Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Thermodynamic properties of LiNiO2, LiCoO2, and LiMnO2 using density-functional theory

Lucas Tosin Paese *, Philippe Zeller , Sylvie Chatain and Christine Guéneau
Université Paris-Saclay, CEA, Service de Recherche en Corrosion et Comportement des Matériaux, 91191, Gif-sur-Yvette, France. E-mail: lucas.tosinpaese@cea.fr

Received 18th April 2023 , Accepted 13th July 2023

First published on 17th July 2023


Abstract

The formation energies of LiCoO2, LiNiO2 and LiMnO2 were calculated using a combination of adequately selected Hess cycles and DFT computations. Several exchange–correlation functionals were tested and PBE for solids (PBEsol) turned out to be the most accurate. The enthalpies of formation at 0 K are −168.0 kJ mol at−1 for LiCoO2, −173.2 kJ mol at−1 for LiNiO2, −209.9 kJ mol at−1 for o-LiMnO2 and −208.8 kJ mol at−1 for r-LiMnO2. In comparison to experimental formation energy data, a difference of 1.6 and 0.01 kJ mol at−1 was obtained for LiCoO2 and LiMnO2, respectively. By contrast, a much larger discrepancy, around 24 kJ mol at−1, was obtained for LiNiO2 and confirmed by using an additional and independent Hess cycle. The influence of slight crystallographic distortions associated with magnetism and/or the Jahn–Teller effect on energy was carefully searched for and taken into account, as well as corrections arising from vibrational contributions. Hence, these results should motivate future measurements of the thermodynamic properties of LiNiO2, which are currently scarce. Vibrational contributions to the structural and energetic properties were computed within the harmonic and the quasi-harmonic approximations. The LiCoO2 heat capacity at constant pressure is in excellent agreement with experimental data, with a difference of only 3.3% at 300 K. In the case of LiNiO2 the difference reaches 17% at 300 K, which could also motivate further investigation. The Cp(T) value for the orthorhombic phase o-LiMnO2, for which no previous data were available, was computed. Structural properties such as specific mass, bulk modulus and coefficient of thermal expansion are presented.


Introduction

Most of the lithium-ion batteries (LIBs) found on the market use NMC (which stands for nickel, manganese and cobalt) as the cathode active material. The NMC crystals have the general formula LiNixMnyCozO2 (x + y + z = 1), with compositions that vary from LiNi0.33Mn0.33Co0.33O2 to LiNi0.8Mn0.1Co0.1O2,1 systematically increasing in Ni content. The growing need for the storage of electrical energy in the world is a strong motivation to optimize NMC-containing batteries. The thermal stability of the cathode material is related to its enthalpy of formation and the voltage profile during delithiation is related to the chemical potential of lithium.2 Hence, good knowledge of the thermodynamic properties of the NMC system is essential for the development of these materials.

The CALPHAD (CALculation PHAase Diagrams) method3,4 is a proven approach for predicting the thermodynamic, kinetic, and other properties of multicomponent materials systems. It couples phase diagram data and thermodynamic properties. The key thermodynamic property is the Gibbs energy from which all other thermodynamic properties can be derived. For each phase, a model is selected and the Gibbs energy is expressed with a polynomial function versus temperature and composition. The parameters of the models are fitted from selected experimental data such as the melting point, phase transition temperature, enthalpy of formation, enthalpy of mixing, heat capacity, and others. The functions and parameters obtained are stored in a database. Once the database is finalized, it provides the thermodynamic properties of the system for any composition. A CALPHAD model is usually refined as new data pertaining to the system are published, which may take place over periods of several years. The CALPHAD approach to the NMC system may be expected to provide a continuous description of properties related to the material according to composition and during the battery cycle, as demonstrated in the studies by Chang et al. in their CALPHAD descriptions of the pseudo-binary equilibria LiCoO2–CoO25 and LiNiO2–NiO2.6

The NMC crystals have a rhombohedral structure that belongs to the R[3 with combining macron]m space group.7 We can represent them as a distorted rocksalt structure made of an FCC network of oxygen anions and another FCC network occupied by cations, in which Li planes alternate with planes containing transition metal cations along the [111] direction. The CALPHAD modelling of the LiNixMnyCozO2 solid solution includes the thermodynamic quantities of its three end-members which by definition are the compounds obtained assuming x = 1, y = 1 or z = 1, i.e. LiCoO2, LiNiO2 and LiMnO2. In this work, we will refer to these compounds generically as LiMO2, given M = Co, Ni or Mn. At room temperature, while LiCoO2 and LiNiO2 are stable and share the same rhombohedral structure as NMC, LiMnO2 is only stable in its orthorhombic structure in the space group Pmmn.8 Additionally, LiNiO2 is particularly complex experimentally, as it is almost impossible to synthesize the perfect stoichiometric compound.9 The thermodynamic properties of these end-members are necessary input data for the CALPHAD model which will eventually allow the optimized design of new NMC compositions.

Several measurements of the enthalpies of formation and heat capacities for the end-member LiCoO2 have been published.10–16 By contrast the data regarding LiNiO2 are very scarce and limited to the enthalpy of formation of LiNiO217 and heat capacity at low temperatures.18 The end-member LiMnO2 is not stable as a crystal belonging to the space group R[3 with combining macron]m19 and, therefore, there are no experimental measurements of its thermodynamic properties.

In the last few years, the boost in computational power allowed the ab initio calculation of physical properties of materials that agree with experimental data and permit the prediction of properties hitherto not known, which provides an invaluable source of input data for the construction of CALPHAD databases. Density-functional theory (DFT)20,21 has been used by many authors to calculate the physical properties of lithium-ion battery cathode materials.22–30 The heat capacity of LiCoO2 at constant volume (Cv) was obtained by means of DFT by Wu et al.22 In a similar work, Du et al.23 calculated the heat capacity of LiMO2 as a function of temperature using density-functional perturbation theory (DFPT)31 and a harmonic approximation (HA), although it is mentioned that a quasi-harmonic approximation (QHA) was employed. This approach may be acceptable at low temperatures, where the calculated Cv agrees well with the experimental Cp values of LiCoO2.

Regarding the study of formation energies, Chang et al.24 calculated the formation energies of LiCoO2 and LiNiO2 from pure elements with a combination of DFT and an empirical method, which consists in adding or subtracting a certain amount of energy of a MyOx species, depending on the oxidation state of the transition metal. Xu et al.25 used the Generalized Gradient Approximation (GGA) to compute the formation energies of LiCoO2, LiMnO2, LiMn2O4 and LiFePO4 using the PBE description of GGA. Kramer and Ceder26 studied the morphology of LiCoO2 by means of DFT+U. Longo et al.27 focused on the study of layered and over-lithiated Mn oxides, Tuccillo et al.28 studied the stability of the compounds LiMO2 in three different structures and Tsebesebe et al.29 assessed the thermodynamic, electronic, and mechanical properties of the same family of compounds. Finally, Lee et al.30 computed the formation energy of LiCoO2 with the support of molecular dynamics. A compilation of the results obtained by these authors is given in Table 1.

Table 1 Formation energies from elements obtained in previous works. The relative error is calculated based on standard enthalpy of formation experimental data10,15,17,32,33
Ref. Calculation method Compound Space group Calculated formation energy (kJ mol at−1) Absolute error (kJ mol at−1) Relative error (%) Transition metal 3d electrons U value (eV)
24 DFT + empirical method LiNiO2 R[3 with combining macron]m −150.5 2.2 1.5
LiCoO2 R[3 with combining macron]m −173.4 3.7 2.2
25 GGA LiCoO2 R[3 with combining macron]m −91.5 78.1 46.1
LiMnO2 C2/m −178.9 31.0 14.7
26 GGA+U LiCoO2 R[3 with combining macron]m −171.8 2.1 1.3 3.3
27 GGA+U LiMnO2 R[3 with combining macron]m −192.6 17.3 8.2 5.2
28 GGA+U LiCoO2 R[3 with combining macron]m −183.5 13.9 8.2 4
LiNiO2 C2/m −154.8 6.4 4.4 4
LiMnO2 Pmmn −209.0 0.9 0.4 4
29 GGA+U LiCoO2 R[3 with combining macron]m −41.7 127.9 71.9 6.1
LiNiO2 R[3 with combining macron]m −43.1 105.3 71.0 2.45
LiMnO2 R[3 with combining macron]m −32.0 178.0 78.5 4.25
30 Molecular dynamics LiCoO2 R[3 with combining macron]m −165.4 4.2 2.5


Table 1 displays a great variability of the results. A likely cause of this variability is the diversity of formation reactions used in the calculations. In some works, the reactants are the pure elements Li and M in their crystal forms and molecular O2 while, in other works, they are combinations of various oxides where a redox reaction takes place. The involvement of molecular O2 in the reaction requires a correction since the GGA underestimates the O2 binding energy by about 150 kJ mol−1.34 Additionally, the choice of Hubbard-U values is not straightforward35 and this is obviously yet another source of variability. The U value adds an energy correction to selected electron states in order to obtain a more localized electron wavefunction and is commonly used in calculating 3d transition metals oxide energies.36–39 Some studies, however, find better agreement with experiments without using U values.40 Additionally, the Hubbard U approach is especially problematic for electrochemical processes.39

The main objective of this study is to compute the thermodynamic properties of the three end-members as a function of temperature using DFT, which is part of an effort aiming at building a CALPHAD model of the NMC system.

