D.
Liu
a,
W.
Zhu
a,
J.
Trottier
a,
C.
Gagnon
a,
F.
Barray
a,
A.
Guerfi
a,
A.
Mauger
b,
H.
Groult
b,
C. M.
Julien
c,
J. B.
Goodenough
d and
K.
Zaghib
*a
aEnergy Storage and Conversion, Research Institute of Hydro-Québec, 1800 Bd Lionel-Boulet, Varennes, Québec J3X 1S1, Canada. E-mail: karim.zaghib@ireq.ca
bUniversité Pierre et Marie Curie – Paris6, Institut de Minéralogie et Physique de la Matière Condensée (IMPMC), 4 place Jussieu, 75005 Paris, France
cUniversité Pierre et Marie Curie – Paris6, Physicochimie des Electrolytes, Colloïdes et Sciences Analytiques (PECSA), 4 place Jussieu, 75005 Paris, France
dThe University of Texas at Austin, 1 University Station C2220, Austin, USA
First published on 6th November 2013
The success of lithium-ion batteries in small-scale applications translates to large-scale applications, with an important impact in the future of the environment by improving energy efficiency and reduction of pollution. In this review, we present the progress that allows lithium-insertion compounds with the spinel structure to become the active cathode element of a new generation of Li-ion batteries, namely the 5 V cathodes, which promise to improve the technologies of energy storage and electric transportation, thereby addressing the replacement of the gasoline engine and the increasing demand for green energy power sources. The compounds considered here include the spinel LiNi0.5Mn1.5O4 and its related Cr-doped structure. Emphasis is placed on the control of physical properties that is needed to guarantee the reliability and the optimum electrochemical performance of these materials as the active cathode element of Li-ion batteries. We also report the structural evolution of the spinel phase in both charge (Li extraction) and discharge (Li insertion) reactions.
Terada et al.15 revealed that the plateau at around 4.7 V is due to the Ni2+/Ni4+ redox, whereas a small plateau at 4.1 V arises from the Mn3+/Mn4+ redox couple, as observed by in situ X-ray absorption fine-structure spectroscopy (XAFS) analysis for the Li1−xNi0.31Mn1.69O4 material. For the ideal LiNi0.52+Mn1.54+O4 composition, i.e. in the absence of any deviation from stoichiometry, the oxidation state of Mn should be fixed at +4, resulting in only the Ni2+/Ni4+ redox couple during the charge–discharge process. The access to two nickel redox couples, Ni3+/Ni2+ and Ni4+/Ni3+, at ∼4.7 eV below the Fermi energy of a lithium metal anode with a negligible voltage step between the two couples leads to a theoretical gravimetric capacity of 146.7 mA h g−1. At high potential, LNM delivers an energy density equivalent to ∼650 W h kg−1 of active material. This value is the highest among commercially available cathode materials such as LiCoO2 (518 W h kg−1), LiMn2O4 (400 W h kg−1), LiFePO4 (495 W h kg−1) and LiCo1/3Ni1/3Mn1/3O2 (576 W h kg−1). Advantages and drawbacks of LiNi0.5Mn1.5O4 have been pointed out: this cathode material includes inexpensive and environmentally benign manganese and it has high electronic and Li+ ion conductivities, excellent rate capability and good safety, but a severe capacity fade at elevated temperatures (∼60 °C) is the main disadvantage.16
The first member of the Li–Mn oxide family that has been commercialized is LiMn2O4, which crystallizes in the spinel structure (S.G. Fd
m). Li and Mn cations occupy tetrahedral (8a) and octahedral (16d) sites, respectively, in a cubic close-packed array of oxygen atoms (32e sites). LiNi0.5Mn1.5O4 crystallizes in two possible crystallographic structures: the face-centered spinel (S.G. Fd
m), called “disordered spinel”, and the simple cubic phase (S.G. P4332) called “ordered spinel” (Table 1).17–19
m S.G.) and ordered (P4332 S.G.) phase
| Space group | Atom | Wyckoff position | x | y | z |
|---|---|---|---|---|---|
Fd m |
Li | 8a | 1/8 | 1/8 | 1/8 |
| Ni | 16d | 1/2 | 1/2 | 1/2 | |
| Mn | 16d | 1/2 | 1/2 | 1/2 | |
| O | 32e | 0.263 | 0.263 | 0.263 | |
| P4332 | Li | 8c | 0.012 | 0.012 | 0.012 |
| Ni | 4a | 5/8 | 5/8 | 5/8 | |
| Mn | 12d | 1/8 | 0.3791 | −0.1291 | |
| O1 | 8c | 0.3863 | 0.3863 | 0.3863 | |
| O2 | 24e | 0.1492 | −0.1467 | 0.1313 |
The diffraction patterns of the cubic P4332 symmetry is characterized by additional weak Bragg lines located at 2θ = 15.3, 39.7, 45.7 and 57.5° due to the 1
