Mechanism studies of LiFePO4 cathode material: lithiation/delithiation process, electrochemical modification and synthetic reaction

Feng Yu aef, Lili Zhang b, Yingchun Li c, Yongxin An d, Mingyuan Zhu a and Bin Dai *a
aKey Laboratory for Green Processing of Chemical Engineering of Xinjiang Bingtuan, School of Chemistry and Chemical Engineering, Shihezi University, Shihezi 832003, P.R. China. E-mail: yufeng05@mail.ipc.ac.cn; daibin_bce@shzu.edu.cn
bInstitute of Chemical and Engineering Sciences, Agency for Science, Technology and Research (A*STAR), Jurong Island 627833, Singapore
cKey Laboratory of Xinjiang Phytomedicine Resources of Ministry of Education, School of Pharmacy, Shihezi University, Shihezi 832002, P.R. China
dGraphene & Energy Storage Technology Research Center, China Energine International (Holdings) Limited, Beijing 100176, P.R. China
eEngineering Research Center of Materials-Oriented Chemical Engineering of Xinjiang Production and Construction Corps, Shihezi 832003, P.R. China
fKey Laboratory of Materials-Oriented Chemical Engineering of Xinjiang Uygur Autonomous Region, Shihezi 832003, P.R. China

Received 10th August 2014 , Accepted 30th September 2014

First published on 1st October 2014


Abstract

Olivine-structured lithium ion phosphate (LiFePO4) is one of the most competitive candidates for fabricating energy-driven cathode material for sustainable lithium ion battery (LIB) systems. However, the high electrochemical performance is significantly limited by the slow diffusivity of Li-ion in LiFePO4 (ca. 10−14 cm2 s−1) together with the low electronic conductivity (ca. 10−9 S cm−1), which is the big challenge currently faced by us. To resolve the challenge, many efforts have been directed to the dynamics of the lithiation/delithiation process in LixFePO4 (0 ≤ x ≤ 1), mechanism of electrochemical modification, and synthetic reaction process, which are crucial for the development of high electrochemical performance for LiFePO4 material. In this review, in order to reflect the recent progress ranging from the very fundamental to practical applications, we specifically focus on the mechanism studies of LiFePO4 including the lithiation/delithiation process, electrochemical modification and synthetic reaction. Firstly, we highlight the Li-ion diffusion pathway in LixFePO4 and phase translation of LixFePO4. Then, we summarize the modification mechanism of LiFePO4 with high-rated capability, excellent low-temperature performance and high energy density. Finally, we discuss the synthetic reaction mechanism of high-temperature carbothermal reaction route and low-temperature hydrothermal/solvothermal reaction route.


image file: c4ra10899j-p1.tif

Feng Yu

Feng Yu received his B.Sc. degree in applied chemistry from the University of Jinan in 2003, and obtained his Ph.D. degree in physical chemistry from the Technical Institute of Physics and Chemistry, Chinese Academy of Sciences (TIPC, CAS) in 2010, working on olive-typed LiMPO4 (M = Fe, Mn, Co and Ni) electrode materials for lithium ion batteries (LIBs) as electrochemical energy storage (EES) devices. He then joined Nanyang Technological University (NTU) as a research fellow and the Institute of Chemical and Engineering Sciences, Agency for Science, Technology and Research (ICES, A*STAR), as a research scientist working on the synthesis and electrochemical performance of functional electrode materials. In 2013, he joined Shihezi University as an associate professor to undertake research on advanced electrode materials for high power EES devices.

image file: c4ra10899j-p2.tif

Lili Zhang

Lili Zhang received her B. Eng. in Chemical and Biomolecular Engineering from the National University of Singapore (NUS) in 2004. After two years of industrial experience in Micron, she continued her Ph.D. study in the same department at NUS and received her Ph.D. degree in 2011. She worked as a research engineer from 2010 to 2011 at NUS and a research fellow in Professor Ruoff's group at the University of Texas at Austin from 2011 to 2012. Now, she is a research scientist in the Institute of Chemical and Engineering Sciences, Agency for Science, Technology and Research (ICES, A*STAR), Singapore. Her research interest is developing high performance energy storage materials and systems.

image file: c4ra10899j-p3.tif

Yingchun Li

Yingchun Li received her B.Sc. degree from Shihezi University in 2003, and M.Sc. degree from Xi'an Jiaotong University in 2006. She continued her Ph.D. study in Martin Luther University of Halle-Wittenberg and received her Ph.D. degree in 2011. Then, she joined Shihezi University as a scientist of the “Recruitment Program of Global Experts” (1000 Talent Plan). Now, Prof. Li continued her research in Shihezi University, and her research field involves the development of advanced functional materials, especially sp2-type carbon materials including carbon nanotubes and graphene. Additionally, she has developed advanced sensors and molecularly imprinted polymer for serving pharmaceutical analysis, screening of pharmaceutical activity and modernization of traditional Chinese medicine.

image file: c4ra10899j-p4.tif

Yongxin An

Yongxin An obtained his B.Sc. and M. Sc. Degrees from Heilongjiang University in 2002 and 2007, respectively. Then, he received his Ph.D. degree from Harbin Institute of Technology in 2010. He then joined the Institute of Chemical and Engineering Sciences, Agency for Science, Technology and Research (ICES, A*STAR), as a research scientist working on the electrochemical performance of electrolytes for lithium ion batteries (LIBs). In 2013, he joined China Energine International (Holdings) Limited as a director of Graphene & Energy Storage Technology Research Center.

image file: c4ra10899j-p5.tif

Mingyuan Zhu

Mingyuan Zhu received his B.Sc. degree from Shandong University in 2003, and obtained his Ph.D. degree from the Dalian Institute of Chemical Physics, Chinese Academy of Sciences (DICP, CAS) in 2009. He then joined Shihezi University. He is currently an associate professor in Shihezi University. His research focuses on advanced functional materials for powerful energy storage including fuel cells and supercapacitors.

image file: c4ra10899j-p6.tif

Bin Dai

Bin Dai obtained his Ph.D. degree in applied chemistry from the Dalian University of Technology (DUT) in 2002. He is the director of Key Laboratory for Green Processing of Chemical Engineering of Xinjiang Bingtuan and a professor at Shihezi University. Prof. Dai has over 20 years of experience in advanced functional materials, working for government, industry and university. His research interests have included porous micro-nano-structured materials for powerful energy storage, plasma catalysis, catalyst for acetylene hydrochlorination, and heterocyclic carbene catalysis.


1 Introduction

As excellent electrochemical energy storage (EES) devices, lithium ion batteries (LIBs) have recently attracted significant attention, since the reversible lithiation/delithiation reaction of LiCoO2 was discovered in 1980 and the use of LiCoO2 as cathode materials for LIBs in 1990.1,2 As compared to conventional lead–acid, nickel–cadmium and nickel–metal hydride batteries, rechargeable LIBs possess high working voltage and superior energy density.3,4 LIBs are not only widely used in consumer electronics, such as cell phones, cameras, toys and laptops, but also used to power increasingly emerging large-scale applications such as electric vehicles (EVs) and hybrid electric vehicles (HEVs).5,6 Nowadays, LIBs would facilitate the regulation of imbalance in electrical power grids and are the closest to coupling the ultimate requirement of advanced energy storage technologies for renewable energy sources including solar power, wind, and ocean waves.7,8 Noticeably, LIBs are efficiently meeting the exigent demands of modern energy technology (ET), which is urgently needed to the thrust area closely linked to combustion engines and environmental pollution.9,10 When combined with renewable energy sources, LIB-based ET is an imperative step to replace the inevitably vanishing non-renewable fossil fuels and avoid negative effects from the current combustion-based ETs on global energy and environmental problems.11,12 LIBs have become the most viable and promising candidates for EES devices, which strongly minimize the environment impact and maximize energy and resources utilization.13,14

Although LIBs are well positioned to satisfy the needs of modern society and emerging ecological concerns, one of the greatest challenges is unquestionably the cathode materials where the Li-ions extracting/inserting process occurs.15,16 Scientists have been focused on the crystal structures and the electrochemical performance of potential cathode materials, such as olivine-structured LiMPO4 (M = Fe, Co, Ni, Mn), α-NaFeO2 layered LiMO2 (M = Co, Ni, Mn), monoclinic structured Li1+xV3O8, orthogonal structured Li2MSiO4 (M = Fe, Mn), spinel structured LiMn2O4 and NASCION structured Li3V2(PO4)3.17,18 Among the aforementioned cathode materials, layer-structured LiCoO2 with two-dimensional (2D) Li-ion transport has been extensively utilized in LIBs. However, it suffers from high toxicity, inferior safety and high cost.19 Recently, cubic-spinelled LiMn2O4 supporting three-dimensional (3D) Li-ion transport has also been widely used in high powerful EES devices. However, its poor cycle life due to the Jahn–Teller effect remains a big concern.20,21

Because of the groundbreaking work conducted by Goodenough and co-workers,22,23 phosphate polyanionic compound LiFePO4 has attracted tremendous attention and has been used as cathode material in LIBs because of its fantastic performance, such as high theoretical capacity (170 mA h g−1), acceptable operating voltage (3.45 V vs. Li+/Li), long cycle life (>2000 cycles), superior safety, low cost, low toxicity, abundant resources, and environmental benignity.24,25 In the orthorhombic olivine-structured LiFePO4, the oxygen atoms are located in a hexagonal close-packed and slightly distorted arrangement.26,27 The phosphorus atoms are on tetrahedral sites and forming PO4 tetrahedra with oxygen atoms. Lithium atoms from LiO6 octahedra occupy edge-sharing octahedral positions, while iron atoms from FeO6 octahedra occupy corner-sharing octahedral positions. Due to the edge-sharing chains along the [010] direction (i.e., b-axis) created by the LiO6 octahedra, one-dimensional (1D) Li-ion transport is formed in the [010] direction. At the common corners in the bc plane, one FeO6 octahedron is chained with four FeO6 octahedra resulting in zigzag planes parallel to the [001] direction (i.e., c-axis). Each FeO6 octahedron shares one edge with PO4 tetrahedra and two LiO6 octahedra, respectively, while PO4 tetrahedra has two common edges with LiO6 octahedra.28 This special olivine-structured LiFePO4 results in spectacular niches including excellent structural flexibility and superior thermal stability, and it provides excellent cycling capabilities and safe characteristics superior to other cathode materials.29–31

However, there are some limitations of high rate capability due to simplex 1D Li-ion transport of the olive-structured LiFePO4, which is different from the layer-structured LiCoO2 providing 2D Li-ion transport and the cubic-spinelled LiMn2O4 supporting 3D Li-ion transport.18,32 Its high-rate performance is significantly limited by the slow diffusivity of the Li-ion in LiFePO4 (ca. 10−14 cm2 s−1) together with the low electronic conductivity (ca. 10−9 S cm−1).23,33 The poor power density of LiFePO4 cathode for LIBs limits its applications in power-demanding EVs and HEVs.15,34 The development of cathode materials for LIBs with high-rate capability, low-temperature performance, good cycle life and high energy density are still challenging. Efforts have been devoted to the understanding of the dynamics of the lithiation/delithiation process LixFePO4 (0 ≤ x ≤ 1), the electrochemical reaction mechanism, and optimization of synthetic approach, which are crucial for the development of high performance LiFePO4 material.

In this review, we highlight the mechanism studies of LiFePO4 including the lithiation/delithiation process, electrochemical property modification and synthetic approach, which are necessary to fully employ its great potential for practical applications. Firstly, we introduce Li-ion diffusion pathway in LixFePO4 and phase translation of LixFePO4 to provide fundamentals for the development of high electrochemical performance LiFePO4 material. Secondly, we focus on the crucial parameters representing the performances of LiFePO4, which mainly include specific capacity, rate capability, tap density and low temperature performance. Comprehensive improvements of all those parameters are desired and are also the targets for future research. Finally, we discuss the synthetic reaction mechanism of carbothermal reaction (CTR) route and hydrothermal/solvothermal reaction route, which are two of the most important synthesis routes that need to be improved urgently. We believe this review not only benefits the research of LiFePO4 but also provides an additional strategy for other materials potentially used in EES devices.

2 Lithiation/delithiation process in LiFePO4

2.1 Li-ion diffusion pathways of LiFePO4

2.1.1 One-dimensional Li-ion diffusion along [010] direction. Owing to the fairly compact olivine structure, LiFePO4 exhibits intrinsic thermal stability in the fully charged state, which makes a major contribution to LIBs safety. Moreover, LiFePO4 presents excellent structural flexibility, which provides remarkably good cycling stability when compared with cathode materials. However, there are still some obstacle and limitations of its high rate capability. Thus, it has been critical to explore Li-ion diffusion pathway and explain the phase transition during the lithiation/delithiation process of LiFePO4.27,35 In recent years, first principle-based modeling has become an important tool to study the reactions during charge and discharge. In 2004, Ceder and his co-workers36 firstly employed first-principles method to study Li-ion transport direction in LiFePO4. They demonstrated that the Li-ion diffusion coefficient was successfully calculated to be several orders of magnitude higher in the [010] direction (i.e. b-axis) than in the [001] direction (i.e. c-axis). The result of D[001]/D[010] ≈ 10−37 clearly shows that the [001] direction hardly makes any contribution to the Li-ion motion. Li-ion diffuses through 1D channels along the [010] direction with low energy barriers to cross between the channels due to the FeO6 octahedral transition state in the [001] direction being face-sharing with two PO4 tetrahedra. Islam's group26,37 further designed the structural modeling of LiFePO4 and used atomistic simulation method to investigate Li-ion migration energy in LiFePO4. They found that the energy of Li-ion migration is Emig([010]) = 0.55 eV lower than Emig([001]) = 2.89 eV and Emig([101]) = 3.36 eV, which strongly indicates a preference for Li-ion transport along the [010] direction and confirms the results of Ceder's group. Some more theoretical calculations have also been reported and suggested that Li-ion transport is along the [101] direction due to the lower Li-ion migration energy than the other directions.38,39

Moreover, Li-ion migration was preferential down the [010] channels following a curved trajectory, which is confirmed by Yamada's group.40 As shown in Fig. 1, ellipsoids representing Li-ions were refined with 95% probability by Rietveld analysis and motions of Li atoms evolve from vibration to diffusion as an expected curved one-dimensional continuous chain. Moreover, Tse's group41 found that Li-ion diffusion is a dominant process through a series of jumps from one site to another, resulting into a zigzag pathway along the crystallographic [010] direction. As mentioned above, the Li-ion diffusion pathway mainly occur in the [010] direction, which creates simplex one-dimensional Li-ion transport pathway and is different from that of 2D layer-structured LiCoO2 and 3D cubic-spinelled LiMn2O4. A 1D Li-ion transport tunnel is considered as the main cause of slow Li-ion diffusion, which significantly restricts the high rate performance of LiFePO4.42


image file: c4ra10899j-f1.tif
Fig. 1 (a) Anisotropic harmonic lithium vibration in LiFePO4 shown as green thermal ellipsoids and the expected curved one-dimensional diffusion pathway (dashed lines show how the motions of Li atoms evolve from vibrations to diffusion). (b) Three-dimensional Li nuclear density data shown as blue contours. The brown octahedra represent FeO6 and the purple tetrahedral represent PO4 units. (Equi-value 0.15 fm Å−3 of the negative portion of the coherent nuclear scattering density distribution.) (c) Two-dimensional contour map sliced on the (001) plane at z = 0.5 and (d) the (010) plane at y = 0, whereas Fe, P and O atoms remain near their original positions.40 (Copyright © 2008, Rights Managed by Nature Publishing Group.)

In order to experimentally prove whether Li-ion diffusion pathway is one-dimensional along the [010] direction, ionic conductivity (or ionic mobility) is generally measured and used to calculate ionic diffusivity. Various techniques including cyclic voltammetry (CV), galvanostatic intermittent titration technique (GITT), potentiostatic intermittent titration technique (PITT), electrochemical impedance spectroscopy (EIS), etc., can be used to measure ionic conductivity under an applied voltage.43 Vaknin's group44 employed AC impedance spectroscopy to study the Li-ion conductivity of three principal directions (i.e. [100], [010] and [001]) in single crystal LiFePO4 as a function of temperature (325–525 K). Their results indicate that ionic conductivity along the [010] direction is much higher than that in both [100] and [001] directions, which agrees well with the computational analysis. Meanwhile, Yamada's group40 provided clear experimental visualization of Li-ion transport LixFePO4 (x = 0.6) by combining high-temperature (620 K) powder neutron diffraction and the maximum entropy method (MEM), as shown in Fig. 1b–d. To a large extent, lithium delocalizes along the continuous curved 1D chain along the [010] direction on account of the probability density of lithium nuclei. The result is consistent with the aforementioned computational predictions. Therefore, the lithium diffusion constant in nanoparticles with shortened transport paths is much faster than that in bulk.45

2.1.2 Two-dimensional Li-ion diffusion along both [010] and [001] directions. The second possible direction of Li transport has been postulated due to the chains separated by octahedral interstitial sites along the [001] direction at elevated temperatures, which suggest that 2D Li-ion transport exist along both [010] and [001] directions.23,46 As shown in Fig. 2a, Maier's group47 investigated ionic and electronic conductivities in single crystalline LiFePO4 as a function of crystallographic orientation over an extended temperature range, and they found that the activation energies obtained for ionic conductivities σLi+ along the [010] and [001] orientations are (i.e., Eact([001]) = Eact([010]) = 0.62 eV) smaller than that along the [100] direction (i.e., Eact([100]) = 0.74 eV), which corresponds to effectively two-dimensional Li+ conduction. In addition, their study suggested that the activation energies presented for Li-ion diffusion DLiδ along the [010] and [001] directions are comparable (i.e., Eact([001]) = 0.75 eV, Eact([010]) = 0.70 eV) distinctly less than that of the [100] direction (i.e., Eact([100]) = 0.96 eV) (see in Fig. 2b), which indicated a preferential 2D Li-ion chemical diffusion in the bc plane. The phenomenon accounted for the effective two-dimensional pattern of Li-ion conductivity and diffusion (i.e., isotropic in the bc plane). Their study suggested that the ionic conductivity magnitudes in both the [010] and [001] directions and the diffusion coefficient are similar within the temperature range of 140–147 °C, indicating that 2D Li-ion transport must occur within these temperatures.
image file: c4ra10899j-f2.tif
Fig. 2 (a) Comparison of ionic and electronic conductivities along different crystallographic orientations. (b) Chemical diffusion coefficient along different directions.47 (Copyright © 2008 Elsevier B.V. All rights reserved.) (c) For the anti-site defect shown, Fe sits in the channel at site M1 and Li sits in the FePO4 lattice at site M2. Dashed lines represent the diffusion pathways for Li ions in the lattice. (d) Energy landscape for diffusion around the defect resulting in cross-channel diffusion of Li ions (red) and vacancies (blue).49 (Copyright © 2011, American Chemical Society.)