In order to estimate and minimize the uncertainties in our results, we studied in detail several aspects of the methodology: (1) the various possibilities for writing the DFT-computed chemical reaction that gives access to the formation energy; (2) the most appropriate theory level for computing the vibrational contributions; (3) the crystallographic structure used as input data to the DFT computations and (4) The specific influence of the exchange–correlation (x–c) functional on the quality of the results, which has been discussed above, in view of the data in Table 1. In this paper, we report on the calculated thermodynamic properties and also on the methodology, the latter being an essential element on which the CALPHAD input data assessments are based.

We now briefly introduce the first three methodological aspects. The choice of the DFT-computed reaction is a central feature. Any chemical reaction involving the LiMO2 compound of interest and other compounds for which reference enthalpy data are available can be used for this purpose. Extracting the desired standard formation energy of the LiMO2 compound from the DFT results is then performed classically using Hess law.41 Although the availability of the required enthalpy reference data somewhat constrains the choice of the DFT-computed reaction, many reactions can be considered. In the following, the possible options are sorted according to the fact that this DFT-computed reaction involves oxidation and reduction, i.e. a redox reaction, and more precisely according to the change in the oxidation state of the transition metal. Since the scheme with no-redox reactions comes out as the best, we shall present it first.

As for the vibrational contributions, we compute the corrections to the energetic properties, namely zero-point energy and enthalpy increments with temperature, both within the harmonic approximation (HA) and the quasi-harmonic approximation (QHA). At no additional computational cost, this gives us access first to the heat capacities at constant volume (Cv) and constant pressure (Cp), within the QHA, and second to structural properties such as specific mass, bulk modulus and thermal expansion coefficient.

The choice for the crystallographic description of the DFT input structure sometimes has a significant influence on the energy results so that it is also a major aspect of our methodology. Ideally, the only input data for a DFT study should be the chemical composition and the total number of electrons. However, most studies of crystalline compounds use the experimentally determined structure as input. This makes the DFT result sensitive to possible errors or uncertainties in the experimental data and, consequently, less “ab initio”. In our work, we depart from this standard procedure to some extent by retaining a structure that leads to a lower energy, in case we find one, while remaining compatible with the available structural data.

In the next section, we present the first step of our methodology, which consists of the selection of suitable chemical no-redox reactions according to available reference data, the design of Hess cycles and the choice of options regarding the DFT levels. The following section describes the computational details. The calculation results are reported and discussed subsequently, where we compare and assess the various aspects of the methodology. Finally, we draw conclusions.

Methodology

We calculate the formation energy of each end-member using a Hess cycle that mixes a DFT-computed chemical reaction with several auxiliary reactions for which reference reaction energy data are available in the literature. We classify the DFT-computed reactions into three categories. Type 1 reactions involve only oxide compounds and no oxidation or reduction of the various chemical elements, referred to as “no-redox reactions” in this work. Type 2 reactions are all-solid-state reactions involving charge transfer. Type 3 reactions are standard formation reactions involving all elements in their standard state at 300 K (gaseous molecular oxygen, metallic Li and transition elements). Type 2 and 3 reactions involve electron transfer and are referred to as “redox reactions”. This section is restricted to Type 1 reactions, which are the cornerstone of our methodology, while Type 2 and Type 3 reactions will only be used in the following sections for comparison purposes. The energy of each compound taking part in the reaction, reactant or product, is computed using DFT to derive the reaction energy.

Table 2 shows the DFT-computed formation reactions evaluated for each end member.

Table 2 Type 1 DFT-computed reactions used to calculate the formation energy of NMC end-members
End-member Reactions assessed Reaction #
LiCoO2 Li2O + Co3O4 → 2LiCoO2 + CoO 1
LiMnO2 Li2O + Mn2O3 → 2LiMnO2 2
LiNiO2 Li2O + 2NaNiO2 → 2LiNiO2 + Na2O 3
Li2O + 2NiTiO3 → 2LiNiO2 + Ti2O3 4


Experimental energies of formation are usually expressed using the pure elements in their standard state. Thus, we choose an appropriate Hess cycle in order to compare experimental and calculated values. Table 3 lists all Hess cycles used for each reaction and all reactions making up each cycle, as well as the associated reaction energies.

Table 3 Thermodynamic cycles employed to obtain formation energies of LiMO2 (M = Co, Ni, Mn) from pure elements in their standard state. Each cycle includes a Type 1 DFT-computed formation reaction (i.e. without any redox process)
Compound Reaction # Hess’ cycle Energy
LiCoO2 1 Li2O + Co3O4 → 2LiCoO2 + CoO ΔER1, calculated with DFT
2Li + 0.5O2 → Li2O ΔELi2O, experimental
3Co + 2O2 → Co3O4 ΔECo3O4, experimental
Co + 0.5O2 → CoO ΔECoO, experimental
Li + Co + O2 → LiCoO2 ΔfE,R1,LiCoO2 = 0.5(ΔER1 + ΔELi2O + ΔECo3O4 − ΔECoO)
LiMnO2 2 Li2O + Mn2O3 → 2LiMnO2 ΔER2, calculated with DFT
2Li + 0.5O2 → Li2O ΔELi2O, experimental
2Mn + 1.5O2 → Mn2O3 ΔEMn2O3, experimental
Li + Mn + O2 → LiMnO2 ΔfE,R2,LiMnO2 = 0.5(ΔER2 + ΔELi2O + ΔEMn2O3)
LiNiO2 3 Li2O + 2NaNiO2 → 2LiNiO2 + Na2O ΔER3, calculated with DFT
2Li + 0.5O2 → Li2O ΔELi2O, experimental
Na + Ni + O2 → NaNiO2 ΔENaNiO2, experimental
2Na + 0.5O2 → Na2O ΔENa2O, experimental
Li + Ni + O2 → LiNiO2 ΔfE,R3,LiNiO2 = 0.5(ΔER3 + ΔELi2O + 2ΔENaNiO2 – ΔENa2O)
4 Li2O + 2NiTiO3 → 2LiNiO2 + Ti2O3 ΔER4, calculated with DFT
2Li + 0.5O2 → Li2O ΔELi2O, experimental
Ni + Ti + 1.5O2 → NiTiO3 ΔENiTiO3, experimental
2Ti + 1.5O2 → Ti2O3 ΔETi2O3, experimental
Li + Ni + O2 → LiNiO2 ΔfE,R4,LiNiO2 = 0.5(ΔER4 + ΔELi2O + 2ΔENiTiO3 − ΔETi2O3)


We computed the energies of the reactions listed in Table 2 using various choices of DFT parameters. We shall first report results obtained with our reference parameter set which consists of the GGA-PBEsol42 x–c function at 0 K and does not include vibrational contributions, a condition denoted “static”. From then on, in an effort to minimize the uncertainty, we shall evaluate various modeling options by comparison with this reference set of results. As the reaction energies of the auxiliary reactions listed in Table 3 are actually experimental enthalpies of formation at 300 K, in a second step, we shall report the DFT computed reaction energies also at 300 K, which were obtained by complementing static data with phonon calculations within the harmonic (HA) or quasi-harmonic (QHA) approximations. The fundamental mathematical formulae used for obtaining the thermodynamic quantities calculated in this work are listed in the ESI.

Due to the significant computational cost of the latter correction, however, we compute it only for LiCoO2 and show that the error introduced by ignoring this correction is small. In a later section, we shall also compare these reference results with static formation energies obtained with other functionals, namely PBE,43 PW9144 and rSCAN,45 as well as with reactions involving a redox process. Then, we also compare them with GGA+U calculations using different U values in ranges that have been recommended in previous works. For Co-3d, we set the Hubbard-U value as 2.6, 3.3 and 4.0 eV. For Ni-3d, we chose 4.0, 6.4 and 8.0 eV. The DFT+U calculations involving species with Mn could not be performed because the calculations did not converge using the set of parameters listed in the Computational details section. The second value of each set of Hubbard-U values corresponds to a value found in the studies by Wang et al.,36 where the U values were fitted according to the energy of the oxidation reactions 6 CoO + O2 → 2 Co3O4 and 2 NiO + O2 → 2 NiO2. We did not have enough reliable data to reasonably assign different U values to different oxidation states of a given chemical element, leading to inaccuracies when computing the energy of Co3O4, for instance, a mixed-valence compound that contains both Co+2 and Co+3. Table 1 illustrates the variety of U values among different studies that sometimes use an average of values found in the literature, and sometimes set a value that best fits a specific property. Thus, we arbitrarily choose a larger and a lower value for each transition metal in order to quantify its influence on the results.

