Next Article in Journal
Quality-by-Design-Based Development of n-Propyl-Gallate-Loaded Hyaluronic-Acid-Coated Liposomes for Intranasal Administration
Next Article in Special Issue
The Fascinating World of Low-Dimensional Quantum Spin Systems: Ab Initio Modeling
Previous Article in Journal
Are Biogenic and Pyrogenic Mesoporous SiO2 Nanoparticles Safe for Normal Cells?
Previous Article in Special Issue
Unusual Spin Exchanges Mediated by the Molecular Anion P2S64−: Theoretical Analyses of the Magnetic Ground States, Magnetic Anisotropy and Spin Exchanges of MPS3 (M = Mn, Fe, Co, Ni)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Towards Reversible High-Voltage Multi-Electron Reactions in Alkali-Ion Batteries Using Vanadium Phosphate Positive Electrode Materials

by
Edouard Boivin
1,2,3,
Jean-Noël Chotard
1,3,4,
Christian Masquelier
1,3,4 and
Laurence Croguennec
2,3,4,*
1
Laboratoire de Réactivité et de Chimie des Solides, CNRS-UMR 7314, Université de Picardie Jules Verne, CEDEX 1, F-80039 Amiens, France
2
CNRS, Université Bordeaux, Bordeaux INP, ICMCB UMR 5026, F-33600 Pessac, France
3
RS2E, Réseau Français sur le Stockage Electrochimique de l’Energie, FR CNRS 3459, CEDEX 1, F-80039 Amiens, France
4
ALISTORE-ERI European Research Institute, FR CNRS 3104, CEDEX 1, F-80039 Amiens, France
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(5), 1428; https://doi.org/10.3390/molecules26051428
Submission received: 30 January 2021 / Revised: 24 February 2021 / Accepted: 1 March 2021 / Published: 6 March 2021

Abstract

:
Vanadium phosphate positive electrode materials attract great interest in the field of Alkali-ion (Li, Na and K-ion) batteries due to their ability to store several electrons per transition metal. These multi-electron reactions (from V2+ to V5+) combined with the high voltage of corresponding redox couples (e.g., 4.0 V vs. for V3+/V4+ in Na3V2(PO4)2F3) could allow the achievement the 1 kWh/kg milestone at the positive electrode level in Alkali-ion batteries. However, a massive divergence in the voltage reported for the V3+/V4+ and V4+/V5+ redox couples as a function of crystal structure is noticed. Moreover, vanadium phosphates that operate at high V3+/V4+ voltages are usually unable to reversibly exchange several electrons in a narrow enough voltage range. Here, through the review of redox mechanisms and structural evolutions upon electrochemical operation of selected widely studied materials, we identify the crystallographic origin of this trend: the distribution of PO4 groups around vanadium octahedra, that allows or prevents the formation of the vanadyl distortion (OV4+=O or OV5+=O). While the vanadyl entity massively lowers the voltage of the V3+/V4+ and V4+/V5+ couples, it considerably improves the reversibility of these redox reactions. Therefore, anionic substitutions, mainly O2− by F, have been identified as a strategy allowing for combining the beneficial effect of the vanadyl distortion on the reversibility with the high voltage of vanadium redox couples in fluorine rich environments.

1. Introduction

In the 2000s, the research on polyanion compounds as positive electrode materials was mainly motivated by the interesting properties of the low cost triphylite LiFePO4 [1,2,3,4,5] (olivine-type structure) providing long-term structural stability, essential for extensive electrochemical cycling and safety issues. In this material, the high voltage for the Fe2+/Fe3+ redox couple delivered in LiFePO4 (i.e., 3.45 V vs. Li+/Li vs. ca. 2.2 V in oxides) is due to the inductive effect of the phosphate group. Further exploitation of the inductive effect with fluorine and/or sulfate has led to materials such as LiFeSO4F (Tavorite or Triplite structure) delivering an even higher voltage than LiFePO4 (i.e., 3.6 V and 3.9 V vs. Li+/Li for Tavorite and Triplite structures, respectively) [6,7]. However, these materials suffer from a deficit of capacity compared to the current best commercially available Li-ion positive electrode materials. Li2FeSiO4 has been proposed to overcome this issue by triggering both Fe2+/Fe3+ and Fe3+/Fe4+ redox couples but the strong structural changes involved seem to be detrimental to long-term performances [8]. To the best of our knowledge, this material is the only one providing a multi-electron reaction (i.e., exchange of more than one electron per transition metal) in iron-based polyanion systems while vanadium phosphate materials offer numerous such examples. Indeed, the ability of vanadium to be stabilized in the large range of oxidation states, from V2+ to V5+ (e.g., from V3+ in Li2VPO4O to V5+ in VPO4O) [9,10,11,12,13] combined with the rather high voltage of the corresponding redox couples (e.g., 4.25 V vs. Li+/Li for V3+/V4+ in LiVPO4F) [14] could allow the achievement of high energy density thanks to reversible high-voltage multi-electron redox in Alkali-ion batteries (Figure 1).
However, depending on the geometry of the VOn polyhedra, the positions of the V3+/V4+ and V4+/V5+ redox couples massively change. For instance, Tavorite LiVPO4F operates at 4.25 V vs. Li+/Li while in the homeotype LiVPO4O, the apparent same V3+/V4+ redox couple is activated at 2.3 V vs. Li+/Li. This large difference cannot be attributed only to the inductive effect, Li site energy or even cation-cation repulsion (i.e., main mechanisms reported to govern the voltage of a given redox couple): it is actually due to the vanadyl distortion observed in LiVPO4O and not in LiVPO4F. The first order Jahn Teller (FOJT) distortion is known to be weak in d1 (V4+) and d2 (V3+) and even inexistent for d3 (V2+) and d0 (V5+) cations (Figure 2). Therefore, the second order Jahn Teller (SOJT) effect drives the distortion of the V4+ and V5+ polyhedra while this distortion can be prevented for V4+ in a mixed O2−/F environment. This wide range of oxidation states for vanadium cations (V2+, V3+, V4+ and V5+) and the extended panel of environments that they can adopt (regular octahedra, distorted octahedra, square pyramids and tetrahedra) with very different electronic configurations depending on the ligand distribution (JT active or JT inactive) confer to the vanadium phosphate a very rich crystal chemistry.
Beyond their fascinating electrochemical properties, vanadium phosphate materials possess very interesting catalytic and magnetic properties. The relation between structures and these properties was already reviewed by Raveau’s group 20 years ago [15] and despite the existence of several reviews on polyanionic structures in Li and Na-ion batteries [16,17,18,19], or even specific to vanadyl phosphates (i.e., Ax(VO)PO4) [20,21], none of them focused on the relation between electrochemical properties and crystallographic structure in vanadium phosphates. Therefore, this article aims at clarifying this relation in order to unveil the structural features that dictate the redox voltage in such compounds. Through the fine description of the redox mechanism and the structural evolution observed during cycling of some widely studied materials (NASICON Na3V2(PO4)3, anti-NASICON Li3V2(PO4)3, Na3V2(PO4)2F3xOx, and Tavorite-like LiVPO4F1−xOx) we propose to sort all the vanadium phosphates reported as positive electrode materials for Li and Na-ion batteries according to the distribution of phosphate groups around the vanadium polyhedra. This classification gives a holistic picture of such systems and allows for identifying the strategies available to tend towards reversible high-voltage multi-electron reactions in Alkali-ion batteries.