Furthermore, Nazar's group48 has also reported that the small polar on the carrier mobility is predicted to be “two-dimensional” with the motions of Li ions as well as electrons being correlated using Mössbauer spectroscopy measurements. They gave experimental evidence for a strong correlation between electron and lithium delocalization events suggesting they are coupled. In 2011, Henkelman's group49 reported the various components of Li-ion kinetics in LiFePO4 calculated from the density functional theory (DFT). As shown in Fig. 2c, there are different kinetic pathways of Li-ion diffusion on the surface, in the bulk, in the presence of defects, and in varying local environments. It is observed that surface diffusion had high barriers resulting into slow kinetics in LiFePO4. Moreover, the slow bulk diffusion was possibly affected by strain and Li concentration. The slow vacancy diffusion in LiFePO4 was explained by anti-site defects, which has a barrier of 0.71 eV (Fig. 2d) as compared to 0.29 eV in defect-free channels. Intriguingly, a concerted Li-ion diffusion in FePO4 exhibited a low barrier of 0.35 eV, allowing for facile cross-channel diffusion at room temperature.

2.2 Phase transition between LiFePO4 and FePO4

2.2.1 Two-phase transformation between LiFePO4 and FePO4. The lithiation/delithiation process in LixFePO4 (0 ≤ x ≤ 1) is commonly proposed as a two-phase reaction mechanism, which includes various models: shrinking core (i.e., core–shell) model,23,50 Laffont's (i.e., new core–shell) model,51 mosaic model,52,53 domino-cascade model,54,55 phase transformation wave model,56,57 and many-particle model.58 Generally, the two-phase growth process involves the coexistence of LiFePO4 and FePO4, particularly for large particles (e.g., particle size >100 nm).

In 1997, Goodenough's group23 firstly reported that Li-ion exhibited a two-phase transport between LiFePO4 and FePO4 due to the flat charge/discharge profile and introduced the “core–shell” model. With delithiation, LiFePO4 changes into FePO4 and yields two phases, forming a two-phase interface. This model is generally named the “core–shell” model. This model can explain the reaction. However, it cannot explain the continuous deviation of the open circuit voltage (OCV) from 3.45 V at the start and end of discharge. Subsequently, Srinivasan's group50 proposed a “shrinking core” model to describe the lithiation of FePO4. This model showed that outside the two-phase coexistence region, there would be a corresponding single phase region, where lithiation proceeds from the surface of a particle moving the two phase interface. With delithiation in the charging process, the FePO4 shell formed and the FePO4/LiFePO4 interface migrated into each particle. Inefficient delithiation from the uncovered LiFePO4 at the center of the larger particles easily leads to capacity loss. Simultaneously, this shrinking-core model can successfully describe the electrochemical charge/discharge profiles at various rates.59

Similar to the shrinking-core model, Thomas's group52,53 proposed a “mosaic model” by introducing a new concept, i.e., lithiation/delithiation starting at different nucleation sites. These two models are generally considered as “the conventional two-phase mechanism”. However, these two models do not take into account any anisotropy arising from the 1D Li-ion motion within LiFePO4, which is powerless to describe the lithiation/delithiation process in LiFePO4. Otherwise, the “mosaic model” is still not supported by direct and convincing experimental evidence.

Laffont's group51 and Richardson's group60 found that the interfacial region between the LiFePO4 and FePO4 domains lie in the ac plane (see in Fig. 3). Then, Laffont's group51 updated the “core–shell” model and proposed a “new core–shell” model based on the studies of thin LixFePO4 platelets (b-axis normal to the surface) with high resolution electron energy loss spectroscopy (HREELS). Different from the previous shrinking core model, the results suggested that the interface is constituted of FePO4 and LiFePO4. There is no LixFePO4 solid solution observed with the gradient of x ranging from 0 to 1. The schematic views of the interfacial region between LiFePO4 and FePO4 phases are provided in Fig. 3. According to this new core model, Li-ion diffusion in the [010] direction is asynchronous, and the LixFePO4 particles always maintain the structure with the shell of FePO4 and core of LiFePO4, as has also been suggested by Prosini's group.61 Moreover, this new core–shell model is unambiguously supported by Lemos's group62 who observed the existence of both FePO4 and LiFePO4 phases at the interface of Li0.11FePO4.


image file: c4ra10899j-f3.tif
Fig. 3 (a) TEM image of a Li0.5FePO4 crystal, showing the domains of LiFePO4 and FePO4 aligned along the c-axis.60 (Copyright 2006, The Electrochemical Society.) (b) TEM and (d) HRTEM images of the delithiated Li0.45FePO4 sample. The phase determination present in the edge, core, or interface of the particle, respectively. (c) Sketch of the Li0.45FePO4 particle. (e–g) Schematic views of the interfacial region between LiFePO4 and FePO4 phases.51 (Copyright © 2006, American Chemical Society.)

However, Delmas' group54,55 found that the interface consisted of a single-domain of LixFePO4 rather than LiFePO4 and FePO4. Based on the results from X-ray diffraction (XRD) and high resolution transmission electron microscope (HRTEM), a “domino-cascade” model was successfully established to explain the phase transformation process of LiFePO4 nanoparticles. The results illustrated that the growth of FePO4 phase at the expense of the LiFePO4 phase is considerably faster than the nucleation of a new domain. According to the “domino-cascade” model, lithiation/delithiation occurred completely and rapidly in some of the LiFePO4 nanoparticles with charging/discharging. Thus, the partially intercalated/de-intercalated LiFePO4 particle generally supported the coexistence of LiFePO4 and FePO4 particles that are single-domain.

In 2013, Zhou's group63 studied the phase transition in LixFePO4 by electrochemical impedance spectroscopy (EIS) and firstly proposed a hybrid phase-transition model combining the core–shell model and domino-cascade model. Based on this model, the delithiation of LixFePO4 starts with the domino-cascade model confirmed by a small angle of 30° for the linear Warburg region of EIS, and then turns into a core–shell model approved by a traditional angle of 45°. This hybrid phase-transition model gives attributions to the strong anisotropy in the LixFePO4 particles and could be potentially extended to some other two-phase active electrode materials.

Recently, a “many-particle model” has also been suggested. Gaberscek's group58 found that Li-ions were inserted/extracted from LiFePO4 particles one by one forming a two-phase system. Meanwhile, Zaghib's group64 found that both LiFePO4 and FePO4 phases exist in the surface layer by using electron microscopy and Raman spectroscopy. This is intended to explain the fact that the lattice coherence length is the same in the process of lithiation/delithiation, while it invalidates both core–shell model and the domino-cascade model. Each particle would be single-phased, either FePO4 or LiFePO4, which explains how one particle chooses to be in one phase while the other one remains in the other phase without violating the causality principle. Subsequently, Orikasa's group65 observed a transient phase change in the two phase reaction during the lithiation/delithiation process by applying the time-resolved X-ray measurement. It is found that the nonequilibrium phase state during the voltage plateau gradually changes and finally reaches the thermodynamically stable state by the analysis of the X-ray absorption near the edge structure (XANES).

Based on the observation and demonstration by Delmas' group54 and Ceder's group,45 Chueh's group66 also observed the overwhelming majority of Li0.5FePO4 particles (i.e., LiFePO4 electrode in a 50% state-of-charge) were either almost completely delithiated LiFePO4 particles or lithiated FePO4 particles in 2013. With the help of scanning transmission X-ray microscopy (STXM), it is clearly found that the delithiation did not appear faster in smaller particles than the larger ones resulting into weak correlation between the particle size and delithiation sequence. Therefore, the mosaic (particle-by-particle) lithiation/delithiation pathway indicated that the rate performance was limited by the rate of phase-transformation initiation rather than the phase-boundary velocity. Moreover, the model can be used to predict the equilibrium between LiFePO4 and FePO4 phases, inherent hysteretic behaviour, sequential charging/discharging mechanism and two-phase disappearance behaviour. Porous LiFePO4 electrodes were also investigated by Bai's group67 using a mathematical phase-field method. It is found that the population dynamics of active LiFePO4 nanoparticles showed non-monotonic transient currents always misinterpreted as the nucleation and growth mechanism by the Kolmogorov–Johnson–Mehl–Avrami (KJMA) theory. The LiFePO4 nanoparticles were not simultaneously transformed. The results decoupled the roles of nucleation and surface reaction, which are always considered to be affected by a special activation rate and the mean active particle-filling speed. Ichitsubo's group68 also applied the phase-field computer simulations to investigate the coherent elastic-strain energy that played a crucial role in the kinetics of phase separation during the lithiation/delithiation process. However, it is found that the nucleation of the new phase is fundamentally unlikely formed in terms of the elastic strain energy, except in the vicinity of the particle surface. The simulation results illustrated that the solid-solution reaction easily occurred by reducing the particle size, but the phase separation by nucleation is quite difficult to carry out. Meanwhile, Dargaville' group69 employed the 2D Cahn–Hilliard-reaction (CHR) equation to examine the intercalation process and suggested that the phase separation only occurred at very low discharge rates.

All these aforementioned models are expected to understand the lithiation/delithiation process in LiFePO4 and provide directions to improve the rate capability of LiFePO4 for high power EES applications. However, these models invalidate each other and are still under debate. Nevertheless, it is still unclear whether the staging phenomenon refers to a thermodynamically metastable or stable situation in LixFePO4 nanoparticles. Theoretical and experimental clarifications should be further undertaken to clarify the formation mechanism of the lithium staging. Thus, as mentioned above, a single unified model is urgently required to depict the natural lithiation/delithiation process in LixFePO4, and a computational model of the lithiation/delithiation process in LiFePO4 is essential to clarify the above question. When compared with conventional generalized gradient approximation (GGA)70,71 and local-density approximation (LDA) methods,72,73 GGA + U method can avoid large errors especially for transition metals with strong localized or f-orbitals metals and can effectively improve the results.74–76 In the future, we anticipate that the lithiation/delithiation process in LiFePO4 will be simulated by a computational model using the GGA + U method. This natural model can not only offer answers to experimental results obtained at moderate or high rates but also provide the direction to prepare LiFePO4 for high power LIBs.

2.2.2 Quasi-single-phase transformation between LiFePO4 and FePO4. Different from two-phase reaction mechanism models, the quasi-single-phase transformation between LiFePO4 and FePO4 has also attracted considerable attention. It is doubtful that an intermediate phase exists during the lithiation/delithiation process of LixFePO4. In 2011, Li's group77 directly observed Li-ions in LiFePO4 at an atomic resolution by an aberration-corrected annular-bright-field scanning transmission electron microscopy (ABF-STEM) technique (Fig. 4a), which is capable of resolving Li-ions directly. It was found that the remaining Li-ions in partially delithiated LiFePO4 preferably occupy every second layer along the [010] direction (i.e., b axis) (Fig. 4b and c). Obviously, this finding challenged the previously proposed LiFePO4/FePO4 two-phase reaction mechanisms. In 2011, Ceder's group78 demonstrated that LixFePO4 may transform through a single-phase path instead of a two-phase progress. As shown in Fig. 4d, the calculated single-particle voltage hysteresis is no more than 30 mV, where 0.05 < x < 0.9 in LixFePO4. It is clearly shown that the LixFePO4 solid solution formed and avoided phase separation in the charging and discharging process. The solid-solution behavior in the LixFePO4 system is also observed by Richardson's group via X-ray powder diffraction (XRD).79 In the Li0.6FePO4 sample, there was an intermediate of a line phase with LixFePO4 (x = 0.60 ± 0.04) composition during the transformation from two-phase mixtures to single phase. In 2011, Bazant's group80 proposed a novel electrochemical phase-field model. Based on the model, nucleation or spinodal decomposition easily leads to moving phase boundaries at small currents, whereas the particles fill homogeneously and the spinodal disappears above a critical current density, as shown in Fig. 5. This model can effectively explain the long cycle life and superior rate capability of LiFePO4 nanoparticles.
image file: c4ra10899j-f4.tif
Fig. 4 (a) Schematic of the annular-bright-field (ABF) imaging geometry. A demonstration of lithium sites within a LiFePO4 crystal is shown in (b) with the corresponding line profile acquired in the box region shown in (c) to confirm the lithium contrast with respect to oxygen.77 (Copyright © 2011, American Chemical Society.) (d) The single-particle voltage within 0.5 < x < 0.9 in LixFePO4.78 (Copyright © 2011, Rights Managed by Nature Publishing Group.) (e) Illustration of a solid solution LixFePO4 upon heating the compound to 485 K. On cooling, the parent LiFePO4 and FePO4 phases recrystallize, and nucleation and growth take place everywhere within the crystal.48 (Copyright © 2006, American Chemical Society.)

image file: c4ra10899j-f5.tif
Fig. 5 Schematic model of a LixFePO4 nanoparticle at low overpotential. (a) Lithium ions are inserted into the particle (blue arrows) from the active (010) facet with fast diffusion and no phase separation in the depth (y) direction, forming a phase boundary of thickness λ between full and empty channels. (b) Numerical simulation of phase transformation triggered by wetting of the particle boundary.80 (Copyright © 2011, American Chemical Society.)

Recently, the single phase of a solid solution LixFePO4 has been found at high temperature above 200 °C.81 As shown in Fig. 4e, Nazar's group48 found that single LixFePO4 solid solution phase formed at an elevated temperature of 212 °C and the lithiation/delithiation process were strongly enhanced. Owing to the single phase, the electrochemical performance is improved by a strong coupled correlation between electron and lithium delocalization events, whereas the power characteristics could be diminished in a two-phase mixture on account of the low mobility of the phase boundary. All these results are in good agreement with the report of Masquelier's group,82,83 who found that the single phase of LixFePO4 becomes multiphase as the temperature decreases and eventually transforms into the phases of LiFePO4 and FePO4 on aging at room temperature.

Excitedly, Masquelier's group84,85 provided in situ experimental evidence that LixFePO4 can be described as a single phase from a well-established two-phase lithiation process by modifying the particle size and cation ordering. It is can be found that the single intermediate phases of Li0.5FePO4 and Li0.75FePO4 might be stabilized at room temperature. Simultaneously, the LixFePO4 solid solution may be stabilized below the critical size of 45 nm, and completely achievable below about 15 nm at room temperature.86 It is worth mentioning that the lithiation/delithiation mechanism is still controversial and has attracted more and more attention. A series of conditions including particle morphologies, synthesis conditions, charge/discharge rates, especially state particles sizes, are still being debated.87,88

3 Electrochemical property modification of LiFePO4

LiFePO4 delivers high capacities of more than 150 mA h g−1 at slow charging/discharging rates (e.g., 0.1 C). However, high-rated charge/discharge properties and low temperature performance are severely limited due to the poor electronic conductivity and the low Li-ion diffusion efficiency. Namely, electronic conductivity and ionic diffusion rate are two of the most important characteristics that need to be improved urgently. It has been demonstrated to enhance the Li-ion diffusion by controlling the crystal growth orientation (along the ac plane) and reducing the particle size of LiFePO4 (e.g., nano-sized89,90 and 3D porous architectures91,92). It is also important to ensure an excellent electronic conductivity by coating conductive materials (e.g., amorphous carbon,93,94 graphene,95 carbon nanotubes (CNTs),96 silver,97 NiP alloy,98 metallic Fe2P,99,100 Li3PO4,101 etc.). Both ionic diffusion and electronic conductivity can be employed to improve the kinetic properties of LiFePO4 by doping in MLi and MFe sites using the cations of metallic elements (e.g., Cu+, Mg2+, Mn2+, Al3+, Ti4+, Zr4+, Nb5+, W6+, etc.).62,102,103 Comprehensive improvements of all the effective strategies are desired and are also the targets for high-rated charge/discharge properties and low temperature performance, which are two important parameters for future research and EES applications (especially for EV/HEV). Moreover, the volume energy density of LiFePO4 is another important parameter, which is critical towards maximizing the space utilization of LIBs. In the following section, we will introduce representative reports on how to improve the performances of LiFePO4. The reports are categorized into the improvement of (a) rate capability, (b) low temperature performance and (c) tap density.