We make use of the fact that for all considered solid-state compounds at a pressure P of 1 atm, the difference between enthalpy and energy is negligible. Thus for data calculated statically at P = 0 or with the QHA at P = 1 atm, we use the terms “formation energies” or “enthalpies of formation” interchangeably. We stick to the term “energy” when computed with the HA at 300 K, since the system is then under much higher pressures or in situations where the pressure is not specified. For the gaseous O2 phase, we take into account the difference between molar energy and enthalpy by assuming a constant value CpCv = R.

As for the vibrational contributions, the heat capacity of LiMO2 at constant volume (Cv) is obtained using the HA and the heat capacity at constant pressure (Cp, P = 1 atm) is obtained within the QHA.46 We calculated the Cp curves in a temperature range from 0 to 1200 K, which corresponds to static pressures varying from −2 to −9.5 GPa, depending on the LiMO2 compound. At each temperature, the volume dependence of the Helmholtz free energy (F) was obtained with phonon calculations over at least four different volumes and it was least-squares fitted to a parabolic equation, as detailed in the ESI. All phonon calculations are performed with the PBEsol functional. In cases where the full QHA model was unusable due to the presence of imaginary modes of vibration at large negative static pressures, we replaced it with the simplified QHA model of Fleche.47 This model is based on DFT data consisting of the static pressure–volume relationship, which we fitted using the Birch–Murnaghan equation of state, and the phonon spectrum computed at zero static pressure. The Fleche model then uses the Debye model to process the acoustic phonons and the Mie–Grüneisen theory to approximate the optical Grüneisen coefficient. In cases where both can be used, the full QHA model and the Fleche's model only show small differences. We noted that for all the compounds studied in this work, we always achieved a complete absence of unstable vibration modes at zero static pressure. To a large extent, this is due to the fact that we have always searched for the structure with the lowest static energy at zero static pressure, starting from the experimentally known structure and allowing for small distortions by considering a variety of supercells. Neumann–Kopp approximations from binary oxides are also presented along with our QHA results. With the QHA, we also obtain the cell specific mass, the bulk modulus and the thermal expansion coefficient of the LiMO2 compounds.

Crystal structures

All the structures were built based on low-temperature experimental information, which includes cell parameters, Wyckoff positions and magnetic structure. Magnetism is taken into account in DFT by spin-polarization. When the antiferromagnetic structure is the most stable, the set of symmetry operations must be changed accordingly compared to the symmetry of the mass distribution. In such cases, or wherever our calculated structure is a slightly distorted supercell of the experimentally known structure, we present the crystallographic details including the spin direction for each of the transition metal orbits in the ESI. This enables a complete description of these structures without ambiguity. The procedure to obtain such distorted structures is a simplified version of the procedure of Togo and Tanaka.48 We build a set of supercells with the original structure taken from the literature for the same compound. In each supercell, we introduce a slight atomic displacement or cell strain and we submit it to a geometry optimization task with CASTEP. The result we reported is the most stable structure according to CASTEP. Compared to the original structure from the literature, it usually contains small distortions of typically 0.05 Å. The geometry-optimized structure is symmetrized using a symmetry finder software tool with a tight tolerance such as 0.005 Å. The Wyckoff positions are then easily read either from the software tool output itself or from e.g. the International Tables of Crystallography Vol A or the Bilbao Crystallographic Server.49 We stop the screening as soon as a significant energy lowering is achieved. We do not claim that our reported structure is the minimum energy state because there is always the possibility that some other supercell, not included in our screening procedure, would have a lower energy.

Table 4 summarizes the experimental structures and enthalpies of formation of all the compounds involved, as well as the structures used for the DFT calculations. The list contains additional species not mentioned in Table 3 that will be employed in a later section when evaluating reactions involving a redox process.

Table 4 List of compounds employed in this study. For each compound, the crystal lattice structure (prototype, Pearson's symbol and space group number), experimental standard formation energy and magnetic structure are given. NSP, FM and AFM stand for “non-spin-polarized”, “ferromagnetic” and “antiferromagnetic”, respectively. LiNiO2, CoO, Mn2O3 and Mn are special cases discussed in the main text
Compound Crystal lattice structure Enthalpy of formation at 300 K (kJ mol at−1) Input structure for the DFT calculation (spin polarization, space group number) and reference for magnetic structure
LiCoO2 NaFeO2, hR12, 166 −169.9 ± 0.615 NSP, 16650
−168.3 ± 0.610
−170.6 ± 0.516
LiMnO2 NaMnO2, oP8, 59 −209.9 ± 0.533 AFM, 1351
LiNiO2 NaFeO2, hR12, 166 −148.3 ± 0.617 FM, 14
Co3O4 Fe3O4, cF56, 227 −130.1 ± 0.252 AFM, 21653
CoO NaCl, cF8, 225 −119.4 ± 0.654 AFM, 122
MnO NaCl, cF8, 225 −191.355 AFM, 16656
Mn2O3 Mn2O3, oP80, 61 −192.3 ± 0.257 AFM, 19
NiO NaCl, cF8, 225 −119.7 ± 0.254 AFM, 16656
NaNiO2 NaNiO2, mS8, 12 −157.358 AFM, 1259
NiTiO3 TiFeO3, hR30, 148 −240.460 AFM, 14661
Ti2O3 Al2O3, hR30, 167 −303.462 FM, 16763
Na2O CaF2, cF12, 225 −138.358 NSP, 225
Li2O CaF2, cF12, 225 −199.1 ± 0.164 NSP, 225
Li2O2 Li2[O2]-a, hP8, 194 −161.165 NSP, 194
Li W, cI2, 229 0 NSP, 229
Co Mg, hP2, 194 0 FM, 194
Mn Mn, cI58, 217 0 AFM, 21566
Ni Cu, cF4, 225 0 FM, 225
O2 0 FM


Among the species considered in this work, the CoO crystal shows a particular inconsistency between experimental and calculated data. The DFT calculations show that CoO in the sphalerite structure with space group I[4 with combining macron]2d (Table S4, ESI) is about 5 kJ mol at−1 more stable than its known rock-salt configuration. This phase is reported in the work of Redman and Steward,67 in which the authors consider it metastable. Sui et al.68 observed that the sphalerite phase is predominant in the early stages of CoO crystal growth and does not disappear in the later stages. It must be noted that the XRD patterns of the rock salt and sphalerite phases are fairly similar, so data with good signal-to-noise ratio and high resolution may be required to differentiate them. This arises the question of which structure was actually associated with the formation enthalpy measurements. The experimental enthalpy of formation of CoO used in this work was obtained in the study by Boyle et al.54 where the X-ray diffraction spectrum data contained additional lines which did not belong to the rock-salt CoO structure. Ghosh et al.69 studied in detail the DFT description of the energy ordering of these phases and their dependence on the x–c functional but did not include PBEsol in their study. The known antiferromagnetic (AFM) structure of the rock-salt CoO is also questionable. We find an AFM arrangement in the P21/m space group (Table S5, ESI) that is 0.3 kJ mol at−1 more stable than the one in the C2/m configuration proposed by Jauch et al.70

The compound LiMnO2 is stable in an orthorhombic configuration (o-LiMnO2, space group Pmmn).71,72 At low temperatures, o-LiMnO2 is antiferromagnetic and a magnetic structure was proposed by Kellerman et al.51 This structure can belong to either P2/c or C2/c space group. We performed DFT calculations for both structures and found similar energies for both. Atom coordinates and cell parameters are given in Table 5 and Table 6. We proceeded with the P2/c structure in our calculations.