:
3 ordering of the Ni and Mn cations.17,19,20 The cubic cell parameter falls from a = 8.243 Å for LiMn2O4 to a = 8.1685 Å for LiNi0.5Mn1.5O4. In the ordered phase, the larger Ni2+ ions (ionic radius 0.69 Å) occupy only the 4b sites that give more room than the 16d sites of the normal spinel structure. The cation distribution in the P4332 symmetry is then Li on 8c, Ni on 4b, Mn on 12d, and O(1) and O(2) oxygen ions occupy the 24e and 8c Wyckoff positions, respectively. The net result is thus a significant optimisation of space occupation leading to a reduced unit cell volume. Oxygen loss in the LNM framework leads to Mn3+ generated to keep the electric neutrality, and the larger ionic radius of Mn3+ (0.645 Å) compared to Mn4+ (0.530 Å) results in a larger cell volume (Fig. 1).
![]() | ||
Fig. 1 The structure of spinel LiNi0.5Mn1.5O4 (Fd m S.G.). Oxygen deficiency partially reduces Mn4+ to Mn3+ to satisfy local charge neutrality. | ||
The present review gives the state-of-the-art understanding of the properties of LNM and Cr-doped LNM materials. Owing to the evolution in this field, these compounds are promising active 5 V cathode elements for the next generation of Li-ion batteries to improve the technology of the energy storage and electric transportation. This progress is the fruit of about a decade of intensive research in the electrochemical community during which chemists, electrochemists, and physicists have added their efforts to understand the properties of the materials, although some obstacles still stay in the way before 5 V batteries can find their commercial use for worldwide applications in the industry.
m.22,26 During the synthesis of the LiNi0.5Mn1.5O4, however, the high calcination temperature sometimes leads to the reduction of the Mn oxidation state from +4 to +3. To cure this oxygen deficiency, some authors employed an annealing process at 700 °C in air after the high-temperature calcination at temperatures as high as 1000 °C, or 500 °C in oxygen atmosphere at high pressure.27,28 The resulting powders delivered flat voltage profiles at around 4.7 V. Idemoto et al.27 reported that LiNi0.5Mn1.5O4 synthesized under O2 atmosphere has the cubic spinel structure with a space group of P4332 instead of Fd
m, as determined by neutron diffraction. Another approach to tune the ordered/disordered phase is an acute control of the cooling rate immediately after high temperature calcinations.29
:
4 molar ratio); this solution heated to 140 °C enables the chelation (reaction of functional carboxyl group of acid with metal ions) process for the esterification of acid with ethylene glycol. The final product is composed by primary crystallites of 50–70 nm size with a surface area of 15.6 m2 g−1.53
The co-precipitation synthetic route provides good crystallization of the Fd
m phase with disordered Ni and Mn.54 This method is described as follows. Stoichiometric amounts of manganese, lithium, and nickel acetates (99% Aldrich) were first dissolved in distilled water and stirred for 1 h. Oxalic-acid solution was then added dropwise under stirring to obtain a green precipitate. The molar ratio of oxalic acid to metal ions was controlled to be 1
:
1. The precipitate solution was continually stirred for 1 h before being dried at 50 °C overnight with constant stirring. The dried precipitate was preheated at 500 °C for 6 h and then ground for 30 min. The preheated powder was pressed into pellets (20 mm dia. and ca. 5 mm thick) and annealed at 900 °C for 24 h in air for better crystal growth. Fig. 2 shows typical SEM and TEM images used to characterize the crystallite size and surface morphology of LNM spinel particles. The samples are not only well crystallized, but the powder particles consist of many small grains, most of the observed grains are >0.2 μm without serious agglomeration after annealing at 900 °C for 24 h in air (Fig. 2a). The grains are well crystallized with sharp corners. Some of the grains are thick disks. However, the TEM images (Fig. 2b and c) indicate a polycrystalline microstructure within the particles with nano-size domains of different orientations, some disordered areas, and bent lattice fringes.