3.1 Morphology control

3.1.1 Crystal growth orientation along the ac plane. The Li-ion diffusion rates vary along different directions in LiFePO4 lattices. If we can deliberately control the synthesis of LiFePO4 crystals with certain exposed facets for faster Li-ion intercalation, the performances, especially rate capabilities, can be significantly improved. Previous reports showed that Li-ion diffused the fastest along the [010] direction in the orthorhombic LiFePO4 lattice.40 Islam's group104 revealed that there were mainly three crystal styles agreeing with the results from Franger's group,105 Nazar's group106 and Richardson's group,60 as shown in Fig. 6a–d. The crystal growth of anisotropic cathode LiFePO4 is addressed as a key factor controlling rapid Li-ion diffusion. Several surface properties of olivine-structure LiFePO4 were investigated via first principles calculations within the GGA + U framework by Ceder's group.107 The calculated Li redox potential for the (010) surface was 2.95 V, which is significantly lower than the bulk value of 3.55 V. The study revealed that it is important to control the morphology of LiFePO4 crystal growth orientation and expose more surface of (010) with high Li+ diffusion rate.108 Thus, if LiFePO4 crystals are fabricated with more (010) facets, such as thin nanoplates with (010) surfaces, superior performances can be expected due to the fast Li+ insertion and extraction. Richardson's group60 reported the hydrothermal synthesis of LiFePO4 microcrystals with 85% (010) surface areas, but no electrochemical testing was performed. For further improvements, like the synthesis of smaller and thinner nanoplates with more (010) surfaces, coating with a conductive carbon layer may be considered to fully explore the potential of those nanoplates.109
image file: c4ra10899j-f6.tif
Fig. 6 (a) Theoretical crystal morphologies of LiFePO4 with equilibrium morphology from relaxed surface energies. Schematic diagrams of the observed crystal morphologies of LiFePO4 produced under different experimental conditions: (b) block-shape, (c) rectangular prism and (d) hexagonal platelet.104 (Copyright © 2008, Royal Society of Chemistry.) SEM images, HRTEM images (inset) and the corresponding SAED patterns (inset) for (e) rectangular prism nanoplates110 (Copyright © 2011, The Electrochemical Society) and (f) hexagonal platelet.112 (Copyright © 2010, Royal Society of Chemistry.) (g) SEM (the inset is an enlarged SEM image) and (h) TEM morphologies of nanoplates.113 (Copyright © 2008, Royal Society of Chemistry.)

In 2011, a combination of LiFePO4 rectangular nanoplates was prepared by a low temperature solvothermal synthesis (STS) route reported by Kim's group,110 as shown in Fig. 6e. It can be seen that the growth of the LiFePO4 nanoplates is identified to be along the [100] and [010] crystallographic directions, resulting into a large surface area of (001) rather than (010). Thus, the as-prepared LiFePO4 nanoplates yielded the first discharge capacity of only 131 mA h g−1 at 0.25 C (Fig. 7), on account of the relatively low (101) surface area. In particular, the as-prepared LiFePO4 nanoplates exhibited a capacity decline of about 38 mA h g−1 until 8 C. Wang's group and Wu's group111 found that the obtained LiFePO4/C composite exhibited high rate capability using carbon coating and yielded 104 mA h g−1 and 95 mA h g−1 at 8 C and 12 C, respectively.


image file: c4ra10899j-f7.tif
Fig. 7 High rate capacity performance of LiFePO4 particles with different observed crystal morphologies.

Fig. 6f shows LiFePO4 hexagonal nanoplates (100 nm thick and 800 nm wide) prepared by a solvothermal method in a H2O-PEG binary solvent.112 When compared with LiFePO4 hexagonal microplates (300 nm thick and 3 mm wide), carbon coated LiFePO4 nanoplates exhibited the calculated Li-ion diffusion coefficient of 4.2 × 10−9 cm2 S−1 instead of 2.2 × 10−9 cm2 S−1, and delivered a discharge capacity of more than 155 mA h g−1 instead of 110 mA h g−1 at 0.1 C. In particular, LiFePO4/C nanoplates could perform a high discharge capacity of 87 mA h g−1 at 60 C (i.e., fully discharged within 30 s), when the content of conductive carbon was increased to 30 wt%.

Recently, Balaya and co-workers113–115 employed a solvothermal method to control the crystal growth orientation along the ac plane and reduce the b-axis to the smallest possible thickness. As shown in Fig. 6g and h, the thickness along the b-axis of LiFePO4 nanoplates is found to be 30–40 nm, while the thickness along the ac plane is in the range of 500–800 nm. The selected area electron diffraction (SAED) pattern viewed along [010] (i.e., b-axis) reveals that the plate is a single crystal with the b-axis along the thinnest direction. It is indicated that the smallest dimension of the nanoplates is the b-axis, which facilitates the diffusion of Li-ions.112 The as-obtained LiFePO4/C nanoplates could store Li-ions comparable to its theoretical capacity of 87 mA h g−1 at 0.1 C. Even at 30 C, they could deliver a discharge capacity of up to 87 mA h g−1. All the results revealed that the excellent high rate performance of LiFePO4 nanoplates is ascertained to the relatively thin thickness along the b-axis and large (101) surface area. In other words, size reduction (similar to 30 nm) at the b-axis and crystal growth orientation along the ac plane could provide fast Li-ion diffusion. Moreover, uniform carbon coating (similar to 5 nm) throughout the LiFePO4 nanoplates favors an electronically conducting path for electrons.

3.1.2 Nanosized LiFePO4 particles. To date, the most widely adopted strategy to improve the rate capability is to use LiFePO4 structures with smaller sizes and coat the structures with carbon to improve the electronic conductivity.108 In particular, nano-structured LiFePO4 particles are in favor of reducing the 1D diffusion length and advantageous for Li-ions transport, as the intrinsic diffusion constant is scale dependent and significantly reduced for large particle sizes.45 Various carbon-modified LiFePO4 nanostructures have been studied, such as nanoparticles, nanorods, nanoflowers, porous microspheres, nanoplates, nanowires/fibers, template mesoporous materials, etc.116 Smaller structures mean shorter diffusion paths for Li-ion, and hence, better utilization of the active materials.117 For example, assuming that for a certain time, Li-ions can diffuse 50 nm within LiFePO4 lattices, if the electrode is made of a single crystal LiFePO4 film with a thickness of 1 mm, then only the 50 nm on the surface can be utilized. However, if the electrodes are made of nanoparticles with less than 100 nm diameters and proper electrolyte wrapping and electron conduction is ensured, the entire nanoparticle, and hence the electrode materials, can be utilized.118 This is why nanosized materials are favored to obtain high specific capacity and high rate capability.119,120 As discussed above, at a high charge/discharge rate, less time is allowed for ion diffusion. Consequently, nanostructures possess a better capability for complete “reactions” with Li-ion within a short time. These nano-sized LiFePO4 materials could provide short diffusion length, yield better rate performances than bulk materials because of smaller diffusion lengths, and be beneficial for Li-ion batteries, as shown in Fig. 8a.
image file: c4ra10899j-f8.tif
Fig. 8 (a) Schematic illustration for active LiFePO4 at different interfaces of electrolyte and various LiFePO4 particles.138 (b) TEM image of LiFePO4/C nanopaticles.122 (Copyright © 2008 WILEY-VCH Verlag GmbH %26 Co. KGaA, Weinheim.) (c) SEM and TEM (inset) images of hollow LiFePO4.123 (Copyright © 2008, American Chemical Society.) (d) SEM and TEM (inset) images of porous and coarse LiFePO4/C composite.136 (Copyright © 2013, American Chemical Society.) (e) SEM image of an individual LiFePO4/C microsphere consisting of nanoplates.92 (Copyright © 2011, American Chemical Society.) (f) FIB images allowing 3D visualization information obtained from the 3D porous LiFePO4/C microsphere. (g) SEM image of A area (indicated by a rectangle in panel f). (g) SEM image of B area (indicated by a rectangle in panel f). (i) Scheme showing the structure of LiFePO4/C nanocomposites in porous microspheres.138

The ball milling method is usually employed to synthesize nano-sized LiFePO4/C composites. It is found that LiFePO4/C from ball milling in acetone displayed a spherical shape with a size of 60 nm, similar to the size of LiFePO4/C observed from dry ball milling (DBM) but with a more irregular morphology.121 Although the LiFePO4/C nanocomposites prepared from wet ball milling (WBM) in acetone and DBM exhibited the similar discharge capacities of 153 mA h g−1 and 120 mA h g−1 at rates of 0.1 C and 10 C, respectively (Fig. 9a), the previous sample yielded much lower polarization resulting in high energy density. A core–shell structured LiFePO4/C nanocomposite was also successfully designed and prepared by Zhou's group.122 An in situ polymerization restriction method was firstly used for the synthesis of FePO4/PANI. Then, a highly crystalline LiFePO4 core was formed with a size of about 20–40 nm, and the polymer shell was transformed into a semi-graphitic carbon shell with a thickness of about 1–2 nm during the heat treatment process at 700 °C, as shown in Fig. 8b. The as-obtained LiFePO4/C yielded excellent rate performance. Even at the high rate of 60 C, it still delivered a capacity of 90 mA h g−1. Additionally, Cho's group123 reported that LiFePO4 nanowires were synthesized using the two-dimensional hexagonal SBA-15 silica as the hard template. The LiFePO4 nanowires delivered excellent rate performance. Even at 10 C and 15 C, LiFePO4 nanowires yielded 144 mA h g−1 and 137 mA h g−1, respectively, corresponding to 93% and 89% retention of the initial capacity.


image file: c4ra10899j-f9.tif
Fig. 9 High rate capacity performance of LiFePO4 particles with (a) nano-sized and (b) 3D porous morphologies.

Taniguchi's group124 employed a combination technique of spray pyrolysis (SP) with WBM to prepare LiFePO4/C nanoparticles. It is found that the LiFePO4/C nanoparticles possessed a geometric mean diameter of 146 nm. A thin carbon layer endowed LiFePO4 with excellent electrochemical performance, because LiFePO4 nanoparticles exhibited an extremely high stability with the carbon coating.125 LiFePO4/C yielded 118 mA h g−1 and 105 mA h g−1 at 10 C and 20 C, respectively. Even at 60 C, LiFePO4/C yielded 75 mA h g−1 without capacity fading after 100 cycles. Amazingly, Ceder's group126 found LiFe1−2yP1−yO4−δ/C (y = 0.05) nanoparticles prepared by WBM in acetone exhibited outstanding high rate performance. The LiFe0.9P0.95O4-δ/C off-stoichiometry nanoparticles showed a specific discharge capacity of 163 mA h g−1 and 152 mA h g−1 at 10 C and 20 C, respectively. Significantly, the specific capacity still remains above 120 mA h g−1 at a discharge rate as high as 197 C, when changing the working electrodes fabrication of the active material, carbon black and binder in a weight ratio from 80[thin space (1/6-em)]:[thin space (1/6-em)]15[thin space (1/6-em)]:[thin space (1/6-em)]5 to 30[thin space (1/6-em)]:[thin space (1/6-em)]65[thin space (1/6-em)]:[thin space (1/6-em)]5. This result is a remarkable progress over the previous studies. This presents a new direction on the improvement of LFP cathode performances, although more work is needed to further improve the repeatability and mechanism understanding of this method.

3.1.3 Three-dimensional porous architectures with micro-nano structures. 3D porous LiFePO4 materials not only enhance the surface-to-volume ratio and reduce the Li-ion transport length but also offer a potential solution to reduce the interfacial energy and agglutinant in electrode, all of which are much better than those measured for the nano-sized LiFePO4 and hollow micro-spherical LiFePO4 (Fig. 8c), improve the electrochemical performance and practical applications of LiFePO4.123,127 In particular, LiFePO4 cathodes with 3D porous architectures, spherical architectures, and micro-nano-structures have been extensively and intensively studied for high powerful LIBs. With the aid of template technology, a variety of 3D porous LiFePO4 materials have been synthesized.128–130 In 2013, Zhang's group131 successfully prepared 3D mesoporous LiFePO4 using Baker's yeast cells both as a structural template and a biocarbon source. The as-obtained LiFePO4 exhibited a high discharge capacity of about 153 mA h g−1 at 0.1 C.

Besides the soft template, Li3PO4/graphene oxide (GO) microspheres are also employed as sacrificial templates to synthesize porous LiFePO4/graphene microspheres.132 The as-obtained LiFePO4/graphene microspheres (2 μm) were assembled by nanoparticles and wrapped by graphene nanosheets. The graphene has been proven to be advantageous for EES applications on account of its superior electrical conductivities and high surface area. Moreover, the porous microspherical structure exhibited an effective way to achieve high power density for LiFePO4. The as-obtained LiFePO4/graphene yielded excellent rate performance, i.e., 101.8 mA h g−1 at 10 C, which holds 72% of the initial capacity at 0.1 C, as shown in Fig. 9b.

Without employing templates, hydrothermal synthesis (HTS) as a low-temperature liquid phase thermal (LPT) synthesis has also been employed to prepare 3D porous LiFePO4 microspheres. In the HTS process, the LiFePO4 precursor nanoparticles deposit on account of their small solubility in the solution. Then, the nano-sized LiFePO4 precursor particles self-assemble to form densely packed microspheres necessary for reducing the surface tension of the dispersed particles. Subsequently, with a dissolution-deposition process, the agglomerated LiFePO4 precursors evolve into 3D porous microspheres.133 The HTS yields several advantages such as simplicity, morphology control, homogeneous particle size distribution, and low cost.134,135 In 2010, Cao's group133 first prepared 3D porous LiFePO4 microspheres, which consisted of nanoparticles, by the hydrothermal process. These LiFePO4/C microspheres exhibited high Brunauer–Emmett–Teller (BET) surface areas of 38.6 m2 g−1 and delivered excellent high rate capability of 115 mA h g−1 at 10 C. Even at 20 C and 30 C, they yielded 93 mA h g−1 and 71 mA h g−1, respectively. In 2011, Goodenough's group92 successfully accomplished the self-assembly of LiFePO4 nanoplates for the porous microspheres using the STS route, as shown in Fig. 8e. It can be seen that the 3D porous LiFePO4/C microspheres (1–3 μm) consist of nanoplates (80 nm), which interweave together forming a flowerlike structure giving a high BET surface area of 32.9 m2 g−1. The as-obtained flowerlike LiFePO4/C microspheres exhibited a discharge capacity of 72 mA h g−1 at 10 C. Due to the conductive polymer PPy processing macromolecule sp2-type carbon, LiFePO4/(C + PPy) yielded excellent rate performance, e.g., high discharge capacity of 92 mA h g−1 at 10 C. Additionally, Eom's group and Kwon's group136 provided a growth technology to synthesize porous and coarse LiFePO4/C composites (5–10 μm) using LiFePO4 nanoparticles (100–200 nm) as seed crystals for the 2nd crystallization process, as shown in Fig. 8d. The SEM and TEM images of LiFePO4/C composites are shown in Fig. 8d. The as-obtained 3D porous LiFePO4/C composites delivered a discharge capacity of 100 mA h g−1 at 10 C, which is 65% of the discharge capacity at 0.1 C. All the excellent rate performances strongly fulfil the requirements of rechargeable lithium batteries for high power applications.

It is worth noting that a versatile spray drying methodology is usually employed to prepare 3D porous spheres with micro-nano-superstructures.137 Zhang's group138 reported a novel and simple template-free concept and the synthesis of 3D porous LiFePO4/C microspheres via a sol–gel-spray drying (sol–gel-SD) method. As shown in Fig. 8f–i, it is clearly seen that the as-obtained LiFePO4/C had a large specific surface area of 20.2 m2 g−1, an average nano-size of 32 nm and a main pore diameter of 45 nm. The as-obtained 3D porous LiFePO4/C microspheres easily contact with the electrolyte, which facilitate the Li-ion diffusion, gave a high Coulombic efficiency of 97.2%, and presented an excellent capacity retention rate close to 100% after 50 cycles. However, it only provided 100 mA h g−1 at the high rate of 10 C, due to lower carbon coating and low porosity.139 When using citric acid as the carbon source, 3D porous LiFePO4/C microspheres have been successfully synthesized.140,141 Polyvinyl alcohol (PVA) gel has been also chosen as the carbon source for its excellent film-forming properties in the spray drying process.142 The carbon layers transformed from the carbon sources were effectively coated around LiFePO4 and showed excellent cyclability and superior rate capability.

3.2 Low temperature performance

LiFePO4 cathode with an excellent power output makes it a superior candidate for HEV and EV applications. However, one key challenge is the poor low temperature performances due to LiFePO4 itself and the electrolyte in LiFePO4-based batteries, as compared to the performances at room temperature. In 2013, Shin's group143 found that pristine LiFePO4 exhibited a higher cycling stability at a lower operating temperature of −20 °C than room temperature. It is suggested that LiFePO4 experiences much milder degradation of the surface and the bulk at low temperatures. To date, LiFePO4 itself and the electrolyte have been both responsible for the poor low-temperature performance, which is the key challenge for wider applications of LiFePO4 in EES devices.144

Generally, carbon coating is used to improve the electron conductivity of the electrodes. The processes are associated with electron conductions.27,145 Good electronic conductivity is of crucial importance to facilitate electron conduction and achieve high performances.146,147 Carbon is a common low-cost material with good conductivity and is widely used for conductivity improvement. Carbon layers can be facilely coated on the nanostructure surfaces by many low-cost and scalable methods, such as organic chemical polymer coating and annealing.148 Carbon coating is quite successful in improving the cathode conductivity, but the weight ratio of carbon has to be carefully controlled.149,150 To be extreme, a cathode made of pure carbon will simply turn the battery into a double-layer supercapacitor, which has very low energy density and unstable output voltage. Higher carbon content also results in lower volumetric energy density due to the smaller mass density of carbon. Apart from the enhancement of conductivity, carbon also serves as a capping material that can effectively reduce the particle growth during crystallization processes. Simultaneously, the reducing atmosphere of carbon also prevents the oxidation of Fe(II) during high temperature annealing.

In 2011, Tu's group151 successfully prepared carbon-coated LiFePO4 materials using polystyrene (PS) spheres (50–300 nm) as the carbon source. The results illustrated that LiFePO4/C with 3.0 wt% carbon content exhibited excellent electrochemical capability at a low temperature of −20 °C, which yielded 147 mA h g−1 and 79.3 mA h g−1 at 0.1 C and 1 C, respectively (Fig. 10). Even after 100 cycles at 1 C, the LiFePO4/C still exhibited almost 100% capacity retention. It can be attributed to the optimal carbon coating thickness (2.5 nm) and good carbon coating morphology. Using citric acid as a carbon source, Wang's group152 reported a kind of 3D LiFePO4/C microspheres. The electronic conductivity of LiFePO4/C microspheres is between 7.5 × 10−2 and 10−1 S cm−1 from −40 °C to 23 °C, which yielded attractive low temperature discharge capacity of 110 mA h g−1 and 64 mA h g−1 at −20 °C and −40 °C. Alternatively, Scrosati's group and Sun's group153 provided double-carbon-coated LiFePO4/C porous microspheres, with an excellent specific discharge capacity of 70 mA h g−1 (1 C) at −20 °C. Additionally, polyacene (PAS) was also employed to optimize LiFePO4.154 The as-obtained LiFePO4/PAS microspheres yielded outstanding discharge capacity of 88 mA h g−1 at 1 C at the low temperature of −20 °C. This favorable electrochemical performance is attributed to the homogeneous morphology, the small particles inside, the porous surface, and the conductive PAS (8.5 wt%).


image file: c4ra10899j-f10.tif
Fig. 10 Rate capacity performance of LiFePO4 at various temperatures.