2. Irreversible Multi-Electron Reactions in NASICON and Anti-NASICON AxV2(PO4)3 (A = Li, Na) Structures

The NASICON (Na-super ionic conductor) and anti-NASICON structures have the general formula AxM2(XO4)3 (with M = Fe, Ti, Sc, Hf, V, Ti, Zr, etc. or mixtures of them and X = W, P, S, Si, Mo or mixtures of them) [22,23]. These versatile structures have provided a great playground for solid state chemists. Manthiram and Goodenough demonstrated experimentally the inductive effect which modulates the voltage of the Fe3+/Fe2+ redox couple in NASICON [24] which is at the origin of all advances on polyanion materials as positive electrode materials for Alkali-ion batteries. The crystallographic arrangements of NASICON and anti-NASICON are closely related. Indeed, they are built on a three-dimensional framework of VO6 octahedra sharing all their corners with PO4 tetrahedra and conversely forming basic V2(PO4)3 repeating units commonly named “lantern units” (Figure 3). The connectivity of the lantern units generates different ion conduction paths, vanadium environments and hence different electrochemical properties.
In the structure of the anti-NASICON polymorph of Li3V2(PO4)3, the lithium ions fully occupy three crystallographic sites (one tetrahedral Li(1)O4 and two pseudo tetrahedral Li(2)O4O and Li(3)O4O sites) [25,26,27]. The electrochemical extraction of lithium from Li3V2(PO4)3 occurs according to several biphasic reactions involving the V3+/V4+ redox couple at 3.6, 3.7 and 4.1 V vs. Li+/Li and then the V4+/V5+ one at 4.5 V vs. Li+/Li (Figure 4).
Nazar and coworkers [27] studied the complex phase diagram involved during lithium extraction from Li3V2(PO4)3 through X-ray and neutron diffraction (XRD and ND) and solid state nuclear magnetic resonance spectroscopy (NMR), as summarized in Figure 5. The first delithiation step leads to the formation of Li2.5V2(PO4)3 with partial depopulation of the pseudo tetrahedral (Li(3)O4O) site and to a complex short range ordering of V3+/V4+ cations [28]. The following delithiation stage affects only the remaining Li(3) ions to yield to Li2V2(PO4)3 characterized by lithium/vacancies and V3+/V4+ orderings as suggested by diffraction. The further oxidation of vanadium allows reaching the V4+-rich LiV2(PO4)3 phase in which only one Li site remains (Li(2)) as fully occupied. In this phase, there are two very similar crystallographic sites for vanadium ((V(1)-O = V(2)-O = 1.91 Å in average).
Figure 4. Electrochemical signature of Li3V2(PO4)3 cycled between (left) 3.0 and 4.3 V vs. Li+/Li or between (right) 3.0 and 4.8 V vs. Li+/Li adapted from ref. [29]. Reproduced with permission from Rui et al., Journal of Power Sources; published by Elsevier, 2014.
Figure 4. Electrochemical signature of Li3V2(PO4)3 cycled between (left) 3.0 and 4.3 V vs. Li+/Li or between (right) 3.0 and 4.8 V vs. Li+/Li adapted from ref. [29]. Reproduced with permission from Rui et al., Journal of Power Sources; published by Elsevier, 2014.
Molecules 26 01428 g004
The last process leading to the V2(PO4)3 composition is kinetically more limited with a large over-potential (around 500 mV) [27]. At this state of charge, the environments of vanadium (with a mixed valence V4+/V5+) become more distorted, although without significant modification compared to the average V-O distances observed in the VIV-rich LiV2(PO4)3 phase. This extraction/insertion process is asymmetrical as the lithium ordering observed for LiV2(PO4)3 during charge is not observed during discharge. A disordered Lithium re-intercalation is observed until the Li2V2(PO4)3 composition is reached [30]. This asymmetrical mechanism is not observed for a lower cut-off voltage (i.e., 4.3 V when the V4+/V5+ redox couple is not activated, see Figure 4). Under this electrochemical cycling conditions the charge and discharge superimposes [31]. That was tentatively explained, in ref. [27], by the occurrence of Lithium/vacancies ordering observed in LiV2(PO4)3 which involves an ordered depopulation of Lithium, whereas from the disordered fully delithiated phase the lithium is free to be inserted randomly until the Li2V2(PO4)3 composition is recovered. More recently, operando XAS at V K-edge investigation of this irreversible mechanism suggested the formation of anti-site Li/V defects at high voltage (Figure 5) providing V5+ with a much more stable tetrahedral environment than its initial distorted octahedral one [32].
Figure 5. Structural evolution during Lithium extraction/insertion from/into Li3V2(PO4)3 [27,32].
Figure 5. Structural evolution during Lithium extraction/insertion from/into Li3V2(PO4)3 [27,32].
Molecules 26 01428 g005
In Li3V2(PO4)3, the electronically insulating phosphate groups isolate the valence electrons of transition metals within the lattices resulting in low intrinsic electronic conductivities—a trend common to all polyanion compounds. Therefore, the use of carbon coating or/and doping elements are required to improve the electrochemical performances: all the works applying these strategies are reviewed in ref [29]. The majority of these studies reports good performances only in the small voltage range (i.e., 3.0–4.3 V vs. Li+/Li, in which only the V3+/V4+ is activated). Indeed, due to the strong distortion of vanadium environments and the Li/V anti-site defects generated, the kinetic limitations of the V4+/V5+ process is difficult to overcome.
The lithium insertion into Li3V2(PO4)3 reveals a complex series of reactions as well, to reach Li5V2(PO4)3 by activating the V2+/V3+ redox couple [33]. The whole lithium insertion process into Li3+xV2(PO4)3 involves four consecutive two-phase regions to reach Li5V2(PO4)3. Approximately 0.5 Li+ is inserted at every potential plateau around 1.95, 1.86, 1.74 and 1.66 V vs. Li+/Li [26]. To the best of our knowledge, the crystallographic details of this complex mechanism have never been fully studied yet.
The anti-NASICON polymorph of Li3V2(PO4)3 is most thermodynamically stable but Gaubicher et al. [34] obtained the NASICON form by Na+/Li+ ionic exchange from Na3V2(PO4)3. This material reveals a similar electrochemical signature compared to the one of Na3V2(PO4)3 with a single plateau until the LiV2(PO4)3 composition at 3.7 V vs. Li+/Li. The crystal structure of Na3V2(PO4)3 was originally reported by Delmas et al. [35] 40 years ago using the standard rhombohedral unit cell, S.G. R-3c. Since then, Na3V2(PO4)3 has almost always been reported to adopt the rhombohedral symmetry with a partial occupancy of both Na(1) (6b Wyckoff position) and Na(2) (18e Wyckoff position) sodium sites. However, a recent article reveals that a C2/c space group is more appropriate to describe this structure at room temperature and below due to Sodium-vacancies ordering [36] within five sites (one 4a and four others 8f) fully occupied. Several transitions between 10 and 230 °C involving four distinct phases (α ordered, β and β′ with incommensurate modulations and γ disordered) were also reported. The transition between the α and β forms occurring close to the ambient temperature (i.e., 27 °C) may impact the sodium diffusion and a fortiori the electrochemical performances while the vanadium environment is hardly impacted by this phase transition. In both cases the VO6 entities are slightly distorted with distances ranging between 1.97 and 2.03 Å for the rhombohedral description (2.00 Å in average on a single vanadium site) or 1.94 and 2.06 Å for the monoclinic one (2.00 Å in average on the three vanadium sites).
The electrochemical sodium extraction from Na3V2(PO4)3 occurs at a 3.4 V vs. Na+/Na according to a biphasic reaction until the NaV2(PO4)3 composition is reached (Figure 6). The structure of this V4+ phase, reported by Jian et al. [37], keeps a NASICON framework (Rhombohedral, R 3 ¯ c) with only one fully occupied sodium site (6b Wyckoff site). During the sodium extraction, the V-O distances in VO6 octahedra undergo an inequivalent shortening leading to distorted VO6 octahedra (with 3 V-O distances at 1.86 Å and three others at 1.95 Å). The electrochemical extraction of the third sodium has never been reported despite the apparent successful chemical extraction realized by Gopalakrishnan et al. [38]. However, they did not report the detailed structure of this mixed-valence V2(PO4)3.
Even though the third Na+ of Na3V2(PO4)3 doesn’t seem electrochemically removable, the V4+/V5+ redox couple in the NASICON was reported to lie at around 4 V vs. Na+/Na thanks to the partial substitution of a part of Vanadium by Aluminum [39], Iron [40] or Chromium [41]. The Aluminum substituted material presents two advantages as it allows an increase in the capacity due to the lower weight of aluminum compared to vanadium (and also iron and chromium) as well as to reach the mixed valence V4+/V5+ state at rather high voltage (i.e., 4.0 V vs. Na+/Na, see Figure 6). However, in the Al3+ substituted compound, the V4+/V5+ capacity is limited contrarily to that observed in Na3VCr(PO4)3 where nearly 1.5 electrons/vanadium are exchanged [42]. At room temperature, this redox process induces a rapid degradation of the performance due to the migration of vanadium into the vacant Na site, while at lower temperature (i.e., −15 °C), vanadium is pinned in its original position leading to a rather reversible process is observed [43]. The V4+/V5+ redox couple has also been reported in Na-rich NASICON such as Na4MnV(PO4)3 [44,45,46,47]. From this compound, ca. 3 Na+ are exchanged based on the V3+/V4+, Mn2+/Mn3+ and then V4+/V5+ redox achieving a capacity of 155 mAh/g. However, this latter appears to be poorly reversible inducing a higher irreversible capacity upon discharge during which the highly polarized S-shape voltage profile contrasts with staircase charge curve [45,46,47].
The replacement of a (PO4)3− group in Na3V2(PO4)3 by 3 F leads to the Na3V2(PO4)2F3 composition, often named as a “NASICON composition” but its crystal structure is fundamentally different.

3. Irreversible Multi-Electron Reactions in Na3V2(PO4)2F3

The first physico-chemical investigation of the Na3M2(PO4)2F3 system was conducted 20 years ago by Le Meins et al. [48]. They demonstrated a great compositional tunability of this framework which can accommodate many trivalent cations in octahedral sites (M = Al, V, Cr, Fe and Ga) and proposed the description of the structure of the vanadium phase in the P42/mnm space group. Later, a combined synchrotron X-ray and neutron diffraction investigation revealed a tiny orthorhombic distortion at room temperature [49].
The Amam space group (i.e., S.G. #63, Cmcm) used leads to a different sodium distribution in the cell with three Na sites, one 4c fully occupied and two 8f partially occupied (approximatively distributed as 1/3:2/3) (Figure 7). The host structure is composed of V2O8F3 bi-octahedra sharing a fluorine aligned along the [001] direction and connected to each other through PO4 tetrahedra aligned in parallel with the (001) plane (Figure 7). The VO4F2 octahedra in this structure are non-centrosymmetric and hence vanadium does not occupy the inversion center. Indeed, a displacement along the c direction leads to two slightly different V-F bonds (V-F(1) = 1.968(6) Å and V-F(2) = 1.981(2) Å).
Slow electrochemical galvanostatic cycling shows the presence of four distinct reversible voltage-composition features at 3.70, 3.73, 4.18 and 4.20 V vs. Na+/Na (Figure 8) suggesting a complex phase diagram upon sodium extraction/reinsertion [50].
The operando synchrotron XRD investigation conducted by Bianchini et al. [52] is summarized in Figure 9. The phase diagram involves several intermediate phases of compositions NaxV2(PO4)2F3 with x = 2.4, 2.2, 2, 1.8 and 1.3 before the NaV2(PO4)2F3 is reached. During extraction of the first sodium, an alternation between ordered and disordered phases (Na+/vacancy and/or V3+/V4+ ordering and disordering) is observed. The superstructure peaks observed for the Na2.4V2(PO4)2F3 disappear for Na2.2V2(PO4)2F3 and the diffraction pattern of Na2V2(PO4)2F3 reveals the reappearance of a series of additional contributions non-indexed in the tetragonal cell. In the V3+-rich phase, the two symmetrically inequivalent V-F bonds are very similar and as the oxidation of vanadium is increased, two kinds of bonds gradually appear as a short one at 1.88 Å and a longer one at 1.94 Å, whereas the equatorial V-O bonds decrease uniformly (from 1.99 to 1.95 Å). The extraction of the second sodium also involves intermediate phases at x = 1.8 and x = 1.3 accompanied by the disappearance of the superstructure peaks and finally leads to the formation of NaV2(PO4)2F3. This phase contains a single Na site and two vanadium sites conferring to vanadium cations two very different environments despite an average oxidation state of V4+. Indeed, the BVS calculation suggests the formation of a V3+-V5+ pair in bi-octahedra at this composition (Figure 9). The investigation of the redox mechanism involved during sodium extraction was conducted by Broux et al. [53] through operando XANES at V K-edge. They evidenced that V4+ starts to disproportionate from Na2V2(PO4)2F3 and hence the formation of V3+-V5+ pairs are confirmed for Na1V2(PO4)2F3.
Kang and coworkers predicted that the extraction of the third Na+ towards the mixed valence V4+/V5+ V2(PO4)2F3 composition would occur only at very high voltage (ca. 4.9 V vs. Na+/Na) [55]. This was confirmed experimentally by Tarascon’s group, under sever oxidative conditions (i.e., potentiostatic step at 4.8 V vs. Na+/Na see Figure 8) in an optimized electrolyte [51]. In this structure, vanadium is displaced from the inversion center of the VO4F2 octahedra in such a way as to generate a short 1.62 Å and a longer 1.92 Å V-F bond lengths within the bi-octahedra. However, these extreme cycling conditions imply an irreversible reaction and only 2 Na+ could be reinserted upon discharge down to 3.0 V, according to a solid solution mechanism, the third Na+ being reinserted at much lower voltage (i.e., 1.6 V vs. 3.7 V for the same composition range upon charge). The subsequent charge/discharge allows for the reversible extraction/insertion of 3 Na+ in a wide voltage range (1.0–4.4 V vs. Na+/Na). This new β-Na3V2(PO4)2F3 polymorph exhibits a different symmetry, different V-X bond lengths and a disordered Na distribution (see Figure 9) which bears strong resemblance with the one of the high temperature phase (T > 400 K) [49,54]. Due to the low voltage associated to the reinsertion of third Na+, only 2 Na+ can be exchanged in a real battery system where the third one acts as an alkali reservoir to compensate for the solid electrolyte interface (SEI) formation at negative electrode [51], which allows offering up to 460 Wh/kg in full cell vs. hard carbon (+18% compared with a conventional α-Na3V2(PO4)2F3), corresponding to the highest energy density reported so far in Na-ion battery [56].
Most of the Na3V2(PO4)2F3 materials reported as stoichiometric in the literature actually present various amounts of vanadyl-type defects (i.e., partial substitution of F by O2− with a charge compensation by partial oxidation of V3+-F into vanadyl V4+=O) impacting on the electrochemical performance. Several authors studied in detail the crystallographic changes generated by this substitution in Na3V2(PO4)2F3−xOx (with 0 ≤ x ≤ 0.5 [54], 0 ≤ x ≤ 2 [55] and x = 1.6 [57,58]). This oxidation has strong effects on the local environments of vanadium and on the sodium distribution and appears to be beneficial for enhancing the charge rates of the battery. Kang’s group [57] was the first to investigate the performance of Na3V2(PO4)2F1.4O1.6 (i.e., V3.8+) as a positive electrode material and reported high charge and discharge rate capabilities, assigned to a low activation energy for Na+ diffusion (~350 meV) inside the framework, despite the poor electronic conductivity (~2.4 × 10−12 S.cm−1) and its great cycling stability was assigned to the small volume change during sodium extraction/insertion (~3%). The same group later published a promising result about the computed voltage for the extraction of the third sodium around 4.7 V for Na3V2(PO4)2F1.5O1.5 [55] (lower than the one computed up to 4.9 V in Na3V2(PO4)2F3) and experimentally realized the reversible exchange of more than one electron per vanadium at high voltage (ca. 3.8V vs. Na+/Na in average) with a symmetrical charge/discharge profile and an improved capacity retention.