Table 5 Cell parameters of the AFM structure of o-LiMnO2 with two different configurations of identical energy. Cell dimensions a, b and c in Å
Cell parameter P2/c C2/c
a 5.341 9.132
b 5.686 11.367
c 5.341 5.34
α and γ 90° 90°
β 117.515° 148.745°


Table 6 Atom coordinates of the AFM structure of o-LiMnO2 with two different configurations of identical energy
Structure Elements Wyckoff position x y z
P2/c Li 2e 0 0.120 0.25
Li 2f 0.5 0.880 0.25
Mn 2e 0 0.634 0.25
Mn 2f 0.5 0.366 0.25
O 4g 0.25 0.398 0
O 4g 0.25 0.142 0
C2/c Li 4e 0 0.310 0.25
Li 4e 0 0.810 0.25
Mn 4e 0 0.067 0.25
Mn 4e 0.5 0.062 0.25
O 8f 0.25 0.179 0.75
O 8f 0.25 0.449 0.75


The LiNiO2 compound, as LiCoO2, crystallizes in the α-NaFeO2 rhombohedral system.73,74 However, a local Jahn–Teller distortion is observed by EXAFS spectroscopy.75 Neutron diffraction studies showed that the distorted structure is best fitted to the C2/m group at low temperatures.76 Several ab initio studies were carried out to better understand the possible cooperative Jahn–Teller distortion at low temperatures23,77–81 and all conclude that a distorted structure is more stable than the non-distorted rhombohedral structure. In this work, we evaluated some configurations of Jahn–Teller distortions of the rhombohedral LiNiO2, and found that the structure belonging to the P21/m space group (Table S3, ESI) to be the most stable, with a maximal deviation of 0.09 Å from the original non-distorted structure, in agreement with the results obtained by Chen et al.80 and Sicolo et al..81

The stable configuration obtained for Mn2O3 belongs to the P212121 space group (Table S1, ESI), with a maximal deviation of 0.04 Å from its experimental Pbca structure.

The magnetic structure of Mn metal has not yet been completely solved, in spite of many experimental and theoretical efforts over the past 70 years (see e.g. Manago et al.82 and references therein). Since we are only interested in the energy, the finer details of the magnetic structure may be neglected. We followed the description given by Hobbs et al.66 for collinear antiferromagnetism which can be computed in a space group P[4 with combining macron]3m. rSCAN is a special case as it leads to strong overmagnetization.83 Since the magnetic moments in α-Mn are strongly dependent on the geometry of the crystal, our rSCAN energy result may be off by up to 3 kJ mol at−1. As a matter of fact, rSCAN finds that a ferrimagnetic arrangement in I[4 with combining macron]3m more stable than the AFM arrangement in P[4 with combining macron]3m, by about 7 kJ mol at−1.

The O2 molecule was built in a cell with a, b and c parameters equal to 20, 21 and 22 Å, respectively and α, β and γ equal to 90° and a spin polarization corresponding to its triplet electronic ground state.

Computational details

Most DFT computations were performed using the CASTEP84–87 code implementing the plane-wave pseudopotential method. The BIOVIA Materials Studio software88 was employed to build all the structures and to analyze the results. Energies of formation were computed by performing geometry optimizations of all structures involved. The generalized gradient approximation (GGA) was used with and without the GGA+U correction for the strong correlations of d electrons. Several x–c functionals were used. Formation energies and phonons were performed with the PBEsol42 functional. The formation energy was also computed with PBE,43 PW9144 and rSCAN45 functionals. The “2017R2” ultrasoft pseudopotentials library was used. For the computation of the static formation energies, the values of the numerical parameters were chosen after a convergence study so as to reach a precision of 0.1 kJ mol at−1 on reaction energies. The energy cutoff was set to 630 eV and the self-consistent field (SCF) convergence threshold to 1 × 10−6 eV per atom. The k-point sampling was generated by Monkhorst and Pack's scheme89 with a k-point interval of 0.04 Å−1 on all axes. All calculations are spin-polarized unless stated otherwise. For formation energy calculations, the geometry of all crystals was fully relaxed up to a maximal force of 0.02 eV Å−1, maximal stress of 0.02 GPa, maximal atom displacement of 0.001 Å and a relaxation energy convergence of 1 × 10−5 eV atom−1. We assumed the experimental configuration as the starting point, with the exception of a few compounds found to have more stable configurations at 0 K, in which case their crystallography information is given in the ESI.

Phonon calculations were performed using CASTEP with the finite displacement method on LixMxO2x (M = Co, Ni, Mn) supercells relaxed with an energy convergence of 5 × 10−6 eV atom−1, a maximal force of 0.003 eV Å−1, a maximal atom displacement of 5 × 10−4 Å and a maximal stress of 0.01 GPa. We use a SCF convergence of 1 × 10−11 eV standard and fine grids of 1.75 and 2.1, respectively, and supercell sizes consistent with a cutoff of 4.5 Å for the force constants. The supercells are described in the ESI. PBEsol was used to obtain both the Birch–Murnaghan equations of state and the lists of phonon frequencies. In cases where the phonon spectrum contained imaginary frequencies, we checked whether these modes actually reflected the instability of the structure, as opposed to numerical uncertainties associated with the calculation of force constants by recomputing the phonon spectra using the ABINIT code90–94 with PAW pseudopotentials and linear response theory (DPFT).

Results

Static formation energies

The enthalpies of formation obtained with Type 1 no-redox reactions and the PBEsol functional for the three LiMO2 compounds are outlined in Table 7:
Table 7 Enthalpies of formation in kJ mol at−1 of LiMO2 compounds computed with the static method. The two values assigned for LiNiO2 refer to the two reactions reported in Table 2
End-member Calculated enthalpy of formation Experimental value δ th-exp Relative error (%)
LiCoO2 −167.97 −169.9 ± 0.615 1.93 1.1
−168.3 ± 0.610 0.33 0.2
−170.6 ± 0.516 2.63 1.5
o-LiMnO2 −209.86 −209.87 ± 0.533 0.01 0.003
r-LiMnO2 −208.82
LiNiO2 −173.05 −24.75 −24.75 16.7
−173.32 −25.02 −25.02 16.9


It should be noted that the calculated enthalpy of formation of LiCoO2 was obtained with the sphalerite crystal structure of CoO. The result obtained with the monoclinic structure reported by Jauch et al.70 is −166.79 kJ mol at−1, 1.18 kJ mol at−1 above the value reported in Table 7.

We observe very good agreement with experimental data for LiCoO2 and o-LiMnO2. However, we find a striking discrepancy between the calculated and experimental values of LiNiO2, which is an order of magnitude larger than what we obtained for the other end-members. The fact that the calculated values for LiNiO2 agree within the two proposed reactions (#3 and #4 in Table 3) motivated some further investigation about this discrepancy with the experiment. To this end, in the first step, we searched for a possible issue with the DFT computation of Ni(III), as opposed to Co(III) and Mn(III). In the second step, we studied the influence of the x–c functional, and, in the third step, we recomputed the LiMO2 formation energies using other formation reactions (i.e. Type 2 and Type 3 as defined above). All of these results are presented in the following subsections of this paper. First, we applied our method to compute the formation energy of NiTiO3 according to the Hess’ cycle in Table 8, which involves the same compounds already employed in Table 2 except the Li-containing ones.

Table 8 Hess’ cycle used for obtaining the formation energy of NiTiO3
Hess’ cycle Energy Value (kJ mol−1)
Ti2O3 + 2NaNiO2 → 2NiTiO3 + Na2O ΔER, calculated with DFT −41.23
2Ti + 1.5O2 → Ti2O3 ΔETi2O3, experimental −1517.0562
Na + Ni + O2 → NaNiO2 ΔENaNiO2, experimental −629.2858
2Na + 0.5O2 → Na2O ΔENa2O, experimental −414.8258
Ni + Ti + O2 → NiTiO3 ΔEf,NiTiO3 = 0.5(ΔER + ΔETi2O3 + 2ΔENaNiO2 − ΔENa2O) −1201.03


The calculated NiTiO3 enthalpy of formation is −240.21 kJ mol at−1. A difference of 0.19 kJ mol at−1 is found in comparison with the experimental value of −240.4 kJ mol at−1 by Kubaschewski et al.60 This excellent agreement supports our attribution of the origin of the discrepancy between the calculated and experimental values for LiNiO2 to the Li containing compounds, and more precisely to LiNiO2 itself since Li2O was validated by the study of LiCoO2 and LiMnO2. This issue is analyzed further in the Discussion section.

To wrap up the static formation energies section, we now go back to LiMnO2. As mentioned in the Methodology section, the stable LiMnO2 adopts an orthorhombic structure (o-LiMnO2) and this is used in our calculations. However, the LiMnO2 compound of interest as an end-member of the NMC system has a rhombohedral configuration (r-LiMnO2), which is metastable. Among the different rhombohedral structures that we tested, including FM and some AFM configurations, a distorted configuration belonging to the P2/c space group was the most stable. The calculated enthalpy of formation of this structure is about 1.0 kJ mol at−1 above that of the stable orthorhombic o-LiMnO2. The crystallographic description of the r-LiMnO2 structure is included in the ESI in Table S2.