Besides coating conductive carbon, metal ion doping is another common method to improve the conductivity of LiFePO4. Various metal dopants have been studied, such as Mg2+, Al3+, Cr3+, Ti4+, Zr4+, Nb5+, W6+, etc. A metal dopant will replace the Li+ (M1 doping) or Fe2+ (M2 doping) in LiFePO4 lattice to generate an n-type semiconductor.37 The conductivity can be enhanced by over seven orders of magnitude after doping. It was believed that the dopants that replaced metal ions in LiFePO4 lattice could enhance the electronic and ionic conductivities.24 In 2011, Liao's group155 optimized LiFePO4 by slight Mn-substitution. The as-obtained LiFe0.98Mn0.02PO4/C delivered 99.8 mA h g−1 (1 C) at −20 °C, which is superior to 90.7 mA h g−1 for LiFePO4/C. Even at the low temperature of −40 °C, LiFe0.98Mn0.02PO4/C still gave 70.5 mA h g−1 at 1 C.

Recently, the mixtures of LiFePO4 and Li3V2(PO4)3 (i.e., xLiFePO4 + yLi3V2(PO4)3, x > 0, y > 0) exhibited excellent electrochemical performance and have recently attracted much attention.156,157 Chen's group144 found that LiFePO4/C delivered a discharge capacity of only 45.4 mA h g−1 (0.3 C) at −20 °C, being 31.5% of the capacity obtained at 23 °C. When compared with LiFePO4/C, the Li3V2(PO4)3/C could give a discharge capacity of 108.1 mA h g−1 (0.3 C) at −20 °C, which is 86.7% of the capacity at 23 °C. It is found that the activation energies of LiFePO4 and Li3V2(PO4)3 are calculated to be 47.48 kJ mol−1 and 6.57 kJ mol−1, respectively. The Li-ion transformation in Li3V2(PO4)3 is much easier than that in LiFePO4. Recently, many mixtures of LiFePO4 and Li3V2(PO4)3 have been studied and given excellent rate performances, such as 100 mA h g−1 at 10 C for 2LiFePO4/C + Li3V2(PO4)3/C,158 116.8 mA h g−1 at 10 C for 3LiFePO4/C + Li3V2(PO4)3/C,159 114 mA h g−1 at 10 C for 5LiFePO4 + Li3V2(PO4)3,160 and 93.3 mA h g−1 at 10 C for 9LiFePO4/C + Li3V2(PO4)3/C.161 However, up to now, the low temperature performance has still been rarely reported. It is worthwhile to examine the influence of the combination of LiFePO4 and Li3V2(PO4)3 on the performance of Li-ion battery, and the low temperature of LiFexV2/3-2x/3PO4/C (0 < x < 1) solid state materials.

An electrolyte is another main reason in addition to LiFePO4 itself. The main reason might be the poor performances of conventional LiPF6/EC + DMC + EMC electrolyte system at low temperatures. The electrolyte becomes more viscous at low temperatures, resulting in a low diffusion coefficient of lithium ions in the electrolyte. Performance degradation can be expected since fast Li-ion diffusion (i.e., shorter intercalation time and better utilization of the active material) is of crucial importance for high performance cathodes. The inherent properties of olivine LiFePO4 might also contribute to the degradations, but it is less likely to be the main reason. In the near future, more efforts are needed to improve the low temperature performances. Huang's group35 reported that LiPF6 salt in EC (the high freezing point of 36.4 °C) easily leads to poor Li-ion diffusion at low temperatures, particularly below −20 °C. Thus, the exploration of novel electrolyte systems with superior low temperature properties might be a promising research direction. The viscosity increase of the electrolyte at low temperatures may be the reason.27 Ma's group162 discussed the 1.0 M LiPF6/EC + DMC + DEC + EMC (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]3, v/v) electrolyte (Table 1). LiFePO4/C composite could yield 90 mA h g−1 and 69 mA h g−1 at −20 °C and −40 °C, respectively. Moreover, Zhang's group163 found that the LiFePO4/C composite could operate over a wide temperature range (−50 to 80 °C) using new lithium salts LiBF4–LiBOB in a solvent mixture of PC + EC + EMC (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]3). Although the LiBOB-based electrolyte has high conductivity above −10 °C and high temperature performance even up to 90 °C, but it fails to perform below −40 °C. Fortunately, the mixture of LiBF4 and LiBOB helps the electrolyte process a large working temperature range from −50 °C to 90 °C due to the high conductivity of LiBF4 below −10 °C. With the help of LiBF4–LiBOB-based electrolytes, LiFePO4 delivered 62 mA h g−1 and 43 mA h g−1 even at −40 °C and −50 °C.

Table 1 Various LiFePO4 and the corresponding electrolyte responsible for the excellent low-temperature performance
Materials Synthesis method Low temperature range/°C Electrolyte Ref.
LiFePO4/C Solid state reaction method −20 1 M LiPF6/EC + DEC (1[thin space (1/6-em)]:[thin space (1/6-em)]1) 144
LiFePO4/C Solid state reaction method −20 1 M LiPF6/EC + DEC (1[thin space (1/6-em)]:[thin space (1/6-em)]1) 151
LiFePO4/C Solid state reaction method −40 1 M LiPF6/EC + DEC (1[thin space (1/6-em)]:[thin space (1/6-em)]1) 152
LiFePO4/C Co-precipitation method −20 1 M LiPF6/EC + DEC (1[thin space (1/6-em)]:[thin space (1/6-em)]1) 153
LiFePO4/PAS Co-precipitation method −20 1 M LiPF6/EC + DEC (1[thin space (1/6-em)]:[thin space (1/6-em)]1) 154
LiFe0.98Mn0.02PO4/C Solid state reaction method −40 1 M LiPF6/EC + DEC (1[thin space (1/6-em)]:[thin space (1/6-em)]1) 155
LiFePO4/C Solid state reaction method −40 1 M LiPF6/EC + DEC + DMC + EMC (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]3) 162
LiFePO4/C Solid state reaction method −50 1 M (0.9LiBF4 + 0.1LiBOB)/PC + EC + EMC (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]3) 163


3.3 Volumetric energy density

The high volumetric energy density of batteries can make the batteries smaller in volume while exhibiting the same performances.35,149 When compared with the mass density (5.1 g cm−3) of the dominant LiCoO2 cathode materials, the theoretical density of LiFePO4 is only 3.6 g cm−3.164 However, the tap density of LiFePO4 usually lies in the range of 0.8–1.5 g cm−3, which seriously limits the volumetric energy density and becomes one main obstacle preventing LiFePO4 to be acceptable in commercial applications.35,165 Optimizing the particle morphology and distribution has proven to be helpful for superior performance. In fact, the spherical architecture of LiFePO4 is preferred to optimize the tap density. When compared with irregular particles, spherical LiFePO4 particles can decrease the vacant space between the particles and improve the fluidity of the particles. Furthermore, less binder (polyvinylidene fluoride (PVDF) or PTFE) can be stuck to spherical LiFePO4 particles in fabricating the cathode film, which can enhance the volumetric energy density. In addition, the coating carbon with a density of 2.2 g cm−3 consumes volumetric energy density while efficiently enhancing the gravimetric energy density.27 Thus, it is imperative to consider the weight ratio and optimize the balance to get a high gravimetric energy density without sacrificing the volume energy density for LIBs.

Recently, the co-precipitation method,166 molten-salt method,167 HTS method,168 and microwave assisted water-bath reaction (MW-WBR)169 have been employed to prepare LiFePO4 with high tap density, as shown in Table 2. The co-precipitation method (including the controlled crystallization method) is considered as one of the most versatile techniques to produce spherical LiFePO4 particles. Firstly, spherical amorphous FePO4·xH2O powders are prepared by the co-precipitation method. Ying's group166 adopted a reactor (Fig. 11a) to prepare microspherical amorphous FePO4·xH2O powders (Fig. 11b) with 8 μm by the controlled crystallization method following the reaction: Fe(NO3)3 + H3PO4 + 3NH3 + xH2O = FePO4·xH2O + 3NH4NO3. It was found that the pH value should be fixed on 2.1, avoiding the hydrolyzation of PO43− into HPO42− and H2PO4 as the solution's pH decreases and protecting Fe3+ from hydrolyzing into Fe(OH)2+, Fe(OH)2+ and Fe(OH)3 when the solution's pH increases. Secondly, the Li source, carbon source and some doping elements were added with FePO4·xH2O and then heat treated at high temperatures. Ying's group166 synthesized spherical Li0.97Cr0.01FePO4/C (Fig. 11c) with a tap density of 1.8 g cm−3, which is superior to 1.37 g cm−3 prepared by ball milling170,171 and similar to 1.82 g cm−3 synthesized by a secondary ball milling method.172 The Li0.97Cr0.01FePO4/C composite exhibited 163 mA h g−1 at 0.005 C, but had low discharge capacity of 142 mA h g−1 at 0.1 C and unsatisfactory rate capability of 110 mA h g−1 at 1 C (Fig. 12a) with the corresponding volumetric discharge capacity of 198 mA h cm−3 (Fig. 12b). This is due to the large amount of inert Li0.97Cr0.01FePO4 at the heart of the solid Li0.97Cr0.01FePO4 microsphere, which contacted poorly with electrolytes, as shown in Fig. 8a.138

Table 2 Various samples with high volumetric energy density
Materials Synthesis methods Tap density (g3 cm−3) Ref.
LiFePO4/Fe2P Ball milling 1.37 170 and 171
LiFePO4/C Ball milling 1.82 172
Li0.97Cr0.01FePO4/C Controlled crystallization method 1.8 166
LiFePO4/PAS Controlled crystallization method 1.6 154
LiFePO4/C Co-precipitation method 1.5–1.6 153, 173 and 175
LiFePO4/C Co-precipitation method 1.8 174
LiFePO4 Molten-salt method 1.55 167
LiFePO4/C Hydrothermal method 1.4 168
LiFePO4 Spray drying method 1.7 176
LiFePO4 Microwave assisted water-bath reaction 2.0 169



image file: c4ra10899j-f11.tif
Fig. 11 (a) Schematic diagram of the reactor for controlled crystallization process. The SEM images panorama and individual (inset) for (b) FePO4·xH2O and (c) Li0.97Cr0.01FePO4/C particles prepared by controlled crystallization method.166 (Copyright © 2005 Elsevier B.V. All rights reserved.) (d) Porous sponge like LiFePO4 and the corresponding cross-sectional SEM image (inset).175 (Copyright 2009, The Electrochemical Society.) (e) SEM images of LiFePO4/C containing 7.0 wt% carbon synthesized by coprecipitation method using self-produced high-density FePO4 pressure-filtrated at 20 MPa.174 (Copyright © 2009 Elsevier Ltd. All rights reserved.) (f) SEM images for panorama and individual (inset) quasi-spherical FePO4·2H2O precursor synthesized by the hydrothermal method.168 (Copyright © 2011, Royal Society of Chemistry.) (g) SEM micrographs (the inset is an enlarged SEM image) of LiFePO4 with a fairly high tap density of 2.0 g cm−3 obtained by microwave assisted water-bath reaction method.169 (Copyright © 2013 World Scientific Publishing Co.)

image file: c4ra10899j-f12.tif
Fig. 12 (a) Rate capacities for various samples and (b) their volumetric energy densities.

In order to simultaneously achieve a high gravimetric energy density and tap density, the conductive polymer PAS was employed to improve the rate capability of LiFePO4 using the controlled crystallization method.154 They prepared LiFePO4/PAS with a tap density of 1.6 g cm−3, which exhibited a rate capability of 129 mA h g−1 and 97 mA h g−1 at 1 C and 5 C respectively, corresponding to the volumetric discharge capacities of 206.4 mA h cm−3 and 155.2 mA h cm−3 at 1 C and 5 C, respectively. In 2010, Sun's group and co-workers173 prepared spherical LiFePO4/C using polyvinylpyrrolidone (PVP) as a carbon source with a particle size of 6 μm and high tap density of 1.6 g cm−3. The as-obtained LiFePO4/C yielded discharge capabilities of 132 mA h g−1 (211.2 mA h cm−3) and 108 mA h g−1 (172.8 mA h cm−3) at 1 C and 5 C, respectively. Chang's group174 improved the carbon content of LiFePO4/C to 7.0 wt% and synthesized LiFePO4/C with a high tap density of 1.8 g cm−3 by the co-precipitation method using self-produced high-density FePO4 pressure-filtrated at 20 MPa (Fig. 11e). Due to the denser structure, LiFePO4/C delivered volumetric discharge capability of 300.6 mA h cm−3 at 0.1 C. Even at 1 C and 5 C, the LiFePO4/C also exhibited high capacities of 257.4 mA h cm−3 and 176.9 mA h cm−3, respectively.

To further improve the high rate performance, 3D porous sponge-like LiFePO4/C particles with a high tap density of 1.5 g cm−3 were successfully prepared by the co-precipitation method using pitch as the carbon source (Fig. 11d).175 Although each sponge-like LiFePO4/C particle with a micron size of 6 μm consisted of nanoscale 200–300 nm primary LiFePO4/C particles, the sponge-like LiFePO4/C particle gave volumetric discharge capacity of 214.5 mA h cm−3 and 186.0 mA h cm−3 at 1 C and 5 C, respectively, which is 2.5 times that of nano-sized LiFePO4/C with a tap density of 0.6 g cm−3. In 2010, double carbon coating (sucrose and pitch as the carbon source) was employed to synthesize LiFePO4/C, which possessed a tap density of 1.5 g cm−3 and yielded volumetric discharge capacities of 225.0 mA h cm−3 and 193.5 mA h cm−3 at 1 C and 5 C, respectively.153 Even at 10 C and 20 C, the LiFePO4/C delivered 116 mA h g−1 (174.0 mA h cm−3) and 79 mA h g−1 (118.5 mA h cm−3), respectively.

In addition to the co-precipitation method, Zhou's group167 prepared microspherical LiFePO4 with a tap density of 1.55 g cm−3 via a molten-salt method, which showed accelerated reaction rate and controllable particle morphology. However, the LiFePO4 product delivered a low capacity of 130.3 mA h g−1 (201.5 mA h cm−3) at 0.1 C due to the large amounts of inert LiFePO4 at the heart of the solid LiFePO4 microsphere.138 In 2011, quasi-microspheres of LiFePO4/C composed of many densely compact nanoplates (with 100 nm size and 30 nm thickness) were synthesized by Zhang's group (Fig. 11f).168 Due to the FePO4·2H2O nanoplates assembled quasi-microspherical precursors via the hydrothermal process, the as-obtained LiFePO4/C quasi-microspheres possessed a tap density of 1.4 g cm−3 and showed excellent high rate performance. Even at 30 C current rate, the discharge capacities can reach 75 mA h g−1 (105 mA h cm−3). In 2014, Kim's group176 successfully prepared 3D porous LiFePO4 microspheres with a high tap density of 1.7 g cm−3 using a spray drying (SD) method. The as-obtained 3D porous LiFePO4 microspheres with the carbon content of 3.3% delivered high initial discharge capacity of 160 mA h g−1 (272 mA h cm−3) at 0.1 C, corresponding to 94% of the theoretical capacity, as shown in Fig. 12.

Interestingly, the MW-WBR method was employed to synthesize spherical LiFePO4/C with a fairly high tap density of 2.0 g cm−3, which will strongly benefit the enhancement of volumetric energy density (Fig. 11g).169 This spherical LiFePO4/C exhibited a tap density higher than 1.0 g cm−3 (Tianjin Sterlan-Energy Ltd. China), 1.4 g cm−3 (Phostech Lithium Inc., Canada), and 1.5 g cm−3 (Valence Technology Inc., US).174 Furthermore, spherical LiFePO4/C yielded excellent discharge capacities of 131 mA h g−1 (262.0 mA h cm−3) and 105 mA h g−1 (210 mA h cm−3) at 1 C and 5 C, respectively.

4 Synthetic reaction mechanism

4.1 Carbothermal reduction method

Recently, synthesis approaches and equipment have been used towards fabricating cathode material LiFePO4, such as solid phase thermal (SPT) routes and liquid phase thermal (LPT) routes.177,178 It is well known that the carbothermal reduction (CTR) method is one of the promising SPT methods. The CTR method is considered as an effective approach for improving the electrochemical performance of LiFePO4 materials.127,179 The CTR method not only provides a special environment favorable for the reduction of Fe(III) and the formation of the LiFePO4/C material but also provides a continuous conductive carbon film as an electron conductor that enhances the electronic conductivity of LiFePO4.94 Moreover, the evenly distributed carbon can also prevent particle coalescence. Recently, the CTR method has been defined as a feasible, low-cost and environment-friendly method for the fabrication of LiFePO4/C composites and has been largely employed to synthesize high performance LiFePO4/C composites.

