its flexible nature it can accommodate the volume strain and maintain the integrity, prevent the aggregation between the particles. The carbon shell has high electronic conductivity that leads to enhance the lithium ion diffusion rate. The shell thickness is also important factor that influences the electrochemical performance of core particles. So to make a proper thickness of carbon coating on the active material enhances the electrochemical performances.4,214 Binders are another important strategy that can overcomes the core–shell structure and graphene composite, because binders such as CMC, PAA, and sodium alginate are capable of completely covering the active material and flexibly accommodate the swelling of the electrode material.219,233 They also restrict the peel of active material from the current collector, facilitating the tight attachment of the active material on the current collector and enhancing the electronic conductivity. However, the binder PVDF does not.219 The aforementioned key factors improve the electrochemical performance by accommodating the volume strain and enhancing the electronic conductivity and adhesion to the current collector. However, several problems are also persistent; stable SEI-film formation and large irreversible capacity loss mainly cause poor cycling stabilities and low coulombic efficiencies in the first cycles, thus leading to the safety problems of using LIBs. But it can be moderately overcome by using electrolyte additives such as fluoroethylene carbonate. FEC electrolyte additives facilitate to thin SEI film formation and reducing the irreversible capacity. Therefore, it is increasing the cycle life by means of modifying the surface chemistry of the SEI film.87,234

In summation, it is concluded that the electrochemical performance of the mixed-transition-metal oxides, including characteristics such as the cycling stability and rate capability, are enhanced in the following ways: synthesized mixed-metal oxides with large surface areas with porosity, composites with carbonaceous materials (core–shells and graphene), highly flexible binders (CMC, PAA, and sodium alginate), and electrolyte additives such as fluoroethylene carbonate that reduce both the irreversible capacity loss and the formation of stable SEI films. The above modifications not only make mixed-metal oxides promising high-energy-density anode materials for LIBs, but due to their multivalent states and better electronic conductivities compared to corresponding simple metal oxides, they may also be excellent candidates for bifunctional electrocatalysts and supercapacitors. Therefore, mixed-metal oxides are some of the most important forthcoming materials for the next generation of energy-storage devices.

Acknowledgements

One of the authors (RKS) is grateful to UGC (No. 41-838/2012 (SR)) for their financial support under UGC-MRP. S. Yuvaraj would like to thank UGC-BSR (No. G2/5357/2013) for providing the fellowship to carry out this work successfully.

Notes and references

  1. I. Romieu, F. Menesses, S. Ruiz, J. J. Sienra, J. Huetra, M. C. White and R. A. Etzel, Am. J. Respir. Crit. Care Med., 1996, 154, 300–307 CrossRef CAS PubMed.
  2. A. Haines, R. S. Kovats, D. Campbell-Lendrum and C. Corvalan, Publ. Health, 2006, 120, 585–596 CrossRef CAS PubMed.
  3. P. Knippertz, M. J. Evans, P. R. Field, A. H. Fink, C. Liousse and J. H. Marsham, Nat. Clim. Change, 2015, 5, 815–822 CrossRef CAS.
  4. L. Ji, Z. Lin, M. Alcoutlabi and X. Zhang, Energy Environ. Sci., 2011, 4, 2682–2699 CAS.
  5. J. W. Long, B. Dunn, D. R. Rolison and H. S. White, Chem. Rev., 2004, 104, 4463–4492 CrossRef CAS PubMed.
  6. M. G. Kim and J. Cho, Adv. Funct. Mater., 2009, 19, 1497–1514 CrossRef CAS.
  7. M. Winter and R. J. Brodd, Chem. Rev., 2004, 104, 4245–4270 CrossRef CAS PubMed.
  8. J. B. Goodenough and Y. Kim, Chem. Mater., 2010, 22, 587–603 CrossRef CAS.
  9. C. M. Hayner, X. Zhao and H. H. Kung, Annu. Rev. Chem. Biomol. Eng., 2012, 3, 445–471 CrossRef CAS PubMed.
  10. M. M. Thackerey, C. Wolverton and E. D. Isaacs, Energy Environ. Sci., 2012, 5, 7854–7863 Search PubMed.
  11. K. Persson, V. A. Sethuraman, L. J. Hardwick, Y. Hinuma, Y. S. Meng, A. Van der Ven, V. Srinivasan, R. Kostecki and G. Ceder, J. Phys. Chem. Lett., 2010, 1, 1176–1180 CrossRef CAS.
  12. J. M. Tarascon and M. Armand, Nature, 2001, 414, 359–367 CrossRef CAS PubMed.
  13. P. Poizot, S. Laruelle, S. Grugeon, L. Dupont and J. M. Tarascon, Nature, 2000, 407, 496–499 CrossRef CAS PubMed.
  14. J. Carbana, L. Monconduit, D. Larcher and M. R. Palacin, Adv. Mater., 2010, 22, E170–E192 CrossRef PubMed.
  15. F. Cheng, Z. Tao, J. Liang and J. Chen, Chem. Mater., 2008, 20, 667–681 CrossRef CAS.
  16. A. S. Arico, P. Bruce, B. Scrosati, J. M. Tarascon and W. V. Schalkwijk, Nat. Mater., 2005, 4, 366–377 CrossRef CAS PubMed.
