H.-N. Girish
and
G.-Q. Shao
*
State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China. E-mail: gqshao@whut.edu.cn; Fax: +86 27 87210216; Tel: +86 27 87210216
First published on 30th October 2015
The orthosilicate, Li2MSiO4 (M = Mn, Fe, Co, Ni, …), is a competitive cathode material for next generation high-capacity (∼330 mA h g−1 with a possible 2 Li+ extraction per formula unit) rechargeable lithium-ion batteries. It is an alternative to other polyanionic compounds because the raw materials of Fe/Mn-containing oxidates and silica are abundant, cost effective, very safe, environmentally friendly, and its Si–O bond is at least as stable as other polyoxyanion groups with an excellent cycling performance. This review highlights the research progress related to the Li2MSiO4 cathode materials in terms of the phase transition of the structure, synthetic methods and techniques for improving the electrochemical performance.
Calculations of the electrochemical potential of lithium insertion or extraction show two processes:
Li2M2+SiO4 → LiM3+SiO4 + Li+ | (1) |
LiM3+SiO4 → M4+SiO4 + Li+ | (2) |
Depending on the TM type, as shown in Fig. 1, the potential in reaction (1) (green square) varies from 3.1 V (for Fe) to 4.1–4.6 V (for Mn, Co and Ni), while that in reaction (2) (red circle) is in the range 4.5–5.1 V. The extraction of the 2nd Li+ requires the application of a high potential above 4.5 V which is a challenge for electrolytes.
![]() | ||
Fig. 1 Variation of Li insertion voltage by M2+/M3+ and M3+/M4+ in Li2MSiO4.18 |
Li2MnSiO4 has more advantages than Li2FeSiO4 with regards to cell safety, ease of preparation and cost effectiveness. The Mn3+ ↔ Mn4+ conversion provides a higher potential than Mn2+/Mn3+ and Fe2+/Fe3+ (it is yet to be clarified whether this involved an Fe3+/Fe4+ redox couple which might just be an electrolyte degradation18,38,40,41). During oxidation, the short Mn–O bonds in the MnO4 tetrahedron decrease in correlation with the increase in the Mn valance state, while the long Mn–O bonds remain unchanged.42 In practical applications, Li2MnSiO4 is limited as a cathode by its low electronic conductivity of ∼5 × 10−16 S cm−1 at room temperature (RT) (and ∼3 × 10−14 S cm−1 at 60 °C), which is 5–6 orders of magnitude smaller than that of LiFePO4 at RT. The preparation of pure Li2MnSiO4 even at temperatures up to 800 °C is not trivial due to the possible presence of mixed phases and impurities such as MnO, MnSiO3, Li2SiO3, etc.43–46
Co- or Ni-containing salts for preparing Li2CoSiO4 and Li2NiSiO4 are commercially expensive and environmentally toxic. Li2NiSiO4, with the lowest band-gap, does not have the shortcomings of Mn derivatives18 but the delithiation potentials (>4.6 V) of Li2CoSiO4 and Li2NiSiO4 are too high to make the present electrolytes suitable Li+ diffusion media. In the same way, their kinetic properties make preparation difficult and cause poor delithiation–lithiation during the charge–discharge process.18,36,47–49
Otherwise, defects such as poor reversibility that are associated with Li2MSiO4 have become a huge obstacle to their extensive application in electric vehicles (EV) and plug-in hybrid electric vehicles (PHEV).50–53 Various methods have been developed to improve their electrochemical properties, such as controlling the particle size and morphology (e.g., nanosheet/rod-like particles),45,54–57 cation substitution (Mg2+, Al3+, Cr2+, etc.)56,58–65 and surface-coating with conductive materials.66–80 For surface-coating methods, an appropriate carbon coating has been widely used,81–84 because it not only improves the electronic conductivity but also reduces the particle size, which are favourable for enhancing the electrochemical properties.
This review will explain the polymorphic structure of Li2MSiO4 and its phase transitions and summarize progress in its research, methods for its synthesis and techniques for improving its electrochemical performance.
Li3PO4 crystallizes in the “α-phase formed in low temperature”, “β-phase in high temperature” and “γ-phase above 1170 °C”.45,85 Several temperature/pressure-dependent polymorphs of Li2MSiO4, namely high temperature monoclinic γo/γs (P21/n) and orthorhombic γII (Pmna, Pmnb, Cmma, γ-Li3PO4), and low temperature orthorhombic βI (Pbn21, Pna21, β-Li3PO4)/βII (Pmn21, β-Li3PO4), can be stabilised/influenced significantly by the synthetic conditions and also the electrochemical cyclic behavior.86 These phases may be quenched to RT and exhibit long-term stability. Metastable monoclinic (Pn) Li2MnSiO4 prepared by ion-exchange from Na2MnSiO4 was also reported.83 Different phase structures of Li2MSiO4 polymorphs are shown in Fig. 2 and discussed below.38,45,83,86–92
![]() | ||
Fig. 2 Structures of Li2MSiO4: (a) orthorhombic βII (Pmn21); (b) orthorhombic β1 (Pbn21, Pna21); (c) orthorhombic γII (Pmna, Pmnb, Cmma); (d) monoclinic γs (P21/n, P21); (e) monoclinic γ0 (P21/n); (f) monoclinic Pn.38,45,83,86–92 SiO4 (yellow); MO4 (brown/green); LiO4 (blue/orange); light and dark blue tetrahedra represent crystallographically distinct Li sites. |
(a) Orthorhombic βII structure (Pmn21) and modified-/inverse-βII structure (Pmn21-I): all the tetrahedra point in the same direction, perpendicular to the close-packed planes, and share only corners with each other. Chains of LiO4 are along the a-axis and parallel to chains of alternating MO4 and SiO4 (Fig. 2a).45
In the modified-/inverse-βII structure (e.g. the cycled polymorph of Li2FeSiO4), all tetrahedra point in the same direction along the c-axis/b-axis and are linked only by corner-sharing. The SiO4 tetrahedra are isolated from each other, sharing corners with LiO4 and (Li/Fe)O4 tetrahedra.86,90,91 This the more stable configuration at low temperature.
(b) Orthorhombic βI structure (Pbn21, Pna21): all tetrahedra point in same direction with chains of alternating LiO4 and MO4 tetrahedra along the a-axis, parallel to chains of alternating LiO4 and SiO4 tetrahedra (Fig. 2b).38,87–90,92
Quoirin et al. prepared an orthorhombic Li2FeSiO4 (Pna21, βI) by precipitation at 100 °C in air which was optionally in the space group of Pccn. It might be an intermediate between the orthorhombic Pmn21 (βII) and monoclinic P21/n (γs) phases.87–89
(c) Orthorhombic γII structure (Pmna, Pmnb, Cmma): the tetrahedra are arranged in groups of three with the central tetrahedron pointing in the opposite direction to the outer two, with which it shares edges; the group of 3 edge-sharing tetrahedra consist of the sequence Li–M–Li (Fig. 2c and the inset).45
(d) Monoclinic γs structure (P21/n, previously designated P21 (ref. 93–95)): half of the tetrahedra point in reverse directions, containing pairs of LiO4/MO4 and LiO4/LiO4 edge-sharing tetrahedra (Fig. 2d and the inset).45
(e) Monoclinic γ0 structure (P21/n): similar to the γII structure, the group of 3 edge-sharing tetrahedra consist of the sequence Li–Li–M (Fig. 2e).45
(f) Monoclinic structure (Pn): similar to the orthorhombic βII structure (Pmn21), it is a metastable form of Li2MnSiO4 and can convert to βII above 370 °C. Here the MnO4 and SiO4 tetrahedra point in the same direction parallel to the c-axis (Fig. 2f).83
Fig. 3 shows a schematic diagram of the temperature/pressure-dependent polymorphs of Li2MSiO4 with phase transitions. Table 1 shows the reported lattice parameters of Li2MSiO4 polymorphs (M = Fe, Mn, Co, Ni).