There are mainly two steps in the CTR method. At the first step, carbon sources are evenly distributed in precursor aggregation using various routes, such as ball milling,180,181 sol–gel,182,183 co-precipitation,184,185 spray drying,138,186 etc. Carbon sources include inorganic carbon (e.g., carbon black + surface activator), organic sucrose, small molecular acid (e.g., citric acid), big molecular polymer (e.g., PEG10000)187 and sp2-type carbon (e.g., carbon nanotube (CNT) and graphene).188,189 At the second step, LiFePO4/C is formed due to vigorous gas evolution (mainly CO and CO2) during the degradation and carbonization of the carbon sources. During the heat treatment of an appropriate precursor, elementary carbon is deposited on the walls of primary nanoparticles as a degradation product. Meanwhile, the continuous conductive carbon film can overcome the electronic and lithium-diffusion limitations, improve the rate of insertion/extraction and optimize the electrochemical performance under high-current regimes.

Noticeably, the electrochemical performance of LiFePO4/C material is influenced by various carbon sources and raw materials. Iron phosphides, such as FeP and Fe2P, were reported to have effects on the improvement of the rate performance. Obviously, the impurities, especially for Fe2P, easily appeared as by-products in the CTR synthetic process. It is difficult to control the content of Fe2P, which has effects on the electrochemical performance of LiFePO4. Why Fe2P was synthesized by carbothermal reaction at a different temperature? It is noteworthy that the phase of Fe2P could take place in the following conditions, such as the self-deoxidization reaction of LiFePO4/C composite, reducing agent iron nanoparticles, and stoichiometric excess of Fe2O3. Moreover, Fe2P could also appear when using the microwave assisted CTR (MW-CTR) method. In this part, we focus on the synthetic reaction mechanism in CTR synthetic conditions.

4.1.1 Effects of Fe(II) and Fe(III) raw material sources. Usually, Fe(III) sources including FePO4, Fe2O3, Fe(NO3)3 and Fe(NH4)2(SO4)2·6H2O are employed to prepare LiFePO4 via the CTR routes.190 In our previous research,94 it is found that LiFePO4/C can be strongly formed at around 450 °C by an in situ CTR method whether using various carbon sources except for inorganic carbon sources (postponed to 539 °C due to the preparation of Li3Fe2(PO4)3, as shown in Fig. 13a and b), similar to the reports by Garbarczyk's group191 and Jiao's group.192 The self-deoxidation reaction of LiFePO4/C takes place at 840 °C (Table 3), which was in accordance with the results of Wang's group.193 The crystallization of Fe2P and Li4P2O7 occurs at 930 °C, suggesting the severe decomposition of LiFePO4/C. Normally, the in situ CTR synthetic temperature should be carefully controlled below 840 °C to avoid the Fe2P by-product during the sintering process. However, it was found that a suitable content of Fe2P would improve the electrochemical performance of LiFePO4. Kim's group194,195 and Lee's group196 found that the Fe2P content increased with the increasing carbon content at the temperature of 900 °C, as shown in Fig. 13c. As the carbon content was above 12 wt%, a large amount of LiFePO4/C disappeared and Fe2P was predominantly produced.
image file: c4ra10899j-f13.tif
Fig. 13 (a) TG-DTA curves of the Fe(III) precursor using carbon black as carbon sources, and (b) corresponding XRD patterns of the precursor and LiFePO4/C samples after heating at various temperatures.94 (Copyright © 2009 Elsevier Ltd. All rights reserved.) (c) XRD profiles of the LiFePO4/Fe2P composites synthesized using different amounts of excess carbon.195 (Copyright © 2006 Elsevier B.V. All rights reserved.) (d) XRD patterns of samples prepared from a slightly oxidized Fe(II) precursor using various reductive heat treatments.208 (Copyright © 2003 Elsevier Science B.V. All rights reserved.)
Table 3 Reaction mechanism of deoxidization reaction of LiFePO4 in various conditions
Reduction condition Chemical reaction equation Ref.
Self-deoxidization reaction of LiFePO4/C composite (at 840 °C) 4LiFePO4 + 7C (or 14C) = 2Fe2P + 2Li2O + 2P + 7CO2 (or 14CO) 94 and 195
Severe decomposition of LiFePO4/C composite (at 938 °C) 4LiFePO4 + C (or 2C) = 2Fe2P + 2Li4P2O7 + P + CO2 (or 2CO) 94
Reducing agents (Fe3O4, Fe3C, Fe nanoparticles) from FeC2O4·2H2O (at 675 °C) 8FeC2O4·2H2O + 4LiFePO4 + 24C (or 48C) = 4Fe2P + 16H2O + 24CO2 (or 48CO) 199 and 200
Reducing agent iron nanoparticles 8Fe + 4LiFePO4 + 7C (or 14C) = 4Fe2P + 2Li2O + 7CO2 (or 14CO) 202
Reductive H2 atmosphere 6LiFePO4 + 16H2 = 2Fe2P + 2Li3PO4 + 2FeP + 16H2O 99
Microwave assisted reaction 8LiFePO4 = 2Li4P2O7 +4Fe2P +9O2 215
Microwave assisted reaction (without atmosphere of inert gases) 12LiFePO4 + 3O2 = 4Li3Fe2(PO4)3 +2Fe2O3 213 and 222


Besides Fe(III) raw material source, FeC2O4 was generally employed as a Fe(II) raw material source to synthesize LiFePO4. It can be seen that the presence of Fe2P appeared at 675 °C.197,198 Zboril's group199 and Molenda's group200 found that the reducing agents including α-Fe nanoparticles, magnetic Fe3O4 and Fe3C could be obtained during the thermal decomposition of FeC2O4·2H2O in a dry argon flow. These reducing agents, especially for Fe particles, could react with LiFePO4/C and generate Fe2P around 700 °C.201,202 It was worth noting that iron phosphides increased with the increase of excessively stoichiometric iron source and temperature.203,204

4.1.2 Reductive atmosphere of hydrogen. In addition to the self-deoxidization reaction of LiFePO4/C composite and reducing agents from Fe raw materials sources, the reductive atmosphere of hydrogen also reduced LiFePO4 to prepare Fe2P at 600 °C according to the work of Nazar's group.99 The Fe2P can be observed after the sintering process at 600 °C for 4 h in a gas atmosphere of N2/H2 = 97/3 (vol%) reported by Taniguchi's group.205,206 Furthermore, the phase of FeP and Fe3P could also be observed.207 Wohlfahrt-Mehrens's group208 studied various gas atmospheres and considered that some traces of Fe(III) could be confirmed when using Fe(II) raw sources but insufficient reductive additives (e.g. N2/H2 = 99/1 (vol%), no carbon as the reductive additive). According to the results, the authors suggested that a considerable amount of Fe2P was generated as the hydrogen raised to 10 vol%, as shown in Fig. 13d. The amount of Fe2P would increase with the carbon addition. Additionally, it is important to note that LiFePO4 will decompose into Fe2P, Li3PO4, Li4P2O7 and Li3P7 when the heating temperature reaches 850° C under a stream of a gas mixture of Ar/H2 = 95/5 (vol%).209
4.1.3 Microwave assisted carbothermal reduction method. When compared with a conventional CTR method, MW-CTR method has recently attracted much attention.177 Microwave irradiation (2.54 GHz) can provide “inert and instant heating”, and exhibit outstanding superiority such as fast response, time effectiveness, low energy consumption, and environmentally benignity.210,211 Moreover, microwave irradiation was also adopted to optimize the physical and electrochemical performance of LiFePO4.212 Under Ar atmosphere, Higuchi's group213 found that iron acetate acts as a microwave susceptor rather than iron lactate. This result illustrates that LiFePO4 was successfully prepared using iron acetate as the Fe sources at 500 W for 10 min, whereas LiFePO4 was not observed using iron lactate as Fe raw sources even up to 30 min. Furthermore, the microwave irradiation time has an important effect on the Fe2P by-products, which could increase with the increasing time due to the work of Kwon's group.214 All the phases of LiFePO4, Fe2P and Li4P2O7 were observed when the amount of Fe2P is above 1.0 wt%.215 In addition to Ar atmosphere, carbon materials and carbon sources were both employed to produce vigorous CO gas, which is necessary for the preparation of LiFePO4.216,217 Because of this strategy, the electrochemical performance of LiFePO4 improved considerably by the MW-CRT method using carbon coating,218 multi-walled CNTs (MWCNTs) doping,219 Mo and La doping.220,221 However, it is worthwhile to note that most of the carbon could be consumed after extended heating treatment. The transformation from Fe(II) to Fe(III) would take place, leading to impurity phases of Fe2O3 and Li3Fe2(PO4)3.222,223 Therefore, the microwave-derived and phase-pure LiFePO4 without impurities could be successfully synthesized via the MW-CTR method under certain proper conditions.

4.2 Low-temperature liquid phase thermal synthesis

In addition to the high-temperature CTR method, the low-temperature LPT routes, such as HTS, STS and ionothermal synthesis (ITS), provided more effective pathways to synthesize LiFePO4. HTS and STS are generally LPT synthetic methods, which exhibit several advantages including morphology control, homogeneous particle size distribution, relatively low temperature (≤300 °C), low cost and simplicity.224,225 Recently, HTS and STS were both extensively employed to prepare special morphologic LiFePO4, such as crystalline growth, nanostructured LiFePO4 and 3D porous LiFePO4 architectures with micro-nano-structures (see Section 3.1.1 in this article).127,226 In order to improve traditional intermittent reactors for HTS and STS, Teja's group227 and Aimable's group228 designed the continuous hydrothermal synthesis, which poise great promise in practical industrial applications. Different from HTS (using water as the main mineralizer) and STS (using organic solvents as the main mineralizers), room temperature ionic liquids (RTILs) are regarded as a novel class of mineralizers that have been employed in the ITS route to synthesize cathode materials, especially LiFePO4 with nano-structures.229–231 Moreover, microwave irradiation is expected to assist HTS, STS and ITS due to energy- and time-saving advantages of microwaves. In this part, we focus on the effects of stoichiometric ratio of raw sources, pH values of solution, additive mediations (e.g., organic acids, surfactants and solvents), and microwave irradiation.
4.2.1 Stoichiometric ratio of raw sources. Usually, the crystalline powders of LiFePO4 were synthesized by mixing amounts of the reactants LiOH, FeSO4 and H3PO4 in the stoichiometric ratio of nLi[thin space (1/6-em)]:[thin space (1/6-em)]nFe[thin space (1/6-em)]:[thin space (1/6-em)]nP = 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1, as shown in Table 4.232 An advisable addition of LiOH should be changed with replacing the H+ with NH4+ in H3PO4. Kanamura's group233,234 found that the stoichiometric ratio of 2.5[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 would be better when using (NH4)3PO4 instead of H3PO4. The stoichiometric ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 and 2[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 were both appropriate for using (NH4)2HPO4. As shown in Fig. 14, our group investigated the products when using NH4H2PO4 as the PO43− source. It can be seen that a large amount of Fe3(PO4)2·H2O as an impurity was obtained in the case of x = 1.0. As the x value changed to 1.5, additional crystallinity of LiFePO4 was observed. The value of x = 2.0 is obviously in favor of the hydrothermal crystallization of pure LiFePO4. The further addition of LiOH (e.g., x = 2.5 and 3.0) induced impurities in Li3PO4. All these results strongly suggest that the addition of LiOH should be changed as the PO43− source was changed. However, the stoichiometric ratio should be changed back to 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 when the other Li sources including LiCl and CH3COOLi were used instead of LiOH.235,236 It is noteworthy that LiFePO4 with a pure phase was also prepared by using other Fe(II) salts such as FeCl2 and (NH4)2Fe(SO4)2 instead of FeSO4, but the stoichiometric ratio remained unchanged.237,238 Furthermore, LiFePO4 was also successfully synthesized using fresh Fe3(PO4)2·5H2O and Li3PO4.239
Table 4 Summary of representative reaction mechanism of LiFePO4 prepared by low-temperatured LPT synthesis. In the table, mediator includes organic acids or surfactant compounds
Synthesis type Reaction reagents Stoichiometric ratio Mediator Solvent Condition Morphology Ref.
HTS LiOH, FeSO4, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 H2O 120 °C, 5 h Hexagonal platelets 232
HTS LiOH, FeSO4, (NH4)3PO4 2.5[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 H2O 170 °C, 12 h 233
HTS LiOH, FeSO4, (NH4)2HPO4 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1&2[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 H2O 170 °C, 12 h Nanoplates 234
HTS LiOH, FeSO4, NH4H2PO4 2[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 H2O 180 °C, 5 h Spindle-like particles This work
HTS LiOH, FeCl2·4H2O, P2O5 6[thin space (1/6-em)]:[thin space (1/6-em)]3[thin space (1/6-em)]:[thin space (1/6-em)]2 H2O 170 °C, 3 days Hexagonal platelets 237
HTS LiOH, FeSO4, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 CTAB H2O 120 °C, 5 h Nanoparticles 251–253
HTS LiOH, (NH4)2Fe(SO4)2·6H2O, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 P123, D230 H2O 220 °C, 24 h Nanoparticles 106
HTS LiOH, (NH4)2Fe(SO4)2·6H2O, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 Citric acid H2O 180 °C, 10 h Spindle-like particles 250
HTS LiOH, FeSO4, NH4H2PO4 2[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 NTA H2O 180 °C, 20 h Nanowires 254
HTS LiAc, FeCl3, NH4H2PO4 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 SDS + EN H2O 180 °C, 10 h Spheres assembled from nanorods This work
STS LiOH, FeSO4, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 TEG + H2O 190 °C, 5 h Rectangular nanoplatelets 111
STS LiOH, FeSO4, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 Citric acid MPG + H2O 140 °C, 12 h Needles assembled nanorods 255
STS LiOH, FeSO4, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 PEG400 + H2O 150 °C, 3 h Needle-like particles 256
STS LiOH, FeSO4, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 PEG400 + H2O 180 °C, 9 h Hexagonal platelet 112
ITS LiH2PO4, FeC2O4·2H2O 1[thin space (1/6-em)]:[thin space (1/6-em)]1 CN-based IL 250 °C, 24 h Needles assembled Lego blocks 229
ITS LiOH, FeSO4, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 SDBS + ascorbic acid IL + H2O 240 °C, 20 h Nanorods 258
MW-HTS LiOH, FeSO4, H3PO4 3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 H2O 200 °C, 5 min Globular particles 259
MW-STS LiOH, FeAc2, H3PO4 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 TEG 300 °C, 5 min Nanorods 262



image file: c4ra10899j-f14.tif
Fig. 14 XRD patterns of the samples prepared using different values of x in a molar ratio of x[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 for the starting materials including LiOH, FeSO4 and NH4H2PO4.

Additionally, Yu's group240 suggested that LiFePO4 should only be obtained under neutral and slightly basic conditions. Fe disorder onto the Li sites (ca. 3.5%) generally happened at a pH value of 6.30. Even at a synthetic condition temperature below 175 °C, there would be some amounts of Fe on Li sites.241,242 Moreover, the synthetic time also had effects on the LiFePO4 nanocrystalline structure. In 2013, Ou's group243 found that the peak intensities of (020) X-ray diffraction line of LiFePO4 nanoplates increased with the reaction time prolonging from 0 to 6 h. Meanwhile, the relative peak intensities of (020) to (111) also increased due to the results calculated from the XRD patterns of LiFePO4 nanoplates. All these results demonstrated that the LiFePO4 nanoplates have a preferred crystal orientation with the large (010) face, which facilitates the good electrochemical performance of LiFePO4.244

4.2.2 Additive mediations of organic acids, surfactants and solvents. In addition to the synthetic temperature, residence time, reactant concentration and pH value,245 additive mediations including organic acids, surfactants and solvents also have effects on the purity, morphology control and electrochemical performance of LiFePO4.246 Firstly, organic acid not only affects the pH value but also can be used as a reduction mediator. Whittingham's group247 provided that both of the reduction from Fe(III) into Fe(II) and the formation of the desired LiFePO4 compound carried out with the addition of L-ascorbic acid. Moreover, it was found that the pH value and reaction time played several roles in self-assembled mesoporous LiFePO4 with hierarchical spindle-like architectures.248 When the value of pH increased from 7.0 to 10.0, primary nanocrystals disappeared and spindle-like hierarchical LiFePO4 particles formed. The full transformation from Li3PO4 to LiFePO4 nanocrystal needs a reaction time of over 5 h. Furthermore, in the presence of an organic acid, e.g. citric acid or ascorbic acid, well-crystallized LiFePO4 nanoparticles have been directly synthesized.249 Chung's group and Grey's group250 successfully controlled the morphological transformations of LiFePO4 particles during the hydrothermal reaction in the presence of citric acid and ammonium ions (NH4+).

Secondly, organic surfactants or polymers including O,O-bis(2-aminopropyl)polypropyleneglycol (D230), EO20PO70EO20 (P123), cyltrimethylammonium bromide (CTAB), sodium dodecyl sulfate (SDS), poly(ethylene glycol) (PEG), etc. have been successfully employed to control the particle growth and size distribution. The results from Nazar's group106 showed the particle size of LiFePO4 became smaller and much more homogeneous as compared to that of products using molecular reducing agents, if non-ionic surfactants, including pluronic P123 and Jeffamine D230 were added to the reactors. Recently, Gerbaldi and his co-workers251–253 investigated the effects of the organic surfactant compound CTAB on LiFePO4 characteristics and electrochemical behavior. It is clearly found that CTAB considerably influences the synthesis and electrochemical properties of LiFePO4. Moreover, the CTAB micelles are beneficial for controlling the grain size and surface area. Additionally, nitrilotriacetic acid (NTA) surfactant has been also employed to synthesize LiFePO4 nanowires.254 In 2010, Zhang's group168 used SDS to assist the synthesis of LiFePO4 via a hydrothermal process. The as-obtained LiFePO4/C microspheres are composed primarily of LiFePO4/C nanoplates with the size of 100 nm and thickness of 30 nm. With the addition of SDS and ethylenediamine (EN), our group successfully synthesized LiFePO4/C microspheres assembled from nanorods, as shown in Fig. 15c and d.


image file: c4ra10899j-f15.tif
Fig. 15 (a) SEM images of a LiFePO4 hemisphere, demonstrating the microsphere consisting of nanofibers.255 (Copyright © 2013, Royal Society of Chemistry.) (b) SEM images of LiFePO4.111 (Copyright © 2011, Springer Science+Business Media, LLC.) (c and d) SEM images of a LiFePO4 hemisphere, and the LiFePO4 nanorods. (e and f) SEM image and lattice fringe (the insets of ED patterns and TEM image) of LiFePO4 nanorods.258 (Copyright © 2011 Elsevier B.V. All rights reserved.) (g–i) TEM images of the large LiFePO4 needle assembling like Lego blocks, and the corresponding SAED pattern. The white arrow shows the “cementing” zone.229 (Copyright © 2009, American Chemical Society.)