  17. B. T.-Y. Wei, C. H. Chen, H.-C. Chien, S.-Y. Lu and C. C. Hu, Adv. Mater., 2010, 22, 347–351 CrossRef PubMed.
  18. L. Qiao, X. H. Wang, L. Qiao, X. L. Sun, X. W. Li, Y. X. Zheng and D. Y. He, Nanoscale, 2013, 5, 3037–3042 RSC.
  19. B. Das, M. V. Reddy, S. Tripathy and B. V. R. Chowdari, RSC Adv., 2014, 4, 33883–33889 RSC.
  20. C. N. Chinnasamy, A. Narayanasamy, N. Ponpandian, K. Chattopadhyay, K. Shinoda, B. Jeyadevan, K. Tohji and K. Nakatsuka, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 63, 184108–184114 CrossRef.
  21. D. Carta, M. F. Casula, A. Falqui, D. Loche, G. Mountjoy, C. Sangregorio and A. Corrias, J. Phys. Chem. C, 2009, 113, 8606–8615 CAS.
  22. N. Guigue Millot, S. Begin-Colin, Y. Champion, M. H'tch, G. Le Caer and P. Perriat, J. Solid State Chem., 2003, 170, 30–38 CrossRef CAS.
  23. M. Mouallem-Bahout, S. Bertrand and O. Pena, J. Solid State Chem., 2005, 178, 1080–1086 CrossRef CAS.
  24. C. M. Park, J. H. Kim, H. Kim and H. J. Sohn, Chem. Soc. Rev., 2010, 39, 3115–3141 RSC.
  25. W. J. Zhang, J. Power Sources, 2011, 196, 877–885 CrossRef CAS.
  26. T. P. Kumar, R. Ramesh, Y. Y. Lin and G. T. K. Fey, Electrochem. Commun., 2004, 6, 520–525 CrossRef CAS.
  27. S. Jin and C. Wang, Nano Energy, 2014, 7, 63–71 CrossRef CAS.
  28. C. T. Cherian, M. Zheng, M. V. Reddy, B. V. R. Chowdari and C. H. Sow, ACS Appl. Mater. Interfaces, 2013, 5, 6054–6060 CAS.
  29. S. Jin, G. Yang, H. Song, H. Cui and C. Wang, ACS Appl. Mater. Interfaces, 2015, 7, 24932–24943 CAS.
  30. F. Zhang, R. Zhang, Z. Zhang, H. Wang and J. Feng, Electrochim. Acta, 2014, 150, 211–217 CrossRef CAS.
  31. J. Jiang, J. Luo, X. Huang, J. Liu and T. Xu, Nanoscale, 2013, 5, 8105–8113 RSC.
  32. C. W. Jung and P. Jacobs, Magn. Reson. Imaging, 1995, 13, 661–674 CrossRef CAS PubMed.
  33. J. P. Liu, Nanoscale Mag. Mater. and Appl., Springer Verlag, 2009 Search PubMed.
  34. N. Ikenaga, Y. Ohgaito, H. Matsushima and T. Suzuki, Fuel, 2004, 83, 661–669 CrossRef CAS.
  35. Y. Yin, B. Zhang, X. Zhang, J. Xu and S. Yang, J. Sol-Gel Sci. Technol., 2013, 66, 540–543 CrossRef CAS.
  36. S. M. Hoque, M. Abdul Hakin, A. Mamun, S. Akhter, Md. Tanvir Hasan, D. Prasad Paul and K. Chattopadhayay, Mater. Sci. Appl., 2011, 2, 1564–1571 CAS.
  37. Y. Pan, Y. Zhang, X. Wei, C. Yuan, J. Yin, D. Cao and G. Wang, Electrochim. Acta, 2013, 109, 89–94 CrossRef CAS.
  38. C. Gong, Y.-J. Bai, Y.-X. Qi, N. Luna and J. Feng, Electrochim. Acta, 2013, 90, 119–127 CrossRef CAS.
  39. A. K. Rai, T. V. Thi, J. Gim and J. Kim, Mater. Charact., 2014, 95, 259–265 CrossRef CAS.
  40. D. S. Mathew and R.-S. Juang, Chem. Eng. J., 2007, 129, 51–65 CrossRef CAS.
  41. A. Shanmugavani, R. Kalai Selvan, S. Layek and C. Sanjeeviraja, J. Magn. Magn. Mater., 2014, 354, 363–371 CrossRef CAS.
  42. Y. Sharma, N. Sharma, G. V. Subba Rao and B. V. R. Chowdari, Electrochim. Acta, 2008, 53, 2380–2385 CrossRef CAS.
  43. P. F. Teh, Y. Sharma, S. S. Pramana and M. Srinivasan, J. Mater. Chem., 2011, 21, 14999–15008 RSC.
  44. W. Song, J. Xie, S. Liu, G. Cao, T. Zhu and X. Zhao, New J. Chem., 2012, 36, 2236–2241 RSC.
  45. J. Sui, C. Zhang, D. Hong, J. Li, Q. Cheng, Z. Li and W. Cai, J. Mater. Chem., 2012, 22, 13674–13681 RSC.
  46. N. N. Wang, H. Xu, L. Chen, X. Gu, J. Yang and Y. Qian, J. Power Sources, 2014, 247, 163–169 CrossRef CAS.
  47. Y. Deng, Q. Zhang, S. Tang, L. Zhang, S. Deng, Z. Shi and G. Chen, Chem. Commun., 2011, 47, 6828–6830 RSC.
  48. Z. Xing, Z. Ju, J. Yang, H. Xu and Y. Qian, Nano Res., 2012, 5, 477–485 CrossRef CAS.
  49. F. Mueller, D. Bresser, E. Paillard, M. Winter and S. Passerini, J. Power Sources, 2013, 236, 87–94 CrossRef CAS.
  50. J. M. Won, S. H. Choi, Y. J. Hong, Y. Na Ko and Y. C. Kang, Sci. Rep., 2014, 4(5857), 1–5 Search PubMed.
  51. H. Xia, Y. Qian, Y. Fu and X. Wang, Solid State Sci., 2013, 17, 67–71 CrossRef CAS.
  52. L. Yao, X. Hou, S. Hu, J. Wang, M. Li, C. Su, M. O. Tade, Z. Shao and X. Liu, J. Power Sources, 2014, 258, 305–313 CrossRef CAS.