![]() | ||
Fig. 3 Schematic diagram of temperature/pressure-dependent polymorphs of Li2MSiO4 with phase transitions. |
Li2MSiO4 | x | Space group | Crystal system | a/Å | b/Å | c/Å | α/° | β/° | γ/° | Refs. |
---|---|---|---|---|---|---|---|---|---|---|
Li2FeSiO4 | — | Pbn21 (βI) | Orthorhombic | 6.2652(1) | 10.8121(1) | 4.9504(1) | 90 | 90 | 90 | 88 and 89 |
— | Pna21 (βI) | Orthorhombic | 10.724(5) | 6.260(7) | 4.958(6) | 90 | 90 | 90 | 87–89 | |
— | Pmn21 (βII) | Orthorhombic | 6.2709(8) | 5.3382(7) | 4.9651(7) | 90 | 90 | 90 | 88 and 89 | |
— | Pmn21 (βII) | Orthorhombic | 6.2661 | 5.3295 | 5.0148 | 90 | 90 | 90 | 23 | |
— | Pmn21 (βII) | Orthorhombic | 6.271(1) | 5.336(1) | 4.9607(9) | 90 | 90 | 90 | 37 | |
— | Pmn21 (βII) | Orthorhombic | 6.3246 | 5.3817 | 4.9967 | 90 | 90 | 90 | 18 | |
— | Pmn21 (βII) | Orthorhombic | 6.313 | 5.393 | 4.979 | 90 | 90 | 90 | 39 | |
— | Pmnb (γII) | Orthorhombic | 6.2853(5) | 10.6592(8) | 5.0367(4) | 90 | 90 | 90 | 103 | |
— | Pnma (γII) | Orthorhombic | 10.656 | 6.285 | 5.038 | 90 | 90 | 90 | 88 | |
— | Cmma (γII) | Orthorhombic | 10.664 | 12.540 | 5.022 | 90 | 90 | 90 | 88 | |
— | P21/n (γs) | Monoclinic | 8.2253(5) | 5.0220(1) | 8.2381(4) | 90 | 99.230 | 90 | 93 | |
— | P21/n (γs) | Monoclinic | 6.2835(7) | 10.6572(1) | 5.0386(5) | 90 | 89.941 | 90 | 103 | |
— | P21 (γs) | Monoclinic | 8.2278(1) | 5.0204(6) | 8.2295(1) | 90 | 99.231 | 90 | 95 | |
Li2MnSiO4 | — | Pmn21 (βII) | Orthorhombic | 6.3109(9) | 5.3800(9) | 4.9662(8) | 90 | 90 | 90 | 37 |
— | Pmn21 (βII) | Orthorhombic | 6.308(3) | 5.377(3) | 4.988(9) | 90 | 90 | 90 | 107 | |
— | Pmn21 (βII) | Orthorhombic | 6.3666 | 5.4329 | 4.0368 | 90 | 90 | 90 | 18 | |
— | Pmn21 (βII) | Orthorhombic | 6.308(9) | 5.385(9) | 4.999(6) | 90 | 90 | 90 | 108 | |
— | Pmn21 (βII) | Orthorhombic | 6.3064 | 5.3893 | 4.9715 | 90 | 90 | 90 | 109 | |
— | Pmn21 (βII) | Orthorhombic | 6.3141 | 5.3702 | 4.9650 | 90 | 90 | 90 | 44 | |
— | Pmnb (γII) | Orthorhombic | 6.3126 | 10.7657 | 5.0118 | 90 | 90 | 90 | 44 | |
— | P21/n (γs) | Monoclinic | 6.3368(1) | 10.9146(2) | 5.0730(1) | 90 | 90.98 | 90 | 106 | |
— | P21/n (γs) | Monoclinic | 6.3363(5) | 10.8950(7) | 5.07506(3) | 90 | 90.99 | 90 | 44 | |
— | Pn (metastable) | Monoclinic | 6.593(5) | 5.402(1) | 5.090(2) | 90 | 89.7(3) | 90 | 83 | |
Li2CoSiO4 | — | Pbn21 (βI) | Orthorhombic | 6.253 | 10.685 | 4.929 | 90 | 90 | 90 | 92 |
— | Pmn21 (βII) | Orthorhombic | 6.2599 | 10.6892 | 4.9287 | 90 | 90 | 90 | 85 | |
— | Pmn21 (βII) | Orthorhombic | 6.159 | 5.440 | 4.988 | 90 | 90 | 90 | 110 | |
— | Pmn21 (βII) | Orthorhombic | 6.287(6) | 5.353(1) | 4.937(1) | 90 | 90 | 90 | 47 | |
— | Pmn21 (βII) | Orthorhombic | 6.267(9) | 5.370(8) | 4.939(4) | 90 | 90 | 90 | 47 | |
— | Pmn21 (βII) | Orthorhombic | 6.2015 | 5.4378 | 4.9909 | 90 | 90 | 90 | 18 | |
— | Pmnb (γII) | Orthorhombic | 6.20 | 10.72 | 5.03 | 90 | 90 | 90 | 85 | |
— | P21/n (γo) | Monoclinic | 6.284 | 10.686 | 5.018 | 90 | 90.60 | 90 | 43 | |
Li2NiSiO4 | — | Pmn21 (βII) | Orthorhombic | 6.2938 | 5.3702 | 4.9137 | 90 | 90 | 90 | 18 |
— | Pmn21 (βII) | Orthorhombic | 6.274 | 5.365 | 4.921 | 90 | 90 | 90 | 111 | |
— | Pmn21 (βII) | Orthorhombic | 6.290 | 5.299 | 4.856 | 90 | 90 | 90 | 112 | |
Li2(Fe1−xMnx)SiO4 (x = 0, 0.3, 0.5, 0.7, 1) | x = 0 | Pmn21 (βII) | Orthorhombic | 6.2512 | 5.3425 | 5.0188 | 90 | 90 | 90 | 113 |
x = 0.3 | 6.2117 | 5.381 | 4.975 | |||||||
x = 0.5 | 6.2581 | 5.3747 | 5.006 | |||||||
x = 0.7 | 6.2851 | 5.381 | 5.0012 | |||||||
x = 1 | 6.3099 | 5.394 | 5.0192 | |||||||
Li2(Mn1−xFex)SiO4/C (x = 0, 0.2, 0.5, 0.8) | x = 0 | Pmn21 (βII) | Orthorhombic | — | — | — | 90 | 90 | 90 | 109 |
x = 0.2 | — | — | — | |||||||
x = 0.5 | — | — | — | |||||||
x = 0.8 | — | — | — | |||||||
Li2(Fe1−xMnx)SiO4 (x = 0, 0.2, 0.5, 1) | x = 0 | P21/n (γs) | Monoclinic | 8.244(8) | 5.004(4) | 8.244(8) | 90 | 99.25 | 90 | 102 |
x = 0.2 | P21/n (γs) | Monoclinic | 8.264(7) | 5.020(7) | 8.264(4) | 90 | 99.00 | 90 | ||
Pmnb (γII) | Orthorhombic | 6.591(6) | 9.462(7) | 4.988(1) | 90 | 90 | 90 | |||
x = 0.5 | Pmn21 (βII) | Orthorhombic | 6.284(0) | 5.391(9) | 5.015(1) | 90 | 90 | 90 | ||
P21/n (γs) | Monoclinic | 6.986(8) | 8.030(9) | 5.990(4) | 90 | 95.19 | 90 | |||
x = 1 | Pmn21 (βII) | Orthorhombic | 6.299(9) | 5.381(5) | 4.974(2) | 90 | 90 | 90 | ||
Pmnb (γII) | Orthorhombic | 5.929(5) | 10.622(0) | 4.961(1) | 90 | 90 | 90 | |||
P21/n (γs) | Monoclinic | 6.305(7) | 11.098(7) | 5.114(9) | 90 | 92.44 | 90 | |||
Li2(Fe1−xMnx)SiO4/C (x = 0, 0.05, 0.1, 0.2, 0.3) | — | Pmn21 (βII) | Orthorhombic | — | — | — | 90 | 90 | 90 | 114 |
Li2(Fe0.5Mn0.5)SiO4 | — | Pmnb (γII) | Orthorhombic | — | — | — | 90 | 90 | 90 | 115 |
Li2(Fe0.5Mn0.5)SiO4/C | — | Pmn21 (βII) | Orthorhombic | — | — | — | 90 | 90 | 90 | 106 |
The differences among these polymorph structures mainly depend on the different arrangements of the cationic tetrahedra and the local order (order probes such as solid-state Magic Angle Spinning Nuclear Magnetic Resonance (MAS NMR) are very useful for identifying the obtained polymorphs and also their mixtures).96,97
Orthorhombic βII-Li2MSiO4 (Pmn21) can be formed at low temperature (150–200 °C) via hydrothermal methods,100,117–120 and at 600–800 °C via solid-state,37,73,81,99,121–128 sol–gel,41,57,78,84,108,129–138 spray pyrolysis,139,140 microwave141 and combustion142 methods, etc. It is very stable and easy to synthesize via all these techniques. For example, βII-Li2CoSiO4 was obtained via a hydrothermal technique at 150 °C and treatment at 700 °C,143 via a combustion route at 950 °C (ref. 144) and via a supercritical technique at 350 °C/38 MPa.145
Orthorhombic βI-Li2MSiO4 (Pbn21, Yamaguchi et al. named it βIl-Pbn21 previously92) is formed at a higher temperature than βIl-Pmn21 and little data have been reported for this phase.143–145 To the best of our knowledge, this phase is not yet reported for Li2MnSiO4.