L = 0.9λ/B cos θ, | (1) |
![]() | ||
| Fig. 3 (a) XRD patterns of the spinel samples at different steps of the synthesis in the temperatures range 700–900 °C before annealing, and for the final sample, post-annealed at 600 °C for 48 h. Inset shows the magnified Bragg lines between 2θ = 35–45° with impurity peaks marked by arrows (b) Rietveld refinement profiles of the XRD data for the LNM electrode film (Si peak comes from sample holder). Reproduced with permission.64 | ||
Note that similar results were obtained for LiNi0.45Mn1.45Cr0.1O4. The same post-annealing strategy at 600 °C as in the case of pristine LNM was used to modify the oxygen deficiency during synthesis of the LiNi0.45Mn1.45Cr0.1O4 cathode for lithium-ion batteries. Structural analyses revealed that post-annealing is an effective way to eliminate the impurity phase without changing the space group. In addition, we found that substitution of Cr3+ for Ni2+ and Mn4+ not only helps to keep the Mn4+ oxidation state unchanged (2Cr3+ = Ni2+ + Mn4+), but also introduces greater disorder of the B-site cations in LNM,62 as we shall see in Section 3.2. This feature will explain why the small amount of Cr substituted for Ni and Mn leads to better rate performance along with cyclability at room temperature, as we shall see in Section 5. The variation with thermal treatment of the particle size and morphology were examined by SEM.64 Images of the disordered LiNi0.45Mn1.45Cr0.1O4 (Fig. 4) represent the well-crystallized material with a particle size in the range of 0.5–2 μm. The micrographs reveal that the post-annealing at 600 °C promotes well-faced grains of regular shape characteristic of the cubic spinel morphology.
The degree of Ni/Mn ordering in the surface layers of the three samples was also investigated with Raman scattering spectroscopy.38,54,68Fig. 5b shows a typical Raman spectrum of crystallized Li[Ni0.5Mn1.5]O4 sample. The octahedral Mn(Ni)O6 structural units have Oh symmetry where the 625 cm−1 peak assigned to the symmetric Mn–O stretching vibration of the MnO6 octahedra in Li[Mn2]O4 is shifted to 635 cm−1. The new features at 399 ± 1 and 490 ± 2 cm−1 are strong and can, therefore, be assigned unequivocally to the Ni–O stretching mode. Two effects contribute to the frequency shift of the Mn–O stretching mode: (a) the increased mean valence state of the Mn ions and (2) a smaller unit-cell volume. None of the peaks characteristic [18] of ordering of the Ni(II) and Mn(IV) in space group P4332 (or P4132), viz. at 218, 237 and 607 cm−1, were detected, which is another indication that the Ni(II) and Mn(IV) of LNM samples were disordered over the 16d sites of the spinel.
Although the vibrational spectroscopy proved very useful to evaluate the degree of cation disorder, it failed to give details on the concentration of oxygen vacancies. In particular, the Mn3+–O stretching mode well observed at circa 518 cm−1 in LiMn2O4 is not detected in our samples. The determination of the concentration of Mn3+ ions then requires another means of investigation. We show in the next section that the investigation of magnetic properties fulfils this purpose.