4. Low Voltage Multi-Electron Reactions in Tavorites LixVPO4Y (Y = O, F and/or OH)

Tavorite-type compositions of general formula AxMXO4Y are a third class of very interesting polyanion structures in which A is an alkali cation (i.e., Li, Na and 0 ≤ x ≤ 2) and M a metal (i.e., Mg, Al, Ti, V, Fe, Mn, Zn or mixture of them). The polyanionic group, XO4, is either PO4 or SO4 and the bridging anion, Y, is a halide, hydroxide, oxygen, H2O group or a mixture of them [18]. The multiple redox center combined with this double inductive effect bring a strong interest at both practical and fundamental levels as it allows scanning a wide range of working voltages, from 1.5 V for Ti3+/Ti4+ in LiTiPO4O [59] to 4.26 V for the V3+/V4+ redox couple in LiVPO4F [60]. The high voltages provided by the V3+/V4+ and V4+/V5+ redox couples confer high theoretical energy densities to the vanadium-based Tavorite compositions.
The crystal structure of Tavorite-like materials can be described in either triclinic (P-1, with Z = 2 or Z = 4) or monoclinic (C2/c or P21/c) systems [14,61,62]. Tavorite-like therefore gathers Tavorite (P-1, Z = 2, LiVPO4F and LiVPO4OH), Montebrasite (P-1, Z = 4, LiVPO4O), Maxwellite (C2/c, NaVPO4F and HVPO4.OH) and even Talisite (P21/c, NaVPO4O) structures. Their crystallographic arrangements present common features which can be broadly described as vanadium octahedra (VO4Y2) sharing a bridging anion Y in order to form infinite chains Y VO 4 Y   . These chains are connected to each other through PO4 tetrahedra sharing their four oxygen atoms with four vanadium octahedra belonging to three different chains. This 3D framework accommodates Li+ or Na+ in hexagonal channels. The symmetry of vanadium octahedra is dictated by the nature of the Vn+-Y bond. Indeed, in V3+-rich LiVPO4F, NaVPO4F and VPO4·H2O, the vanadium is located on an inversion center of the VO4Y2 octahedra, whereas in V4+-rich NaVPO4O and LiVPO4O a loss of the centrosymmetry of the vanadium environment is observed. Indeed, in the Tavorite-like structure, for an oxidation state of vanadium superior to +3, vanadium likely forms the vanadyl bond resulting from the Jahn Teller activity of V4+ (d1 t2g1eg0). This strongly covalent V=O bond can be formed only with oxygen atoms which are not already involved in a covalent PO4 group. Only the bridging oxygen, Y, fulfils these requirements and, therefore, in the V4+ compounds, an ordering between short and long bonds takes place along the chains (Figure 10). This ordering generates a change of space group (from C2/c for NaVPO4F to P21/c for NaVPO4O) or a doubling of the cell size (Z = 2 for LiVPO4F to Z = 4 for LiVPO4O).
The lithium content in LixVPO4O can vary from 0 to 2, leading to a capacity of 300 mAh/g at an average voltage of ca. 3.1 V allowing the achievement a stable energy density > 900 Wh/kg using surface engineering and nanosizing strategies [9,10,11,12,13]. However, the large difference between the voltage for oxidation of V4+ into V5+ (i.e., 3.95 V vs. Li+/Li) and that for the reduction in V4+ to V3+ (around 2.3 V vs. Li+/Li) makes this multi-electron reaction unsuitable for a real battery system (Figure 11).
In the high voltage region (i.e., 3.0–4.6 V vs. Li+/Li involving the V4+/V5+ redox couple), the oxidation process occurs via a biphasic mechanism between LiVPO4O and VPO4O [14]. The crystal structure of this V5+ phase (ε-VPO4O) is described in a Cc space group allowing the formation of vanadyl bonds appearing as shorter than the ones observed in LiVPO4O (i.e., 1.59 vs. 1.67 Å, Figure 10). Conversely, the antagonist V5+…O bond along the chains elongates from 2.2 Å in LiVPO4O to 2.5 Å in VPO4O leading to an unconventional increase in the cell volume during lithium extraction (ΔV/V = 4.1%) [63]. This VPO4O polymorph can also be obtained while deintercalating the homeotype LiVPO4OH (and also VPO4·H2O), according to an original mechanism [64,65]. Indeed, VPO4OH appears instable vs. LiVPO4OH and VPO4O as this V4+-rich phase is not formed upon Li+ deintercalation from LiVPO4OH. On the contrary, VPO4O is formed showing that the V3+-O/V5+=O redox couple is activated at a constant equilibrium voltage of 3.95 V vs. Li+/Li [65]. Indeed, in the VPO4OH hypothetical phase the competition between the two highly covalent bonds, V4+=O on one side and O-H bond on the other side, would destabilize the VIV-O-H sequence, leading to the concomitant extraction of Li+ and H+ and to the atypical two-electron V3+/V5+=O redox reaction at a constant voltage. Unfortunately, on the contrary to the two-electron reaction observed in LixVPO4O over 3.2 V, which is reversible but not practical, this one observed at a constant high voltage leads to an irreversible phase transformation.
The Li+ insertion within LiVPO4O involves two intermediate phases, Li1.5VPO4O and Li1.75VPO4O, before reaching the Li2VPO4O [66]. Although this V3+-rich composition is described in a triclinic (P-1, Z = 4) system allowing the formation of a vanadyl-type distortion along the chains, the refined V-O distances do not reveal significant differences between them [63], in agreement with the weak Jahn Teller activity of V3+ (d2 t2g2eg0). Lin et al. [67] studied in detail the structural evolutions at the local scale during the lithium insertion in Li1+xVPO4O, and V K-edge EXAFS shows the disappearance of vanadyl bond for Li1.5VPO4O and the persistence of the longer antagonist until Li1.75VPO4O in good agreement with the phase transitions observed (Figure 11).
The investigation of LiVPO4F started in 2003 with a series of studies conducted by Barker and co-workers [60,68,69] who highlighted the promising performance of this material. Indeed, the high voltage delivered for the Lithium extraction (4.25 V vs. Li+/Li for the V3+/V4+ redox voltage, Figure 12) and a capacity very close to the theoretical one even at high C-rate confer to this material a higher practical energy density compared to the ones of commercially available LiFePO4 and LiCoO2 (655 vs. 585 and 525 Wh/kg respectively).
The Lithium extraction from LiVPO4F involves an intermediate phase, Li2/3VPO4F, and then VPO4F through two biphasic reactions. The crystal structure of VPO4F was reported by Ellis et al. [70], its C2/c space group involving centrosymmetric vanadium octahedra with V-F distances of 1.95 Å (Figure 13) whereas the actual nature of the Li2/3VPO4F phase is still unclear, although superstructure peaks have been identified and indexed by doubling the b parameter [71]. This intermediate phase is not formed during discharge where a biphasic reaction between the end-member compositions VPO4F and LiVPO4F takes place [72]. This asymmetric charge/discharge mechanism is not understood at the moment even though it was first attributed by Ellis et al. to the presence of two lithium sites partially occupied (0.8/0.2) in the starting LiVPO4F. Nevertheless, this hypothesis was ruled out later by Ateba Mba et al. [67] who localized Lithium in a single fully occupied site. Piao et al. [73] conducted operando V-K edge XANES in order to probe the redox mechanism during delithiation of LiVPO4F. By a principal component analysis, three components were required to fit the series of spectra recorded upon charge. This might suggest at least a V3+/V4+ ordering for Li2/3VPO4F. The lithium insertion into LiVPO4F occurs at low voltage, typical for the V3+/V2+ redox couple (i.e., 1.8 V vs. Li+/Li) through a biphasic reaction leading to the formation of Li2VPO4F (Figure 12). The structure of Li2VPO4F is described in a C2/c space group with V2+ sitting in a centrosymmetric VO4F2 octahedra with V-F distances at 2.10 Å and equatorial V-O ones at 2.13 Å in average [70] (Figure 13) while Li+ ions are distributed between two 8f Wyckoff sites half occupied in LiO3F2 environments.
Various chemical routes to obtain polycrystalline powders of Tavorite LiVPO4F were reported: sol-gel-assisted carbo thermal reduction (CTR) [74], ionothermal [75]. The majority of these reports highlight the difficulty to obtain pure powders (i.e., without anti-NASICON Li3V2(PO4)3 impurity) or vanadyl-free compounds. Indeed, a series of recent papers demonstrated, by 7Li NMR (and its 2D analogue) and DFT calculations, the presence of various amounts of vanadyl-type defects in crystallographically pure “LiVPO4F” [76,77]. Recently, B. Kang and co-workers [78] reported on an ingenious strategy to avoid the fluorine loss during synthesis, using PTFE as an additional fluorine source. The material thus obtained reveals high electrochemical performance with a stable discharge capacity of 120 mAh/g at 10C over 500 cycles. The same group also published for the first time the electrochemical properties of the mixed valence V3+/V4+ LiVPO4F0.25O0.75 [79]. This strategy aimed at decreasing the difference in voltage between Li insertion and extraction reactions, conferring to the material a high energy density (i.e., 820 Wh/kg) in a reduced voltage range (i.e., 2.0–4.5 V vs. Li+/Li) with the activation of the V3+/V4+ and V4+/V5+ redox couples, respectively. Further investigation of the LiVPO4F-LiVPO4O tie-line has allowed several compositions to stabilize in which the competition between ionicity of the V3+-F bond and covalency of the V4+=O bond distorts the structure, freezes the framework upon Li extraction and hence allows for improved rate capabilities compared with the end-member phases [80,81]. Interestingly, upon Li deintercalation from these materials, the V4+=O/V5+=O redox couple is triggered first before the V3+/V4+ is activated in fluorine rich environments leading to the formation of a mixed valence V3+-V5+ phase at half charge [81]. Although surprising, this redox mechanism is in full agreement with the operating voltage of the end-member phases, the V4+/V5+ redox couple being activated at 3.95 V in LiVPO4O and the V3+/V4+ redox couple at 4.25 V in LiVPO4F due to the absence of vanadyl distortion in LiVPO4F and VPO4F.
Most of the vanadium phosphates discussed above operate at a rather high V3+/V4+ redox voltage, suggesting a massive improvement of the energy density delivered while triggering the V4+/V5+ redox couple. However, materials that operate at such a high V3+/V4+ voltage are usually unable to reversibly exchange several electrons in a narrow enough voltage range. In the following section we will clarify the crystallographic origin of this trend and identify the strategies able to overcome this issue.