  53. D. Bresser, E. Pailard, R. Kloepsch, S. Krueger, M. Fiedler, R. Schmitz, D. Baither, M. Winter and S. Passerini, Adv. Energy Mater., 2013, 3, 513–523 CrossRef CAS.
  54. X. Yao, J. Kong, D. Zhou, C. Zhao, R. Zhou and X. Lu, Carbon, 2014, 79, 493–499 CrossRef CAS.
  55. L. Lin and Q. Pan, J. Mater. Chem. A, 2015, 3, 1724–1729 CAS.
  56. B. Wang, S. Li, B. Li, J. Liu and M. Yu, New J. Chem., 2015, 39, 1725–1733 RSC.
  57. R. M. Thankachan, M. M. Rahman, I. Sultana, A. M. Glushenkov, S. Thomas, N. Kalarikkal and Y. Chen, J. Power Sources, 2015, 282, 462–470 CrossRef CAS.
  58. Y. Ding, Y. Yang and H. Shao, Solid State Ionics, 2012, 217, 27–33 CrossRef CAS.
  59. L. Luo, R. Cui, H. Qiao, K. Chen, Y. Fei, D. Li, Z. Pang, K. Liu and Q. Wei, Electrochim. Acta, 2014, 144, 85–91 CrossRef CAS.
  60. L. Jin, Y. Qiu, H. Deng, W. Li, H. Li and S. Yang, Electrochim. Acta, 2011, 56, 9127–9132 CrossRef CAS.
  61. M. Bomio, P. Lavela and J. Tirado, ChemPhysChem, 2007, 8, 1999–2007 CrossRef CAS PubMed.
  62. Z. Xing, Z. Ju, J. Yang, H. Xu and Y. Qian, Electrochim. Acta, 2013, 102, 51–57 CrossRef CAS.
  63. B. Aslibeiki, P. Kameli, H. Salamati, M. Eshraghi and T. Tahmasebi, J. Magn. Magn. Mater., 2010, 322, 2929–2934 CrossRef CAS.
  64. X. Cui, S. Belo, D. Kruger, Y. Yan, R. T. M. Rosales, M. J. Osoro, H. Ye, S. Su, D. Mathe, N. Kovacs, I. Horvath, M. Semjeni, K. Sunssee, K. Szigeti, M. A. Green and P. J. Blower, Biomaterials, 2014, 35, 5840–5846 CrossRef CAS PubMed.
  65. J. Dobson, Drug Dev. Res., 2006, 67, 55–60 CrossRef CAS.
  66. Q. A. Pankhurst, J. Connolly, S. Jones and J. Dobson, J. Phys. D: Appl. Phys., 2003, 36, 167–181 CrossRef.
  67. Y. P. Lin and N. L. Wu, J. Power Sources, 2011, 196, 851–854 CrossRef CAS.
  68. Z. Zhang, Y. Wang, Q. Tan, Z. Zhong and F. Su, J. Colloid Interface Sci., 2013, 398, 185–192 CrossRef CAS PubMed.
  69. S. W. Oh, H. J. Bang, Y. C. Bae and Y. K. Sun, J. Power Sources, 2007, 173, 502–509 CrossRef CAS.
  70. X. W. Lou, L. A. Archer and Z. C. Yang, Adv. Mater., 2008, 20, 3987–4019 CrossRef CAS.
  71. A. Latz and J. Zausch, J. Power Sources, 2011, 196, 3296–3302 CrossRef CAS.
  72. Y. Xiao, J. Zai, L. Tao, B. Li, Q. Han, C. Yu and X. Qian, Phys. Chem. Chem. Phys., 2013, 15, 3939–3945 RSC.
  73. H. Xiang, K. Zhang, G. Ji, J. Y. Lee, C. Zou, X. Chen and J. Wu, Carbon, 2011, 49, 1787–1796 CrossRef CAS.
  74. H. Zhao, Z. Zheng, K. W. Wong, S. Wang, B. Huang and D. Li, Electrochem. Commun., 2007, 9, 2606–2610 CrossRef CAS.
  75. C. Vidal-Abarca, P. Lavela and J. L. Tirado, J. Phys. Chem. C, 2010, 114, 12828–12832 CAS.
  76. P. Lavela and J. L. Tirado, J. Power Sources, 2007, 172, 379–387 CrossRef CAS.
  77. C. T. Cherian, J. Sundaramurthy, M. V. Reddy, P. S. Kumar, K. Mania, D. Pliszka, C. H. Sow, S. Ramakrishna and B. V. R. Chowdari, ACS Appl. Mater. Interfaces, 2013, 5, 9957–9963 CAS.
  78. Y. Ding, Y. Yang and H. Shao, J. Power Sources, 2013, 244, 610–613 CrossRef CAS.
  79. Y. Fu, Y. Wan, H. Xia and X. Wang, J. Power Sources, 2012, 213, 338–342 CrossRef CAS.
  80. G. Zhang, L. Yu, H. B. Wu, H. E. Hoster and X. W. Lou, Adv. Mater., 2012, 24, 4609–4613 CrossRef CAS PubMed.
  81. G. Zhou, D. W. Wang, F. Li, L. Zhang, N. Li, Z. S. Wu, L. Wen, G. Q. Lu and H. M. Cheng, Chem. Mater., 2010, 22, 5306–5313 CrossRef CAS.