![]() | ||
Fig. 6 (a) Inverse of the magnetic susceptibility H/M measured at H = 10 kOe, for the LiNi0.45Mn1.45Cr0.1O4 samples after post annealing at 600 °C and the LiNi0.45Mn1.45Cr0.1O4 crystallized in the same Fd m space group. (b) Magnetization curves of the LiNi0.5Mn1.5O4 sample crystallized in the Fd m space group. Reproduced with permission.64 | ||
Electron transfer from the half-filled Ni2+ σ-bonding eg orbital to a half-filled Mn4+ π-bonding tg orbital in a 90° Ni2+–O–Mn4+ interaction via a common pσπ orbital is restricted by the Pauli exclusion principle to give an antiferromagnetic Ni2+–Mn4+ interaction.71 Ordering of the Ni2+ and Mn4+ ions in the P4332 Li[Ni0.5Mn1.5]O4 long-range-ordered phase has been shown to give a ferrimagnetic phase at TC with antiferromagnetic coupling between the Ni2+ and Mn4+ sublattices.19 Frustrated magnetic interactions in a completely disordered phase would give a much lower magnetic ordering temperature than in the ordered phase. Therefore, the existence of long-range magnetic order below a TC similar to that of the ordered P4332 phase is a signature of a strong correlation function Cij = 〈P(Mn)iP(Ni)j〉 with P(Mn)iP(Ni)j the probability that if site i is occupied by Mn, site j is occupied by Ni and Cij is large if i and j are nearest 16d-sites. The fact that the peaks of the XRD spectra reported above are well-described in the framework of the disordered phase means that the correlation length Cij is smaller than the length scale probed by XRD, typically a few nm (only an analysis of the diffusive X-ray scattering would give access to the correlation function at shorter length scale). The FTIR in the previous section, which is a probe at the molecular scale, already gave evidence of short-range ordering at this scale. The short-range order is confirmed by the magnetic experiments, which are a probe at the atomic scale and show that the crystal has a large degree of order at the scale of the nearest neighbours.
For long-range order, the saturation magnetization at low temperature results from the difference between the magnetic moment carried by Mn4+ and Ni2+. The orbital moment at Mn4+ and Ni2+ is quenched by the crystal field, which makes a spin-only atomic magnetic moment a good approximation. The spins of Mn4+ and Ni2+ are S = 3/2 and S = 1, respectively, so the magnetic moment at saturation for the ordered LiNi0.5Mn1.5O4 should be (3 × 1.5 − 2 × 0.5) = 3.50 μB per formula unit. The experimental value of the saturation magnetization 3.40 μB/formula unit at 4.2 K of the phase (Fig. 6b) is in reasonable agreement with the theoretical value for a fully ordered phase and shows the absence of Ni3+ ions that would have raised the magnetic momentum to a larger value, as it has been observed in some cases in LNM, ordered or not. The saturation magnetization in the LiNi0.45Mn1.45Cr0.1O4 sample is only reduced to 3.2 μB per formula unit. This result suggests that the Cr3+ are not distributed randomly, but tend to form antiferromagnetically coupled dimers randomly distributed in the lattice. This is actually expected and implicit in the notation 2Cr3+ = Ni2+ + Mn4+. Two Cr3+ on nearest 16d sites form the most favourable configuration to lower the energy for two reasons: first, 2Cr3+ insures charge neutrality at the molecular scale when substituting for Ni2+ + Mn4+, thus minimizing the cost in Coulomb energy; second, the ionic radii r satisfy approximately the relation 2r(Cr3+) ≈ r(Mn4+) + r(Ni2+), so that this substitution also minimizes the lattice distortion. The concentration of Cr is only 5% of the metal ions, i.e. far smaller than the percolation threshold of the 16d-site sublattice, and thus too small to destroy the long-range ferrimagnetic ordering at finite temperature; it is, however, responsible for a decrease in the short-range Mn4+, Ni2+ order and, therefore of the Néel temperature by 12 K that is observed with respect to LiMn1.5Ni0.5O4. In addition, if we subtract the magnetic moment carried by the Cr3+, 0.3 μB per formula, from the Mn and Ni contribution to the magnetic moment at saturation in the antiferromagnetic phase of LiNi0.45Mn1.45Cr0.1O4, we find 1.45 × 3 − 0.45 × 2 − 0.3 = 3.15 μB in agreement with the experimental value 3.2 μB.