Vibrational contribution to the formation energy

The corrections to the zero-point energy (ZPE) and vibrational energy at 300 K are often overlooked in computations of formation energies because they are usually marginal in comparison to the DFT precision with respect to the experimental data and they require calculations that are computationally heavy in contrast to static calculations. In the present study, however, where the discrepancy between static DFT and experiment was found to reach up to a few kJ mol at−1, it becomes fully justified to study this correction. Thus, we compute this contribution to the formation energy of LiCoO2.

The energy corrections obtained with the HA for all the compounds involved in the DFT-computed reaction (reaction 1 of Table 2) are given in Table 9. We choose energy references for the three chemical elements Li, Co and O in such a way that the static total energies of Li2O, CoO and Co3O4 are all equal to zero. The fourth column (internal energy at 300 K) is the sum of the three other columns (static total energy + zero point energy + energy gain from 0 to 300 K).

Table 9 Corrections to internal energies of species involved in Reaction 1 of Table 2, and (last line) corrections to the reaction energy, arising from phonon contributions computed within the HA. Values in kJ mol−1
Compound Static total energy Zero point energy Energy gain from 0 to 300 K Internal energy at 300 K
Li2O 0 23.28 7.55 30.83
CoO 0 10.15 5.11 15.26
Co3O4 0 51.71 19.32 71.03
LiCoO2 −36.77 29.71 10.56 3.49
Reaction 1 −73.55 −5.42 −0.64 −79.61


While the total HA contribution to the internal energy of each compound turns out to be of the order of several tens of kJ mol at−1, the reaction energy of reaction 1 changes by only about 8% or, in absolute terms, −6.1 kJ mol−1. This is in line with the common tendency of the heat of solid-state reactions to be little dependent on temperature. From the point of view of the differences reported in Table 7 however, this correction is significant. Thus, we shall take it into account below in the assessment of the precision of our calculation method.

We also computed the vibrational contributions to the LiCoO2 energy within the QHA. At a pressure of 1 bar, the energy gain from 0 to 300 K is equal to 2.8 kJ mol at−1, which is only 0.2 kJ mol at−1 greater than that obtained with HA. Assuming that the difference between HA and QHA for Li2O, CoO and Co3O4 is of the same order of magnitude and that some compensation between reactants and products also occurs, we may presume that the QHA does not bring any significant improvement for computing formation energies, compared to the HA.

As reported in the thermodynamic assessment of Chen et al.,95 CoO undergoes an AFM to paramagnetic phase transition at a temperature close to 300 K. This transition affects its Cp as it shows an abrupt increase starting at around 200 K with a peak at 290 K. Integration of this experimental Cp gives an enthalpy increment from 0 to 300 K of 4.7 kJ mol at−1. As a result of the difference with the value in Table 9, LiCoO2 has a formation energy of 0.05 kJ mol at−1 which is lower than that obtained without this correction. All these corrections to the formation energy of LiCoO2 are reported in Table 10.

Table 10 Formation energy of LiCoO2 in kJ mol at−1 according to the static method and the corrections of ZPE and energy gain at 300 K. The energy difference and relative error are calculated in relation to the average of available experimental data, −169.6 ± 1.7 kJ mol at−1.10,15,16
Method Formation energy δ th-exp Relative error (%)
Static −167.97 1.63 1.0
Including ZPE + T = 300 K, HA −168.73 0.87 0.5
Including ZPE + T = 300 K, HA + real CoO Cp −168.77 0.82 0.5


Influence of x–c functional and Hubbard U on the formation energy

As mentioned in the Methodology section, we tested not only reactions with no redox process (Type 1), listed in Table 3, but also other reactions with oxides involving redox (Type 2) and the reactions with pure elements in their standard states at 300 K (Type 3). These additional reactions are listed in Table 11. Reactions 5, 7 and 9 belong to Type 2 whereas reactions 6, 8 and 10 are of Type 3 (with gaseous molecular oxygen, metallic Li and transition element).
Table 11 Thermodynamic cycles employed to obtain formation energies of LiMO2 from reactions with redox processes. The experimental enthalpies of formation are listed inTable 4
Compound Reaction # Hess’ cycle and reactions Energy Transition metal electronic transfer
LiCoO2 5 Li2O2 + 2CoO → 2LiCoO2 ΔER5, calculated with DFT 2Co+2 → 2Co+3 + 2e
2Li + O2 → Li2O2 ΔELi2O2, experimental
Co + 0.5O2 → CoO ΔECoO, experimental
Li + Co + O2 → LiCoO2 ΔfE,R5,LiCoO2 = 0.5(ΔER5 + ΔELi2O2 + 2ΔECoO)
6 Li + Co + O2 → LiCoO2 ΔfE,R6,LiMnO2, calculated with DFT Co → Co+3 + 3e
LiMnO2 7 Li2O2 + 2MnO → 2LiMnO2 ΔER7, calculated with DFT 2Mn+2 → 2Mn+3 + 2e
2Li + O2 → Li2O2 ΔELi2O2, experimental
Mn + 0.5O2 → MnO ΔEMnO, experimental
Li + Mn + O2 → LiMnO2 ΔfE,R7,LiMnO2 = 0.5(ΔER8 + ΔELi2O2 + 2ΔEMnO)
8 Li + Mn + O2 → LiMnO2 ΔfE,R8,LiMnO2, calculated with DFT Mn → Mn+3 + 3e
LiNiO2 9 Li2O2 + 2NiO → 2LiNiO2 ΔER9, calculated with DFT 2Ni+2 → 2Ni+3 + 2e
2Li + O2 → Li2O2 ΔELi2O2, experimental
Ni + 0.5O2 → NiO ΔENiO, experimental
Li + Ni + O2 → LiNiO2 ΔfE,R9,LiNiO2 = 0.5(ΔER + ΔELi2O2 + 2ΔENiO)
10 Li + Ni + O2 → LiNiO2 ΔfE,R10,LiNiO2 calculated with DFT Ni → Ni+3 + 3e


We also computed the same reaction energies with other x–c functionals, including PBE + U with different U values.

Fig. 1 shows all our results of computed enthalpies of formation with the various x–c functionals and several Hubbard U values. For CALPHAD applications, we consider a precision of ±4 kJ mol at−1 to be acceptable with respect to the experimental value. This choice is of course somewhat arbitrary, but realistic. The calculations using PBE + U = 4.0 eV did not converge for CoO and, therefore, the enthalpy of formation of LiCoO2 with reactions that use this compound could not be computed.


image file: d3cp01771k-f1.tif
Fig. 1 Formation energy of (a) LiCoO2 (b) o-LiMnO2 and (c) LiNiO2 according to the DFT-computed reaction and the x–c functional. The dashed line is the experimental value. The numbered label on each point is the absolute value of the energy difference between the result and the experimental value.10,15–17,33 The dotted lines represent the precision threshold level of ± 4 kJ mol at−1 for CALPHAD applications.

Type 1 data are represented by triangles, Type 2 by squares and Type 3 by circles.

Type 1 data obtained with PBEsol, PW91, rSCAN, PBE or PBE + U with the lowest value of U (i.e. 2.6 eV for Co and 4.8 eV for Ni) show a very high degree of internal consistency as they differ by at most 1.2 kJ mol at−1.

In contrast, each of the other two sets of data points, Type 2 and Type 3, clearly shows more scattering with respect to the x–c functional. While in the Type 2 data set, this scattering remains limited between 5 and 20 kJ mol at−1 depending on the compound, and in the Type 3 data set, it becomes very large, ranging from 22 to 120 kJ mol at−1.

As for the comparison to experimental data is concerned, for each of the three studied compounds, the experimental value falls more or less in the middle of the whole set of data points which covers an interval of width ranging from about 62 kJ mol at−1 in the case of LiNiO2 to 120 kJ mol at−1 for LiMnO2. In other words, an uninformed choice of the DFT-computed reaction, the x–c functional and the Hubbard U correction might result in an estimate of the LMO formation energy with an error that is most likely around ±30 kJ mol at−1 and possibly up to ±60 kJ mol at−1. This is indeed the picture that is shown in Table 1.

If however we focus on the Type 1 data points, we are led to two very different conclusions. First, in the case of LiCoO2 and o-LiMnO2, the agreement with experimental data is extremely good, with discrepancies ranging from 0.01 to 2.8 kJ mol at−1. A small superior prediction capacity of PBEsol can be observed with a maximum difference of 1.6 kJ mol at−1 w.r.t experiment. We note that the good quality of PBEsol could be expected since this functional is designed for crystals, whereas PBE and PW91 have more generic applications. Second, the discrepancy of about 24 kJ mol at−1 in the case of LiNiO2 already noted from the examination of Table 7 becomes very striking here, considering once again the consistency of the DFT data across the various x–c functionals.