Monoclinic γs-Li2MSiO4 (P21/n) is formed at a higher temperature than the βI- and βIl-phases in the temperature range 650–900 °C via solid-state,43,65,79,82,105 sol–gel56,61,63,129,146 and combustion43,147,148 methods. When the temperature was further increased to 800–1000 °C, an orthorhombic γII-phase (Cmma,87,88 Pnma149 or Pmnb44,85,87,88,100,103,129,150) resulted.
The β-phase is more stable than the γ-phase because the latter has a larger volume and requires a higher temperature for preparation and quenching.44
Nyten et al. obtained pure orthorhombic βII-Li2FeSiO4 (Pmn21) via a solid-state method.23,99,152,153 Quoirin et al. synthesized orthorhombic βI-Li2FeSiO4 (Pna21) via a wet chemical method at 100 °C, and orthorhombic γII-Li2FeSiO4 at 800 °C (Cmma)/900 °C (Pnma) via a solid-state method.87–89 Nishimura et al. prepared monoclinic γs-Li2FeSiO4 at 800 °C (P21/n).93–95 Sirisopanaporn et al. obtained orthorhombic γII-Li2FeSiO4 (Pmnb) at 900 °C that differs from the monoclinic γs-phase obtained by quenching from 800 °C.93,103 Bini et al. synthesized pure monoclinic γs-Li2FeSiO4 (P21/n) at 650 °C and pure orthorhombic γII-Li2FeSiO4 (Pmnb) at 900 °C via a sol–gel method.129
At 580 °C, complete phase transformation (starting at 400 °C) from orthorhombic (Pmn21, βII) to monoclinic (P21/n, γs) was observed and it was maintained up to 820 °C. Above 900 °C, it phase transformed again into an orthorhombic phase (Pmna, γII). This involved a statistical distribution of Li and Fe over the three (2/3 Li and 1/3 Fe) cation sites. By quenching the orthorhombic phase (Pmna, γII) from 900 °C, the monoclinic phase (P21/n, γs) was fully regained at 600 °C and maintained down to RT.101 Nanorods of orthorhombic Li2FeSiO4 (Pmn21, βII) were prepared at 200 °C over 6 days and phase transformed to a monoclinic phase (P21/n, γs) by heating at 600 °C.57
Politaev et al. synthesized monoclinic (P21/n) Li2MnSiO4 at 950–1050 °C via a solid-state method.106 As mentioned before, the monoclinic Pn phase would transform to a stable orthorhombic phase (Pmn21) above 370 °C.83 By rate-cooling, a reversible phase transition can be observed in Li2MnSiO4. The sample (Pmnb) prepared at 800 °C via a solid-state method transformed to a P21/n phase when it cooled to RT at 200 K min−1, but the sample prepared at 700 °C formed the Pmn21 phase. This strongly indicated that the Pmnb phase transformed to orthorhombic disordered wurtzite-type structures.61
Bini et al. synthesized pure orthorhombic βII-Li2MnSiO4 (Pmn21) at 650 °C and mixed phases of orthorhombic γII (Pmnb)/monoclinic γs (P21/n) at 900 °C via a sol–gel method.129
By a hydrothermal technique and heating at 700 °C, orthorhombic Li2MnSiO4 (Pmn21, βII) was synthesized. Above 750 °C, it changed into a monoclinic (P21/n, γs) phase due to discontinuity in the c lattice parameter.101 By high pressure and high temperature processes, phase transitions can be observed with minimal impurities. The mixed phase of Li2MnSiO4 polymorphs, i.e. Pmnb and P21/n (63:
13 w/w) was prepared at 900 °C over 10 h by slowly heating (2 °C min−1) with small impurities of MnO and Li2SiO3. At 60 kbar and 600 °C, the mixed phase transformed to a Pmn21 phase, being stable at 80 kbar and 900 °C with a new impurity of Mn2SiO4.43
The low-temperature orthorhombic forms are more stable (due to their large volume) than the monoclinic form, which can only be prepared above 900 °C.44,97 By increasing the process pressure/temperature, polymorphic transformations (Pmn21 → Pmnb → P21/n) will take place.44,144
Monoclinic Li2CoSiO4 (P21/n, γo) was prepared at 950 °C and could transform (at 900 °C) into orthorhombic Pmn21 (βIl) and Pbn21 (βI) phases under 40/60 kbar, respectively. Both cell volumes are smaller than the monoclinic one.43 Orthorhombic Li2CoSiO4 (Pmn21, βII) was synthesized hydrothermally (150 °C/72 h). It transformed to the orthorhombic (Pbn21, βI) phase when heated at 700 °C over 2 h and to a monoclinic (P21/n, γo) phase at 1100 °C over 2 h.36 Otherwise, a considerably slow phase transition from βII ↔ γII at 170 °C has been determined by DTA studies.85
Li2(Fe1−xMnx)SiO4 (x = 0, 0.2, 0.5, 1) was prepared via a modified sol–gel route using polyvinylpyrrolidone (PVP) as the chelating agent and carbon source. A pure monoclinic phase (P21/n) was obtained when x = 0. Monoclinic (P21/n) and orthorhombic (Pmnb) mixed phases were obtained when x = 0.2. Major orthorhombic (Pmn21) and minor monoclinic (P21/n) mixed phases were obtained when x = 0.5. Major orthorhombic (Pmn21), minor orthorhombic (Pmnb) and minor monoclinic (P21/n) mixed phases were obtained when x = 1.102
However, Li2(Fe1−xMnx)SiO4 (x = 0, 0.3, 0.5, 0.7, 1) obtained via a citric acid assisted sol–gel technique could be indexed to the orthorhombic phase (Pmn21), and the lattice parameters were similar.113 All samples of Li2(Fe1−xMnx)SiO4/C (x = 0, 0.05, 0.1, 0.2, 0.3) obtained via a solution route exhibited an orthorhombic phase (Pmn21) with a small amount of impurities such as Fe3O4 and Li4SiO4.114