Fig. 7 reports the schematic density of states and Fermi energies for the LixNi0.5−yMn1.5−yCr2yO4 spinel cathode. In this case, it may be possible to access two formal valences on the active cation without a voltage step on passing from one formal valence state to the next. However, when EFC of the host falls below the HOMO of the electrolyte, a passivating, Li permeable SEI layer must form on the surface of the active particle if a reversible reaction is to be obtained. If the passivating SEI layer is not permeable to Li+ ions, it blocks insertion/extraction into/from the host. The HOMO of a liquid carbonate electrolyte is about 4.3 eV below the Fermi energy EFA of a lithium anode, and the decomposition voltage of the electrolyte is at about 5 eV below it. Access to Ni(III) and Ni(IV) valence states is possible in cation disordered LiNi0.5Mn1.5O4 spinel; the SEI layer formed at voltage V >4.3 V is self-limiting and Li-permeable.
![]() | ||
| Fig. 7 Schematic density of states and Fermi energies for LixNi0.5−yMn1.5−yCr2yO4 spinel cathode. The Li permeable SEI layer formed on the electrode surface preserves the overall reversible reaction. | ||
![]() | ||
| Fig. 8 Typical voltage profile of charge of a Li//Cr-doped LNM cell showing the various redox reactions Mn3+/4+ (at 4.1 V), Ni2+/4+ (at 4.7 V) and Cr3+/4+ (at 4.8 V). | ||
In the case of Cr-doped spinel, the last step of the charge, after all the Mn have been oxidized to Mn4+ and Ni to Ni4+, proceeds with the oxidation of Cr3+ ions at an upper voltage plateau according to the reaction:
![]() | (2) |
Fig. 9 shows the change in the occupation of the energy levels in LixNi0.5−yMn1.5−yCr2yO4 spinel during the charge process (x ≈ 0.08) according to this model. Considering the electronic states, the energy difference ΔE = 0.5 eV between the Mn eg level and the Cr t2g level corresponds to the potential difference between the two plateaus of the discharge profile for the Li//LiMn2−yCryO4 cell (see Fig. 10).
:
1 ethylene carbonate–diethylene carbonate (EC–DEC) as electrolyte. Fig. 10 shows the initial charge–discharge cycle at a low rate of C/5 for both cathode materials. The characteristic 4.1 V Mn3+/Mn4+ redox couple is always observed in the pristine or metal-doped LNM cathodes before post-annealing as a result of oxygen loss at high-temperature synthesis.61–63
However, in Fig. 10 obtained for the samples post-annealed at 600 °C, no obvious 4.1 V step is detected in LiNi0.45Mn1.45Cr0.1O4 spinel (red curves), confirming that most of the residual Mn3+ ions have been re-oxidized to Mn4+ after re-annealing at 600 °C in agreement with the analysis of magnetic properties. This is also consistent with the Rietveld refinement of the XRD spectra. However, a small anomaly near 4.1 V can be detected on the discharge curves, pronounced in the case of the un-doped LNM, which suggests that Cr-doping can reduce, but not avoid entirely a loss of oxygen with time during cycling. We shall return to this problem in Section 7. In addition, LiNi0.45Mn1.45Cr0.1O4 shows two distinct plateaus at around 4.7 V. In contrast, the un-doped sample only exhibits a single flat voltage profile at about 4.7 V. No obvious great change was found after 100 cycles, suggesting good reversibility for both samples.