  82. E. M. Hetdari, B. Zhang, M. H. Sohi, A. Ataie and J. K. Kim, J. Mater. Chem. A, 2014, 2, 8314–8322 Search PubMed.
  83. A. Magasinski, B. Zdyrko, I. Kovalenko, B. Hertzberg, R. Burtovyy, C. F. Huebner, T. F. Fuller, I. Luzinov and G. Yushin, ACS Appl. Mater. Interfaces, 2010, 2, 3004–3010 CAS.
  84. I. Kovalenko, B. Zdyrko, A. Magasinski, B. Hertzberg, Z. Millicev, R. Burtovyy, I. Luzinov and G. Yushin, Science, 2011, 334, 75–79 CrossRef CAS PubMed.
  85. P. Ramesh Kumar and S. Mitra, RSC Adv., 2013, 3, 25058–25064 RSC.
  86. A. M. Chockla, T. D. Bogart, C. M. Hessel, K. C. Klavetter, C. B. Mullins and B. A. Korgel, J. Phys. Chem. C, 2012, 116, 18079–18086 CAS.
  87. A. M. Chockla, K. C. Klavetter, C. B. Mullins and B. A. Korgel, Chem. Mater., 2012, 24, 3738–3745 CrossRef CAS.
  88. Y. Xiao, J. Zai, X. Li, Y. Gong, B. Li, Q. Han and X. Qian, Nano Energy, 2014, 6, 51–58 CrossRef CAS.
  89. T. Yoon, C. Chae, Y. K. X. Zhao, H. H. Kung and J. K. Lee, J. Mater. Chem., 2011, 21, 17325–17330 RSC.
  90. P. Lavela, J. L. Tirado, M. Womes and J. C. Jumas, J. Phys. Chem. C, 2009, 113, 20081–20087 CAS.
  91. P. R. Kumar, P. Kollu, C. Santhosh, K. E. V. Rao, D. K. Kim and A. N. Grace, New J. Chem., 2014, 38, 3654–3661 RSC.
  92. P. Fen Teh, S. S. Pramana, Y. Sharma, Y. Wen Ko and S. Madhavi, ACS Appl. Mater. Interfaces, 2013, 5, 5461–5467 Search PubMed.
  93. J. Cabana, L. Monconduit, D. Larcher and M. R. Palacin, Adv. Mater., 2010, 22, E170–E192 CrossRef CAS PubMed.
  94. S. Li, B. Wang, L. Liu and M. Yu, Electrochim. Acta, 2014, 129, 33–39 CrossRef CAS.
  95. Z. Zhang, W. Li, R. Zou, W. Kang, Y. S. Cui, M. F. Yuen, C.-S. Lee and W. Zhang, J. Mater. Chem. A, 2015, 3, 6990–6997 CAS.
  96. Z. Zhang, Y. H. Wang, M. J. Zhang, Q. Q. Tan, X. Lv, Z. Y. Zhong and F. B. Su, J. Mater. Chem. A, 2013, 1, 7444–7450 CAS.
  97. M. Zhang, X. Yang, X. Kan, X. Wang, L. Ma and M. Jia, Electrochim. Acta, 2013, 112, 727–734 CrossRef CAS.
  98. H. Xia, D. Zhu, Y. Fu and X. Wang, Electrochim. Acta, 2012, 83, 166–174 CrossRef CAS.
  99. Q. Q. Xiong, J. P. Tu, S. J. Shi, X. Y. Liu, X. L. Wang and C. D. Gu, J. Power Sources, 2014, 256, 153–159 CrossRef CAS.
  100. Y. J. Mai, X. H. Xia, R. Chen, C. D. Gu, X. L. Wang and J. P. Tu, Electrochim. Acta, 2012, 67, 73–78 CrossRef CAS.
  101. W. Wei, J. L. Wang, L. J. Zhou, J. Yang, B. Schumann and Y. N. Nuli, Electrochem. Commun., 2012, 13, 742–745 Search PubMed.