5. Towards Reversible High-Voltage Multi-Electron Reactions

Many other vanadium phosphates (as well as pyrophosphates and phosphites, see Table 1) have been stabilized and studied as positive electrode materials for Li(Na)-ion batteries. This article does not aim at providing an exhaustive review of all of them, however careful descriptions of selected systems, provided above, now allow us to generalize and predict part of their properties (especially working voltages, redox mechanisms and structural evolutions) from the only consideration of their crystal structures in their pristine state.
Table 1 highlights the significant divergence in the V3+/V4+ redox voltages which cannot be attributed to the inductive effect, the cation-cation repulsions or even Li site energy, which are the main reported features impacting the voltage in polyanions [103,104]. Indeed, the voltage for the V3+/V4+ redox couple in the Tavorite system LixVPO4Y (with Y = O or F) varies from 2.4 V for Li1+xVPO4O (0 ≤ x ≤ 1) to 4.26 V in Li1-xVPO4F (0 ≤ x ≤ 1). This is attributed to the effect of the highly covalent vanadyl bond which is observed for oxidation states of vanadium strictly superior to 3 in Li1-xVPO4O, Na4VO(PO4)2 … These structures present a common crystallographic feature: at least one oxygen around vanadium is not involved in a covalent P-O bond and hence could be engaged in a vanadyl bond. In the compounds where the VO6 octahedra share all their oxygen atoms with PO4 (or P2O7) groups, the structure of the de-alkalinated V4+ phases are vanadyl free with VO6 octahedra slightly distorted. The corresponding vanadyl free V3+/V4+ redox couple is located at 3.9 V in monoclinic Li3-xV2(PO4)3 and 4.2 V in Li1-xVP2O7, a much higher voltage than the V3+/V4+ couple involved in Li1+xVPO4O polymorphs (around 2.3 V vs. Li+/Li).
Boudin et al. [15] proposed a classification of the vanadium phosphates into three groups according to size of the “clusters” of vanadium polyhedra ([VOx]n with 1 < n < ∞). Although this classification is pertinent to the discussion of the catalytic or magnetic properties of vanadium phosphates, it does not really make sense for a discussion of electrochemical properties. Therefore, we chose to sort these materials considering vanadyl-forbidden (type I and type II) and vanadyl-allowed (type III) structures (summarized in Table 1 and Figure 14):
  • For type I materials (e.g., Li3V2(PO4)3), in which the vanadyl bond cannot appear due to the involvement of each oxygen atom of VO6 octahedra in a PO4-type entity, the typical evolution of the vanadium environment upon oxidation (from V2+ to V5+) follows a quasi-homogeneous shortening of V-O bonds from V2+ to V4+ and a strong increase in VO6 distortion to reach the V5+ state with corresponding voltages of 1.8 V vs. Li+/Li for V2+/V3+, 3.9 V vs. Li+/Li for V3+/V4+ and 4.4 V vs. Li+/Li for V4+/V5+ redox couples (on average for all the type I materials reported in Table 1).
  • In type II materials (e.g., LiVPO4F), at least one of the ligands around vanadium is unshared with a phosphate group and hence would be available to form the vanadyl bond. However, in that case, the nature of this ligand (F instead of O2−) inhibits its formation. From V2+ to V4+, the evolution of the vanadium environment follows a similar trend with slightly higher voltages than for type I due to the higher ionicity of V-F versus V-O. For V5+, for instance in deintercalated Na3V2(PO4)2F3, a “vanadyl-like” distortion appears with V-F bond length of 1.6 Å and 1.9 Å. Such an FV-F sequence has never been reported elsewhere and the precise nature of the V-F bonds formed is still to be clarified.
  • Type III group (e.g., LiVPO4O) gathers the structures having at least one oxygen belonging to VO6 octahedra available to form the covalent vanadyl bond for vanadium oxidation states higher than 3. In this class of materials, the V3+ environments are quasi undistorted. As the oxidation state of vanadium is increased, vanadium leaves the inversion center of the VO6 octahedra in order to form the vanadyl bond. The formation of a distorted VIVO6 octahedra (with typical distances ranging between 1.6 and 2.4 Å along dz2 and quasi equivalent equatorial distances around 2 Å) and VVO5 pyramids (in which the short V=O bond is about 1.6 Å and a shortening of the equatorial distances is observed around 1.8–1.9 Å) are observed. The corresponding voltages appear completely different to those of type I and type II materials: 2.4 V vs. Li+/Li for the V3+-O/V4+=O and 3.95 V vs. Li+/Li for the V4+=O/V5+=O redox couples.
Note that type II materials are crystallographically pseudo type III ones in which the oxygen involved in the vanadyl bond is replaced by Fluorine. Therefore, partial substitution of this fluorine by oxygen leads to mixed type II/III materials—which is actually the case for most of the type III materials, difficult to obtain as vanadyl-free. Extended oxy-fluorine solid solutions were investigated for Na3V2(PO4)2F3−yOy [54,55,57,58] and for LiVPO4F1-yOy [79,80,81,105,106]. The particularity of these compounds resides in the redox paradox of vanadium where the V4+=O/V5+=O is activated at lower voltage than the V3+-F/V4+-F [55,81,107]. Depending on the distribution of ligands around V, it behaves as type II (V3+O4F2), type III (V4+O4O2) or mixed type II/III (V3/4+O4OF) [81]. For this latter environment, the V5+=O vanadyl-like distortion is allowed upon cycling, promoting the reversibility of the process, but is observed at higher voltage than type III materials thanks to the antagonist fluorine. This highlights the importance of the heteroleptic units formed in statistically distributed or in peculiar O/F ordered compounds which are somewhat difficult to obtain due to the different nature of the V4+=O and V3+-F bonds promoting their clusterization [108].
This classification makes further sense regarding the ability of each type of vanadium phosphates to reversibly exchange several electrons per transition metal at high voltage and in a narrow enough voltage range. In type III materials multi-electron redox through V3+-O/V4+=O and V4+=O/V5+=O couples have often been reported [9,10,11,12]. The oxygen atoms unshared with PO4 facilitate the formation of the vanadyl bond allowing for two rather reversible electron processes and thus allow the achievement of cycling stability with energy density higher than 900 Wh/kg [9]. However, this multi-electron reaction cannot be used in a real battery system due to the large voltage difference between the V3+-O/V4+=O and V4+=O/V5+=O (ca. 2.5 V) redox couples. Substituting oxygen by fluorine in such a way to obtain LiVPO4F0.75O0.25 allows raising the voltage of the V3+/V4+ redox and hence reversibly intercalating 1.6 electrons per vanadium in a reduced voltage range [79]. However, this material suffers from rapid capacity fading under such conditions. Since then, the possibility to stabilize multiple compositions along the LiVPO4F-LiVPO4O tie-line has been demonstrated and a systemic investigation of substitution ratio (i.e., y) vs. the voltage range could allow fixing this issue by controlling the Δx in Li1±xVPO4F1−yOy.
In type I and type II materials, while the low voltage V2+/V3+ (≈1.8 V vs. Li+/Li) and high voltage V3+/V4+ (3.9–4.2 V vs. Li+/Li) redox are easily triggered, the V4+/V5+ redox is rarely reported (see Table 1). Moreover, this latter is often kinetically limited and/or irreversible, most probably due to the structural rearrangements required to provide to V5+ cations a satisfying environment. Indeed, the V5+ cations are stable either in a pyramidal ([1+4] coordination) or in a tetrahedral ([2+2] coordination) or even in a very distorted octahedral ([2+2+2] coordination) [15] environments. In each case, at least one covalent vanadyl bond must be formed, but this formation is not privileged by the crystallographic arrangements adopted by type I and II materials. In order to provide to the V5+ cations with a more stable environment than this distorted octahedral one, migration of V5+ in tetrahedral site has been proposed [32]. Therefore, kinetic limitations and/or an irreversible capacity, which can be compensated only at very low voltage (as seen in Na3V2(PO4)2F3, Li3V2(PO4)3, Li5V(PO4)2F2 and Li9V3(P2O7)3(PO4)2), appear. Although V5+ migration in tetrahedral sites has been reported only in Li3V2(PO4)3 so far, analyzing the electrochemical response upon subsequent discharge for other compounds gives insight about the nature of the irreversible reaction taking place. For instance, in Na3V2(PO4)2F3, the Na re-insertion into V2(PO4)F3 (i.e., V4.5+) occurs at 3.9 V vs. Na+/Na in average, until the composition Na2V2(PO4)2F3 (V4.5+ to V3.5+): this voltage range is associated to the Na3(V3+)-Na1(V4+) composition range during the previous charge. The further Na insertion occurs at 1.6 V vs. Na+/Na until the composition Na3V2(PO4)2F3 is recovered. Moreover, the length of this low voltage plateau is proportional to the amount of vanadium oxidized above V4+ during the previous charge. Therefore, this low voltage feature is more likely to correspond to the reduction in V3+ (i.e., ≈1.5 V vs. Na+/Na for V3+/V2+ redox in type I and type II materials) rather than to the reduction in the V4+ into V3+ (i.e., 3.6–3.9 V vs. Na+/Na). This behavior could agree with the presence of V5+ in Td sites. Indeed, as seen in transition metal vanadates used as anode in alkali-ion batteries [109], V5+Td is not reduced above 1.5 V vs. Li+/Li without migrating back in an octahedral site. Therefore, the V3+ reduction would occur at a higher voltage than the V5+Td reduction. The presence of oxygen in the fluorine site would help in accommodating V5+ cation in distorted octahedral site in the charged state. Indeed, it has been shown by theoretical calculations that the partial substitution of fluorine by oxygen in such a way to obtain Na3V2(PO4)2F3−xOx composition tends to decrease the voltage of extraction of the third Na+ cations (from 4.9 V to 4.7 V vs. Na+/Na from pure fluoride to oxy-fluoride) leading to the reversible exchange of more than one electron per vanadium with an excellent rate capability [55].
Finally, this review reveals that the versatility of the vanadium chemistry with a large number of stable oxidation states stabilized in very different environments opens the road towards the formation of new materials whose strains imposed by the crystal field give attractive electrochemical properties. While, in the battery field, the search for new polyanion positive electrode materials slows down for few years, maintaining the efforts towards the stabilization of new phases is crucial. LiVPO4F1–xOx and Na3V2(PO4)2F3–xOx are the only vanadium phosphate oxy-fluorides studied as positive electrode materials and have shown very promising properties. Further playing with anionic substitution, not only with vanadium phosphate oxy-fluorides but also oxy-nitrides (as recently reported with Na3V2(PO3)3N [91]) and even oxy-sulfides, would offer new degrees of flexibility for such versatile polyanion systems and could allow the achievement of high energy density (ca. 1 kWh/kg of active positive electrode material corresponding to ca. 400 Wh/kg at the cell level) through reversible high-voltage multi-electron redox.

6. Conclusions

This review has identified the vanadyl distortion as the main feature governing the operating voltage in vanadium phosphates and their ability to reversibly store several electrons per transition metal. The classification of such materials in three groups, according to the nature of the ligands in the vanadium octahedra and to the distribution of PO4 around them, has allowed to unveil the strategies to increase their energy density. Indeed, anionic substitutions have led to vanadium phosphate oxy-fluorides which allow to combine the beneficial effect of the vanadyl distortion on the reversibility with the high voltage of vanadium redox couples in fluorine rich environments. Further investigation of these anionic substitutions could allow to tend towards reversible high-voltage multi-electron reactions in Alkali-ion batteries.