The Type 2 data set displays a systematic discrepancy w.r.t. experiment which is much larger than the scattering associated to the x–c functional. For the three LiMO2 compounds, this error is of the order of −20 to −30 kJ mol at−1.

No systematic trend can be drawn for the Type 3 data set as the scattering across the x–c functional is large and the shift w.r.t. the experiment does not have a definite sign. The formation energies estimates however seem to order almost systematically with Type 2 being the lowest and Type 3 the highest.

We shall come back to this in the Discussion section. In conclusion, the results for LiCoO2 and LiMnO2 very clearly reveal the different precision levels obtained with the no-redox reactions compared to reactions involving electron transfer. While the latter lead to overestimates and underestimates with offsets from the experimental value in the order of several tens of kJ mol at−1, the no-redox reactions consistently yield results within the ±4 kJ mol at−1 acceptable uncertainty margin.

The influence of the Hubbard U over the enthalpy of formation of LiCoO2 and LiNiO2 is quite large among all tested reactions. We address this subject in the Discussion section.

Heat capacity

The heat capacity of LiMO2 compounds was obtained within the HA and QHA. The HA only provides access to the heat capacity at constant volume (Cv). We calculate it using the static volume obtained at 0 K (Cv (T, V = Vstatic)). An important consequence is that pressure varies with temperature. Its value is out of reach of the HA but is an output of the QHA. For example, in the case of LiCoO2, it reaches 2.5 GPa at 300 K and 7.5 GPa at 1000 K. Thus, Cv is arguably not the most appropriate quantity to compare with the experimental heat capacity, which, for solid state, is always obtained at constant pressure and varying volume. By employing the QHA, we obtain a heat capacity at constant pressure (Cp (T, p = 1 atm)). Both Cv and Cp are plotted along with experimental data and Neumann–Kopp approximation. The supercells used for phonon calculations are described in Tables S6, S7 and S8 (ESI).

The calculated Cp for LiCoO2 is in very good agreement with experimental data, with a maximal error of 10% around 150 K. The results are given in Fig. 2.


image file: d3cp01771k-f2.tif
Fig. 2 Heat capacity of LiCoO2 and comparison with experimental measurements.10–14 A fit of experimental data was used to calculate the relative error of the calculated Cp. Two Neumann–Kopp approximations were derived: from Li2O2 + CoO and from Li2O + CoO + Co3O4.96

Regarding LiNiO2 (Fig. 3), we are limited to experimental measurements at low temperatures and we find an error of about 17% at 300 K, which also motivates further investigations.


image file: d3cp01771k-f3.tif
Fig. 3 Heat capacity of LiNiO2 and comparison with experimental data18 and with a Neumann–Kopp approximation from oxides Li2O2 and NiO.96

We have not found any report of experimental heat capacity measurements of LiMnO2 in the literature. For the phonon calculations, we chose the P2/c structure. Some imaginary vibration frequencies were present for values of static pressure starting between −3 and −4 GPa, which corresponds to a temperature in the range 250–300 K. Imaginary frequencies in general can be due to either an imprecision in the calculations or a crystal instability. There is no definitive solution on how to deal with them.97,98 In order to shed light on the origin of the problem we recomputed the phonon spectrum using the ABINIT code90–92 using a linear response. The imaginary modes were still present, which lead us to the conclusion that the structure is unstable. To bypass this problem, we calculate the Cp of o-LiMnO2 with a QHA proposed by Fleche,47 in which a single phonon calculation at null pressure is necessary. The results are displayed in Fig. 4.


image file: d3cp01771k-f4.tif
Fig. 4 Heat capacity of o-LiMnO2 in P2/c symmetry calculated with a QHA from Fleche47 and comparison with a two Neumann–Kopp approximation: from Li2O + Mn2O3 and from Li2O2 + MnO.57,96,99

Structural properties

Analogously to the enthalpy of formation, the cell parameters are properties measured at room temperature whereas the static DFT calculations correspond to 0 K and do not include zero-point vibrational (ZPV) corrections. We have corrected the ZPV and the volume expansion between 0 and 300 K as obtained with the QHA and we derived a specific mass. The results for LiCoO2 are presented in Table 12.
Table 12 LiCoO2 specific mass computed with the static and QHA methods at 300 K and difference with respect to the experimental value of Hertz et al.50
Method Specific mass (g cm−3) Difference (g cm−3) Relative error (%)
Static, 0 K and no ZPV 5.137 0.091 1.8
QHA, 300 K 5.048 0.002 0.04
Experimental value50 5.046


The good agreement between the experimental and the QHA values shows that static DFT computations of the cell volume must be compared to a corrected experimental volume of 96.7–1.7 = 95.0 Å3 rather than to the exact experimental volume at 300 K. We use this corrected value to compare with the static volume obtained using other functionals and the results are presented in Fig. 5.


image file: d3cp01771k-f5.tif
Fig. 5 Calculated cell volume of LiCoO2 for different x–c and comparison with an estimated experimental static volume derived from ref. 50.

The calculated bulk modulus, the coefficient of volumetric thermal expansion (CTE) and the specific mass at 300 K are listed in Table 13.

Table 13 Structural properties of LiMO2 end-members at 300 K obtained with the QHA
Compound Bulk modulus (GPa) CTE (K−1) Specific mass (g cm−3)
LiCoO2 146.4 3.39 × 10−5 5.048
o-LiMnO2 83.9 6.25 × 10−5 4.213
LiNiO2 111.0 4.76 × 10−5 4.774


The experimental bulk modulus determined by Hu et al.100 for LiCoO2 is 159.5 ± 2.2 GPa.

Discussion

Static formation energy

As illustrated by Fig. 1, the performance of DFT in computing the formation energies of LiCoO2 and LiMnO2 strongly depends on the type of DFT-computed reaction. Type 1 no-redox reactions give significantly better agreement between DFT and experiment than Type 2 and Type 3 reactions. Another striking feature displayed in Fig. 1 is that formation energies obtained with Type 1 reactions are much less dependent on the x–c functional than Type 3 data, while Type 2 data show intermediate sensitivity. A third point of relevance in this discussion is that, as shown by Table 14, the reaction energy of Type 1 reactions is much closer to zero than that of Type 2 or Type 3 reactions.
Table 14 Comparison of the reaction energy ΔER (in kJ per mol of reaction) of the DFT-computed reaction and the error δth-exp (in kJ mol at−1) on the formation energy of LiMO2
DFT-computed reaction x–c functional ΔER δ th-exp
Li2O + Co3O4 → 2LiCoO2 + CoO (no-redox) PBEsol −73.6 1.6
PW91 −67.8 2.4
rSCAN −64.2 2.8
PBE −67.4 2.4
Li2O2 + 2CoO → 2LiCoO2 PBEsol −452.8 31.2
PW91 −412.2 26.1
rSCAN −402.9 10.2
PBE −409.0 25.7
Li + Co + O2 → LiCoO2 PBEsol −607.8 21.6
PW91 −629.0 16.4
rSCAN −647.6 26.4
PBE −576.2 29.6
Li2O + Mn2O3 → 2LiMnO2 (no-redox) PBEsol −122.9 0.01
PW91 −115.3 1.0
rSCAN −114.5 1.1
PBE −114.7 1.0
Li2O2 + 2MnO → 2LiMnO2 PBEsol −503.0 29.7
PW91 −464.6 24.9
rSCAN −435.2 21.3
PBE −463.0 24.7
Li + Mn + O2 → LiMnO2 PBEsol −527.8 77.9
PW91 −803.4 9.0
rSCAN −990.9 37.8
PBE −728.1 27.8


Lastly, Table 14 illustrates that the discrepancy δth-exp between calculation and experiment on the LiMO2 formation energy ΔrELiMO2 is much larger for Type 2 and Type 3 reactions than for Type 1 reactions. In summary, we point out a striking correlation between four properties of the DFT-computed reaction: (P1) constancy of the formal charges, (P2) insensitivity to the x–c functional, (P3) smallness of the energy of the DFT-computed reaction, and (P4) good agreement between calculation and experiment on ΔrELiMO2. Type 1 and Type 3 reactions best illustrate the extreme cases, while Type 2 reactions are intermediate. Although we have not made any additional theoretical calculations to support this, this correlation lends itself to a simple qualitative interpretation based on the approximate character of the x–c functionals (w.r.t. the universal—and unknown—exact functional) and on a partition of the electron density among ions reflecting more or less their formal charge. In a Type 1 reaction, if we suppose that the sets of ions do not change much between reactants and products, the errors intrinsic to the x–c functionals on each ion separately compensate between reactants and products. The differences between the various functionals are hidden behind this compensation effect (P2) and the formation energy based on such a reaction is in good agreement with the experiment (P4). As for property P3, it is intuitively consistent with the similarities in ionic charges and coordination. In a Type 3 reaction, equivalent but opposite arguments may be given: large changes in formal charges reflect large electron density changes within each ion, so that the associated energy values differ according to the x–c functionals which implies that not all of these values can agree with the experimental value. What remains puzzling in this qualitative interpretation is that it is based on formal charges, which are difficult to reconcile with the details of the electronic structure, e.g. they hide the fact that bonding always includes a significant covalent contribution. In fact, the population analysis output by CASTEP in our calculations reveals changes in Mulliken (both charges and bond orders) and Hirshfeld populations during Type 1 reactions that are not necessarily smaller than in Type 2 reactions. Thus, the constancy of formal charges may just reflect a global compensation of various microscopic evolutions in ionic charge, covalency, bond lengths, etc, and this remains to be investigated more closely. As to the partitioning of the electron density among ions, it could be investigated using differences in charge density maps and Bader charges and this will be considered for future work.