  102. F. M. Courtel, H. Duncan, Y. A. Lebdeh and I. J. Davidson, J. Mater. Chem., 2011, 21, 10206–10218 RSC.
  103. L. Zhou, H. B. Shao, J. M. Wang, L. Liu, J. Q. Zhang and C. N. Cao, J. Mater. Chem., 2012, 22, 827–829 RSC.
  104. C. Yuan, H. B. Wu, Y. Xie and X. Wen (David) Lou, Angew. Chem., Int. Ed., 2014, 53, 1488–1504 CrossRef CAS PubMed.
  105. L. Hu, H. Zhong, X. Zheng, Y. Huang, P. Zhang and Q. Chen, Sci. Rep., 2012, 2(986), 1–8 Search PubMed.
  106. L. Zhou, D. Zhao and X. W. Lou, Adv. Mater., 2012, 24, 745–748 CrossRef CAS PubMed.
  107. G. Zhang, H. B. Wu, H. E. Hoster and X. W. (David) Lou, Energy Environ. Sci., 2014, 7, 302–305 CAS.
  108. L. Wang, B. Liu, S. Ran, L. Wang, L. Gao, F. Qu, D. Chen and G. Shen, J. Mater. Chem. A, 2013, 1, 2139–2143 CAS.
  109. G. Yang, X. Xu, W. Yan, H. Yang and S. Ding, Electrochim. Acta, 2014, 137, 462–469 CrossRef CAS.
  110. J. Li, S. Xiong, X. Li and Y. Qian, Nanoscale, 2013, 5, 2045–2054 RSC.
  111. W. Kang, Y. Tang, W. Li, X. Yang, H. Xue, Q. Yang and C.-S. Lee, Nanoscale, 2015, 7, 225–231 RSC.
  112. M. Li, D. Zhou, W.-L. Song, X. Li and L.-Z. Fan, J. Mater. Chem. A, 2015, 3, 19907–19912 CAS.
  113. Y. Yang, Y. Zhao, L. Xiao and L. Zhang, Electrochem. Commun., 2008, 10, 1117–1120 CrossRef CAS.
  114. L. Xiao, Y. Yang, J. Yin, Q. Li and L. Zhang, J. Power Sources, 2009, 194, 1089–1093 CrossRef CAS.
  115. Y. Deng, S. Tang, Q. Zhang, Z. Shi, L. Zhang, S. Zhan and G. Chen, J. Mater. Chem., 2011, 21, 11987–11995 RSC.
  116. F. M. Courtel, Y. A. Lebdeh and I. J. Davidson, Electrochim. Acta, 2012, 71, 123–127 CrossRef CAS.
  117. K. Xu, Chem. Rev., 2004, 104, 4303–4418 CrossRef CAS PubMed.
  118. J. G. Kim, S. H. Lee, Y. Kim and W. B. Kim, ACS Appl. Mater. Interfaces, 2013, 5, 11321–11328 CAS.
  119. N. Wang, X. Ma, H. Xu, L. Chen, J. Yue, F. Niu, J. Yang and Y. Qian, Nano Energy, 2014, 6, 193–199 CrossRef CAS.
  120. Y. Liu, J. Bai, X. Ma, J. Li and S. Xiong, J. Mater. Chem. A, 2014, 2, 14236–14244 CAS.
  121. Z. Bai, N. Fan, C. Sun, Z. Ju, C. Guo, J. Yang and Y. Qian, Nanoscale, 2013, 5, 2442–2447 RSC.
  122. P. Xiong, B. Liu, V. Teran, Y. Zhao, L. Peng, X. Wang and G. Yu, ACS Nano, 2014, 8, 8610–8616 CrossRef CAS PubMed.
  123. L. Zhou, H. B. Wu, T. Zhu and X. W. (David) Lou, J. Mater. Chem., 2012, 22, 827–829 RSC.
  124. W. Yao, J. Xu, W. J. Luo, Q. Shi and Q. Zhang, ACS Sustainable Chem. Eng., 2015, 3, 2170–2177 CrossRef CAS.
  125. S. Vijayanand, P. A. Joy, H. S. Potdar, D. Patil and P. Patil, Sens. Actuators, B, 2011, 152, 121–129 CrossRef CAS.
  126. M. Prabu, K. Ketpang and S. Shanmugam, Nanoscale, 2014, 6, 3173–3181 RSC.
  127. Q. Lu, Y. Chen, W. Li, J. G. Chen, J. Q. Xiao and F. Jiao, J. Mater. Chem. A, 2013, 1, 2331–2336 CAS.
  128. Y. Sharma, N. Sharma, G. V. Subba Rao and B. V. R. Chowdari, J. Power Sources, 2007, 173, 495–501 CrossRef CAS.
  129. Y. Sharma, N. Sharma, G. V. Subba Rao and B. V. R. Chowdari, Solid State Ionics, 2008, 179, 587–597 CrossRef CAS.
  130. S. Sun, Z. Wen, J. Jin, Y. Cui and Y. Lu, Microporous Mesoporous Mater., 2013, 169, 242–247 CrossRef CAS.
  131. B. Y. Sharma, N. Sharma, G. V. Subba Rao and B. V. R. Chowdari, Adv. Func. Mater., 2007, 17, 2855–2861 CrossRef.
  132. W. Luo, X. Hu, Y. Sun and Y. Huang, J. Mater. Chem., 2012, 22, 8916–8921 RSC.
  133. B. Liu, J. Zhang, X. Wang, G. Chen, D. Chen, C. Zhou and G. Shen, Nano Lett., 2012, 12, 3005–3011 CrossRef CAS PubMed.
  134. Q. Xie, F. Li, H. Guo, L. Wang, Y. Chen, G. Yue and D. L. Peng, ACS Appl. Mater. Interfaces, 2013, 5, 5508–5517 CAS.
  135. H. Long, T. Shi, S. Jiang, S. Xi, R. Chen, S. Liu, G. Liao and Z. Tang, J. Mater. Chem. A, 2014, 2, 3741–3748 CAS.
  136. L. Hu, B. Qu, C. Li, Y. Chen, L. Mei, D. Lei, L. Chen, Q. Li and T. Wang, J. Mater. Chem. A, 2013, 1, 5596–5602 CAS.
  137. J. Li, J. Wang, D. Wexler, D. Shi, J. Liang, H. Liu, S. Xiong and Y. Qian, J. Mater. Chem. A, 2013, 1, 15292–15299 CAS.