Funding

The authors acknowledge FEDER, the Reégion Hauts-de-France and the RS2E Network for the funding of EB’s PhD thesis, as well as the financial support of Région Nouvelle Aquitaine, of the French National Research Agency (STORE-EX Labex Project ANR-10-LABX-76-01) and of the European Union’s Horizon 2020 research and innovation program under grant agreement No 875629 (NAIMA project).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Padhi, A.K.; Nanjundaswamy, K.S.; Goodenough, J.B. Phospho-Olivines as Positive-Electrode Materials for Rechargeable Lithium Batteries. J. Electrochem. Soc. 1997, 144, 1188–1194. [Google Scholar] [CrossRef]
  2. Andersson, A.S.; Thomas, J.O.; Kalska, B.; Häggström, L. Thermal Stability of LiFePO4-Based Cathodes. Electrochem. Solid State Lett. 2000, 3, 66–68. [Google Scholar] [CrossRef]
  3. Andersson, A.S.; Kalska, B.; Haggstrom, L.; Thomas, J.O. Lithium Extraction/Insertion in LiFePO4: An X-Ray Diffraction and Mössbauer Spectroscopy Study. Solid State Ion. 2000, 130, 41–52. [Google Scholar] [CrossRef]
  4. Yamada, A.; Chung, S.C.; Hinokuma, K. Optimized LiFePO4 for Lithium Battery Cathodes. J. Electrochem. Soc. 2001, 148, 224–229. [Google Scholar] [CrossRef]
  5. Huang, H.; Yin, S.; Nazar, L.F. Approaching Theoretical Capacity of LiFePO4 at Room Temperature at High Rates. Electrochem. Solid State Lett. 2001, 4, 170–172. [Google Scholar] [CrossRef]
  6. Recham, N.; Chotard, J.-N.; Jumas, J.-C.; Laffont, L.; Armand, M.; Tarascon, J.-M. Ionothermal Synthesis of Li-Based Fluorophosphates Electrodes. Chem. Mater. 2010, 22, 1142–1148. [Google Scholar] [CrossRef]
  7. Barpanda, P.; Ati, M.; Melot, B.C.; Rousse, G.; Chotard, J.-N.; Doublet, M.-L.; Sougrati, M.T.; Corr, S.A.; Jumas, J.-C.; Tarascon, J.-M. A 3.90 V Iron-Based Fluorosulphate Material for Lithium-Ion Batteries Crystallizing in the Triplite Structure. Nat. Mater. 2011, 10, 772–779. [Google Scholar] [CrossRef]
  8. Lv, D.; Bai, J.; Zhang, P.; Wu, S.; Li, Y.; Wen, W.; Jiang, Z.; Mi, J.; Zhu, Z.; Yang, Y. Understanding the High Capacity of Li 2 FeSiO 4: In Situ XRD/XANES Study Combined with First-Principles Calculations. Chem. Mater. 2013, 25, 2014–2020. [Google Scholar] [CrossRef]
  9. Siu, C.; Seymour, I.D.; Britto, S.; Zhang, H.; Rana, J.; Feng, J.; Omenya, F.O.; Zhou, H.; Chernova, N.A.; Zhou, G.; et al. Enabling Multi-Electron Reaction of ε-VOPO4 to Reach Theoretical Capacity for Lithium-Ion Batteries. Chem. Commun. 2018, No. 54, 7802–7805. [Google Scholar] [CrossRef]
  10. Shi, Y.; Zhou, H.; Seymour, I.D.; Britto, S.; Rana, J.; Wangoh, L.W.; Huang, Y.; Yin, Q.; Reeves, P.J.; Zuba, M.; et al. Electrochemical Performance of Nanosized Disordered LiVOPO4. ACS Omega 2018, 3, 7310–7323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Shi, Y.; Zhou, H.; Britto, S.; Seymour, I.D.; Wiaderek, K.M.; Omenya, F.; Chernova, N.A.; Chapman, K.W.; Grey, C.P.; Whittingham, M.S. A High-Performance Solid-State Synthesized LiVOPO4 for Lithium-Ion Batteries. Electrochem. Commun. 2019, 105, 106491. [Google Scholar] [CrossRef]
  12. Chung, Y.; Cassidy, E.; Lee, K.; Siu, C.; Huang, Y.; Omenya, F.; Rana, J.; Wiaderek, K.M.; Chernova, N.A.; Chapman, K.W.; et al. Nonstoichiometry and Defects in Hydrothermally Synthesized ϵ-LiVOPO4. ACS Appl. Energy Mater. 2019, 2, 4792–4800. [Google Scholar] [CrossRef]
  13. Rana, J.; Shi, Y.; Zuba, M.J.; Wiaderek, K.M.; Feng, J.; Zhou, H.; Ding, J.; Tianpin, W.; Cibin, G.; Balasubramanian, M.; et al. Role of Disorder in Limiting the True Multi-Electron Redox in ε-LiVOPO4. J. Mater. Chem. A 2018, 42, 1–10. [Google Scholar] [CrossRef]
  14. Ateba Mba, J.; Masquelier, C.; Suard, E.; Croguennec, L. Synthesis and Crystallographic Study of Homeotypic LiVPO4F and LiVPO4O. Chem. Mater. 2012, 24, 1223–1234. [Google Scholar] [CrossRef]
  15. Boudin, S.; Guesdon, A.; Leclaire, A.; Borel, M.M. Review on Vanadium Phosphates with Mono and Divalent Metallic Cations: Syntheses, Structural Relationships and Classification, Properties. Int. J. Inorg. Mater. 2000, 2, 561–579. [Google Scholar] [CrossRef]
  16. Whittingham, M.S. Ultimate Limits to Intercalation Reactions for Lithium Batteries. Chem. Rev. 2014, 114, 11414–11443. [Google Scholar] [CrossRef] [PubMed]
  17. Croguennec, L.; Palacin, M.R. Recent Achievements on Inorganic Electrode Materials for Lithium-Ion Batteries. J. Am. Chem. Soc. 2015, 137, 3140–3156. [Google Scholar] [CrossRef]
  18. Masquelier, C.; Croguennec, L. Polyanionic (Phosphates, Silicates, Sulfates) Frameworks as Electrode Materials for Rechargeable Li (or Na) Batteries. Chem. Rev. 2013, 113, 6552–6591. [Google Scholar] [CrossRef] [PubMed]
  19. Yabuuchi, N.; Kubota, K.; Dahbi, M.; Komaba, S. Research Development on Sodium-Ion Batteries. Chem. Rev. 2014, 114, 11636–11682. [Google Scholar] [CrossRef] [PubMed]
  20. Shi, H.-Y.; Jia, Z.; Wu, W.; Zhang, X.; Liu, X.-X.; Sun, X. The Development of Vanadyl Phosphate Cathode Materials for Energy Storage Systems: A Review. Chem. A Eur. J. 2020, 26, 8190–8204. [Google Scholar] [CrossRef] [PubMed]
  21. Beneš, L.; Melánová, K.; Svoboda, J.; Zima, V. Intercalation Chemistry of Layered Vanadyl Phosphate: A Review. J. Incl. Phenom. Macrocycl. Chem. 2012, 73, 33–53. [Google Scholar] [CrossRef]
  22. Jian, Z.; Hu, Y.-S.; Ji, X.; Chen, W. NASICON-Structured Materials for Energy Storage. Adv. Mater. 2017, 29, 1601925. [Google Scholar] [CrossRef]
  23. Anantharamulu, N.; Koteswara Rao, K.; Rambabu, G.; Vijaya Kumar, B.; Radha, V.; Vithal, M. A Wide-Ranging Review on Nasicon Type Materials. J. Mater. Sci. 2011, 46, 2821–2837. [Google Scholar] [CrossRef]
  24. Manthiram, A.; Goodenough, J.B. Lithium Insertion into Fe2(SO4)3 Frameworks. J. Power Sources 1989, 26, 403–408. [Google Scholar] [CrossRef]
  25. Huang, H.; Yin, S.C.; Kerr, T.; Taylor, N.; Nazar, L.F. Nanostructured Composites: A High Capacity, Fast Rate Li3V2(PO4)3/Carbon Cathode for Rechargeable Lithium Batteries. Adv. Mater. 2002, 14, 1525–1528. [Google Scholar] [CrossRef]
  26. Patoux, S.; Wurm, C.; Morcrette, M.; Rousse, G.; Masquelier, C. A Comparative Structural and Electrochemical Study of Monoclinic Li3Fe2(PO4)3 and Li3V2(PO4)3. J. Power Sources 2003, 119, 278–284. [Google Scholar] [CrossRef]
  27. Yin, S.-C.; Grondey, H.; Strobel, P.; Anne, M.; Nazar, L.F. Electrochemical Property: Structure Relationships in Monoclinic Li3-YV2(PO4)3. J. Am. Chem. Soc. 2003, 125, 10402–10411. [Google Scholar] [CrossRef] [PubMed]
  28. Yin, S.C.; Strobel, P.S.; Grondey, H.; Nazar, L.F. Li2.5V2(PO4)3: A Room-Temperature Analogue to the Fast-Ion Conducting High-Temperature γ-Phase of Li3V2(PO4)3. Chem. Mater. 2004, 16, 1456–1465. [Google Scholar] [CrossRef]
  29. Rui, X.; Yan, Q.; Skyllas-Kazacos, M.; Lim, T.M. Li3V2(PO4)3 Cathode Materials for Lithium-Ion Batteries: A Review. J. Power Sources 2014, 258, 19–38. [Google Scholar] [CrossRef]
  30. Kang, J.; Mathew, V.; Gim, J.; Kim, S.; Song, J.; Im, W.B.; Han, J.; Lee, J.Y.; Kim, J. Pyro-Synthesis of a High Rate Nano-Li3V2 (PO4)3/C Cathode with Mixed Morphology for Advanced Li-Ion Batteries. Sci. Rep. 2014, 4, 1–9. [Google Scholar] [CrossRef] [Green Version]
  31. Saïdi, M.Y.; Barker, J.; Huang, H.; Swoyer, J.L.; Adamson, G. Performance Characteristics of Lithium Vanadium Phosphate as a Cathode Material for Lithium-Ion Batteries. J. Power Sources 2003, 119, 266–272. [Google Scholar] [CrossRef]
  32. Kim, S.; Zhang, Z.; Wang, S.; Yang, L.; Cairns, E.J.; Penner-Hahn, J.E.; Deb, A. Electrochemical and Structural Investigation of the Mechanism of Irreversibility in Li3V2(PO4)3 Cathodes. J. Phys. Chem. C 2016, 120, 7005–7012. [Google Scholar] [CrossRef] [Green Version]
  33. Rui, X.H.; Yesibolati, N.; Chen, C.H. Li3V2(PO4)3/C Composite as an Intercalation-Type Anode Material for Lithium-Ion Batteries. J. Power Sources 2011, 196, 2279–2282. [Google Scholar] [CrossRef]
  34. Gaubicher, J.; Wurm, C.; Goward, G.; Masquelier, C.; Nazar, L. Rhombohedral Form of Li3V2(PO4)3 as a Cathode in Li-Ion Batteries. Chem. Mater. 2000, 12, 3240–3242. [Google Scholar] [CrossRef]
  35. Delmas, C.; Olazcuaga, R.; Cherkaoui, F.; Brochu, R.; Leflem, G. New Family of Phosphates with Formula Na3M2(PO4)3 (M= Ti,V,Cr,Fe). C. R. Seances Acad. Sci. 1978, 287, 169–171. [Google Scholar]
  36. Chotard, J.-N.; Rousse, G.; David, R.; Mentré, O.; Courty, M.; Masquelier, C. Discovery of a Sodium-Ordered Form of Na3V2(PO4)3 below Ambient Temperature. Chem. Mater. 2015, 27, 5982–5987. [Google Scholar] [CrossRef]
  37. Jian, Z.; Yuan, C.; Han, W.; Lu, X.; Gu, L.; Xi, X.; Hu, Y.; Li, H.; Chen, W.; Chen, D.; et al. Atomic Structure and Kinetics of NASICON NaxV2(PO4)3 Cathode for Sodium-Ion Batteries. Adv. Funct. Mater. 2014, 24, 4265–4272. [Google Scholar] [CrossRef]