In order to understand the difference in the electrochemical properties of the LiNi0.5Mn1.5O4 and LiNi0.45Mn1.45Cr0.1O4 spinel cathodes, Fig. 11 compares the incremental capacity dQ/dV vs. V graphs, where Q = ∫Idt from t = 0 at 3.5 V to t at V–3.5 calculated from curves in Fig. 10. Removal of Li from the tetrahedral sites of the spinel LNM framework initially probes the oxidation reaction of Ni2+ → Ni3+ just below 4.7 V (typically ∼4.69 V) for the disordered Fd
m and above 4.7 V (typically ∼4.72 V) for the ordered P4332 spinels.68 Ordering of the Ni and Mn raises by ∼0.02 eV the V(x) profile of LNM. From Fig. 11a, two anodic peaks at 4.663 and 4.731 V plus two cathodic peaks at 4.638 and 4.704 V were observed for the LiNi0.45Mn1.45Cr0.1O4, which is in agreement with two voltage plateaus for disordered LNM.20,68 Kim et al. suggested that, as the crystallographic structure changed from Fd
m to P4332, the voltage gaps between the two plateaus became narrower at around 4.75 V and resulted in a flatter voltage profile.67 This separation is known to decrease from about 60 mV to around 20 mV, depending on the degree of ordering.66–68 The average potential difference, ΔV, between these two peaks is ca. 60 mV for the Cr-doped LNM material and 30 mV for the commercial sample. Therefore these results corroborate the much stronger cation disorder induced by the Cr-doping. Additionally, small redox peaks were observed in the LiMn1.45Cr0.1Ni0.45O4 at about 4.85 V, showing the redox reaction between Cr3+ and Cr4+ and confirming the electrochemical activity of Cr in the Cr-substituted LNM.
![]() | ||
| Fig. 11 Differential capacity curves, dQ/dV vs. V, of the un-doped LNM (top) and LiNi0.45Mn1.45Cr0.1O4 (bottom). The values at the peaks are given in volt. Reproduced with permission.64 | ||
The smaller ΔV in un-doped LNM compared to that of LiNi0.45Mn1.45Cr0.1O4 suggests faster lithium insertion/extraction kinetics in the former.72 This is, however, controversial because it is well known that the cation disorder and the decrease of the concentration of defects such as oxygen vacancies is beneficial to the electrochemical performance. Therefore, the LiNi0.45Mn1.45Cr0.1O4 should have the best properties, and actually, this was the motivation for Cr-doping. Some insight into the insertion/extraction of Li is provided by the modified Peukert plot in Fig. 12 for both un-doped and Cr-doped LNM cathodes. The capacity loss after 25 cycles at 5 C rate is 32.4% and 16.3%, respectively. It is necessary to note that only 6% of conductive carbon was used to make the cathode electrode in this work.
![]() | ||
Fig. 12 Modified Peukert curves of LNM and LiNi0.45Mn1.45Cr0.1O4 spinel cathodes cycled between 3.5 and 4.9 V in CR2032 coin cell with LiPF6 in EC–DEC (1 : 1) as electrolyte. | ||
From Fig. 12, the LiNi0.45Mn1.45Cr0.1O4 exhibits better rate capability and capacity retention than LiNi0.5Mn1.5O4. For instance, the LiNi0.45Mn1.45Cr0.1O4 delivered a reversible capacity of ∼115, 104, 95 and 40 mA h g−1 at C/5, 1 C, 2 C and 5 C, respectively; the un-doped LNM offered a lower reversible capacity of ∼110, 98, 85 and 12 mA h g−1 at the same C rates. When returned to C/5 from 5 C after 100 cycles, about 99% of reversible capacity was retained for LiMn1.45Cr0.1Ni0.45O4vs. 94% for un-doped LNM. These results give evidence that the electrochemical properties of the Cr-doped sample are definitely better than those of the un-doped sample at high C-rate, as expected. Therefore, the lower value of ΔV in un-doped LNM cannot be attributed to lower kinetics in the Cr-doped sample. To understand this difference in ΔV between the two samples, the phase diagram must be investigated; this is the purpose of the next section.