  138. B. Qu, L. Hu, Q. Li, Y. Wang, L. Chen and T. Wang, ACS Appl. Mater. Interfaces, 2014, 6, 731–736 CAS.
  139. Y. Wang, J. Ke, Y. Zhang and Y. Huang, J. Mater. Chem. A, 2015, 3, 24303–24308 CAS.
  140. A. K. Rai, T. V. Thi, B. J. Paul and J. Kim, Electrochim. Acta, 2014, 146, 577–584 CrossRef CAS.
  141. X. Song, Q. Ru, Y. Mo, S. Hu and B. An, J. Alloys Compd., 2014, 606, 219–225 CrossRef CAS.
  142. X. Song, Q. Ru, B. Zhang, S. Hu and B. An, J. Alloys Compd., 2014, 585, 518–522 CrossRef CAS.
  143. N. Du, Y. Xu, H. Zhang, J. Yu, C. Zhai and D. Yang, Inorg. Chem., 2011, 50, 3320–3324 CrossRef CAS PubMed.
  144. R. Wu, X. Qian, K. Zhou, J. Wei, J. Lou and P. M. Ajayan, ACS Nano, 2014, 8, 6297–6303 CrossRef CAS PubMed.
  145. S. H. Choi and Y. C. Kang, ChemSusChem, 2013, 6, 2111–2116 CrossRef CAS PubMed.
  146. S. G. Mohamed, T.-F. Hung, C.-J. Chen, C. K. Chen, S.-F. Hu, R.-S. Liu, K.-C. Wang, X.-K. Xing, H.-M. Liu, A.-S. Liu, M.-H. Hsieh and B.-J. Lee, RSC Adv., 2013, 3, 20143–20149 RSC.
  147. L. Li, Y. Q. Zhang, X. Y. Liu, S. J. Shi, X. Y. Zhao, X. Ge, G. F. Cai, C. D. Gu, X. L. Wang and J. P. Tu, Electrochim. Acta, 2014, 116, 467–474 CrossRef CAS.
  148. S. G. Mohamed, T.-F. Hung, C.-J. Chen, C. K. Chen, S.-F. Hu and R.-S. Liu, RSC Adv., 2014, 4, 17230–17235 RSC.
  149. X. Hou, X. Wang, B. Liu, Q. Wang, T. Luo, D. Chen and G. Shen, Nanoscale, 2014, 6, 8858–8864 RSC.
  150. C. Fu, G. Li, D. Luo, X. Huang, J. Zheng and L. Li, ACS Appl. Mater. Interfaces, 2014, 6, 2439–2449 CAS.
  151. J. Li, S. Xiong, X. Li and Y. Qian, J. Mater. Chem., 2012, 22, 23254–23259 RSC.
  152. A. Shanmugavani and R. Kalai Selvan, Electrochim. Acta, 2016, 189, 283–294 CrossRef CAS.
  153. J. Xu, L. He, W. Xu, H. Tang, H. Liu, T. Han, C. Zhang and Y. Zhang, Electrochim. Acta, 2014, 145, 185–192 CrossRef CAS.
  154. L. Shen, L. Yu, X.-Y. Yu, X. Zhang and X. W. (David) Lou, Angew. Chem., 2015, 54, 1868–1872 CrossRef CAS PubMed.
  155. J. Li, S. Xiong, Y. Liu, Z. Ju and Y. Qian, ACS Appl. Mater. Interfaces, 2013, 5, 981–988 CAS.
  156. H. S. Jadhav, R. S. Kalubarme, C.-N. Park, J. Kim and C.-J. Park, Nanoscale, 2014, 6, 10071–10076 RSC.
  157. L. Hu, L. Wu, M. Liao, X. Hu and X. Fang, Adv. Funct. Mater., 2012, 22, 998–1004 CrossRef CAS.
  158. A. K. Mondal, D. Su, S. Chen, K. Kretschmer, X. Xie, H.-J. Ahn and G. Wang, ChemPhysChem, 2015, 16, 169–175 CrossRef CAS PubMed.
  159. A. K. Mondal, D. Su, S. Chen, X. Xie and G. Wang, ACS Appl. Mater. Interfaces, 2014, 6, 14827–14835 CAS.
  160. F. Zheng, D. Zhu and Q. Chen, ACS Appl. Mater. Interfaces, 2014, 6, 9256–9264 CAS.
  161. X. Yao, C. Zhao, J. Kong, D. Zhou and X. Lu, RSC Adv., 2014, 4, 37928–37933 RSC.
  162. H. Guo, L. Liu, T. Li, W. Chen, J. Liu, Y. Guo and Y. Guo, Nanoscale, 2014, 6, 5491–5497 RSC.
  163. Y. Mo, Q. Ru, X. Song, S. Hu, L. Guo and X. Chen, Electrochim. Acta, 2015, 176, 575–585 CrossRef CAS.
  164. L. Peng, H. Zhang, Y. Bai, J. Yang and Y. Wang, J. Mater. Chem. A, 2015, 3, 22094–22101 CAS.
  165. S. G. Mohamed, C.-J. Chen, C. K. Chen, S.-F. Hu and R.-S. Liu, ACS Appl. Mater. Interfaces, 2014, 6, 22701–22708 CAS.
  166. X. Shen, D. Mu, S. Chen, B. Wu and F. Wu, ACS Appl. Mater. Interfaces, 2013, 5, 3118–3125 CAS.
  167. Y. M. Lin, K. C. Klavetter, A. Heller and C. B. Mullins, J. Phys. Chem. Lett., 2013, 4, 999–1004 CrossRef CAS PubMed.
  168. J. K. Feng, M. O. Lai and L. Lu, Electrochem. Commun., 2011, 13, 287–289 CrossRef CAS.
  169. F. Zou, X. Hu, L. Qie, Y. Jiang, X. Xiong, Y. Qiao and Y. Huang, Nanoscale, 2014, 6, 924–930 RSC.
  170. W. Li, Y. X. Yin, S. Xin, W. G. Song and Y. G. Guo, Energy Environ. Sci., 2012, 5, 8007–8013 CAS.
  171. W. W. Li, X. F. Wang, B. Liu, S. J. Luo, Z. Liu, X. J. Hou, Q. Y. Xiang, D. Chen and G. Z. Shen, Chem.–Eur. J., 2013, 19, 8650–8656 CrossRef CAS PubMed.