In summary, we propose that the success of our methodology based on no-redox DFT-computed reactions relies on a compensation mechanism of the error due to the approximate character of the exchange–correlation (x–c) functional.

If we now go back to Fig. 1 and focus on Type 1 (triangles) data points, the question arises why agreement with experimental data is excellent for LiCoO2 and LiMnO2, with a discrepancy δth-exp ≈ 1 kJ mol at−1, while in the case of LiNiO2, the disagreement is very clear. Since stable oxides of nickel are limited to NiO with Ni(II), we first chose to base the DFT contribution on a reaction with Li2O, NaNiO2 and Na2O, which does not involve any redox process. With this reaction, we have obtained an enthalpy of formation that is about 24 kJ mol at−1 lower than the experimental value. We then studied another no-redox reaction with Li2O, NiTiO3 and Ti2O3 to help locate the source of this large discrepancy between DFT and experiment. Since this reaction leads to a very similar result as the previous one, we tested the accuracy of the proposed method in computing the formation energy of NiTiO3, revealing a satisfactory result (Table 8) with a discrepancy of 0.19 kJ mol at−1. This very good agreement, together with that obtained previously in cycles involving Li2O, suggests LiNiO2 as being the only source of the disagreement observed in Table 2. It is also worth noting that all the studies reported in Table 1 which calculated the LiNiO2 formation energy using DFT and obtained satisfactory results used either an empirical method24 or an adjustment of the Ni-3d Hubbard U value.28 Thus, we conclude that our results motivate a reassessment of the experimental data for the LiNiO2 heat of formation.

In our methodology, the choice of the no-redox DFT-computed LiMO2 formation reaction is largely determined by the availability of experimental data for the formation energy of the other compounds taking part in the Hess’ cycle. While this might, at first sight, be considered as a weakness hindering a wide use of the methodology, in practice, this has so far not been a limitation. For instance, a straightforward no-redox reaction from binary oxides and LiCoO2 is not possible, since Co2O3 is metastable, but we could easily bypass this issue by employing other stable binary oxides (CoO and Co3O4). Further, within this no-redox constraint, we have even been able to check that the choice of Hess’ cycle does not influence the final result. There remains to uncover the source of the discrepancy regarding the formation enthalpy of LiNiO2, but overall we are confident that our methodology based on no-redox reactions deserved to be pursued further.

DFT+U results

Regarding the DFT+U results, we originally expected the influence of the U value on the Type 1 results to be small. This is true to some extent, given that the Hubbard correction shifts the energy of each of the compounds by several hundreds of kJ mol at−1 while the resulting LiMO2 formation energies only change by at most a few tens of kJ mol at−1. But this remainder appears quite large on the scale of Fig. 1, which also displays the better results obtained without the Hubbard correction. We observe that the residual sensitivity to U is rather large compared to the precision required by our target application of CALPHAD modeling. Due to the uncertainty affecting the correct value of the input parameter U, this example also illustrates how difficult it is to improve the precision of DFT+U formation enthalpies below the average 5 kJ mol at−1 level obtained by semi-empirical fitting methods such as those of Wang et al.36 or Stevanovic et al.101 Further, the scattering of the LiCoO2 Type 1 results in Fig. 1a, spread over about 25 kJ mol at−1, may be due to some additional reason. We think that there may be an issue with Co3O4 since it contains both Co(II) and Co(III) whereas we assigned one and the same U value to all Co ions. Metastable electronic states arising with DFT+U are a known and complex problem.102 This possible issue with the computation of Co3O4 with GGA+U is out of the scope of the present work. Overall, the data in Fig. 1 illustrate a sensitivity of ΔE to U which, combined with the uncertainty about the value of U and the potential pitfall of metastable electronic states, may explain the rather poor performance of DFT+U in the present case. In comparison, the straightforward and parameter-free combination of standard DFT, PBEsol and no-redox reactions leads to a good agreement on thermodynamic and structural properties.

Vibrational contribution to the formation energy

The corrections due to zero point energy and temperature increments result in a formation energy of LiCoO2 0.76 kJ mol at−1 closer to the average of the experimental data, while the discrepancy remains in the range of 1 to 4 kJ mol at−1 depending on the source of experimental data. The smallness of this value is consistent with the widespread point of view that vibrational corrections to the formation energy are in generally below the so-called 1 kcal mol at−1 chemical accuracy, but we think that here it is mostly due to the compensation effect behind the use of a no-redox reaction. Nevertheless, considering that the uncertainty in the experimental measurement of the LiCoO2 enthalpy of formation was estimated by their authors to be 0.7 kJ mol at−1,10 we conclude that the ZPE and the vibrational contribution at 300 K obtained within the HA make up a relevant and valuable complement to static energy estimation. The additional precision brought by the QHA seems unnecessary for the Type 1 DFT-computed reaction.

Structural properties

Even though the QHA does not contribute significantly to the formation energies obtained with Type 1 Hess’ cycles, it remains essential to reach a proper description of the structural data. For example, the LiCoO2 specific mass output by the QHA model at 300 K has a discrepancy of only 0.002 g cm−3, as compared to 0.091 g cm−3 when only the static method is used (Table 12). Among the studied x–c functionals, PBEsol provided the most accurate specific mass. We recall that the differences in cell volumes obtained with various x–c functionals are of the same order of magnitude as the difference between the static volume and the QHA estimate at 300 K, so that a meaningful comparison with experimental data can only be done if the latter correction is taken into account, as we did in Fig. 5. Within the QHA, we calculate the bulk modulus B(300 K) of LiCoO2 and an underestimation of 8% in relation to the experimental value was found.

Heat capacity

The QHA is also essential for heat capacity data at higher temperatures, as shown by the differences between Cp and Cv (Fig. 2, Fig. 3 and Fig. 4). Our computation method for the heat capacity could be validated with LiCoO2 for which numerous measurements are available from 0 to 1200 K. As for LiNiO2, the heat capacity agrees very well with experimental data in a temperature range of approximately 0–200 K. An increasing deviation is observed between 200 K and 300 K, the maximum temperature at which experimental Cp data are available. The maximal divergence occurs at 300 K, with a relative error of 17.2%. Our phonon data allowed us to compute Cp up to 1200 K. The discrepancy at 300 K requires a validation that will require new measurements to be made, in order to give more confidence in our prediction for the higher temperatures. As for o-LiMnO2 for which no experimental data are available, we compute a Cp curve that might be employed as an entry in a thermodynamic database. We note however that the presence of unstable vibration modes beyond a specific volume corresponding to a temperature of about 300 K also needs experimental validation. Note that a significant structural cationic disorder has been observed in LiMnO2 by Kellerman et al.51 and the thermodynamic contributions arising from this disorder remain to be assessed. For the CALPHAD application, as a first approximation, it is reasonable to assume that the Cp value of r-LiMnO2 is the same as that obtained for o-LiMnO2. Such an assumption is in fact a kind of Neumann–Kopp approximation, more precise than that obtained from binary oxides or pure elements. For LiMnO2, the calculated Cp agrees with the two different Neumann–Kopp approximations proposed. For LiCoO2, however, the calculated Cp value is in clear disagreement with two Neumann–Kopp approximations based on binary oxides, even with the one based on the Type 1 no-redox reaction, and these two N–K approximations also disagree with each other. Thus, more investigations are needed on the heat capacity of these compounds. This is also a reminder that the N–K approximation should always be used with caution.

Conclusion

In the present study, a new methodology for computing the structural and thermodynamic properties of LiMO2 (M = Co, Mn and Ni) was proposed and assessed using experimental results available for comparison. This paves the way for the study of the NMC system of which they are end-members. The properties were computed by means of DFT and using three levels of approximation: static, harmonic and quasi-harmonic.