Since the disordered LNM outperforms the ordered LNM as a cathode element for Li-ion batteries, we found it desirable to clarify the phase diagram of LNM in the disordered phase. The purpose of this work is to report the study of the structural changes of LNM during lithium extraction and insertion by in situ XRD measurements. Since, in addition, oxygen vacancies have undesirable effects mentioned above; this investigation has been made both on un-doped and Cr-doped LNM that is free of oxygen vacancy.64,77 The differences in phase evolution were compared and the effect of Cr substitution is discussed.77
The evolution of in situ XRD patterns of the Cr-doped LNM electrode during the cycling at C/24 rate is reported in Fig. 13, in which (311) and (511) Bragg lines have been selected to understand better the phase evolution. Overall, all the diffraction peaks shifted to the higher 2θ angles as the Li+ ions were removed from the two host structure. The analysis of the XRD pattern shows that all the phases are cubic, and the lattice parameters for each phase as a function of x over one charge–discharge cycle are reported in Fig. 14.
![]() | ||
| Fig. 13 In situ XRD patterns of the (311) and (511) peaks as a function of x(Li) of Cr-doped LNM spinel during the charge at C/24 rate. | ||
![]() | ||
| Fig. 14 Variation of the lattice parameters of the different cubic phases in the Cr-doped LNM sample as a function of the Li concentration x during the charge (a) and discharge (b) during cycling at C/24 rate. Reproduced with permission.77 | ||
As charge proceeds from x = 1, a solid solution is observed in the whole range 1 ≤ x ≤ 0.72, the decrease in Li resulting only in the decrease of the lattice parameter of the cubic phase that we label as phase-1. A new cubic spinel phase (phase-2) started to emerge and the system thus undergoes a first-order transition at x = 0.72, to enter in a two-phase domain 0.72 ≤ x ≤ 0.54, in which the phase-2 grows at the expense of the phase-1, so that the phase-1 disappears at the Li concentration x = 0.54. Below this concentration, we find a small but finite range of concentration 0.54 ≤ x ≤ 0.37, in which the Cr-doped is again a solid solution in the phase-2 only. The system re-enters a two-phase region in the range 0.37 ≤ x ≤ 0.13 with the coexistence of phase-2 and a new phase (phase-3) that grows at the expense of phase-2. Finally, a solid solution with phase-3 only is observed in the range x ≤0.13. The experimental XRD spectra have been measured along the cycle with steps Δx ∼ 0.1, so that the uncertainty in the estimates of the Li-concentrations of the phase boundaries can be estimated to ±0.05. Upon discharge, phase-3 is found to be a solid solution up to x = 0.25. This result gives evidence that, even at the low C-rate of the experiment, thermodynamic equilibrium at the end of the charge was not reached. At the end of charge a very small faction (few% only) of phase-2 is still detectable. However, during the period of time the sample has been let in open-circuit before the cell was discharged, this phase-2 has been converted into phase-3 only. This can also be seen as the discontinuity between the end of discharge and the beginning of charge. Therefore, at equilibrium, the Cr-doped sample will be in a solid solution in phase-3 at low values of x ≤0.25. Upon discharging, the phase-2/phase-3 system is found in the range 0.26 ≤ x ≤ 0.43. The difference between 0.37 and 0.43 does not exceed the experimental uncertainty and is thus negligible. The onset of phase-1 in phase-2 is also the same at charge and discharge. Therefore, within this experimental uncertainty, we can deduce that the phase transformations are totally reversible, without any hysteresis, except for a small shift concerning the limit in the solid solution in phase-3 that, however, is not an intrinsic effect; it is clearly an artefact due to deviation from equilibrium.