  38. Golalakrishnan, J.; Kasthuri Rangan, K. NASICON-Type Vanadium Phosphate Synthesized by Oxidative Deintercalation of Sodium from Sodium Vanadium Phosphate (Na3V2(PO4)3). Chem. Mater. 1992, 4, 745–747. [Google Scholar] [CrossRef]
  39. Lalère, F.; Seznec, V.; Courty, M.; David, R.; Chotard, J.N.; Masquelier, C. Improving the Energy Density of Na3V2(PO4)3-Based Positive Electrodes through V/Al Substitution. J. Mater. Chem. A 2015, 3, 16198–16205. [Google Scholar] [CrossRef]
  40. Mason, C.W.; Gocheva, I.; Hoster, H.E.; Dennis, Y.W. Activating Vanadium’s Highest Oxidation State in the NASICON Structure. ECS Trans. 2014, 58, 41–46. [Google Scholar] [CrossRef]
  41. Aragón, M.J.; Lavela, P.; Ortiz, G.F.; Tirado, J.L. Benefits of Chromium Substitution in Na3V2(PO4)3 as a Potential Candidate for Sodium-Ion Batteries. ChemElectroChem 2015, 2, 995–1002. [Google Scholar] [CrossRef]
  42. Liu, R.; Xu, G.; Li, Q.; Zheng, S.; Zheng, G.; Gong, Z.; Li, Y.; Kruskop, E.; Fu, R.; Chen, Z.; et al. Exploring Highly Reversible 1.5-Electron Reactions (V3+/V4+/V5+) in Na3VCr(PO4)3 Cathode for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2017, 9, 43632–43639. [Google Scholar] [CrossRef]
  43. Liu, R.; Zheng, S.; Yuan, Y.; Yu, P.; Liang, Z.; Zhao, W.; Shahbazian-Yassar, R.; Ding, J.; Lu, J.; Yang, Y. Counter-Intuitive Structural Instability Aroused by Transition Metal Migration in Polyanionic Sodium Ion Host. Adv. Energy Mater. 2020, 2003256, 1–9. [Google Scholar] [CrossRef]
  44. Zhou, W.; Xue, L.; Gao, H.; Li, Y.; Xin, S.; Fu, G.; Cui, Z.; Zhu, Y.; Goodenough, J.B. MV(PO4)3 (M= Mn, Fe, Ni) Structure and Properties for Sodium Extraction. Nano Lett. 2016, 16, 7836–7841. [Google Scholar] [CrossRef]
  45. Zakharkin, M.V.; Drozhzhin, O.A.; Tereshchenko, I.V.; Chernyshov, D.; Abakumov, A.M.; Antipov, E.V.; Stevenson, K.J. Enhancing Na+ Extraction Limit through High Voltage Activation of the NASICON-Type Na4MnV(PO4)3 Cathode. ACS Appl. Energy Mater. 2018, 1, 5842–5846. [Google Scholar] [CrossRef]
  46. Zakharkin, M.V.; Drozhzhin, O.A.; Ryazantsev, S.V.; Chernyshov, D.; Kirsanova, M.A.; Mikheev, I.V.; Pazhetnov, E.M.; Antipov, E.V.; Stevenson, K.J. Electrochemical Properties and Evolution of the Phase Transformation Behavior in the NASICON-Type Na3+xMnxV2-x(PO4)3 (0≤x≤1) Cathodes for Na-Ion Batteries. J. Power Sources 2020, 470, 1–8. [Google Scholar] [CrossRef]
  47. Chen, F.; Kovrugin, V.M.; David, R.; Mentré, O.; Fauth, F.; Chotard, J.N.; Masquelier, C. A NASICON-Type Positive Electrode for Na Batteries with High Energy Density: Na4MnV(PO4)3. Small Methods 2019, 3, 1–9. [Google Scholar] [CrossRef]
  48. Le Meins, J.; Crosnier-Lopez, M.-P.; Hemon-Ribaud, A.; Courbion, G. Phase Transitions in the Na3M2(PO4)2F3 Family (M= Al3+,V3+,Cr3+,Fe3+,Ga3+): Synthesis, Thermal, Structural, and Magnetic Studies. J. Solid State Chem. 1999, 148, 260–277. [Google Scholar] [CrossRef]
  49. Bianchini, M.; Brisset, N.; Fauth, F.; Weill, F.; Elkaim, E.; Suard, E.; Masquelier, C.; Croguennec, L. Na3V2(PO4)2F3 Revisited: A High-Resolution Diffraction Study. Chem. Mater. 2014, 26, 4238–4247. [Google Scholar] [CrossRef]
  50. Shakoor, R.A.; Seo, D.-H.; Kim, H.; Park, Y.-U.; Kim, J.; Kim, S.-W.; Gwon, H.; Lee, S.; Kang, K. A Combined First Principles and Experimental Study on Na3V2(PO4)2F3 for Rechargeable Na Batteries. J. Mater. Chem. 2012, 22, 20535. [Google Scholar] [CrossRef]
  51. Yan, G.; Mariyappan, S.; Rousse, G.; Jacquet, Q.; Deschamps, M.; David, R.; Mirvaux, B.; Freeland, J.W.; Tarascon, J.M. Higher Energy and Safer Sodium Ion Batteries via an Electrochemically Made Disordered Na3V2(PO4)2F3 Material. Nat. Commun. 2019, 10, 1–12. [Google Scholar] [CrossRef] [PubMed]
  52. Bianchini, M.; Fauth, F.; Brisset, N.; Weill, F.; Suard, E.; Masquelier, C.; Croguennec, L. Comprehensive Investigation of the Na3V2(PO4)2F3−NaV2(PO4)2F3 System by Operando High Resolution Synchrotron X-ray Diffraction. Chem. Mater. 2015, 27, 3009–3020. [Google Scholar] [CrossRef]
  53. Broux, T.; Bamine, T.; Simonelli, L.; Stievano, L.; Fauth, F.; Ménétrier, M.; Carlier, D.; Masquelier, C.; Croguennec, L. VIV Disproportionation Upon Sodium Extraction From Na3V2(PO4)2F3 Observed by Operando X-ray Absorption Spectroscopy and State NMR. J. Phys. Chem. C 2017, 121, 4103–4111. [Google Scholar] [CrossRef]
  54. Broux, T.; Bamine, T.; Fauth, F.; Simonelli, L.; Olszewski, W.; Marini, C.; Ménétrier, M.; Carlier, D.; Masquelier, C.; Croguennec, L. Strong Impact of the Oxygen Content in Na3V2(PO4)2F3-YOy (0 ≦ y ≦ 0.5) on Its Structural and Electrochemical Properties. Chem. Mater. 2016, 28, 7683–7692. [Google Scholar] [CrossRef] [Green Version]
  55. Park, Y.U.; Seo, D.H.; Kim, H.; Kim, J.; Lee, S.; Kim, B.; Kang, K. A Family of High-Performance Cathode Materials for Na-Ion Batteries, Na3(VO1-XPO4)2 F1+2x (0 ≤ x ≤ 1): Combined First-Principles and Experimental Study. Adv. Funct. Mater. 2014, 24, 4603–4614. [Google Scholar] [CrossRef]
  56. Mariyappan, S.; Wang, Q.; Tarascon, J.M. Will Sodium Layered Oxides Ever Be Competitive for Sodium Ion Battery Applications? J. Electrochem. Soc. 2018, 165, 3714–3722. [Google Scholar] [CrossRef]
  57. Park, Y.U.; Seo, D.H.; Kwon, H.S.; Kim, B.; Kim, J.; Kim, H.; Kim, I.; Yoo, H.I.; Kang, K. A New High-Energy Cathode for a Na-Ion Battery with Ultrahigh Stability. J. Am. Chem. Soc. 2013, 135, 13870–13878. [Google Scholar] [CrossRef]
  58. Serras, P.; Alonso, J.; Sharma, N.; Miguel, J.; Kubiak, P.; Gubieda, M.-L.; Rojo, T. Electrochemical Na Extraction/Insertion of Na3V2O2x(PO4)2F3−2x. Chem. Mater. 2013, 25, 4917–4925. [Google Scholar] [CrossRef]
  59. Morimoto, H.; Ito, D.; Ogata, Y.; Suzuki, T.; Sakamaki, K.; Tsuji, T.; Hirukawa, M.; Matsumoto, A.; Tobishima, S. Charge/Discharge Behavior of Triclinic LiTiOPO4 Anode Materials for Lithium Secondary Batteries. Electrochem. Soc. Jpn. 2016, 84, 878–881. [Google Scholar] [CrossRef] [Green Version]
  60. Barker, J.; Saidi, M.Y.; Swoyer, J.L. Electrochemical Insertion Properties of the Novel Lithium Vanadium Fluorophosphate, LiVPO4F. J. Electrochem. Soc. 2003, 150, 1394–1398. [Google Scholar] [CrossRef]
  61. Badraoui, A.E.; Pivan, J.-Y.; Maunaye, M.; Pena, O.; Louer, M.; Louer, D. Order-Disorder Phenomena in Vanadium Phosphates. Structures and Properties of Tetragonal and Monoclinic VPO4·H2O. Ann. Chim. Sci. Matériaux 1998, 23, 97–101. [Google Scholar] [CrossRef]
  62. Lii, K.H.; Li, H.; Cheng, C.; Wang, S. Synthesis and Structural Characterization of Sodium Vanadyl (IV) Orthophosphate NaVOPO4. Z. Krist. Cryst. Mater. 2010, 197, 67–73. [Google Scholar] [CrossRef]
  63. Bianchini, M.; Ateba-Mba, J.M.; Dagault, P.; Bogdan, E.; Carlier, D.; Suard, E.; Masquelier, C.; Croguennec, L. Multiple Phases in the ε-VPO4O–LiVPO4O–Li2VPO4O System: A Combined Solid State Electrochemistry and Diffraction Structural Study. J. Mater. Chem. A 2014, 2, 10182–10192. [Google Scholar] [CrossRef]
  64. Song, Y.; Zavalij, P.Y.; Whittingham, M.S. ε-VOPO4: Electrochemical Synthesis and Enhanced Cathode Behavior. J. Electrochem. Soc. 2005, 152, 721–728. [Google Scholar] [CrossRef]
  65. Boivin, E.; Chotard, J.-N.; Ménétrier, M.; Bourgeois, L.; Bamine, T.; Carlier, D.; Fauth, F.; Suard, E.; Masquelier, C.; Croguennec, L. Structural and Electrochemical Studies of a New Tavorite Composition LiVPO4OH. J. Mater. Chem. A 2016, 4, 11030–11045. [Google Scholar] [CrossRef]
  66. Lee Harrison, K.; Bridges, C.; Segre, C.; Varnado, D.; Applestone, D.; Bielawski, C.; Paranthaman, M.; Manthiram, A. Chemical and Electrochemical Lithiation of LiVOPO4 Cathodes for Lithium-Ion Batteries. Chem. Mater. 2014, 26, 3849–3861. [Google Scholar] [CrossRef]
  67. Lin, Y.; Wen, B.; Wiaderek, K.; Sallis, S.; Liu, H.; Lapidus, S.; Borkiewicz, O.; Quackenbush, N.; Chernova, N.; Karki, K.; et al. Thermodynamics, Kinetics and Structural Evolution of ε -LiVOPO4 over Multiple Lithium Intercalation. Chem. Mater. 2016, 28, 1794–1805. [Google Scholar] [CrossRef]
  68. Barker, J.; Saidi, M.Y.; Swoyer, J.L. A Comparative Investigation of the Li Insertion Properties of the Novel Fluorophosphate Phases, NaVPO4F and LiVPO4F. J. Electrochem. Soc. 2004, 151, 1670–1677. [Google Scholar] [CrossRef]
  69. Barker, J.; Gover, R.K.B.; Burns, P.; Bryan, A.; Saidi, M.Y.; Swoyer, J.L. Structural and Electrochemical Properties of Lithium Vanadium Fluorophosphate, LiVPO4F. J. Power Sources 2005, 146, 516–520. [Google Scholar] [CrossRef]
  70. Ellis, B.L.; Ramesh, T.N.; Davis, L.J.M.; Goward, G.R.; Nazar, L.F. Structure and Electrochemistry of Two-Electron Redox Couples in Lithium Metal Fluorophosphates Based on the Tavorite Structure. Chem. Mater. 2011, 23, 5138–5148. [Google Scholar] [CrossRef]
  71. Boivin, E. Crystal Chemistry of Vanadium Phosphates as Positive Electrode Materials for Li-Ion and Na-Ion Batteries. Ph.D. Thesis, University of Picardie Jules Verne, Amiens, France, 2017. [Google Scholar]