  172. R. Yi, J. K. Feng, D. P. Lv, M. L. Gordin, S. R. Chen, D. W. Choi and D. H. Wang, Nano Energy, 2013, 2, 498–504 CrossRef CAS.
  173. F. Zou, X. Hu, Y. Sun, W. Luo, F. Xia, L. Qie, Y. Jiang and Y. Huang, Chem.–Eur. J., 2013, 19, 6027–6033 CrossRef CAS PubMed.
  174. R. Wang, S. Wu, Y. Lv and Z. Lin, Langmuir, 2014, 30, 8215–8220 CrossRef CAS PubMed.
  175. Q. Li, X. Miao, C. Wang and L. Yin, J. Mater. Chem. A, 2015, 3, 21328–21336 CAS.
  176. W. Chen, L. Lu, S. Maloney, Y. Yang and W. Wang, Phys. Chem. Chem. Phys., 2015, 17, 5109–5114 RSC.
  177. W. Chen, S. Maloney and W. Wang, Electrochim. Acta, 2015, 176, 96–102 CrossRef CAS.
  178. Y. Feng, X. Li, Z. Shao and H. Wang, J. Mater. Chem. A, 2015, 3, 15274–15279 CAS.
  179. C. H. Kim, Y. S. Jung, K. T. Lee, J. H. Ku and S. M. Oh, Electrochim. Acta, 2009, 54, 4371–4377 CrossRef CAS.
  180. P. A. Connor and J. T. S. Irvine, J. Power Sources, 2001, 97–98, 223–225 CrossRef CAS.
  181. P. A. Connor and J. T. S. Irvin, Electrochim. Acta, 2002, 47, 2885–2892 CrossRef CAS.
  182. R. Alcantara, G. F. Ortiz, P. Lavela and J. L. Tirado, Electrochem. Commun., 2006, 8, 731–736 CrossRef CAS.
  183. S. Lei, K. Tang, C. Chen, Y. Jin and L. Zhou, Mater. Res. Bull., 2009, 44, 393–397 CrossRef CAS.
  184. G. Wang, X. P. Gao and P. W. Shen, J. Power Sources, 2009, 192, 719–723 CrossRef CAS.
  185. Y. Qi, N. Du, H. Zhang, P. Wu and D. Yang, J. Power Sources, 2011, 196, 10234–10239 CrossRef CAS.
  186. S. Yuvaraj, S. Amaresh, Y. S. Lee and R. Kalai Selvan, RSC Adv., 2014, 4, 6407–6416 RSC.
  187. G. Wang, Z. Y. Liu and P. Liu, Electrochim. Acta, 2011, 56, 9515–9519 CrossRef CAS.
  188. J. Zhang, J. Liang, Y. Zhu, D. Wei, L. Fan and Y. Qian, J. Mater. Chem. A, 2014, 2, 2728–2734 CAS.
  189. J. M. Schnorr and T. M. Swagar, Chem. Mater., 2011, 23, 646–657 CrossRef CAS.
  190. B. J. Landi, M. J. Ganter, C. D. Cress, R. A. Dileo and R. P. Raffaelle, Energy Environ. Sci., 2009, 2, 638–654 CAS.
  191. F. F. Cao, Y. G. Guo, S. F. Zheng, X. L. Wu, L. Y. Jiang, R. R. Bi, L. J. Wan and J. Maier, Chem. Mater., 2010, 22, 1908–1914 CrossRef CAS.
  192. S. F. Feng, J. S. Hu, L. S. Zhong, W. G. Song, L. J. Wan and Y. G. Guo, Chem. Mater., 2008, 20, 3617–3622 CrossRef.
  193. H. C. Liu and S. K. Yen, J. Power Sources, 2007, 166, 478–484 CrossRef CAS.
  194. X. Hou, Q. Cheng, Y. Bai and W. F. Zhang, Solid State Ionics, 2010, 181, 631–634 CrossRef CAS.
  195. N. Feng, S. Peng, X. Sun, L. Qiao, X. Li, P. Wang, D. Hu and D. He, Mater. Lett., 2012, 76, 66–68 CrossRef CAS.
  196. J. Yu and G. Choi, Sens. Actuators, B, 2001, 72, 141–148 CrossRef CAS.
  197. B. Tan, E. Toman, Y. Li and Y. Wu, J. Am. Chem. Soc., 2007, 129, 4162–4163 CrossRef CAS PubMed.
  198. Y. Lin, S. Lin, M. Luo and J. Liu, Mater. Lett., 2009, 63, 1169–1171 CrossRef CAS.
  199. J. W. Lee and C. H. Lee, J. Supercrit. Fluids, 2010, 55, 252–258 CrossRef CAS.
  200. F. Belliard, P. A. Connor and J. T. S. Irvine, Solid State Ionics, 2000, 135, 163–167 CrossRef CAS.
  201. A. Rong, X. P. Gao, G. R. Li, T. Y. Yan, H. Y. Zhu, J. Q. Qu and D. Y. Song, J. Phys. Chem. B, 2006, 110, 14754–14760 CrossRef CAS PubMed.
  202. W. S. Yuan, Y. W. Tian and G. Q. Liu, J. Alloys Compd., 2010, 506, 683–687 CrossRef CAS.
  203. K. Kim, A. Annamalai, S. H. Park, T. H. Kwon, M. W. Pyeon and M. J. Lee, Electrochim. Acta, 2012, 76, 192–200 CrossRef CAS.