To understand the link between the phase diagrams determined in the previous section and the electrochemical properties. We have reported in Fig. 15 the variation of the voltage with x for the two samples together with the range of existence of the different phases. First, we observe that the voltage of the battery with the Cr-doped sample is smaller than with the un-doped sample at any Li concentration x, except in the small region 0.1 ≤ x ≤ 0.2. The unexpected larger value of the voltage in this small region is clearly attributable to the fact that a larger proportion of phase-2 is still observed in the un-doped LNM sample in this range of concentration. Since the voltage associated to the phase-2 is lower than that of phase-3, the persistence of this phase lowers the potential, which becomes lower than in the Cr-doped sample where only the phase-3 is observed. Second, we also note that the voltage is strongly dependent of the composition in this small range of composition. This is true for the case of the Cr-doped sample, and this is expected since the Cr-doped sample is a solid solution in phase-3. In case of a two-phase regime, however, the Gibbs rule tells us that V(x) should be a flat plateau, while the potential varies very fast with x in this region. Therefore, the strong variation of V(x) in the un-doped sample in the small range 0.1 ≤ x ≤ 0.2 confirms that it is not in a two-phase regime. Instead, it should be considered as a solid solution in phase-3. The presence of the phase-2 simply means the existence of some isolated phase-2 clusters attributable to the difficulty to reach thermodynamic equilibrium.
![]() | ||
| Fig. 15 Variation of the voltage as a function of the Li concentration x in un-doped and Cr-doped LNM samples during discharge of cells cycled at C/24 rate in relation to the phase diagram. Reproduced with permission.77 | ||
More recently, we have successfully used LiFePO4 as a coating of LNM particles.86 This process allowed us to take benefit from the fact that LiFePO4 (LFP) shows very good rate performance not only at room temperature but also at 60 °C.87,88 In addition, in contrast with the transition metal oxides, the oxygen forms covalent bonding with P, to form PO4 phosphate units that are very stable, so that LFP shows an excellent thermal stability and is also stable in the battery up to a high voltage of 5.4 V. As usual with LFP, its low electrical conductivity implies that is must be carbon coated, so that the particles are multi composite, with a LNM core surrounded by a LFP coat, with conductive carbon on top of it.86Fig. 16 shows the modified Peukert plot before and after C–LFP coating. The performance at high C-rate is improved importantly in the multi-composite. This is confirmed in Fig. 17 showing a major improvement induced by the C–LFP coat on the aging of the cell upon cycling. Therefore, the C–LFP coating is a promising technology to protect the LNM particles against the reaction with the electrolyte and to prevent the oxygen loss.
![]() | ||
| Fig. 16 Modified Peukert plot for the Li//LMN and Li//LFP-coated LMN cells between 3.0 and 4.9 V vs. Li+/Li0. | ||
![]() | ||
| Fig. 17 Cyclability of the Li//LMN and Li//LFP-coated LMN cells at 1 C between 3.0 and 4.9 V vs. Li+/Li0. Reproduced with permission.86 | ||
m) phase because the electronic conductivity in this phase is several orders of magnitude larger than in the ordered phase, and because the diffusivity of lithium ions is higher. The larger electronic conductivity was attributed to the presence of oxygen vacancies in the disordered phase, but it is more likely an intrinsic property. In the stoichiometric composition of LiNi0.45Mn1.45Cr0.1O4, the cations exist as Ni2+ and Mn4+, so that they do not carry the same electronic charge. Therefore, the ordering between Mn and Ni is also a charge ordering that will result in the opening of an electrostatic gap.
The structural evolution of un-doped and Cr-doped LNM spinels in the Fd
m phase as a function of the Li content (x) investigated by in situ XRD analyses shows that both charge (Li extraction) and discharge (Li insertion) reactions occur with the existence of three phases that form alternatively solid solutions and two-phase regions. The results have been understood on the basis of a model77 that takes strain effects into account, also explaining the fact that the phase diagram is sample dependent, differs between the different results reported in the literature. The analysis of the phase diagram confirms the faster dynamics of the Li-insertion/de-insertion in the Cr-doped sample, evidenced by the improved capacity retention at high C-rates. The other benefit of the Cr-substitution is the increase of the stability of the lattice. The drawback is a decrease in the energy density that is not due to a loss of capacity, but a smaller redox potential of the nickel vs. Li+/Li. Although the coating of the particles with C–LiFePO4 efficiently protects LNM against the surface reactions with the electrolyte, which was the main limiting factor for this material, making LNM the most promising material to obtain high-voltage Li-ion batteries.
| This journal is © The Royal Society of Chemistry 2014 |