  72. Ateba Mba, J.-M.; Croguennec, L.; Basir, N.I.; Barker, J.; Masquelier, C. Lithium Insertion or Extraction from/into Tavorite-Type LiVPO4F: An In Situ X-Ray Diffraction Study. J. Electrochem. Soc. 2012, 159, 1171–1175. [Google Scholar] [CrossRef]
  73. Piao, Y.; Qin, Y.; Ren, Y.; Heald, S.M.; Sun, C.; Zhou, D.; Polzin, B.J.; Trask, S.E.; Amine, K.; Wei, Y.; et al. A XANES Study of LiVPO4F: A Factor Analysis Approach. Phys. Chem. Chem. Phys. 2014, 16, 3254–3260. [Google Scholar] [CrossRef] [PubMed]
  74. Zhong, S.; Chen, W.; Li, Y.; Zou, Z.; Liu, C. Synthesis of LiVPO4F with High Electrochemical Performance by Sol-Gel Route. Trans. Nonferrous Met. Soc. China 2010, 20, 275–278. [Google Scholar] [CrossRef]
  75. Rangaswamy, P.; Shetty, V.R.; Suresh, G.S.; Mahadevan, K.M.; Nagaraju, D.H. Enhanced Electrochemical Performance of LiVPO4F/f-Graphene Composite Electrode Prepared via Ionothermal Process. J. Appl. Electrochem. 2017, 47. [Google Scholar] [CrossRef]
  76. Messinger, R.J.; Ménétrier, M.; Salager, E.; Boulineau, A.; Duttine, M.; Carlier, D.; Ateba Mba, J.-M.; Croguennec, L.; Masquelier, C.; Massiot, D.; et al. Revealing Defects in Crystalline Lithium-Ion Battery Electrodes by Solid-State NMR: Applications to LiVPO4F. Chem. Mater. 2015, 27, 5212–5221. [Google Scholar] [CrossRef] [Green Version]
  77. Bamine, T.; Boivin, E.; Boucher, F.; Messinger, R.J.; Salager, E.; Deschamps, M.; Masquelier, C.; Croguennec, L.; Ménétrier, M.; Carlier, D. Understanding Local Defects in Li-Ion Battery Electrodes through Combined DFT/NMR Studies: Application to LiVPO4F. J. Phys. Chem. C 2017, 121, 3219–3227. [Google Scholar] [CrossRef] [Green Version]
  78. Kim, M.; Lee, S.; Kang, B. Fast-Rate Capable Electrode Material with Higher Energy Density than LiFePO4: 4.2 V LiVPO4F Synthesized by Scalable Single-Step Solid-State Reaction. Adv. Sci. 2015, 3, 1500366. [Google Scholar] [CrossRef] [Green Version]
  79. Kim, M.; Lee, S.; Kang, B. High Energy Density Polyanion Electrode Material: LiVPO4O1-XFx (x ≈ 0.25) with Tavorite Structure. Chem. Mater. 2017, 29, 4690–4699. [Google Scholar] [CrossRef]
  80. Boivin, E.; David, R.; Chotard, J.N.; Bamine, T.; Iadecola, A.; Bourgeois, L.; Suard, E.; Fauth, F.; Carlier, D.; Masquelier, C.; et al. LiVPO4F1-YOy Tavorite-Type Compositions: Influence of the Vanadyl-Type Defects’ Concentration on the Structure and Electrochemical Performance. Chem. Mater. 2018, 30, 5682–5693. [Google Scholar] [CrossRef] [Green Version]
  81. Boivin, E.; Iadecola, A.; Fauth, F.; Chotard, J.N.; Masquelier, C.; Croguennec, L. Redox Paradox of Vanadium in Tavorite LiVPO4F1-YOy. Chem. Mater. 2019, 31, 7367–7376. [Google Scholar] [CrossRef]
  82. Wurm, C.; Morcrette, M.; Rousse, G.; Dupont, L.; Masquelier, C. Lithium Insertion/Extraction into/from LiMX2O7 Compositions (M = Fe, V.; X = P, As) Prepared via a Solution Method. Chem. Mater. 2002, 14, 2701–2710. [Google Scholar] [CrossRef]
  83. Barker, J.; Gover, R.K.B.; Burns, P.; Bryan, A. LiVP2O7: A Viable Lithium-Ion Cathode Material? Electrochem. Solid State Lett. 2005, 8, 446–448. [Google Scholar] [CrossRef]
  84. Deng, C.; Zhang, S. 1D Nanostructured Na7V4(P2O7)4(PO4) as High-Potential and Superior-Performance Cathode Material for Sodium-Ion Batteries. Appl. Mater. Interfaces 2014, 6, 9111–9117. [Google Scholar] [CrossRef] [PubMed]
  85. Kovrugin, V.; Chotard, J.-N.; Fauth, F.; Jamali, A.; David, R.; Christian, M. Structural and Electrochemical Studies of Novel Batteries. J. Mater. Chem. A 2017, 5, 14365–14376. [Google Scholar] [CrossRef]
  86. Kim, J.; Yoon, G.; Kim, H.; Park, Y.U.; Kang, K. Na3V(PO4)2: A New Layered-Type Cathode Material with High Water Stability and Power Capability for Na-Ion Batteries. Chem. Mater. 2018, 30, 3683–3689. [Google Scholar] [CrossRef]
  87. Liu, R.; Liang, Z.; Xiang, Y.; Zhao, W.; Liu, H.; Chen, Y. Recognition of V3+/V4+/V5+ Multielectron Reactions in Na3V(PO4)2: A Potential High Energy Density Cathode for Sodium-Ion Batteries. Molecules 2020, 25, 1000. [Google Scholar] [CrossRef] [Green Version]
  88. Sandineni, P.; Madria, P.; Ghosh, K.; Choudhury, A. A Square Channel Vanadium Phosphite Framework as High Voltage Cathode for Li- and Na- Ion Batteries. Mater. Adv. 2020. [Google Scholar] [CrossRef]
  89. Kuang, Q.; Xu, J.; Zhao, Y.; Chen, X.; Chen, L. Layered Monodiphosphate Li9V3(P2O7)3(PO4)2: A Novel Cathode Material for Lithium-Ion Batteries. Electrochim. Acta 2011, 56, 2201–2205. [Google Scholar] [CrossRef]
  90. Deng, C.; Zhang, S.; Zhao, B. First Exploration of Ultra Fine Na7V3(P2O7)4 as a High-Potential Cathode Material for Sodium-Ion Battery. Energy Storage Mater. 2016, 4, 71–78. [Google Scholar] [CrossRef]
  91. Reynaud, M.; Wizner, A.; Katcho, N.A.; Loaiza, L.C.; Galceran, M.; Carrasco, J.; Rojo, T.; Armand, M.; Casas-Cabanas, M. Sodium Vanadium Nitridophosphate Na3V(PO3)3N as a High-Voltage Positive Electrode Material for Na-Ion and Li-Ion Batteries. Electrochem. Commun. 2017, 84, 14–18. [Google Scholar] [CrossRef]
  92. Boivin, E.; Chotard, J.-N.; Bamine, T.; Carlier, D.; Serras, P.; Veronica, P.; Rojo, T.; Iadecola, A.; Dupont, L.; Bourgeois, L.; et al. Vanadyl-Type Defects in Tavorite-like NaVPO4F: From the Average Long Range Structure to Local Environments. J. Mater. Chem. A 2017, 5, 25044–25055. [Google Scholar] [CrossRef]
  93. Fedotov, S.S.; Khasanova, N.R.; Samarin, A.S.; Drozhzhin, O.A.; Batuk, D.; Karakulina, O.M.; Hadermann, J.; Abakumov, A.M.; Antipov, E.V. AVPO4F (A = Li, K): A 4 V Cathode Material for High-Power Rechargeable Batteries. Chem. Mater. 2016, 28, 411–415. [Google Scholar] [CrossRef] [Green Version]
  94. Zhang, B.; Dugas, R.; Rousse, G.; Rozier, P.; Abakumov, A.M.; Tarascon, J.M. Insertion Compounds and Composites Made by Ball Milling for Advanced Sodium-Ion Batteries. Nat. Commun. 2016, 7, 1–9. [Google Scholar] [CrossRef] [PubMed]
  95. Makimura, Y.; Cahill, L.S.; Iriyama, Y.; Goward, G.R.; Nazar, L.F. Layered Lithium Vanadium Fluorophosphate, Li 5 V(PO 4 ) 2 F 2: A 4 V Class Positive Electrode Material for Lithium-Ion Batteries. Chem. Mater. 2008, 20, 4240–4248. [Google Scholar] [CrossRef]
  96. Liang, Z.; Zhang, X.; Liu, R.; Ortiz, G.F.; Zhong, G.; Xiang, Y.; Chen, S.; Mi, J.; Wu, S.; Yang, Y. New Dimorphs of Na5V(PO4)2F2 as an Ultrastable Cathode Material for Sodium-Ion Batteries. ACS Appl. Energy Mater. 2020, 3, 1181–1189. [Google Scholar] [CrossRef] [Green Version]
  97. Barker, J.; Saidi, M.Y.; Swoyer, J.L. Electrochemical Properties of β-LiVOPO4 Prepared by Carbothermal Reduction. J. Electrochem. Soc. 2004, 151, 796–800. [Google Scholar] [CrossRef]
  98. He, G.; Kan, W.; Manthiram, A. β-NaVOPO4 Obtained by a Low-Temperature Synthesis Process: A New 3.3 V Cathode for Sodium-Ion Batteries. Chem. Mater. 2016, 28, 1503–1512. [Google Scholar] [CrossRef]
  99. He, G.; Bridges, C.A.; Manthiram, A. Crystal Chemistry of Electrochemically and Chemically Lithiated Layered α-LiVOPO4. Chem. Mater. 2015, 27, 6699–6707. [Google Scholar] [CrossRef]
  100. Satya Kishore, M.; Pralong, V.; Caignaert, V.; Malo, S.; Hebert, S.; Varadaraju, U.V.; Raveau, B. Topotactic Insertion of Lithium in the Layered Structure Li4VO(PO4)2: The Tunnel Structure Li5VO(PO4)2. J. Solid State Chem. 2008, 181, 976–982. [Google Scholar] [CrossRef]
  101. Kim, J.; Kim, H.; Lee, S. High Power Cathode Material Na4VO(PO4)2 with Open Framework for Na Ion Batteries. Chem. Mater. 2017, 29, 3363–3366. [Google Scholar] [CrossRef]
  102. Kishore, M.S.; Pralong, V.; Caignaert, V.; Varadaraju, U.V.; Raveau, B. A New Lithium Vanadyl Diphosphate Li2VOP2O7: Synthesis and Electrochemical Study. Solid State Sci. 2008, 10, 1285–1291. [Google Scholar] [CrossRef]
  103. Manthiram, A.; Goodenough, J.B. Lithium Insertion into Fe2(MO4)3 Frameworks: Comparison of M = W with M = MO. J. Solid State Chem. 1987, 71, 349–360. [Google Scholar] [CrossRef]
  104. Ben Yahia, M.; Lemoigno, F.; Rousse, G.; Boucher, F.; Tarascon, J.-M.; Doublet, M.-L. Origin of the 3.6 V to 3.9 V Voltage Increase in the LiFeSO4F Cathodes for Li-Ion Batteries. Energy Environ. Sci. 2012, 5, 9584. [Google Scholar] [CrossRef]
  105. Parapari, S.S.; Ateba Mba, J.M.; Tchernychova, E.; Mali, G.; Arčon, I.; Kapun, G.; Gülgün, M.A.; Dominko, R. Effects of a Mixed O/F Ligand in the Tavorite-Type LiVPO4O Structure. Chem. Mater. 2020, 32, 262–272. [Google Scholar] [CrossRef] [Green Version]
  106. Boivin, E.; Chotard, J.N.; Ménétrier, M.; Bourgeois, L.; Bamine, T.; Carlier, D.; Fauth, F.; Masquelier, C.; Croguennec, L. Oxidation under Air of Tavorite LiVPO4F: Influence of Vanadyl-Type Defects on Its Electrochemical Properties. J. Phys. Chem. C 2016, 120, 26187–26198. [Google Scholar] [CrossRef]