Li2(MnxFe1−x)SiO4 (x = 0, 0.2, 0.5, 0.8) was prepared via a combination of spray pyrolysis at 400 °C and wet ball milling followed by annealing at 600 °C in N2 for 4 h. They exhibited a major orthorhombic phase (Pmn21) and small amounts of impurities such as FeO and SiO2 were found in materials with high Fe contents of 0.5 and 0.8.109
The thermal behavior of Li2(Fe0.5Mn0.5)SiO4 obtained via a sol–gel route, from RT to 950 °C, exhibited a major Pmnb phase (determined by in situ XRD) over the explored temperature range. A pure and stable Pmnb phase was formed at 950 °C.115
Li2(Fe0.5Mn0.5)SiO4/C was prepared using the precursors of Li2FeSiO4 and Li2MnSiO4 via a sol–gel route followed by heating at 600 °C for 10 h. It exhibited an orthorhombic phase (Pmn21) but the atomic ratio was not maintained.106
When the lithium ions were fully extracted from LixMSiO4 (M = Mn, Fe, Co, Ni, …), a larger volume expansion was found in the Fe-/Co-/Ni-systems compared to the Mn-system.111
Based on the structural characterization of the Pmn21-delithiated materials, Thomas et al. suggested that the voltage shift was due to a structural transformation of the host compound. Armstrong et al. suggested a Li2FeSiO4 phase transition from the monoclinic (P21) polymorph to the orthorhombic (Pmn21-I, inverse-βII) phase at 50 °C and at C/16 (1C = 160 mA g−1).174 Chen et al. confirmed the same phase transition happened at RT and at C/20.102 But the monoclinic phase could be preserved even after cycling.155
Nanocrystalline Li2(Fe1−xMnx)SiO4/C (x = 0, 0.2, 0.5, 1) powders were tested electrochemically, and characterized using XRD, 7Li MAS NMR spectroscopy, 57Fe Mössbauer spectroscopy and in situ XAS to study the structural evolution. Results showed the materials exhibited partially reversible structural changes upon cycling. Amorphization and structural rearrangements from the initial P21/n polymorph to Pmn21 occurred during the 1st charge–discharge cycle. Pmn21-Li2FeSiO4 exhibited a stable cycling performance over the subsequent 100 cycles due to the Fe2+/Fe3+ process. A competitive redox reaction between the Mn and Fe species was deduced for Li2Fe0.8Mn0.2SiO4 and Li2Fe0.5Mn0.5SiO4. A high conversion rate existed in the 1st charge from Fe2+/Mn2+ to Fe3+/Mn3+, while the presence of a small fraction (<7.5 mol%) of cations with higher oxidation states (Fe4+/Mn4+) could not be excluded.102
Methods | Temp./time/pressure (°C/h/MPa) | Morphology | Size (nm) | Specific capacity (mA h g−1)/current density | Retention (%)/cycles | Refs. |
---|---|---|---|---|---|---|
Solid-state | 650/10/— | Agglomerate (Li2FeSiO4/C) | — | 144.8/0.1 | 94.27/13 | 188 |
800/12/— | Agglomerate (Li2MnSiO4/C) | 30–80 | 129/0.06 | —/10 | 128 | |
650/8/— | Agglomerate (LiFe0.9Mn0.1SiO4/C) | — | 158.1/0.03 | 94.3/30 | 114 | |
900/6/— | Agglomerate (Li2FeSiO4/C/rGO) | 28.5 | 191.6/0.1 | 91.3/60 | 127 | |
700/10/— | Agglomerate (Li1.95FeSiO4/C/CNTs) | 30–70 | 148/0.2 | 99.2/100 | 159 | |
700/8/— | Spherical (Li2MnSiO4/C/graphene) | 50 | 215.3/0.05 | 81.3/40 | 125 | |
650/10/— | Agglomerate (LiFe0.97Co0.03SiO4/C) | 100–500 | 199/3.0 | 71.6/100 | 65 | |
700/15/— | Rounded (Li2FeSiO4/C) | 200 | 102/— | 91.8/10 | 126 | |
700/10/— | Agglomerate (Li1.95FeSiO4/C) | 100 | 142/1.0/100 | 95.1/100 | 122 | |
650/10/— | Agglomerate (LiFe0.95V0.05SiO4/C) | 7–10 | 220.4/0.1 | 78.7/50 | 161 | |
700/10/— | Agglomerate (Li2MnSiO4/C) | 13.4–23.3 | 201.8/— | 73.5/15 | 163 | |
Sol–gel | 800/—/— | Spherical (Li2CoSiO4) | 460 | 32/0.02 | —/2 | 164 |
700/5/— | Spherical (Li2MnSiO4) | 25–30 | 161/0.05 | 77.6/50 | 137 | |
700/10/— | Irregular (Li2MnSiO4/C) | 20–30 | 240/0.02 | 45.4/30 | 130 | |
700/—/— | Spherical (Li2MnSiO4) | 25–30 | 113/— | —/15 | 138 | |
650/10/— | Mesopores (Li2FeSiO4/C) | 20–30 | 163/0.06 | 96/200 | 166 | |
680/10/— | Mesopores (Li2MnSiO4/C) | 10–20 | 164.2/0.06 | 80/60 | 136 | |
650/10/— | Agglomerate (Li1.9Na0.1MnSiO4/C) | 30 | 175/0.01 | 45/40 | 167 | |
650/10/— | Agglomerate LiFeSi0.9V0.1O4/C) | 100–200 | 159/0.06 | 90/30 | 63 | |
700/12/— | Agglomerate (LiFe0.97Mg0.03SiO4) | 100 | 153.2/0.06 | 98.6/50 | 61 | |
700/10/— | Agglomerate (LiMn0.8Fe0.2SiO4/C) | 15–30 | 224/0.05 | 63.8/50 | 189 | |
700/10/— | Honeycomb (Li1.8MnSiO4) | 6–8 | 147.1/0.03 | 83.3/25 | 168 | |
600/10/— | Aggregated (Li2FeSiO4/C-NHT) | 20–30 | 195.5/0.1 | —/50 | 190 | |
700/10/— | Agglomerate (Li0.5Mn0.5SiO4/C) | 220 | 230.1/0.1 | 70.4/20 | 191 | |
700/12/— | Core–shell (Li2.05Mn0.95P0.05Si0.95O4) | 30–60 | 170/0.06 | —/60 | 170 | |
700/12/— | Agglomerate (LiFe0.97Zn0.03SiO4) | 100 | 128/0.1 | 97.5/55 | 169 | |
650/10/— | Mesopores (Li2FeSiO4@CMK-3) | 50 | 160/0.1 | —/80 | 192 | |
600/10/— | Agglomerate (Li2FeSiO4/C) | 15 | 187/0.1 | —/50 | 193 | |
700/10/— | Agglomerate (Li2.05Fe0.95P0.05Si0.95O4/C) | 50–100 | 213/0.3 | —/50 | 194 | |
650/6/— | Spherical (Li2MnSiO4/C) | 20–50 | 253.4/0.03 | 77/20 | 133 | |
600/10/— | Nanowires (Li2MnSiO4/C/V2O5) | 30 | 277.0/0.1 | —/50 | 195 | |
700/12/— | Nanofibers (LMn0.94Cr0.06SiO4/C) | 70–150 | 295/— | 65.8/50 | 64 | |
650/10/— | Agglomerate (LiMn0.09Ti0.01SiO4) | ∼50 | 211/0.01 | —/50 | 196 | |