Thirdly, the hybrids of water and organic solvents have also been employed in the STS route. In 2013, Liao's group255 successfully prepared high-performance LiFePO4 microspheres consisting of nanofibers (Fig. 15a) by a high pressure alcohol-thermal approach in a water and 1,2-propanediol (PD) composite solvent. With the effect of the PD solvent, the LiFePO4 crystalline nuclei were linearly elongated to form nanofibers. LiFePO4 nanofibers aggregated together and formed LiFePO4 microspheres. Tetraethylene glycol (TEG) has been also used as a cosolvent with water.111 As shown in Fig. 15b, rectangular LiFePO4 nanoplatelets were formed with a very small thickness of 50–80 nm, favouring fast Li-ion diffusion. Moreover, PEG has been usually used to assist the morphology control of LiFePO4.256,257 Meanwhile, Liu's group112 provided a water-PEG400 binary solvent to assist the solvothermal method and prepare LiFePO4 nanoparticles (50 nm in size), hexagonal nanoplates (800 nm wide and 100 nm thick) and hexagonal microplates (3 mm wide and 300 nm thick).

Recently, RTILs have been also been used as the solvent and template to synthesize LiFePO4. With this new and promising ITS method, Tarascon's group178,229 enabled LiFePO4 crystalline growth with controlled morphology and size at around 250 °C. As shown in Fig. 15g–i, large LiFePO4 needles were assembled similar to Lego blocks. As shown in Fig. 15e and f, Teng and co-workers258 successfully prepared LiFePO4 nanorods in an ionic liquid in the presence of sodium dodecyl benzene sulfonate (SDBS) surfactant and ascorbic acid. Thus, super architectures with micro-nano-structured nanocrystals and morphology control can be obtained with the addition of organic acids, organic surfactants and organic solvents.

4.2.3 Microwave-assisted liquid phase thermal routes. Undoubtedly, microwave irradiation has been employed to assist LPT routes, because microwave irradiation assisted synthesis (MIAS) exhibited overwhelming superiorities such as heating the precursor considerably and quickly, and also provide an efficient and suitable approach for large-scale syntheses in order to be used in practical applications of electrode materials. In our previous work,177 we exhaustively reviewed MW-LPT including MW-HTS,259,260 MW-STS261 and MW-ITS.178 Li-ion diffusion can be improved considerably by controlling the morphology of LiFePO4 via MW-LPT. Moreover, electronic conductivity can be optimized by coating conductive materials such as carbon,262,263 CNT264,265 and conductive polymers, e.g., poly(3,4-ethylenedioxythiophene)(PEDOT).266

5 Conclusion and perspective

In summary, the pathway of Li-ion diffusion in LixFePO4 is mainly along the [010] direction. Sometimes, 2D Li-ion transport occurs along both the [010] and [001] directions at elevated temperatures. Undoubtedly, the phase transformation of LiFePO4 and FePO4 are affected by various conditions, such as particle size, charge/discharge rate, temperature, etc. To date, many phase transformation models, including shrinking core (i.e., core–shell) model, Laffont's (i.e., new core–shell) model, mosaic model, domino-cascade model, phase transformation wave model, and many-particle model, have been proposed to explain the lithiation/delithiation process and the phase transformation in LixFePO4. However, these models are still in disagreement with each other. Generally, the two-phase transformation models can be used to explain a two-phase growth process involving the coexistence of LiFePO4 and FePO4 for the large particle, below a critical current and at a relative low temperature. Simultaneously, the quasi-single-phase transformation models are dominantly used in the solid-solution reaction where LixFePO4 particles do not separate into Li-rich and Li-poor phases during the lithiation/delithiation process.

In order to endow LiFePO4 with excellent electrochemical performance including high rate capability and low temperature performance, reducing the pathway of Li-ion diffusion, particularly the [010] direction, is an effective route. Usually, fast Li-ion diffusion is achieved by morphology controlling, such as crystal growth orientation along the ac plane, nano-sized morphologies and 3D porous architectures with micro-nano-structures. Meanwhile, the electronic conductivity is improved by coating conductive materials, e.g., amorphous carbon, CNT, graphene, PAS, PPy, Fe2P, etc. However, achieving both high rate performance and volumetric energy density is still a big challenge. Recently, carbon coated LiFePO4 microspheres with the highest tap density of 2.0 g cm−3 have been successfully synthesized by the MW-WBR method. They can deliver outstanding discharge capacities of 93 mA h g−1 and 78 mA h g−1 even at 10 C and 20 C, respectively. This additional strategy provides a facile and novel route for obtaining high tap density from LiFePO4, and it holds the potential to be extended to EES devices.

In addition, it is noteworthy that synthetic reactions play an important role to improve the performance of LiFePO4 cathodes. When using the CTR method, the purity of LiFePO4 is significantly affected by the Fe sources, carbon sources and the reductive atmosphere. Obviously, the impurities, particularly for Fe2P, can be easily separated as by-products. To date, controlling the content of Fe2P is still a major obstacle and issue, where Fe2P either provides the high conductive networks for LiFePO4 or blocks the pathway of Li-ion diffusion. Low-temperature LPT synthesis can be also employed to prepare LiFePO4 with much reduced impurities. However, both the crystalline and electrochemical performances of LiFePO4 still need to be enhanced by subsequent heat treatment. Intriguingly, on account of “inert and instant heating”, MIAS can be used to assist both the CTR and LPT routes. We believe that MIAS provides an extremely efficient and suitable route for the large-scale synthesis of LiFePO4 electrode materials.

Acknowledgements

The authors highly acknowledge the authors and publications data of the original materials. We would like to thank Dr Chaoyi Yan, Prof. Haibin Su (School of Materials Science and Engineering, Nanyang Technological University (MSE, NTU)) and Prof. Jianyi Lin (Institute of Chemical and Engineering Sciences, Agency for Science, Technology and Research (ICES, A*STAR)) for their helpful discussions. This work was financially supported by Scientific Research Start-up Fund for High-Level Talents, Shihezi University (no. RCZX201305), the Doctor Foundation of Bingtuan (no. 2014BB004), the Program for Changjiang Scholars and Innovative Research Team in University (no. IRT1161) and the Program of Science and Technology Innovation Team in Bingtuan (no. 2011CC001).