  107. Nguyen, L.H.B.; Iadecola, A.; Belin, S.; Olchowka, J.; Masquelier, C.; Carlier, D.; Croguennec, L. A Combined Operando Synchrotron X-ray Absorption Spectroscopy and First-Principles Density Functional Theory Study to Unravel the Vanadium Redox Paradox in the Na3V2(PO4)2F3−Na3V2(PO4)2FO2 Compositions. J. Phys. Chem. C 2020, 124, 23511–23522. [Google Scholar] [CrossRef]
  108. Bamine, T.; Boivin, E.; Masquelier, C.; Croguennec, L.; Salager, E.; Carlier, D. Local Atomic and Electronic Structure in the LiVPO4(F,O) Tavorite-Type Materials from Solid State NMR Combined with DFT Calculations. Magn. Reson. Chem. 2020, 58, 1109–1117. [Google Scholar] [CrossRef]
  109. Ni, S.; Liu, J.; Chao, D.; Mai, L. Vanadate-Based Materials for Li-Ion Batteries: The Search for Anodes for Practical Applications. Adv. Energy Mater. 2019, 9, 1803324. [Google Scholar] [CrossRef]
Figure 1. Voltage vs. capacity plot for LiFePO4 (purple), LiVPO4F (red), LiVPO4O (blue). Combining the high voltage of LiVPO4F with the high capacity of the multi-electron redox in LiVPO4O could allow the achievement of higher energy density through high voltage multi-electron redox. The dash lines represent constant energy densities in Wh/kg of active positive electrode material.
Figure 1. Voltage vs. capacity plot for LiFePO4 (purple), LiVPO4F (red), LiVPO4O (blue). Combining the high voltage of LiVPO4F with the high capacity of the multi-electron redox in LiVPO4O could allow the achievement of higher energy density through high voltage multi-electron redox. The dash lines represent constant energy densities in Wh/kg of active positive electrode material.
Molecules 26 01428 g001
Figure 2. Stable environments of vanadium according to its oxidation state [15]. The number in square brackets correspond to the number of “equivalent bonds”.
Figure 2. Stable environments of vanadium according to its oxidation state [15]. The number in square brackets correspond to the number of “equivalent bonds”.
Molecules 26 01428 g002
Figure 3. Structural relationship between Nasicon (left) and anti-Nasicon (right) structures adapted from ref. [18].
Figure 3. Structural relationship between Nasicon (left) and anti-Nasicon (right) structures adapted from ref. [18].
Molecules 26 01428 g003
Figure 6. High voltage signature of Nasicon Na3V2(PO4)3 and Na3V1.5Al0.5(PO4)3, adapted from ref. [39]. Reproduced from Ref. [39] with permission from the Centre National de la Recherche Scientifique (CNRS) and The Royal Society of Chemistry.
Figure 6. High voltage signature of Nasicon Na3V2(PO4)3 and Na3V1.5Al0.5(PO4)3, adapted from ref. [39]. Reproduced from Ref. [39] with permission from the Centre National de la Recherche Scientifique (CNRS) and The Royal Society of Chemistry.
Molecules 26 01428 g006
Figure 7. Structure of Na3V2(PO4)2F3 (left) and sodium distribution (right) [49].
Figure 7. Structure of Na3V2(PO4)2F3 (left) and sodium distribution (right) [49].
Molecules 26 01428 g007
Figure 8. Galvanostatic electrochemical voltage-composition data of Na3V2(PO4)2F3 at C/10 per exchanged ion and the corresponding derivative curve in the 3.0–4.4 V or 3.0–4.8 V voltage range adapted from ref. [51]. Reproduced with permission from Guochan Yan et al., Nature Communications; published by Springer Nature, 2019.
Figure 8. Galvanostatic electrochemical voltage-composition data of Na3V2(PO4)2F3 at C/10 per exchanged ion and the corresponding derivative curve in the 3.0–4.4 V or 3.0–4.8 V voltage range adapted from ref. [51]. Reproduced with permission from Guochan Yan et al., Nature Communications; published by Springer Nature, 2019.
Molecules 26 01428 g008
Figure 9. Evolution of vanadium environments and Na/vacancies ordering upon cycling of Na3V2(PO4)2F3 [51,52,54].
Figure 9. Evolution of vanadium environments and Na/vacancies ordering upon cycling of Na3V2(PO4)2F3 [51,52,54].
Molecules 26 01428 g009
Figure 10. Structural relationships between different Tavorite-type materials.
Figure 10. Structural relationships between different Tavorite-type materials.
Molecules 26 01428 g010
Figure 11. Voltage profile of LixVPO4O cycled between 3.0–4.6 V vs. Li+/Li (left) and between 3.0 and 1.5 V vs. Li+/Li in GITT mode (right) adapted from ref. [14]. Reproduced with permission from Ateba Mba et al., Chemistry of Materials; published by American Chemical Society, 2012.
Figure 11. Voltage profile of LixVPO4O cycled between 3.0–4.6 V vs. Li+/Li (left) and between 3.0 and 1.5 V vs. Li+/Li in GITT mode (right) adapted from ref. [14]. Reproduced with permission from Ateba Mba et al., Chemistry of Materials; published by American Chemical Society, 2012.
Molecules 26 01428 g011
Figure 12. Voltage profile of LixVPO4F cycled between 3.0–4.6 V vs. Li+/Li (left) and between 3.0 and 1.5 V vs Li+/Li in GITT mode (right) adapted from ref. [14]. Reproduced with permission from Ateba Mba et al., Chemistry of Materials; published by American Chemical Society, 2012.
Figure 12. Voltage profile of LixVPO4F cycled between 3.0–4.6 V vs. Li+/Li (left) and between 3.0 and 1.5 V vs Li+/Li in GITT mode (right) adapted from ref. [14]. Reproduced with permission from Ateba Mba et al., Chemistry of Materials; published by American Chemical Society, 2012.
Molecules 26 01428 g012
Figure 13. Structural evolution during Lithium extraction/insertion from/into LiVPO4F [72] and LiVPO4O [63,67].
Figure 13. Structural evolution during Lithium extraction/insertion from/into LiVPO4F [72] and LiVPO4O [63,67].
Molecules 26 01428 g013
Figure 14. Typical evolution of vanadium environments according to the oxidation state of vanadium for type I, type II and type III materials.
Figure 14. Typical evolution of vanadium environments according to the oxidation state of vanadium for type I, type II and type III materials.
Molecules 26 01428 g014
Table 1. List of the vanadium phosphate, pyro-phosphate and phosphite materials with their redox voltage and corresponding practical capacity based on vanadium redox. More details about the classification of these materials (Type I, II or III) are provided in the text and at the Figure 14.
Table 1. List of the vanadium phosphate, pyro-phosphate and phosphite materials with their redox voltage and corresponding practical capacity based on vanadium redox. More details about the classification of these materials (Type I, II or III) are provided in the text and at the Figure 14.
As Synthetized
Compositions
Initial Vn+M/P
Ratio
V2+/V3+V3+/V4+V4+/V5+Ref.
E (V vs. Li+/Li)Capacity (mAh/g)E (V vs. Li+/Li)Capacity
(mAh/g)
E (V vs. Li+/Li)Capacity
(mAh/g)
Type I Materials
Na3V2(PO4)3V3+0.671.9 *593.7 *118//[39]
Na3V1.5Al0.5(PO4)3V3+0.671.9 *603.7 *854.3 *28[39]
Na3VCr(PO4)3V3+0.67//3.7 *604.4 *50[42]
Na4VMn(PO4)3V3+0.67//3.7 *604.2 *50[47]
r-Li3V2(PO4)3V3+0.67//3.7131//[34]
m-Li3V2(PO4)3V3+0.671.81313.91314.533[27]
LiVP2O7V3+0.521164.395//[82,83]
Na7V4(P2O7)3(PO4)2V3+0.5//4.2 *90//[84]
Na7V3Al1(P2O7)3(PO4)2V3+0.5//4.2 *774.546[85]
Na3V(PO4)2V3+0.5//3.8 *904.4 *20[86,87]
LiV(HPO3)2V3+0.5//4.175//[88]
Li9V3(P2O7)3(PO4)2V3+0.375//3.7554.555[89]
Na7V3(P2O7)4V3+0.375//4.3 *80//[90]
Na3V(PO3)3NV3+0.33//4.3 *74//[91]
Type II Materials
LiVPO4FV3+11.81564.2156//[14]
NaVPO4FV3+1//≈4.2 *20//[92]
LiVPO4OHV3+11.4155////[65]
(Li,K)VPO4FV3+1//4.0110//[93]
Na3V2(PO4)2F3V3+0.671.5 *644.0 *64≈ 4.8 *64[51,53,94]
Li5V(PO4)2F2V3+0.5//4.15854.785[95]
t-Na5V(PO4)2F2V3+0.5//3.7 *62//[96]
o-Na5V(PO4)2F2V3+0.5//3.9 *65//[96]
Type III Materials
α-LiVPO4OV4+1//2.41553.95150[14]
β-LiVPO4OV4+1//2.21554130[97]
β-NaVPO4OV4+1////3.6 *58[98]
γ-LiVPO4OV4+1//2804150[99]
Li4VO(PO4)2V4+0.5//2944.194[100]
Na4VO(PO4)2V4+0.5////3.8 *77[101]
Li2VOP2O7V4+0.5////4.164[102]
The voltage values are reported vs. Li+/Li even for those obtained in Na-cell (according to E(Na+/Na) = 0.3 V vs. Li+/Li), in that case the voltage is marked by *
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Boivin, E.; Chotard, J.-N.; Masquelier, C.; Croguennec, L. Towards Reversible High-Voltage Multi-Electron Reactions in Alkali-Ion Batteries Using Vanadium Phosphate Positive Electrode Materials. Molecules 2021, 26, 1428. https://doi.org/10.3390/molecules26051428

AMA Style

Boivin E, Chotard J-N, Masquelier C, Croguennec L. Towards Reversible High-Voltage Multi-Electron Reactions in Alkali-Ion Batteries Using Vanadium Phosphate Positive Electrode Materials. Molecules. 2021; 26(5):1428. https://doi.org/10.3390/molecules26051428

Chicago/Turabian Style

Boivin, Edouard, Jean-Noël Chotard, Christian Masquelier, and Laurence Croguennec. 2021. "Towards Reversible High-Voltage Multi-Electron Reactions in Alkali-Ion Batteries Using Vanadium Phosphate Positive Electrode Materials" Molecules 26, no. 5: 1428. https://doi.org/10.3390/molecules26051428

Article Metrics

Back to TopTop