Hydrothermal/solvothermal/supercritical | 150/92/— | Flower (Li2FeSiO4) | 20–50 | 130/0.1 | —/30 | 117 |
700/—/— | Flower (Li2MnSiO4/C) | — | 100/0.05 | —/40 | 162 | |
550/5/— | Spherical (Li2MnSiO4) | 644 | 177/— | 32/20 | 173 | |
650/10/— | Hollow spheres (Li2FeSiO4) | 0.5–2 μm | 152/0.05 | —/100 | 178 | |
600/6/— | Hot-dog (Li2FeSiO4) | 200 | 150/— | —/50 | 180 | |
900/4/— | Hollow spheres (Li2CoSiO4/C) | 300–400 | 33/— | —/50 | 49 | |
600/5/— | Agglomerate (Al–LiFe0.5Mn0.5SiO4/C) | 20–100 | 216/0.01 | —/20 | 62 | |
200/72/— | Spherical (Li2FeSiO4/C) | ∼20 | 136/0.2 | 96.1/100 | 118 | |
700/10/— | Flower (Li2MnSiO4) | 20 | 226/0.5 | —/10 | 116 | |
650/10/— | Shuttle-like (Li2MnSiO4/C) | 0.4–0.5 μm | 206/0.05 | —/50 | 132 | |
180/192/— | Hierarchical shuttle (Li2FeSiO4) | 31.8 | 159.4/0.01 | 97.5/20 | 120 | |
600/2/— | Nanorods (Li2FeSiO4/C) | 80–100 | 155/0.06 | —/50 | 175 | |
650/10/— | Spherical (Li2FeSiO4/C) | 18.3 | 211.3/0.1 | 97.7/1000 | 176 and 181 | |
700/10/— | Plates (LiMn0.8Ni0.2SiO4) | 100 | 100/0.2 | —/10 | 179 | |
600/10/— | 3D porous hierarchical (Li2FeSiO4/C) | ∼60 | 170/0.1 | —/20 | 57 | |
600/06/— | Nanorods (Li2FeSiO4/graphene) | 10–25 | 306.4/0.3 | 95/240 | 97 | |
700/10/— | Agglomerate (Li2MnSiO4@Ni)/C | 20–50 | 274.5/0.05 | —/20 | 185 | |
700/5/— | Spherical (Li2MnSiO4) | 15–500 | 260/0.1 | —/5 | 177 | |
380/30 min/38 | Rod (Li2FeSiO4) | 20–80 | 177/0.01 | —/25 | 145 | |
350/30 min/38 | Spherical (Li2CoSiO4) | 20–15 | 107/0.1 | —/4 | 156 | |
300/5 min/38 | Spherical (Li2MnSiO4) | 5–20 | 313/0.05 | 76.7/20 | 197 | |
400/4 min/38 | Hierarchical nano (Li2MnSiO4) | 4–5 | 291/0.05 | —/50 | 198 | |
Microwave | 650/6/— | Nanosphere (Li2FeSiO4/C) | ∼20 | 148/0.05 | —/20 | 171 |
650/6/— | Agglomerate (Li2MnSiO4/C) | ∼20 | 210/0.05 | —/20 | 171 | |
700/20/— | Agglomerate (Li2FeSiO4) | — | 116.2/0.05 | —/10 | 183 | |
700/20/— | Agglomerate (Li2FeSiO4) | — | 116.9/0.05 | —/10 | 184 | |
Spray pyrolysis/combustion/hydrochemical | 800/4/— | Agglomerate (Li2MnSiO4/C) | ∼50 | 184/0.05 | —/20 | 139 |
700/10/— | Spherical (Li2FeSiO4/C) | 1–5 μm | 123/1.0 | 98.4/300 | 199 | |
600/4/— | Spherical (Li2FeSiO4/C) | 65 | 154/0.05 | —/70 | 200 | |
600/4/— | Agglomerate (LiFe0.5Mn0.5SiO4/C) | 65 | 149/1.0 | —/50 | 109 | |
700/12/— | Agglomerate (Li2FeSiO4/C/C-nanosphere) | ∼200 | 164.7/0.1 | 98.4/60 | 187 | |
650/10/— | Porous (Li2FeSiO4/C) | 28 | 135/0.06 | —/— | 148 | |
800/10/— | Agglomerate (Li2FeSiO4/C) | 29 | 130/0.06 | —/50 | 147 | |
700/10/— | Agglomerate (Li2MnSiO4/C) | 50 | 164/0.01 | —/20 | 142 |
Among the cathode materials obtained via solid-state methods, Li2Fe0.95V0.05SiO4/C exhibited the highest initial discharge capacity of 220.4 mA h g−1 (∼1.33 Li+ extracted LFS) at 0.1 C. After 50 cycles, it decreased to 146.6 mA h g−1 with a retention of 78.7%. It was prepared using CH3COOLi·2H2O, FeC2O4·2H2O, TEOS, V2O5 and sucrose at 650 °C for 10 h.161 An appropriate percentage of V-substitution (5 mol%) could possibly make for >1 Li+ extraction.
Nanospherical Li2MnSiO4/C/graphene exhibited a high initial discharge capacity of 215.3 mA h g−1 (∼1.3 Li+ extracted LMS) at 0.05 C (Fig. 4a). After 40 cycles, this decreased to 175 mA h g−1. It was prepared using MnCO3, LiOH·H2O, nano-SiO2, polyethylene glycol-600 (PEG-600), glucose, cellulose acetate, and graphene oxide at 700 °C for 10 h. Spherical SiO2, mixed carbon sources and PEG-600 form a 3D nest-like carbon network, which are favourable for improving the capacity and cyclic stability.125
![]() | ||
Fig. 4 (a) Charge–discharge profile of Li2MnSiO4/C/graphene (0.05 C) via a solid-state method;125 (b) cyclic performance of Li2MnSiO4/C via a sol–gel technique at different current densities;130 (c) charge–discharge capacity of PEDOT/Li2MnSiO4 via a supercritical technique at 0.05 C and at 40 °C;156 (d) charge–discharge profile of Li2MnSiO4/C via a microwave method at C/20 and 55 °C.171 |
There are some dissatisfying results reported for the Li2MSiO4 (M = Mn, Co, …) series prepared via solid-state methods where >1 Li+ extraction (i.e. capacity > 166 mA h g−1) could not be attained. Li2MnSiO4/C exhibited a low initial discharge capacity of 129 mA h g−1 at 0.06 C, which was prepared using Li2SiO3, Mn(CH3COO)2·4H2O and sucrose at 800 °C for 12 h.128 Another example is Li2MnSiO4 prepared from LiCH3COO·2H2O, Mn(CH3COO)2·4H2O, SiO2 and glucose at 700 °C for 10 h. It exhibited a low initial discharge capacity of 152 mA h g−1.163 The poor performance could be ascribed to impurities in the samples, unsuitability of the carbon source and a process not effectively hampering the crystal structure destruction.
Most of the mesoporous structures of Li2MSiO4 prepared via the sol–gel method exhibited better discharge capacities than the other structures. Hydrochloric acids and propylene oxides used for enhancing the hydrolysis of TEOS could improve the capacity retention more than the other chelating agents and catalysts.137,138 Nanoparticles ranging from 10–200 nm could be attained by this technique with substitution with elements such as V2+,63 Mg2+,61 Na+,167 Cu2+,169 P5+,170 Zn2+ (ref. 169) and Ni2+ (ref. 169) etc.