  204. Y. Zhao, Y. Huang, Q. Wang, K. Wang, M. Zong, L. Wang and X. Sun, Ceram. Int., 2014, 40, 2275–2280 CrossRef CAS.
  205. R. Zhang, Y. He and L. Xu, J. Mater. Chem. A, 2014, 2, 17979–17985 CAS.
  206. H. Huang, Y. Huang, M. Wang, X. Chen, Y. Zhao, K. Wang and H. Wu, Electrochim. Acta, 2014, 147, 201–208 CrossRef CAS.
  207. W. Song, J. Xie, W. Hu, S. Liu, G. Cao, T. Zhu and X. Zhao, J. Power Sources, 2013, 229, 6–11 CrossRef CAS.
  208. Y. Zhao, Y. Huang, X. Sun, H. Huang, K. Wang, M. Zong and Q. Wang, Electrochim. Acta, 2014, 120, 128–132 CrossRef CAS.
  209. K. Wang, Y. Huang, Y. Shen, L. Xue, H. Huang, H. Wu and Y. Wang, Ceram. Interfaces, 2014, 40, 8021–8025 CrossRef CAS.
  210. X. Zheng, Y. Li, Y. Xu, Z. Hong and M. Wei, CrystEngComm, 2012, 14, 2112–2116 RSC.
  211. K. Wang, Y. Huang, T. Han, Y. Zhao, H. Huang and L. Xue, Ceram. Interfaces, 2014, 40, 2359–2364 CrossRef CAS.
  212. Y. Zhao, Y. Huang, L. Xue, X. Sun, Q. Wang, W. Zhang, K. Wang and M. Zong, Polym. Test., 2013, 32, 1582–1587 CrossRef CAS.
  213. G. F. Ma, H. Peng, J. J. Mu, H. H. Huang, X. Z. Zhou and Z. Q. Lei, J. Power Sources, 2013, 229, 72–78 CrossRef CAS.
  214. S. Yuvaraj, W. J. Lee, C. W. Lee and R. Kalai Selvan, RSC Adv., 2015, 5, 67210–67219 RSC.
  215. S. Zhang, M. Lu, Y. Li, F. Sun, J. Yang and S. Wang, Mater. Lett., 2013, 100, 89–92 CrossRef CAS.
  216. F. Mueller, D. Bresser, N. Minderjahn, J. Kalhoff, S. Menne, S. Krueger, M. Winter and S. Passerini, Dalton Trans., 2014, 43, 15013–15021 RSC.
  217. S. Yuvaraj, K. Karthikeyan, L. Vasylechko and R. Kalai Selvan, Electrochim. Acta, 2015, 158, 446–456 CrossRef CAS.
  218. S. Zhang, L. Ren and S. Peng, CrystEngComm, 2014, 16, 6195–6202 RSC.
  219. S.-L. Chou, Y. Pan, J.-Z. Wang, H.-K. Liu and S.-X. Dou, Phys. Chem. Chem. Phys., 2014, 16, 20347–20359 RSC.
  220. S. S. Zhang, K. Xu and T. R. Jow, J. Power Sources, 2004, 138, 226 CrossRef CAS.
  221. J. H. Lee, S. Lee, U. Paik and Y. M. Choi, J. Power Sources, 2005, 147, 249 CrossRef CAS.
  222. R. Zhang, X. Yang, D. Zhang, H. Qiu, Q. Fu, H. Na, Z. Guo, F. Du, G. Chen and Y. Wei, J. Power Sources, 2015, 285, 227–234 CrossRef CAS.
  223. J. S. Bridel, T. Azais, M. Morcrette, J. M. Tarascon and D. Larcher, Chem. Mater., 2010, 22, 1229 CrossRef CAS.
  224. D. Mazouzi, B. Lestriez, L. Roue and D. Guyomard, Electrochem. Solid-State Lett., 2009, 12, A215 CrossRef CAS.
  225. U. Kumar Sen and S. Mitra, ACS Appl. Mater. Interfaces, 2013, 5, 1240–1247 Search PubMed.
  226. S. Mitra, P. S. Veluri, A. Chakraborthy and R. K. Petla, ChemElectroChem, 2014, 1, 1068–1074 CrossRef CAS.
  227. H. B. Wu, J. S. Chen, H. H. Hng and X. Wen (David) Lou, Nanoscale, 2012, 4, 2526–2542 RSC.
  228. J. Xiao, J. Zheng, X. Li, Y. Shao and J.-G. Zhang, Nanotechnology, 2013, 24, 424004–424011 CrossRef PubMed.
  229. H. Zhang, X. Yu and P. Braun, Nat. Nanotechnol., 2011, 6, 277–281 CrossRef CAS PubMed.
  230. H. Li, L. Shen, G. Pang, S. Fang, H. Luo, K. Yang and X. Zhang, Nanoscale, 2015, 7, 619–624 RSC.
  231. D. Li, C. Feng, H. K. Liu and Z. Guo, Sci. Rep., 5, 11326 CrossRef CAS PubMed.
  232. J. Xiao, Nano Lett., 2011, 11, 5071–5079 CrossRef CAS PubMed.
  233. L. Su, Y. Jing and Z. Zhou, Nanoscale, 2011, 3, 3967–3983 RSC.
  234. V. Etacher, O. Haik, Y. Goffer, G. A. Roberts, I. C. Stefen, R. Fasching and D. Aurbach, Langmuir, 2012, 28, 965–976 CrossRef PubMed.

This journal is © The Royal Society of Chemistry 2016