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
First published on 1st October 2014
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.
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.
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
![]() | ||
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
![]() | ||
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.
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 a–c 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.
![]() | ||
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.
![]() | ||
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.) |
![]() | ||
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
![]() | ||
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.
![]() | ||
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 a–c 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 a–c 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 a–c 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.
![]() | ||
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.
![]() | ||
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:
15
:
5 to 30
:
65
:
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.
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.
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%).
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:
1
:
1
:
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
:
1
:
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.
Materials | Synthesis method | Low temperature range/°C | Electrolyte | Ref. |
---|---|---|---|---|
LiFePO4/C | Solid state reaction method | −20 | 1 M LiPF6/EC + DEC (1![]() ![]() |
144 |
LiFePO4/C | Solid state reaction method | −20 | 1 M LiPF6/EC + DEC (1![]() ![]() |
151 |
LiFePO4/C | Solid state reaction method | −40 | 1 M LiPF6/EC + DEC (1![]() ![]() |
152 |
LiFePO4/C | Co-precipitation method | −20 | 1 M LiPF6/EC + DEC (1![]() ![]() |
153 |
LiFePO4/PAS | Co-precipitation method | −20 | 1 M LiPF6/EC + DEC (1![]() ![]() |
154 |
LiFe0.98Mn0.02PO4/C | Solid state reaction method | −40 | 1 M LiPF6/EC + DEC (1![]() ![]() |
155 |
LiFePO4/C | Solid state reaction method | −40 | 1 M LiPF6/EC + DEC + DMC + EMC (1![]() ![]() ![]() ![]() ![]() ![]() |
162 |
LiFePO4/C | Solid state reaction method | −50 | 1 M (0.9LiBF4 + 0.1LiBOB)/PC + EC + EMC (1![]() ![]() ![]() ![]() |
163 |
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
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 |
![]() | ||
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.) |
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.
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.
![]() | ||
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.) |
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
Synthesis type | Reaction reagents | Stoichiometric ratio | Mediator | Solvent | Condition | Morphology | Ref. |
---|---|---|---|---|---|---|---|
HTS | LiOH, FeSO4, H3PO4 | 3![]() ![]() ![]() ![]() |
— | H2O | 120 °C, 5 h | Hexagonal platelets | 232 |
HTS | LiOH, FeSO4, (NH4)3PO4 | 2.5![]() ![]() ![]() ![]() |
— | H2O | 170 °C, 12 h | — | 233 |
HTS | LiOH, FeSO4, (NH4)2HPO4 | 1![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
— | H2O | 170 °C, 12 h | Nanoplates | 234 |
HTS | LiOH, FeSO4, NH4H2PO4 | 2![]() ![]() ![]() ![]() |
— | H2O | 180 °C, 5 h | Spindle-like particles | This work |
HTS | LiOH, FeCl2·4H2O, P2O5 | 6![]() ![]() ![]() ![]() |
— | H2O | 170 °C, 3 days | Hexagonal platelets | 237 |
HTS | LiOH, FeSO4, H3PO4 | 3![]() ![]() ![]() ![]() |
CTAB | H2O | 120 °C, 5 h | Nanoparticles | 251–253 |
HTS | LiOH, (NH4)2Fe(SO4)2·6H2O, H3PO4 | 3![]() ![]() ![]() ![]() |
P123, D230 | H2O | 220 °C, 24 h | Nanoparticles | 106 |
HTS | LiOH, (NH4)2Fe(SO4)2·6H2O, H3PO4 | 3![]() ![]() ![]() ![]() |
Citric acid | H2O | 180 °C, 10 h | Spindle-like particles | 250 |
HTS | LiOH, FeSO4, NH4H2PO4 | 2![]() ![]() ![]() ![]() |
NTA | H2O | 180 °C, 20 h | Nanowires | 254 |
HTS | LiAc, FeCl3, NH4H2PO4 | 1![]() ![]() ![]() ![]() |
SDS + EN | H2O | 180 °C, 10 h | Spheres assembled from nanorods | This work |
STS | LiOH, FeSO4, H3PO4 | 3![]() ![]() ![]() ![]() |
— | TEG + H2O | 190 °C, 5 h | Rectangular nanoplatelets | 111 |
STS | LiOH, FeSO4, H3PO4 | 3![]() ![]() ![]() ![]() |
Citric acid | MPG + H2O | 140 °C, 12 h | Needles assembled nanorods | 255 |
STS | LiOH, FeSO4, H3PO4 | 3![]() ![]() ![]() ![]() |
— | PEG400 + H2O | 150 °C, 3 h | Needle-like particles | 256 |
STS | LiOH, FeSO4, H3PO4 | 3![]() ![]() ![]() ![]() |
— | PEG400 + H2O | 180 °C, 9 h | Hexagonal platelet | 112 |
ITS | LiH2PO4, FeC2O4·2H2O | 1![]() ![]() |
— | CN-based IL | 250 °C, 24 h | Needles assembled Lego blocks | 229 |
ITS | LiOH, FeSO4, H3PO4 | 3![]() ![]() ![]() ![]() |
SDBS + ascorbic acid | IL + H2O | 240 °C, 20 h | Nanorods | 258 |
MW-HTS | LiOH, FeSO4, H3PO4 | 3![]() ![]() ![]() ![]() |
— | H2O | 200 °C, 5 min | Globular particles | 259 |
MW-STS | LiOH, FeAc2, H3PO4 | 1![]() ![]() ![]() ![]() |
— | TEG | 300 °C, 5 min | Nanorods | 262 |
![]() | ||
Fig. 14 XRD patterns of the samples prepared using different values of x in a molar ratio of x![]() ![]() ![]() ![]() |
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
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.
![]() | ||
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.
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 a–c 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.
This journal is © The Royal Society of Chemistry 2014 |