The Li2MnSiO4/C composite exhibited an initial discharge capacity of 240 mA h g−1 (∼1.44 Li+ extracted LMS) at 0.02 C (45.4% retention) and 125 mA h g−1 at 0.4 C (52% retention) as shown in Fig. 4b. The discharge capacity decreased with the rate capacity, while the retention increased. It was prepared using LiCH3COO·2H2O, pre-synthesized Mn3O4 and TEOS (2.040:
0.763
:
2.083 wt%,) as starting materials, adding acetic acid (catalyst) and sucrose (carbon source) and heating at 700 °C for 10 h. Replacing the soluble metal source with manganese oxide improved the cycle performance due to the more compact carbon-coating and more effective prevention of the side reaction between the cathode and the electrolyte.130
The Li2MnSiO4/C nanocomposite exhibited a discharge capacity of 253.4 mA h g−1 (∼1.53 Li+ extracted LMS) at 0.03 C and 149.9 mA h g−1 at 1 C (56.4% retention). It was prepared using LiCH3COO·2H2O (0.02 mol), Mn(CH3COO)2·4H2O (0.01 mol), TEOS (0.02:
0.01
:
0.01, mol%), water–acetic acid solution (1
:
2, wt%) and glucose at 650 °C for 6 h.133 The capacity decrease may be associated with the amorphous tendency causing a volumetric effect.107,133,171
Li2(Mn0.94Cr0.06)SiO4/C nanofibers exhibited a discharge capacity of 295 mA h g−1 (∼1.77 Li+ extracted LMS) at the first cycle and reached the highest capacity of 314 mA h g−1 (∼1.88 Li+ extracted LMS) at the 5th cycle. After 20 cycles, the discharge capacity was maintained at 273 mA h g−1 (∼1.63 Li+ extracted LMS). It was prepared using N,N-dimethylformamide (DMF), Mn(CH3COO)2·4H2O, LiCH3COO·2H2O and TEOS at 700 °C for 12 h. Appropriate Cr-substitution improved its cycling behavior due to the large unit cell volume. The carbon nanofiber matrix contributed to the faster electron and ion transportation, leading to good reversibility for the cathode materials.64 There are also some dissatisfying results reported in the Li2MSiO4 (M = Mn, Co, …) series prepared via the sol–gel method where >1 Li+ extraction (i.e. capacity >166 mA h g−1) could not be attained. Li2MnSiO4 exhibited a low initial discharge capacity of 113 mA h g−1 at the first cycle. It was prepared using LiCH3COO·2H2O, Mn(CH3COO)2·4H2O, TEOS and adipic acid at 700 °C.138 Li2CoSiO4 exhibited only an initial discharge capacity of 32 m h g−1 at 0.02 C in the 1st cycle and it decreased after 10 cycles. It was prepared using LiNO3, Co(NO3)2·6H2O, silica particles and polyacrylic acid (PAA, used as a chelating agent) at 800 °C. The low capacity could be ascribed to poor electronic conductivity, insufficient ball milling and an unsuitable electrolyte.164
By the hydrothermal method, Mali et al. first reported three polymorphs of Li2MnSiO4 and their NMR spectra.97 The flower-like Li2MnSiO4 exhibited an initial discharge capacity of 226 mA h g−1 (∼1.36 extracted LMS) at 0.05 C. After 10 cycles, the capacity was stabilized at ∼130 mA h g−1. It was prepared using LiOH·2H2O, Mn(CH3COO)2·4H2O and TEOS at 700 °C for 10 h. The flower-like particles could produce good electrochemical properties due to their large surface area, unique morphology and high interconnectivity with each other.116
Ni-modified Li2MnSiO4 (LMS@Ni/C) prepared by the solvothermal technique exhibited an initial discharge capacity of 274.5 mA h g−1 (∼1.67 Li+ extracted LMS) at 0.05 C. After 20 cycles, this decreased to 119.8 mA h g−1. It was prepared using C2H3O2Li·2H2O, NiC4H6O4·4H2O, TEOS, MnC4H6O4·4H2O, cetrimonium bromide (2.2:
0.05
:
1.0
:
0.95
:
0.1, mol%) and starch at 700 °C for 10 h. The Ni improved the ion diffusion coefficient and electronic conductivity.97
Li2MnSiO4 prepared by the supercritical technique, exhibited a very high discharge capacity of 313 mA h g−1 at 40 °C and 0.05 C (∼1.9 Li+ extracted LMS) as shown in Fig. 4c. After 20 cycles, this decreased to 240 mA h g−1. It was prepared using LiOH·2H2O, MnCl2·2H2O, TEOS and ascorbic acid (1:
4
:
1
:
0.1, mol%) at 300 °C and 38 MPa for 5 min.156 Nano-Li2MSiO4 (M = Fe, Mn, Co) materials were prepared at 300–380 °C and 38 MPa for 5–30 min. They had a high discharge capacity in the 1st cycle at different rates but failed to maintain the stability.145,156,177 Li2CoSiO4 exhibited a discharge capacity of 107 mA h g−1 (∼1.51 Li+ extracted LCS) at 0.01 C. After 3 cycles, this decreased to 80 mA h g−1. It was prepared using LiOH·2H2O, CoCl2·6H2O, TEOS (4
:
1
:
1, mol%) and oleylamine at 300–350 °C and 30 MPa for 10–30 min.145
There are a few results reported in the Li2MSiO4 (M = Mn, Co, Ni, …) series prepared via hydrothermal/solvothermal methods where >1 Li+ extraction (i.e. capacity > 166 mA h g−1) could not be attained. Li2Mn0.8Ni2SiO4 prepared via the solvothermal technique exhibited a low discharge capacity of 100 mA h g−1 (∼0.7 Li+ extracted LMS) at 0.2 C. It was prepared using LiOH·2H2O, NiCl2·6H2O, MnCl2·4H2O, SiO2 (0.386:
0.2
:
0.8
:
0.138, wt%) and sucrose at 700 °C for 10 h. Ni-substitution maintained the structure and improved the charge capacity, but caused no improvement to the discharge capacity.181
By a microwave-assisted hydrothermal technique, Li2MnSiO4/C exhibited an initial discharge capacity of 210 mA h g−1 (∼1.27 Li+ extracted LMS) at RT and 250 mA h g−1 (∼1.5 Li+ extracted LMS) at 55 °C for 0.2 C (Fig. 4d). After 20 cycles, this drastically decreased by 50% (105 mA h g−1) at RT and 15% (240 mA h g−1) at 55 °C. It was prepared using LiOH·2H2O, Mn(CH3COO)2·4H2O, TEOS and sucrose at 650 °C for 6 h.171
Li2MnSiO4 obtained via a microwave-assisted solvothermal synthesis exhibited an initial discharge capacity of 260 mA h g−1 at 50 °C (∼1.57 Li+ extracted LMS) and 130 mA h g−1 at RT at 0.1 C. After 4 cycles, this drastically decreased. It was prepared using LiOH·2H2O, Mn(CH3COO)2·4H2O, TEOS and urea at 700° for 5 h.185
The spray pyrolysis method is used to prepare powders with an uniform chemical composition and a narrow particle size distribution from the nanometre to the micrometer scale, easily used for large-scale applications.186 Li2MnSiO4/C exhibited an initial discharge capacity of 184 mA h g−1 (∼1.1 Li+ extracted LMS) at RT and 0.05 C. After 20 cycles, this decreased to 110.4 mA h g−1 (∼60%). Moreover, it exhibited an initial discharge capacity of 225 mA h g−1 (∼1.36 Li+ extracted LMS) at 60 °C and 0.05 C but this decreased drastically at the end of the 20th cycle.139
Li2MSiO4 (M = Fe, Mn) was prepared via a combustion method, using LiNO3, Mn(CH3COO)2·4H2O, TEOS, citric acid and 20 wt% of acetylene black at 700 °C for 10 h.142 Li2MnSiO4/C exhibited an initial discharge capacity of 161 mA h g−1 (∼0.96 Li+ extracted LMS) at 0.06 C.