Notes and references

  1. K. Mizushima, P. C. Jones, P. J. Wiseman and J. B. Goodenough, Mater. Res. Bull., 1980, 15, 783–789 CrossRef CAS.
  2. T. Nagaura and K. Tazawa, Prog. Batteries Sol. Cells, 1990, 9, 209–217 CAS.
  3. B. Xu, D. N. Qian, Z. Y. Wang and Y. S. L. Meng, Mater. Sci. Eng., R, 2012, 73, 51–65 CrossRef CAS.
  4. M. Wagemaker and F. M. Mulder, Acc. Chem. Res., 2013, 46, 1206–1215 CrossRef CAS PubMed.
  5. J. Cabana, L. Monconduit, D. Larcher and M. R. Palacin, Adv. Mater., 2010, 22, E170–E192 CrossRef CAS PubMed.
  6. X. P. Gao and H. X. Yang, Energy Environ. Sci., 2010, 3, 174–189 CAS.
  7. M. R. Palacin, Chem. Soc. Rev., 2009, 38, 2565–2575 RSC.
  8. S. Chu and A. Majumdar, Nature, 2012, 488, 294–303 CrossRef CAS PubMed.
  9. J. M. Tarascon and M. Armand, Nature, 2001, 414, 359–367 CrossRef CAS PubMed.
  10. R. Marom, S. F. Amalraj, N. Leifer, D. Jacob and D. Aurbach, J. Mater. Chem., 2011, 21, 9938–9954 RSC.
  11. M. Wakihara, Mater. Sci. Eng., R, 2001, 33, 109–134 CrossRef.
  12. A. S. Arico, P. Bruce, B. Scrosati, J. M. Tarascon and W. Van Schalkwijk, Nat. Mater., 2005, 4, 366–377 CrossRef CAS PubMed.
  13. M. Gong and G. Wall, Exergy: An International Journal, 2001, 1, 217–233 CrossRef.
  14. K. Zaghib, A. Mauger and C. M. Julien, J. Solid State Electrochem., 2012, 16, 835–845 CrossRef CAS.
  15. J. B. Goodenough and Y. Kim, Chem. Mater., 2010, 22, 587–603 CrossRef CAS.
  16. M. S. Islam and C. A. J. Fisher, Chem. Soc. Rev., 2014, 43, 185–204 RSC.
  17. M. S. Whittingham, Chem. Rev., 2004, 104, 4271–4302 CrossRef CAS PubMed.
  18. F. Yu, J. J. Zhang, C. Y. Wang, J. Yuan, Y. F. Yang and G. Z. Song, Progr. Chem., 2010, 22, 9–18 CAS.
  19. Y. Gu, D. Chen and X. Jiao, J. Phys. Chem. B, 2005, 109, 17901–17906 CrossRef CAS PubMed.
  20. K. Y. Chung and K. B. Kim, Electrochim. Acta, 2004, 49, 3327–3337 CrossRef CAS.
  21. J. Vetter, P. Novak, M. R. Wagner, C. Veit, K. C. Moller, J. O. Besenhard, M. Winter, M. Wohlfahrt-Mehrens, C. Vogler and A. Hammouche, J. Power Sources, 2005, 147, 269–281 CrossRef CAS.
  22. A. K. Padhi, K. S. Nanjundaswamy, C. Masquelier, S. Okada and J. B. Goodenough, J. Electrochem. Soc., 1997, 144, 1609–1613 CrossRef CAS.
  23. A. K. Padhi, K. S. Nanjundaswamy and J. B. Goodenough, J. Electrochem. Soc., 1997, 144, 1188–1194 CrossRef CAS.
  24. S. Y. Chung, J. T. Bloking and Y. M. Chiang, Nat. Mater., 2002, 1, 123–128 CrossRef CAS PubMed.
  25. J. W. Fergus, J. Power Sources, 2010, 195, 939–954 CrossRef CAS.
  26. C. A. J. Fisher, V. M. H. Prieto and M. S. Islam, Chem. Mater., 2008, 20, 5907–5915 CrossRef CAS.
  27. Y. G. Wang, P. He and H. S. Zhou, Energy Environ. Sci., 2011, 4, 805–817 CAS.
  28. M. S. Islam and P. R. Slater, MRS Bull., 2009, 34, 935–941 CrossRef CAS.
  29. D. D. MacNeil, Z. H. Lu, Z. H. Chen and J. R. Dahn, J. Power Sources, 2002, 108, 8–14 CrossRef CAS.
  30. K. Zaghib, P. Charest, A. Guerfi, J. Shim, M. Perrier and K. Striebel, J. Power Sources, 2004, 134, 124–129 CrossRef CAS.
  31. S. P. Ong, V. L. Chevrier, G. Hautier, A. Jain, C. Moore, S. Kim, X. H. Ma and G. Ceder, Energy Environ. Sci., 2011, 4, 3680–3688 CAS.
  32. W. J. Zhang, J. Power Sources, 2011, 196, 2962–2970 CrossRef CAS.
  33. P. P. Prosini, M. Lisi, D. Zane and M. Pasquali, Solid State Ionics, 2002, 148, 45–51 CrossRef CAS.
  34. D. W. Liu and G. Z. Cao, Energy Environ. Sci., 2010, 3, 1218–1237 CAS.
  35. L. X. Yuan, Z. H. Wang, W. X. Zhang, X. L. Hu, J. T. Chen, Y. H. Huang and J. B. Goodenough, Energy Environ. Sci., 2011, 4, 269–284 CAS.
  36. D. Morgan, A. Van der Ven and G. Ceder, Electrochem. Solid-State Lett., 2004, 7, A30–A32 CrossRef CAS.
  37. M. S. Islam, D. J. Driscoll, C. A. J. Fisher and P. R. Slater, Chem. Mater., 2005, 17, 5085–5092 CrossRef CAS.
  38. C. Y. Ouyang, S. Q. Shi, Z. X. Wang, X. J. Huang and L. Q. Chen, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 69, 104303 CrossRef.
  39. X. F. Ouyang, M. L. Lei, S. Q. Shi, C. L. Luo, D. S. Liu, D. Y. Jiang, Z. Q. Ye and M. S. Lei, J. Alloys Compd., 2009, 476, 462–465 CrossRef CAS.
  40. S. Nishimura, G. Kobayashi, K. Ohoyama, R. Kanno, M. Yashima and A. Yamada, Nat. Mater., 2008, 7, 707–711 CrossRef CAS PubMed.
  41. J. J. Yang and J. S. Tse, J. Phys. Chem. A, 2011, 115, 13045–13049 CrossRef CAS PubMed.
  42. J. B. Goodenough, J. Power Sources, 2007, 174, 996–1000 CrossRef CAS.
  43. M. Park, X. C. Zhang, M. D. Chung, G. B. Less and A. M. Sastry, J. Power Sources, 2010, 195, 7904–7929 CrossRef CAS.
  44. J. Y. Li, W. L. Yao, S. Martin and D. Vaknin, Solid State Ionics, 2008, 179, 2016–2019 CrossRef CAS.
  45. R. Malik, D. Burch, M. Bazant and G. Ceder, Nano Lett., 2010, 10, 4123–4127 CrossRef CAS PubMed.
  46. B. L. Ellis, K. T. Lee and L. F. Nazar, Chem. Mater., 2010, 22, 691–714 CrossRef CAS.
  47. R. Amin, J. Maier, P. Balaya, D. P. Chen and C. T. Lin, Solid State Ionics, 2008, 179, 1683–1687 CrossRef CAS.
  48. B. Ellis, L. K. Perry, D. H. Ryan and L. F. Nazar, J. Am. Chem. Soc., 2006, 128, 11416–11422 CrossRef CAS PubMed.
  49. G. K. P. Dathar, D. Sheppard, K. J. Stevenson and G. Henkelman, Chem. Mater., 2011, 23, 4032–4037 CrossRef CAS.
  50. V. Srinivasan and J. Newman, J. Electrochem. Soc., 2004, 151, A1517–A1529 CrossRef CAS.
  51. L. Laffont, C. Delacourt, P. Gibot, M. Y. Wu, P. Kooyman, C. Masquelier and J. M. Tarascon, Chem. Mater., 2006, 18, 5520–5529 CrossRef CAS.
  52. A. S. Andersson, B. Kalska, L. Haggstrom and J. O. Thomas, Solid State Ionics, 2000, 130, 41–52 CrossRef CAS.
  53. A. S. Andersson and J. O. Thomas, J. Power Sources, 2001, 97–98, 498–502 CrossRef CAS.
  54. C. Delmas, M. Maccario, L. Croguennec, F. Le Cras and F. Weill, Nat. Mater., 2008, 7, 665–671 CrossRef CAS PubMed.
  55. R. Dedryvere, M. Maccario, L. Croguennec, F. Le Cras, C. Delmas and D. Gonbeau, Chem. Mater., 2008, 20, 7164–7170 CrossRef CAS.
  56. D. Burch, G. Singh, G. Ceder and M. Z. Bazant, in Theory, Modeling and Numerical Simulation of Multi-Physics Materials Behavior, ed. V. Tikare, G. E. Murch, F. Soisson and J. K. Kang, 2008, vol. 139, pp. 95–100 Search PubMed.
  57. G. K. Singh, G. Ceder and M. Z. Bazant, Electrochim. Acta, 2008, 53, 7599–7613 CrossRef CAS.
  58. W. Dreyer, J. Jamnik, C. Guhlke, R. Huth, J. Moskon and M. Gaberscek, Nat. Mater., 2010, 9, 448–453 CrossRef CAS PubMed.
  59. N. Meethong, H. Y. S. Huang, S. A. Speakman, W. C. Carter and Y. M. Chiang, Adv. Funct. Mater., 2007, 17, 1115–1123 CrossRef CAS.
  60. G. Y. Chen, X. Y. Song and T. J. Richardson, Electrochem. Solid-State Lett., 2006, 9, A295–A298 CrossRef CAS.
  61. P. P. Prosini, J. Electrochem. Soc., 2005, 152, A1925–A1929 CrossRef CAS.
  62. D. X. Gouveia, V. Lemos, J. A. C. de Paiva, A. G. Souza, J. Mendes, S. M. Lala, L. A. Montoro and J. M. Rosolen, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 72, 024105 CrossRef.
  63. D. Li, T. Zhang, X. Z. Liu, P. He, R. W. Peng, M. Wang, M. Han and H. S. Zhou, J. Power Sources, 2013, 233, 299–303 CrossRef CAS.
  64. C. V. Ramana, A. Mauger, F. Gendron, C. M. Julien and K. Zaghib, J. Power Sources, 2009, 187, 555–564 CrossRef CAS.
  65. Y. Orikasa, T. Maeda, Y. Koyama, H. Murayama, K. Fukuda, H. Tanida, H. Arai, E. Matsubara, Y. Uchimoto and Z. Ogumi, Chem. Mater., 2013, 25, 1032–1039 CrossRef CAS.
  66. W. C. Chueh, F. El Gabaly, J. D. Sugar, N. C. Bartelt, A. H. McDaniel, K. R. Fenton, K. R. Zavadil, T. Tyliszczak, W. Lai and K. F. McCarty, Nano Lett., 2013, 13, 866–872 CrossRef CAS PubMed.
  67. P. Bai and G. Y. Tian, Electrochim. Acta, 2013, 89, 644–651 CrossRef CAS.
  68. T. Ichitsubo, K. Tokuda, S. Yagi, M. Kawamori, T. Kawaguchi, T. Doi, M. Oishi and E. Matsubara, J Mater Chem A, 2013, 1, 2567–2577 CAS.
  69. S. Dargaville and T. W. Farrell, Electrochim. Acta, 2013, 94, 143–158 CrossRef CAS.
  70. L. Wang, T. Maxisch and G. Ceder, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 195107 CrossRef.
  71. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  72. M. S. A. Rasiman, F. W. Badrudin, T. I. T. Kudin, M. K. Yaakob, M. F. M. Taib, M. Z. A. Yahya and O. H. Hassan, Integr. Ferroelectr., 2014, 155, 71–79 CrossRef CAS.
  73. S. Leoni, M. Baldoni, L. Craco, J. O. Joswig and G. Seifert, Z. Phys. Chem., 2012, 226, 95–106 CrossRef CAS.
  74. M. E. A. Y. de Dompablo, N. Biskup, J. M. Gallardo-Amores, E. Moran, H. Ehrenberg and U. Amador, Chem. Mater., 2010, 22, 994–1001 CrossRef.
  75. S. Q. Yang, T. R. Zhang, Z. L. Tao and J. Chen, Acta Chim. Sin., 2013, 71, 1029–1034 CrossRef CAS.
  76. N. Marx, L. Croguennec, D. Carlier, A. Wattiaux, F. Le Cras, E. Suard and C. Delmas, Dalton Trans., 2010, 39, 5108–5116 RSC.
  77. L. Gu, C. B. Zhu, H. Li, Y. Yu, C. L. Li, S. Tsukimoto, J. Maier and Y. Ikuhara, J. Am. Chem. Soc., 2011, 133, 4661–4663 CrossRef CAS PubMed.
  78. R. Malik, F. Zhou and G. Ceder, Nat. Mater., 2011, 10, 587–590 CrossRef CAS PubMed.
  79. G. Y. Chen, X. Y. Song and T. J. Richardson, J. Electrochem. Soc., 2007, 154, A627–A632 CrossRef CAS.
  80. P. Bai, D. A. Cogswell and M. Z. Bazant, Nano Lett., 2011, 11, 4890–4896 CrossRef CAS PubMed.
  81. J. L. Dodd, R. Yazami and B. Fultz, Electrochem. Solid-State Lett., 2006, 9, A151–A155 CrossRef CAS.
  82. C. Delacourt, J. Rodriguez-Carvajal, B. Schmitt, J. M. Tarascon and C. Masquelier, Solid State Sci., 2005, 7, 1506–1516 CrossRef CAS.
  83. C. Delacourt, P. Poizot, J. M. Tarascon and C. Masquelier, Nat. Mater., 2005, 4, 254–260 CrossRef CAS.
  84. A. Yamada, H. Koizumi, S. I. Nishimura, N. Sonoyama, R. Kanno, M. Yonemura, T. Nakamura and Y. Kobayashi, Nat. Mater., 2006, 5, 357–360 CrossRef CAS PubMed.
  85. P. Gibot, M. Casas-Cabanas, L. Laffont, S. Levasseur, P. Carlach, S. Hamelet, J. M. Tarascon and C. Masquelier, Nat. Mater., 2008, 7, 741–747 CrossRef CAS PubMed.
  86. N. Meethong, H. Y. S. Huang, W. C. Carter and Y. M. Chiang, Electrochem. Solid-State Lett., 2007, 10, A134–A138 CrossRef CAS.
  87. J. Maier, Nat. Mater., 2005, 4, 805–815 CrossRef CAS PubMed.
  88. X. L. Wu, L. Y. Jiang, F. F. Cao, Y. G. Guo and L. J. Wan, Adv. Mater., 2009, 21, 2710–2714 CrossRef CAS.
  89. P. G. Bruce, B. Scrosati and J. M. Tarascon, Angew. Chem., Int. Ed., 2008, 47, 2930–2946 CrossRef CAS PubMed.
  90. Y. Wang and G. Z. Cao, Adv. Mater., 2008, 20, 2251–2269 CrossRef CAS.
  91. F. Yu, J. J. Zhang, Y. F. Yang and G. Z. Song, J. Power Sources, 2010, 195, 6873–6878 CrossRef CAS.
  92. C. W. Sun, S. Rajasekhara, J. B. Goodenough and F. Zhou, J. Am. Chem. Soc., 2011, 133, 2132–2135 CrossRef CAS PubMed.
  93. G. X. Wang, L. Yang, S. L. Bewlay, Y. Chen, H. K. Liu and J. H. Ahn, J. Power Sources, 2005, 146, 521–524 CrossRef CAS.
  94. F. Yu, J. J. Zhang, Y. F. Yang and G. Z. Song, Electrochim. Acta, 2009, 54, 7389–7395 CrossRef CAS.
  95. Y. Ding, Y. Jiang, F. Xu, J. Yin, H. Ren, Q. Zhuo, Z. Long and P. Zhang, Electrochem. Commun., 2010, 12, 10–13 CrossRef CAS.
  96. B. Jin, E. M. Jin, K. H. Park and H. B. Gu, Electrochem. Commun., 2008, 10, 1537–1540 CrossRef CAS.
  97. Z. L. Liu, S. W. Tay, L. A. Hong and J. Y. Lee, J. Solid State Electrochem., 2011, 15, 205–209 CrossRef CAS.
  98. C. S. Li, S. Y. Zhang, F. Y. Cheng, W. Q. Ji and J. Chen, Nano Res., 2008, 1, 242–248 CrossRef CAS.
  99. Y. H. Rho, L. F. Nazar, L. Perry and D. Ryan, J. Electrochem. Soc., 2007, 154, A283–A289 CrossRef CAS.
  100. P. S. Herle, B. Ellis, N. Coombs and L. F. Nazar, Nat. Mater., 2004, 3, 147–152 CrossRef CAS PubMed.
  101. S. X. Zhao, H. Ding, Y. C. Wang, B. H. Li and C. W. Nan, J. Alloys Compd., 2013, 566, 206–211 CrossRef CAS.
  102. S. Hamelet, M. Casas-Cabanas, L. Dupont, C. Davoisne, J. M. Tarascon and C. Masquelier, Chem. Mater., 2011, 23, 32–38 CrossRef CAS.
  103. J. Molenda and J. Marzec, Funct. Mater. Lett., 2008, 1, 97–104 CrossRef CAS.
  104. C. A. J. Fisher and M. S. Islam, J. Mater. Chem., 2008, 18, 1209–1215 RSC.
  105. S. Franger, F. Le Cras, C. Bourbon and H. Rouault, J. Power Sources, 2003, 119, 252–257 CrossRef.
  106. B. Ellis, W. H. Kan, W. R. M. Makahnouk and L. F. Nazar, J. Mater. Chem., 2007, 17, 3248–3254 RSC.
  107. L. Wang, F. Zhou, Y. S. Meng and G. Ceder, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 165435 CrossRef.
  108. Y. S. Meng and M. E. Arroyo-de Dompablo, Energy Environ. Sci., 2009, 2, 589–609 CAS.
  109. C. Y. Nan, J. Lu, L. H. Li, L. L. Li, Q. Peng and Y. D. Li, Nano Res., 2013, 6, 469–477 CrossRef CAS.
  110. J. Lim, V. Mathew, K. Kim, J. Moon and J. Kim, J. Electrochem. Soc., 2011, 158, A736–A740 CrossRef CAS.
  111. H. F. Xiang, D. W. Zhang, Y. Jin, C. H. Chen, J. S. Wu and H. H. Wang, J. Mater. Sci., 2011, 46, 4906–4912 CrossRef CAS.
  112. S. L. Yang, X. F. Zhou, J. G. Zhang and Z. P. Liu, J. Mater. Chem., 2010, 20, 8086–8091 RSC.
  113. K. Saravanan, M. V. Reddy, P. Balaya, H. Gong, B. V. R. Chowdari and J. J. Vittal, J. Mater. Chem., 2009, 19, 605–610 RSC.
  114. K. Saravanan, P. Balaya, M. V. Reddy, B. V. R. Chowdari and J. J. Vittal, Energy Environ. Sci., 2010, 3, 457–464 CAS.
  115. P. Balaya, K. Saravanan and S. Hariharan, in Energy Harvesting and Storage: Materials, Devices, and Applications, ed. N. K. Dhar, P. S. Wijewarnasuriya and A. K. Dutta, Spie-Int Soc Optical Engineering, Bellingham, 2010, vol. 7683 Search PubMed.
  116. Z. Y. Bi, X. D. Zhang, W. He, D. D. Min and W. S. Zhang, RSC Adv., 2013, 3, 19744–19751 RSC.
  117. K. T. Lee and J. Cho, Nano Today, 2011, 6, 28–41 CrossRef CAS.
  118. C. M. Julien, A. Mauger and K. Zaghib, J. Mater. Chem., 2011, 21, 9955–9968 RSC.
  119. W. J. Zhang, J. Electrochem. Soc., 2010, 157, A1040–A1046 CrossRef CAS.
  120. X. Zhao, B. M. Sanchez, P. J. Dobson and P. S. Grant, Nanoscale, 2011, 3, 839–855 RSC.
  121. D. Zhang, R. Cai, Y. K. Zhou, Z. P. Shao, X. Z. Liao and Z. F. Ma, Electrochim. Acta, 2010, 55, 2653–2661 CrossRef CAS.
  122. Y. G. Wang, Y. R. Wang, E. J. Hosono, K. X. Wang and H. S. Zhou, Angew. Chem., Int. Ed., 2008, 47, 7461–7465 CrossRef CAS PubMed.
  123. S. Y. Lim, C. S. Yoon and J. P. Cho, Chem. Mater., 2008, 20, 4560–4564 CrossRef CAS.
  124. M. Konarova and I. Taniguchi, J. Power Sources, 2010, 195, 3661–3667 CrossRef CAS.
  125. J. J. Wang, J. L. Yang, Y. J. Tang, J. Liu, Y. Zhang, G. X. Liang, M. Gauthier, Y. C. K. Chen-Wiegart, M. N. Banis, X. F. Li, R. Y. Li, J. Wang, T. K. Sham and X. L. Sun, Nat. Commun., 2014, 5, 3415 Search PubMed.
  126. B. Kang and G. Ceder, Nature, 2009, 458, 190–193 CrossRef CAS PubMed.
  127. F. Yu, S. G. Ge, B. Li, G. Z. Sun, R. G. Mei and L. X. Zheng, Curr. Inorg. Chem., 2012, 2, 194–212 CrossRef CAS.
  128. F. Cheng, Z. Tao, J. Liang and J. Chen, Chem. Mater., 2008, 20, 667–681 CrossRef CAS.
  129. M. R. Hill, G. J. Wilson, L. Bourgeois and A. G. Pandolfo, Energy Environ. Sci., 2011, 4, 965–972 CAS.
  130. L. Dimesso, C. Forster, W. Jaegermann, J. P. Khanderi, H. Tempel, A. Popp, J. Engstler, J. J. Schneider, A. Sarapulova, D. Mikhailova, L. A. Schmitt, S. Oswald and H. Ehrenberg, Chem. Soc. Rev., 2012, 41, 5068–5080 RSC.
  131. X. D. Zhang, X. G. Zhang, W. He, C. Y. Sun, J. Y. Ma, J. L. Yuan and X. Y. Du, Colloids Surf., B, 2013, 103, 114–120 CrossRef CAS PubMed.
  132. Y. Ma, X. L. Li, S. F. Sun, X. P. Hao and Y. Z. Wu, Int. J. Electrochem. Sci., 2013, 8, 2842–2848 CAS.
  133. J. F. Qian, M. Zhou, Y. L. Cao, X. P. Ai and H. X. Yang, J. Phys. Chem. C, 2010, 114, 3477–3482 CAS.
  134. X. Qin, X. H. Wang, H. M. Xiang, J. Xie, J. J. Li and Y. C. Zhou, J. Phys. Chem. C, 2010, 114, 16806–16812 CAS.