By the hydrochemical technique, Li2FeSiO4/C/CNS with a double carbon coating (glucose-derived carbon and carbon nanospheres) exhibited an initial discharge capacity of 164.7 mA h g−1 (∼1.0 Li+ extracted LFS) at 0.1 C. After 60 cycles, 98.4% capacity remained. It was prepared using CH3COOLi·2H2O, Fe(CH3COO)2·4H2O, TEOS, glucose and CNS at 700 °C for 12 h. The double carbon coating increased electron transport among the particles.187
Through DFT calculations for a specific structure, the properties of Li2MSiO4 (M = Mn, Fe, Co, Ni, …) can be accurately predicted and its reaction mechanisms can be fully described. The applications of DFT calculations include crystal structure modeling and stability investigations of delithiated and lithiated phases, averaging of the lithium intercalation voltage, prediction of charge distributions and band structures, and kinetic studies of lithium ion diffusion processes (Li+ vacancy migration barriers86,206), which can provide atomic understanding of the capacity, reaction mechanism, rate capacity, and cycling ability. The results obtained from DFT are valuable for establishing the relationship between the structure and the electrochemical properties, promoting the design of new electrode materials.201
The total energy vs. volume of Li2MnSiO4 polymorphs (orthorhombic βII-Pmn21/γII-Pmnb and monoclinic γs-P21/n) and their electrochemical properties as electrodes for LIBs were investigated by Arroyo-de Dompablo et al.,44 combining experimental and computational methods. Calculation results show that the crystal structure has little effect on the average Li+ intercalation voltage (4.18/4.19/4.08 V for Pmnb/Pmn21/P21/n, respectively). The Pmnb form is 2.4/65 meV/f.u. more stable than the Pmn21/P21/n forms, respectively (Fig. 5, GGA + U results) but one cannot deny that other variants of β- and/or γ-Li3PO4 could be energetically accessible. In fact, a β-Li3PO4 (βI-Pbn21) type Li2MnSiO4 sample had been prepared through the hydrothermal method. DFT calculations also revealed that the denser Pmn21 polymorph can be obtained by a high-pressure high-temperature treatment of the other polymorphs or their mixtures.44
![]() | ||
Fig. 5 Calculated total energy vs. volume curves of Li2MnSiO4 polymorphs; Pmn21 (red), Pmnb (blue) and P21/n (green). The calculated average voltage for the 2 electron process is given in parentheses.44,209 |
6Li MAS NMR spectroscopy combined with DFT calculations was used to study the structural differences between various polymorphs of Li2MSiO4 (M = Mn, Fe, Zn). Results show that fully lithiated Li2MSiO4, delithiated LiMSiO4 and MSiO4 are semiconducting and the band gap of Li2MSiO4 decreases while extracting lithium ions.96,97 The Si–O bonds remain almost unchanged during lithiation–delithiation for all polymorphisms, which contributes significantly to the structural stability.86 Comparisons exhibit a more promising role for the monoclinic P21/n configuration. The corresponding fully delithiated MSiO4 attained a better stability due to the high-spin state from M2+ to M3+ and to M4+ ions.48 In contrast to the LiMPO4 counterpart, the potentials of Li2MSiO4 are largely increased due to the higher valence state of the M3+/M4+ redox couple207 and Li+ conduction exhibits a two-dimensional anisotropic character.19
Kokalj et al. predicted through DFT calculations that it might be possible to obtain a stable material with >1 Li+/f.u. by using a Li2MnxFe1−xSiO4 solid solution. The voltage required is lower by 0.7 V than that for the pure Fe counterpart.202 However, Larsson et al. argued that when the ratio of Mn substitution was lowered to 12.5%, the structural distortion and high voltage would destroy the feasibility of this design.208
Kalantarian et al. showed by DFT calculations that βII-Li2MnSiO4 (Pmn21) should have better electrochemical properties, even after a 2 Li+/f.u. extraction.203 The insertion–extraction mechanism was put into a relation with the voltage–capacity behaviour by considering a diffusion model. They also related the voltage behaviour to relevant parameters such as the reaction energy, Li+ diffusion coefficient and particle size, so suggesting some strategies for optimizing materials.204 Li2M0.5N0.5SiO4 (M, N = Mn, Fe, Co, Ni) compounds with a Pmn21 structure were studied by DFT using GGA (+USIC) and LSDA (+USIC) methods. Mixed compounds score better than pure materials, except for Fe–Mn in comparison to Fe. Considering both the electron conductivity and theoretical lithiation–delithiation reaction voltage, the best properties are shown by Fe–Ni, Mn–Ni, Fe–Co and Mn–Co, in descending order. Fe–Ni is theoretically the most promising material. Based on the gravimetric specific energy, these materials are sorted as: Mn–Fe, Fe, Mn, Mn–Ni, Fe–Co, Mn–Co, Fe–Ni, Co, Co–Ni and Ni, in ascending order (Fig. 6).205
![]() | ||
Fig. 6 Band gaps for delithiated states of Li2M0.5N0.5SiO4 (M, N = Mn, Fe, Co, Ni) calculated via GGA/LSDA + USIC methods.205 |
Many researchers have suggested that the selected carbon precursors are directly involved in the characteristics of the carbon additives, in terms of their structure, distribution and the thickness of the coating layer, which are proportional to the performance of electrodes.212–215
Carbon coating not only provides conductive connections among the active particles which are favourable for electron and ion-transfer (Fig. 7a),216 but also decreases the particle size of the active materials effectively during the heat treatment. It can shorten the diffusion path of the lithium ions and also facilitate good contact between neighbouring particles with a reduction of the polarization.45,61,217 Uniform carbon coating (∼2–5 nm thickness on nanoparticles) increases the capacity and rate performance by providing short pathways and huge electrochemical interfaces for fast ion diffusion and transportation (Fig. 7b).133,218,219 Various carbon precursors directly affected the electrical conductivity of the carbon, which supports the improvement of the electrochemical performance of Li2MSiO4/C cathode materials.136,220 Increased sp2-coordinated carbon promised a better electrochemical performance.221–223 Functional structures or ring-forming structures of the organic precursor have attracted considerable attention as carbon sources for electrodes.212,221,222
![]() | ||
Fig. 7 (a) Schematic presentation of electron and ion transport in a carbon-coated cathode material;216 (b) uniform thickness of carbon coating on a Li2MSiO4 particle.133 |
If the aggregated carbon content is too high, it will block the pathway for electrolyte percolation. Hence the ionic conduction is blocked.224 The total amount of impurities increases with the carbon content, i.e. the content of electrochemically active Li2MSiO4 decreases with increasing carbon content.
Up to now, many carbon sources have been used in synthesizing Li2MSiO4 cathode materials, such as carbon nanospheres (CNSs),187 carbon nanotubes (CNTs),159 reduced graphene oxide (rGO),127 graphene oxide (GO),125 PEG-600,125 cellulose acetate,125 adipic acid,225 acetylene black or Ketjen black,73,160 sucrose,20,114,128,130,131,136,139,158,170,171,180,188,212,226–229 cellulose acetate,158 acetylene black,73,109,118,132,200,226,229–232 cellulose,108 polyethylene–poly(ethylene glycol),125,141,153,233 ascorbic acid,56,156 citric acid,41,56,63,82,132,141,191,216,227,233–236 corn starch,84,166 glucose,62,81,118,125,183,187,212,237 carbon nanosheets/tubes,187,238–241 graphene,125,127 pyromellitic acid,135,242 P123 polymer,243 carbon black,24,106,113,134,147,148,153,187,188,244,245 polyacrylonitrile,189 tartaric acid,80 adipic acid,137,138,160,162,246 pitch81,122 and phenolic resin.211
Li2MnSiO4 obtained via a sol–gel technique using 2.1 wt% sucrose showed a higher cyclic performance than that using 5.9 wt% content.235 Li2MnSiO4 obtained via a sol–gel technique using 0.2 mol% adipic acid (acts as a carbon source and chelating agent) exhibited a better discharge capacity of 113 mA h g−1 than that using 0.2 mol% content (58 mA h g−1).138 Carbon suppression to crystallites and connection with the pores in Li2MnSiO4/C resulted in a carbon-coated compound. This showed a higher discharge capacity of 275.2 mA h g−1 than that in uncoated Li2MnSiO4 (178.3 mA h g−1).136
Li2MnSiO4/C obtained via a combustion method using 0 wt%, 10 wt%, and 20 wt% of acetylene black exhibited the discharge capacity of 2 mA h g−1, 128 mA h g−1 and 164 mA h g−1, respectively.142
Graphitized carbon sources afford a better electrochemical performance than the non-graphitized forms. A Li2MnSiO4/C/graphene composite was prepared using carbon sources like polyethylene glycol 600, glucose, cellulose acetate and graphene oxide. It exhibited an initial discharge capacity of 215.3 mA h g−1 at 0.05 C. The 3D nest-like carbon network and carbon layer are favourable for improving the capacity and cycling stability.125
Li2MnSiO4 particles exhibited a uniform spherical shape and a size of around 50 nm coated with a carbon/graphene layer (Fig. 8a).125 Nanospherical Li2MnSiO4 (20–30 nm) was synthesized using a chelating agent (adipic acid). It exhibited an initial discharge capacity of ∼161 mA h g−1 and was stable up to 50 cycles.137
![]() | ||
Fig. 8 (a) TEM of Li2MnSiO4 particles exhibiting an uniform spherical shape and a size of around 50 nm coated with a carbon/graphene layer;125 (b) SEM of Li2MnSiO4 thin flowers with a flake-type petal;116 (c) TEM of Li2MnSiO4 particles with a hierarchical structure.250 |
The 20 nm thin flower-like morphology with a flake margin of petals (Fig. 8b) was formed by Li2MnSiO4 during the nucleation growth stage in the presence of water under an optimized hydrothermal treatment. The resulting material exhibited an electrochemical stability of ∼125 mA h g−1 due to its large surface area, unique morphology and high interconnectivity beween the nearest neighbouring particles.116
Large numbers of pores in the mesoporous and macroporous structures (also in the hierarchical porous structure) greatly influence the cycling behavior, resulting in a discharge capacity of 275.2 mA h g−1 in Li2MnSiO4/C and 163 mA h g−1 in Li2FeSiO4/C at 0.06 C. The current penetrating electrolyte through the porous structure was of benefit for promoting Li+ transportation.133,136,166,178,249 Honeycomb-like Li1.8MnSO4 with mesopores of 6–8 nm was produced by decomposing raw materials at high temperature (to release CO2 and H2O), and it exhibited a discharge capacity of 110.9 mA h g−1.168
Li2MnSiO4/C/V2O5 in which V2O5 nanowires adhered to Li2MnSiO4/C nanoparticles exhibited a high discharge capacity of 277 mA h g−1 (1.66 Li+ extracted LMS) at 0.1 C, because V2O5 nanowires suppressed the dissolution of manganese and protected Li2MnSiO4 from a corrosive reaction with the electrolyte.195
Li2MnSiO4 with a hierarchical structure (200–400 nm, Fig. 8c) was prepared via a supercritical fluid process using oleic acid as a surfactant. It exhibited an initial discharge capacity of 283 mA h g−1 (1.7 Li+ extracted LMS). The suitable particle size and the special structure enabled nearly 2 Li+ to be extracted.
The substitution of Mn2+ by Cr3+ in Li2(Mn1−xCrx)SiO4/carbon nanofiber composites (x = 0.03, 0.06 and 0.1) increased the crystal unit cell volumes. The crystalline structure was prevented from collapsing and the structural stability was improved during the charge–discharge cycles. Li2(Mn0.94Cr0.06)SiO4 exhibited a discharge capacity of 314 mA h g−1 and an improved cycling behavior due to the maximized unit cell volume.64
The substitution of Mn2+ by Ti4+ in Li2(Mn1−xTix)SiO4 (x = 0, 0.06, 0.1 and 0.2) led to a slight shrinkage of the volume due to the smaller ionic radius of Ti4+ compared to that of Mn2+. The particles had smaller sizes, better monodispersion and larger specific areas compared to pristine Li2MnSiO4. Li2(Mn0.8Ti0.2)SiO4 showed a good cycling performance and maintained a capacity of around 100 mA h g−1 after 50 cycles (Fig. 9a).196
![]() | ||
Fig. 9 (a) Cyclic performance of Li2Mn1−xTixSiO4 (x = 0, 0.06, 0.1 and 0.2);262 (b) cyclic performance of Mg2+, Ga3+ and Al3+ substituted Li2MnSiO4.263 |
After substitution of trivalent ions such as Ga3+, Al3+ and Mg2+ in Li2MnSiO4, small intensity peaks were found in the region of Li/Mn sites but the substitution did not affect the structural arrangement. Only Ga3+ led to a better charge–discharge capacity (Fig. 9b) and exhibited an effect on well-dispersed nanoparticle formation.56 If substitution was on the Si site, charge compensation could be achieved through the creation of Li+ interstitials.19
Li2FeSiO4 suffers low electronic conductivity and slow Li+ diffusion. The substitution of iron with other transition metals is a possible solution to the problem. At the same time, the initial structure of Li2MnSiO4 is progressively amorphized during delithiation, which leads to poor stability during cycles. In order to combine the high discharge capacity of Li2MnSiO4 with the high cycling stability of Li2FeSiO4, partial substitution of Fe2+ with Mn2+ was employed.47,202,233 In this way, substitution of Fe2+ with Mn2+, which undergoes a 2+ → 4+ transition, can facilitate a “>1 electron redox reaction” through the following reaction with the removal of up to 2 Li+, resulting in a capacity increase from 170 mA h g−1 to 340 mA h g−1.264
Li+2(Fe2+1−xMn2+x)SiO4 ↔ Li+1−x(Fe3+1−xMn4+x)SiO4 + (1 + x)Li+ + (1 + x)e− |
The substitution of Fe2+ with Mn2+ in Li2(Fe1−xMnx)SiO4 (x = 0, 0.03, 0.1, 0.2 and 0.3) through a solution route led Li2(Fe0.9Mn0.1)Si04/C to exhibit an initial discharge capacity of 158.1 mA h g−1 (∼0.95 Li+ extracted LFS) with a capacity retention of 94.3% after 30 cycles.114
The substitution of Mn2+ with Fe2+ in Li2(Mn1−xFex)SiO4/C (0 ≤ x ≤ 0.8) via a spray pyrolysis/ball milling/annealing route led Li2(Mn0.5Fe0.5)SiO4/C to exhibit an initial discharge capacity of 149 mA h g−1 (∼0.9 Li+ extracted LFS) at 1 C.109 In contrast, the substitution of Fe2+ with Mn2+ in Li2(Fe1−xMnx)SiO4 (x = 0, 0.3, 0.5, 0.7, 1) led to a discharge capacity with only 61.4% retention after 50 cycles.113 The Mn-substituted materials exhibited a higher redox potential and higher initial discharge capacity but suffered a poor cycling performance and electrochemical reversibility. This might be due to the structural stability and electronic conductivity. Li2(Fe0.8Mn0.2)SiO4 showed good reversibility and exhibited an initial discharge capacity of 230.1 mA h g−1 (∼1.4 Li+ extracted LFS) at 0.1 C and retained 162 mA h g−1 after 20 cycles.
Other efforts involve the replacement of O with N,265 substituting polyanionic SiO4 with AsO4,266 BO3,267 VO4,268 and doping trivalent Al and Ga on Si sites.19
The stable Li2MSiO4 polymorphs include high temperature monoclinic γo/γs (P21/n), high temperature orthorhombic γII (Pmna, Pmnb, Cmma) and low temperature orthorhombic βI (Pbn21, Pna21) and βII (Pmn21), which are influenced by the synthetic conditions and electrochemical behavior. Metastable monoclinic Pn-Li2MnSiO4 would convert to the βII structure (Pmn21) above 370 °C. The densest βII (Pmn21) polymorph attained the best electrochemical performance due to its special structure. It can be prepared from other less dense polymorphs, by controlling the temperature, pressure and/or electrochemical process.
Once the materials’ amorphization or collapse were inhibited during lithium ion insertion/extraction from the host lattice, reversible structural changes and good cycling performances could be attained. Hence, it is suggested that ion-substitution, for stabilizing the structures, is an effective way to achieve high-capacity Li2MSiO4 materials. Carbon-coating and optimization of the particle size/morphology are also applied as exterior modifying methods to enhance the electrochemical performance. Thus understanding the relationship between the polymorph structure, synthesis and phase transition in Li2MSiO4 is most important.
This journal is © The Royal Society of Chemistry 2015 |