  135. D. G. Tong, Y. L. Li, W. Chu, P. Wu and F. L. Luo, Dalton Trans., 2011, 40, 4087–4094 RSC.
  136. D. W. Han, W. H. Ryu, W. K. Kim, S. J. Lim, Y. I. Kim, J. Y. Eom and H. S. Kwon, ACS Appl. Mater. Interfaces, 2013, 5, 1342–1347 CAS.
  137. A. Carné-Sánchez, I. Imaz, M. Cano-Sarabia and D. Maspoch, Nat. Chem., 2013, 5, 203–211 CrossRef PubMed.
  138. F. Yu, J. J. Zhang, Y. F. Yang and G. Z. Song, J. Mater. Chem., 2009, 19, 9121–9125 RSC.
  139. S. Yu, S. Kim, T. Y. Kim, J. H. Nam and W. I. Cho, J. Appl. Electrochem., 2013, 43, 253–262 CrossRef CAS.
  140. F. Yu, J. J. Zhang, Y. F. Yang and G. Z. Song, J. Power Sources, 2009, 189, 794–797 CrossRef CAS.
  141. J. Liu, T. E. Conry, X. Y. Song, M. M. Doeff and T. J. Richardson, Energy Environ. Sci., 2011, 4, 885–888 CAS.
  142. B. F. Zou, Y. Q. Wang and S. M. Zhou, Mater. Lett., 2013, 92, 300–303 CrossRef CAS.
  143. W. Chang, S. J. Kim, I. T. Park, B. W. Cho, K. Y. Chung and H. C. Shin, J. Alloys Compd., 2013, 563, 249–253 CrossRef CAS.
  144. X. H. Rui, Y. Jin, X. Y. Feng, L. C. Zhang and C. H. Chen, J. Power Sources, 2011, 196, 2109–2114 CrossRef CAS.
  145. H. Q. Li and H. S. Zhou, Chem. Commun., 2012, 48, 1201–1217 RSC.
  146. D. Zane, M. Carewska, S. Scaccia, F. Cardellini and P. P. Prosini, Electrochim. Acta, 2004, 49, 4259–4271 CrossRef CAS.
  147. S. Xin, Y. G. Guo and L. J. Wan, Acc. Chem. Res., 2012, 45, 1759–1769 CrossRef CAS PubMed.
  148. J. J. Wang and X. L. Sun, Energy Environ. Sci., 2012, 5, 5163–5185 CAS.
  149. F. Y. Kang, J. Ma and B. H. Li, New Carbon Mater., 2011, 26, 161–170 CrossRef CAS.
  150. M. Inagaki, Carbon, 2012, 50, 3247–3266 CrossRef CAS.
  151. Y. Zhou, C. D. Gu, J. P. Zhou, L. J. Cheng, W. L. Liu, Y. Q. Qiao, X. L. Wang and J. P. Tu, Electrochim. Acta, 2011, 56, 5054–5059 CrossRef CAS.
  152. X. D. Yan, G. L. Yang, J. Liu, Y. C. Ge, H. M. Xie, X. M. Pan and R. S. Wang, Electrochim. Acta, 2009, 54, 5770–5774 CrossRef CAS.
  153. S. W. Oh, S. T. Myung, S. M. Oh, K. H. Oh, K. Amine, B. Scrosati and Y. K. Sun, Adv. Mater., 2010, 22, 4842–4845 CrossRef CAS PubMed.
  154. H. M. Xie, R. S. Wang, J. R. Ying, L. Y. Zhang, A. F. Jalbout, H. Y. Yu, G. L. Yang, X. M. Pan and Z. M. Su, Adv. Mater., 2006, 18, 2609–2613 CrossRef CAS.
  155. L. J. Zeng, Q. Gong, X. Z. Liao, L. He, Y. S. He and Z. F. Ma, Chin. Sci. Bull., 2011, 56, 1262–1266 CrossRef CAS.
  156. J. C. Zheng, X. H. Li, Z. X. Wang, J. H. Li, L. Wu, L. J. Li and H. J. Guo, Acta Phys.-Chim. Sin., 2009, 25, 1916–1920 CAS.
  157. G. Yang, C. Y. Jiang, X. M. He, J. R. Ying and J. Gao, Ionics, 2013, 19, 1247–1253 CrossRef CAS.
  158. M. R. Yang, W. H. Ke and S. H. Wu, J. Power Sources, 2007, 165, 646–650 CrossRef CAS.
  159. L. Wu, J. J. Lu and S. K. Zhong, J. Solid State Electrochem., 2013, 17, 2235–2241 CrossRef CAS.
  160. J. C. Zheng, X. H. Li, Z. X. Wang, D. M. Qin, H. J. Guo and W. J. Peng, J. Inorg. Mater., 2009, 24, 143–146 CrossRef CAS.
  161. Z. P. Ma, G. J. Shao, X. Wang, J. J. Song and G. L. Wang, Ionics, 2013, 19, 1861–1866 CrossRef CAS.
  162. X. Z. Liao, Z. F. Ma, Q. Gong, Y. S. He, L. Pei and L. J. Zeng, Electrochem. Commun., 2008, 10, 691–694 CrossRef CAS.
  163. S. S. Zhang, K. Xu and T. R. Jow, J. Power Sources, 2006, 159, 702–707 CrossRef CAS.
  164. A. Yamada, S. C. Chung and K. Hinokuma, J. Electrochem. Soc., 2001, 148, A224–A229 CrossRef CAS.
  165. Z. H. Li, D. M. Zhang and F. X. Yang, J. Mater. Sci., 2009, 44, 2435–2443 CrossRef CAS.
  166. J. R. Ying, M. Lei, C. Y. Jiang, C. R. Wan, X. M. He, J. J. Li, L. Wang and J. G. Ren, J. Power Sources, 2006, 158, 543–549 CrossRef CAS.
  167. J. F. Ni, H. H. Zhou, J. T. Chen and X. X. Zhang, Mater. Lett., 2007, 61, 1260–1264 CrossRef CAS.
  168. X. M. Lou and Y. X. Zhang, J. Mater. Chem., 2011, 21, 4156–4160 RSC.
  169. P. K. Shen, H. L. Zou, H. Meng and M. M. Wu, Funct. Mater. Lett., 2011, 4, 209–215 CrossRef CAS.
  170. Y. H. Yin, M. X. Gao, J. L. Ding, Y. F. Liu, L. K. Shen and H. G. Pan, J. Alloys Compd., 2011, 509, 10161–10166 CrossRef CAS.
  171. Y. H. Yin, M. X. Gao, H. G. Pan, L. K. Shen, X. Ye, Y. F. Liu, P. S. Fedkiw and X. W. Zhang, J. Power Sources, 2012, 199, 256–262 CrossRef CAS.
  172. Z. P. Ma, G. J. Shao, W. Wang and C. Y. Zhang, Asian J. Chem., 2013, 25, 3579–3582 CrossRef CAS.
  173. S. W. Oh, S. T. Myung, S. M. Oh, C. S. Yoon, K. Amine and Y. K. Sun, Electrochim. Acta, 2010, 55, 1193–1199 CrossRef CAS.
  174. Z. R. Chang, H. J. Lv, H. W. Tang, H. J. Li, X. Z. Yuan and H. J. Wang, Electrochim. Acta, 2009, 54, 4595–4599 CrossRef CAS.
  175. S. W. Oh, S. T. Myung, H. J. Bang, C. S. Yoon, K. Amine and Y. K. Sun, Electrochem. Solid-State Lett., 2009, 12, A181–A185 CrossRef CAS.
  176. J. K. Kim, CrystEngComm, 2014, 16, 2818–2822 RSC.
  177. F. Yu, L. L. Zhang, M. Y. Zhu, Y. X. An, L. L. Xia, X. G. Wang and B. Dai, Nano Energy, 2014, 3, 64–79 CrossRef CAS.
  178. J. M. Tarascon, N. Recham, M. Armand, J. N. Chotard, P. Barpanda, W. Walker and L. Dupont, Chem. Mater., 2010, 22, 724–739 CrossRef CAS.
  179. O. Toprakci, H. A. K. Toprakci, L. W. Ji and X. W. Zhang, KONA Powder Part. J., 2010, 50–73 CrossRef CAS.
  180. S. M. Oh, H. G. Jung, C. S. Yoon, S. T. Myung, Z. H. Chen, K. Amine and Y. K. Sun, J. Power Sources, 2011, 196, 6924–6928 CrossRef CAS.
  181. J. Wang, Z. B. Shao and H. Q. Ru, Ceram. Int., 2014, 40, 6979–6985 CrossRef CAS.
  182. J. J. Tan and A. Tiwari, Nanosci. Nanotechnol. Lett., 2011, 3, 487–490 CrossRef CAS.
  183. Y. Y. Liu, D. W. Liu, Q. F. Zhang and G. Z. Cao, J. Mater. Chem., 2011, 21, 9969–9983 RSC.
  184. D. Jugovic, M. Mitric, M. Kuzmanovic, N. Cvjeticanin, S. Skapin, B. Cekic, V. Ivanovski and D. Uskokovic, J. Power Sources, 2011, 196, 4613–4618 CrossRef CAS.
  185. C. Miao, J. Zhou, S. Q. Sun and X. Y. Wang, Rare Met. Mater. Eng., 2013, 42, 379–382 Search PubMed.
  186. Y. Zhang, Q. Y. Huo, P. P. Du, L. Z. Wang, A. Q. Zhang, Y. H. Song, Y. Lv and G. Y. Li, Synth. Met., 2012, 162, 1315–1326 CrossRef CAS.
  187. F. Yu, J. J. Zhang, Y. F. Yang and G. Z. Song, Chin. J. Inorg. Chem., 2009, 25, 42–46 CAS.
  188. J. P. Jegal and K. B. Kim, J. Power Sources, 2013, 243, 859–864 CrossRef CAS.
  189. J. L. Yang, J. J. Wang, Y. J. Tang, D. N. Wang, X. F. Li, Y. H. Hu, R. Y. Li, G. X. Liang, T. K. Sham and X. L. Sun, Energy Environ. Sci., 2013, 6, 1521–1528 CAS.
  190. P. P. Prosini, M. Carewska, S. Scaccia, P. Wisniewski and M. Pasquali, Electrochim. Acta, 2003, 48, 4205–4211 CrossRef CAS.
  191. P. Jozwiak, J. E. Garbarczyk, M. Wasiucionek, I. Gorzkowska, F. Gendron, A. Mauger and C. M. Julien, Solid State Ionics, 2008, 179, 46–50 CrossRef CAS.
  192. L. Yang, L. F. Jiao, Y. L. Miao and H. T. Yuan, J. Solid State Electrochem., 2009, 13, 1541–1544 CrossRef CAS.
  193. J. Liu, J. W. Wang, X. D. Yan, X. F. Zhang, G. L. Yang, A. F. Jalbout and R. S. Wang, Electrochim. Acta, 2009, 54, 5656–5659 CrossRef CAS.
  194. C. W. Kim, M. H. Lee, W. T. Jeong and K. S. Lee, J. Power Sources, 2005, 146, 534–538 CrossRef CAS.
  195. C. W. Kim, J. S. Park and K. S. Lee, J. Power Sources, 2006, 163, 144–150 CrossRef CAS.
  196. K. T. Lee and K. S. Lee, J. Power Sources, 2009, 189, 435–439 CrossRef CAS.
  197. Y. B. Xu, Y. J. Lu, L. Yan, Z. Y. Yang and R. D. Yang, J. Power Sources, 2006, 160, 570–576 CrossRef CAS.
  198. S. B. Lee, S. H. Cho, S. J. Cho, G. J. Park, S. H. Park and Y. S. Lee, Electrochem. Commun., 2008, 10, 1219–1221 CrossRef CAS.
  199. M. Hermanek, R. Zboril, M. Mashlan, L. Machala and O. Schneeweiss, J. Mater. Chem., 2006, 16, 1273–1280 RSC.
  200. W. Ojczyk, J. Marzec, K. Swierczek, W. Zajac, M. Molenda, R. Dziembaj and J. Molenda, J. Power Sources, 2007, 173, 700–706 CrossRef CAS.
  201. H. Liu, J. Y. Xie and K. Wang, Solid State Ionics, 2008, 179, 1768–1771 CrossRef CAS.
  202. H. W. Liu and D. G. Tang, Solid State Ionics, 2008, 179, 1897–1901 CrossRef CAS.
  203. Y. Lin, J. B. Wu and Y. B. Yu, in Multi-Functional Materials and Structures Iii, Pts 1 and 2, Trans Tech Publications Ltd, Stafa-Zurich, 2010, vol. 123–125, pp. 1227–1230 Search PubMed.
  204. M. X. Gao, Y. Lin, Y. H. Yin, Y. F. Liu and H. G. Pan, Electrochim. Acta, 2010, 55, 8043–8050 CrossRef CAS.
  205. M. Konarova and I. Taniguchi, Mater. Res. Bull., 2008, 43, 3305–3317 CrossRef CAS.
  206. M. Konarova and I. Taniguchi, J. Power Sources, 2009, 194, 1029–1035 CrossRef CAS.
  207. Y. Lin, M. X. Gao, D. Zhu, Y. F. Liu and H. G. Pan, J. Power Sources, 2008, 184, 444–448 CrossRef CAS.
  208. G. Arnold, J. Garche, R. Hemmer, S. Strobele, C. Vogler and A. Wohlfahrt-Mehrens, J. Power Sources, 2003, 119, 247–251 CrossRef.
  209. Y. Z. Dong, Y. M. Zhao, H. Duan, L. Chen, Z. F. He and Y. H. Chen, J. Solid State Electrochem., 2010, 14, 131–137 CrossRef CAS.
  210. S. Balaji, D. Mutharasu, N. Sankara Subramanian and K. Ramanathan, Ionics, 2009, 15, 765–777 CrossRef CAS.
  211. K. Cherian, Adv. Mater. Processes, 2011, 169, 23–27 CAS.
  212. X. F. Guo, H. Zhan and Y. H. Zhou, Solid State Ionics, 2009, 180, 386–391 CrossRef CAS.
  213. M. Higuchi, K. Katayama, Y. Azuma, M. Yukawa and M. Suhara, J. Power Sources, 2003, 119, 258–261 CrossRef.
  214. M. S. Song, Y. M. Kang, J. H. Kim, H. S. Kim, D. Y. Kim, H. S. Kwon and J. Y. Lee, J. Power Sources, 2007, 166, 260–265 CrossRef CAS.
  215. M. S. Song, Y. M. Kang, Y. I. Kim, K. S. Park and H. S. Kwon, Inorg. Chem., 2009, 48, 8271–8275 CrossRef CAS PubMed.
  216. K. S. Park, J. T. Son, H. T. Chung, S. J. Kim, C. H. Lee and H. G. Kim, Electrochem. Commun., 2003, 5, 839–842 CrossRef CAS.
  217. H. L. Zou, G. H. Zhang and P. K. Shen, Mater. Res. Bull., 2010, 45, 149–152 CrossRef CAS.
  218. L. Wang, Y. D. Huang, R. R. Jiang and D. Z. Jia, Electrochim. Acta, 2007, 52, 6778–6783 CrossRef CAS.
  219. L. Wang, Y. D. Huang, R. R. Jiang and D. Z. Jia, J. Electrochem. Soc., 2007, 154, A1015–A1019 CrossRef CAS.
  220. D. Li, Y. D. Huang, N. Sharma, Z. X. Chen, D. Z. Jia and Z. P. Guo, Phys. Chem. Chem. Phys., 2012, 14, 3634–3639 RSC.
  221. D. Li, Y. D. Huang, D. Z. Jia, Z. P. Guo and S. J. Bao, J. Solid State Electrochem., 2010, 14, 889–895 CrossRef CAS.
  222. Y. Zhang, H. Feng, X. B. Wu, L. Z. Wang, A. Q. Zhang, T. C. Xia, H. C. Dong and M. H. Liu, Electrochim. Acta, 2009, 54, 3206–3210 CrossRef CAS.
  223. X. F. Guo, Y. T. Zhang, H. Zhan and Y. H. Zhou, Solid State Ionics, 2010, 181, 1757–1763 CrossRef CAS.
  224. S. Feng and L. Guanghua, in Modern Inorganic Synthetic Chemistry, ed. X. Ruren, P. Wenqin, W. P. Qisheng HuoA2-Ruren Xu and H. Qisheng, Elsevier, Amsterdam, 2011, pp. 63–95 Search PubMed.
  225. X. Q. Ou, S. Z. Xu, G. C. Liang, L. Wang and X. Zhao, Sci. China, Ser. E: Technol. Sci., 2009, 52, 264–268 CrossRef CAS.
  226. M. K. Devaraju and I. Honma, Adv. Energy Mater., 2012, 2, 284–297 CrossRef CAS.
  227. C. B. Xu, J. Lee and A. S. Teja, J. Supercrit. Fluids, 2008, 44, 92–97 CrossRef CAS.
  228. A. Aimable, D. Aymes, F. Bernard and F. Le Cras, Solid State Ionics, 2009, 180, 861–866 CrossRef CAS.
  229. N. Recham, L. Dupont, M. Courty, K. Djellab, D. Larcher, M. Armand and J. M. Tarascon, Chem. Mater., 2009, 21, 1096–1107 CrossRef CAS.
  230. N. Recham, J. N. Chotard, L. Dupont, K. Djellab, M. Armand and J. M. Tarascon, J. Electrochem. Soc., 2009, 156, A993–A999 CrossRef CAS.
  231. P. Barpanda, N. Recham, J. N. Chotard, K. Djellab, W. Walker, M. Armand and J. M. Tarascon, J. Mater. Chem., 2010, 20, 1659–1668 RSC.
  232. M. C. Tucker, M. M. Doeff, T. J. Richardson, R. Finones, J. A. Reimer and E. J. Cairns, Electrochem. Solid-State Lett., 2002, 5, A95–A98 CrossRef CAS.
  233. K. Shiraishi, K. Dokko and K. Kanamura, J. Power Sources, 2005, 146, 555–558 CrossRef CAS.
  234. K. Dokko, S. Koizumi, K. Sharaishi and K. Kanamura, J. Power Sources, 2007, 165, 656–659 CrossRef CAS.
  235. K. Kanamura, S. H. Koizumi and K. R. Dokko, J. Mater. Sci., 2008, 43, 2138–2142 CrossRef CAS.
  236. K. Dokko, S. Koizumi and K. Kanamura, Chem. Lett., 2006, 35, 338–339 CrossRef CAS.
  237. S. F. Yang, P. Y. Zavalij and M. S. Whittingham, Electrochem. Commun., 2001, 3, 505–508 CrossRef CAS.
  238. M. Koltypin, D. Aurbach, L. Nazar and B. Ellis, J. Power Sources, 2007, 174, 1241–1250 CrossRef CAS.
  239. S. Franger, F. Le Cras, C. Bourbon and H. Rouault, Electrochem. Solid-State Lett., 2002, 5, A231–A233 CrossRef CAS.
  240. J. L. Liu, R. R. Jiang, X. Y. Wang, T. Huang and A. S. Yu, J. Power Sources, 2009, 194, 536–540 CrossRef CAS.
  241. J. Chen, S. Wang and M. S. Whittingham, J. Power Sources, 2007, 174, 442–448 CrossRef CAS.
  242. J. J. Chen and M. S. Whittingham, Electrochem. Commun., 2006, 8, 855–858 CrossRef CAS.
  243. X. Q. Ou, H. C. Gu, Y. C. Wu, J. W. Lu and Y. J. Zheng, Electrochim. Acta, 2013, 96, 230–236 CrossRef CAS.
  244. R. G. Mei, X. R. Song, Y. F. Yang, Z. G. An and J. J. Zhang, RSC Adv., 2014, 4, 5746–5752 RSC.
  245. J. Lee and A. S. Teja, J. Supercrit. Fluids, 2005, 35, 83–90 CrossRef CAS.
  246. K. Dokko, S. Koizumi, H. Nakano and K. Kanamura, J. Mater. Chem., 2007, 17, 4803–4810 RSC.
  247. J. J. Chen, M. J. Vacchio, S. J. Wang, N. Chernova, P. Y. Zavalij and M. S. Whittingham, Solid State Ionics, 2008, 178, 1676–1693 CrossRef CAS.
  248. Y. Xia, W. K. Zhang, H. Huang, Y. P. Gan, J. Tian and X. Y. Tao, J. Power Sources, 2011, 196, 5651–5658 CrossRef CAS.
  249. J. F. Ni, M. Morishita, Y. Kawabe, M. Watada, N. Takeichi and T. Sakai, J. Power Sources, 2010, 195, 2877–2882 CrossRef CAS.
  250. Z. G. Lu, H. L. Chen, R. Robert, B. Y. X. Zhu, J. Q. Deng, L. J. Wu, C. Y. Chung and C. P. Grey, Chem. Mater., 2011, 23, 2848–2859 CrossRef CAS.
  251. G. Meligrana, C. Gerbaldi, A. Tuel, S. Bodoardo and N. Penazzi, J. Power Sources, 2006, 160, 516–522 CrossRef CAS.
  252. C. Gerbaldi, J. Nair, C. B. Minella, G. Meligrana, G. Mulas, S. Bodoardo, R. Bongiovanni and N. Penazzi, J. Appl. Electrochem., 2008, 38, 985–992 CrossRef CAS.
  253. S. Bodoardo, C. Gerbaldi, G. Meligrana, A. Tuel, S. Enzo and N. Penazzi, Ionics, 2009, 15, 19–26 CrossRef CAS.
  254. G. X. Wang, X. P. Shen and J. Yao, J. Power Sources, 2009, 189, 543–546 CrossRef CAS.
  255. Y. M. Jiang, S. J. Liao, Z. S. Liu, G. Xiao, Q. B. Liu and H. Y. Song, J. Mater. Chem. A, 2013, 1, 4546–4551 CAS.
  256. S. Tajimi, Y. Ikeda, K. Uematsu, K. Toda and M. Sato, Solid State Ionics, 2004, 175, 287–290 CrossRef CAS.
  257. Z. H. Xu, L. Xu, Q. Y. Lai and X. Y. Ji, Mater. Res. Bull., 2007, 42, 883–891 CrossRef CAS.
  258. F. Teng, M. D. Chen, G. Q. Li, Y. Teng, T. G. Xu, S. I. Mho and X. Hua, J. Power Sources, 2012, 202, 384–388 CrossRef CAS.
  259. G. Yang, H. M. Ji, H. D. Liu, K. F. Huo, J. J. Fu and P. K. Chu, J. Nanosci. Nanotechnol., 2010, 10, 980–986 CrossRef CAS PubMed.
  260. G. Yang, H. M. Ji, X. W. Miao, A. Q. Hong and Y. Yan, J. Nanosci. Nanotechnol., 2011, 11, 4781–4792 CrossRef CAS PubMed.
  261. D. Carriazo, M. D. Rossell, G. B. Zeng, I. Bilecka, R. Erni and M. Niederberger, Small, 2012, 8, 2231–2238 CrossRef CAS PubMed.
  262. A. V. Murugan, T. Muraliganth and A. Manthiram, J. Phys. Chem. C, 2008, 112, 14665–14671 CAS.
  263. A. V. Murugan, T. Muraliganth and A. Manthiram, J. Electrochem. Soc., 2009, 156, A79–A83 CrossRef.
  264. T. Muraliganth, A. V. Murugan and A. Manthiram, J. Mater. Chem., 2008, 18, 5661–5668 RSC.
  265. A. V. Murugan, T. Muraliganth, P. J. Ferreira and A. Manthiram, Inorg. Chem., 2009, 48, 946–952 CrossRef CAS PubMed.
  266. A. V. Murugan, T. Muraliganth and A. Manthiram, Electrochem. Commun., 2008, 10, 903–906 CrossRef.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.