For each of these compounds, we compared several Hess’ cycles involving different DFT-computed formation reactions. The best performance is clearly obtained using a Hess’ cycle in which no redox process occurs in the DFT-computed reaction involving the target compound LiMO2 (M = Co, Ni, Mn). These no-redox reactions consistently give the best accuracy for the formation energy, between 0.01 and 1.6 kJ mol at−1 for LiCoO2 and o-LiMnO2. This scheme naturally avoids the use of DFT estimates of molecular O2. This very high precision and the fact that the same results are reproduced with four different x–c functionals enabled us to uncover an unexplained and large discrepancy of 24 kJ mol at−1 for the LiNiO2 formation energy, confirmed by reproducing the calculation with two different Hess cycles. This motivates further investigation on the experimental value of the LiNiO2 standard enthalpy of formation. The enthalpy of formation of the metastable rhombohedral LiMnO2 phase was computed and it is about 1.0 kJ mol at−1 more positive than that of o-LiMnO2.

The harmonic vibrational contribution to the formation energy brings the result 0.76 kJ mol at−1 closer to the LiCoO2 experimental value, resulting in a prediction in line with the level of accuracy required for thermodynamic CALPHAD calculations. Our results also show that, even though the contributions obtained with the QHA are not significant to formation energy computations when using no-redox DFT-computed reactions, they are relevant for the best agreement with structural experimental data at 300 K and the heat capacity obtained at temperatures above 300 K. We found that PBEsol performed best for all the properties computed in this study. A very good agreement between calculated and experimental heat capacities at constant pressure (P = 1 atm) was obtained for LiCoO2, enabling us to propose estimates for the heat capacity of LiNiO2 at high temperatures as well as the heat capacity of o-LiMnO2, hitherto unknown experimentally. As for the methodology to formation energy calculations using no-redox DFT-computed reactions, which gives very good results for LiCoO2 and LiMnO2, the question arises whether it might be generalized. It may be considered in a wide array of problems provided formal charges may be defined and a reaction without changes of formal charges can be found, involving compounds with known experimental enthalpies of formation. The definition of formal charges is not universally possible, but it is possible at least for the large family of oxides. While we have no theoretical support for this methodology, we trust that the literature will eventually provide more data on which to base an assessment of this methodology.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We are grateful to Dr Jean-Louis Flèche for providing us with his QHA code. This work was supported by the Cross-cutting basic research Program (RTA Program) of the CEA Energy Division.

References

  1. N. Zhang, J. Li, H. Li, A. Liu, Q. Huang, L. Ma, Y. Li and J. R. Dahn, Chem. Mater., 2018, 30, 8852–8860 CrossRef CAS.
  2. W. Zhang, D. M. Cupid, P. Gotcu, K. Chang, D. Li, Y. Du and H. J. Seifert, Chem. Mater., 2018, 30, 2287–2298 CrossRef CAS.
  3. N. Li, D. Li, W. Zhang, K. Chang, F. Dang, Y. Du and H. J. Seifert, Prog. Nat. Sci.: Mater. Int., 2019, 29, 265–276 CrossRef CAS.
  4. A. A. Luo, Calphad, 2015, 50, 6–22 CrossRef CAS.
  5. K. Chang, B. Hallstedt, D. Music, J. Fischer, C. Ziebert, S. Ulrich and H. J. Seifert, Calphad, 2013, 41, 6–15 CrossRef CAS.
  6. K. Chang, B. Hallstedt and D. Music, Calphad, 2012, 37, 100–107 CrossRef CAS.
  7. R. Robert, C. Villevieille and P. Novák, J. Mater. Chem. A, 2014, 2, 8589 RSC.
  8. B. Ammundsen and J. Paulsen, Adv. Mater., 2001, 13, 943–956 CrossRef CAS.
  9. M. Bianchini, M. Roca-Ayats, P. Hartmann, T. Brezesinski and J. Janek, Angew. Chem., Int. Ed., 2019, 58, 10434–10458 CrossRef CAS PubMed.
  10. P. Gotcu-Freis, D. M. Cupid, M. Rohde and H. J. Seifert, J. Chem. Thermodyn., 2015, 84, 118–127 CrossRef CAS.
  11. H. Kawaji, M. Takematsu, T. Tojo, T. Atake, A. Hirano and R. Kanno, J. Therm. Anal. Calorim., 2002, 68, 833–839 CrossRef CAS.
  12. A. L. Emelina, M. A. Bykov, M. L. Kovba, B. M. Senyavin and E. V. Golubina, Russ. J. Phys. Chem., 2011, 85, 357–363 CrossRef CAS.
  13. M. Ménétrier, D. Carlier, M. Blangero and C. Delmas, Electrochem. Solid-State Lett., 2008, 11, A179–A182 CrossRef.
  14. O. Jankovský, J. Kovařík, J. Leitner, K. Ruzicka and D. Sedmidubský, Thermochim. Acta, 2016, 634, 26–30 CrossRef.
  15. M. Wang, A. Navrotsky, S. Venkatraman and A. Manthiram, J. Electrochem. Soc., 2005, 152, J82–J84 CrossRef CAS.
  16. M. Masoumi, D. M. Cupid, T. L. Reichmann, K. Chang, D. Music, J. M. Schneider and H. J. Seifert, Int. J. Mater. Res., 2017, 108, 869–878 CAS.
  17. M. Wang and A. Navrotsky, Solid State Ionics, 2004, 166, 167–173 CrossRef CAS.
  18. H. Kawaji, Solid State Ionics, 2002, 152–153, 195–198 CrossRef CAS.
  19. P. Strobel, J.-P. Levy and J.-C. Joubert, J. Cryst. Grow., 1984, 66, 257–261 CrossRef CAS.
  20. P. Hohenberg and W. Kohn, Phys. Rev., 1964, 136, B864–B871 CrossRef.
  21. W. Kohn and L. J. Sham, Phys. Rev., 1965, 140, A1133–A1138 CrossRef.
  22. L. Wu, W. H. Lee and J. Zhang, Mater. Today: Proc., 2014, 1, 82–93 CrossRef.
  23. T. Du, B. Xu, M. Wu, G. Liu and C. Ouyang, J. Phys. Chem. C, 2016, 120, 5876–5882 CrossRef CAS.
  24. K. Chang, B. Hallstedt and D. Music, Chem. Mater., 2012, 24, 97–105 CrossRef CAS.
  25. H. Xu, W. Xiao, Z. Wang, J. Hu and G. Shao, J. Energy Chem., 2021, 59, 229–241 CrossRef CAS.
  26. D. Kramer and G. Ceder, Chem. Mater., 2009, 21, 3799–3809 CrossRef CAS.
  27. R. C. Longo, F. T. Kong, S. Kc, M. S. Park, J. Yoon, D.-H. Yeon, J.-H. Park, S.-G. Doo and K. Cho, Phys. Chem. Chem. Phys., 2014, 16, 11233–11242 RSC.
  28. M. Tuccillo, O. Palumbo, M. Pavone, A. B. Muñoz-García, A. Paolone and S. Brutti, Crystals, 2020, 10, 526 CrossRef CAS.
  29. N. T. Tsebesebe, K. Kgatwane, R. Ledwaba and P. Ngoepe, J. Phys.: Conf. Ser., 2022, 2298, 012010 CrossRef.
  30. E. Lee, K.-R. Lee and B.-J. Lee, Comput. Mater. Sci., 2018, 142, 47–58 CrossRef CAS.
  31. K. Refson, P. R. Tulip and S. J. Clark, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 155114 CrossRef.
  32. M. Masoumi, Dissertation, KIT, 2019.
  33. M. Wang and A. Navrotsky, J. Solid State Chem., 2005, 178, 1230–1240 CrossRef CAS.
  34. D. C. Patton, D. V. Porezag and M. R. Pederson, Phys. Rev. B: Condens. Matter Mater. Phys., 1997, 55, 7454–7459 CrossRef CAS.
  35. L. Wang, T. Maxisch and G. Ceder, Chem. Mater., 2007, 19, 543–552 CrossRef CAS.
  36. L. Wang, T. Maxisch and G. Ceder, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 195107 CrossRef.
  37. A. Jain, G. Hautier, S. P. Ong, C. J. Moore, C. C. Fischer, K. A. Persson and G. Ceder, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 045115 CrossRef.
  38. C. Franchini, R. Podloucky, J. Paier, M. Marsman and G. Kresse, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 195128 CrossRef.
  39. F. Zhou, M. Cococcioni, C. A. Marianetti, D. Morgan and G. Ceder, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 70, 235121 CrossRef.
  40. A. Chakraborty, S. Kunnikuruvan, S. Kumar, B. Markovsky, D. Aurbach, M. Dixit and D. T. Major, Chem. Mater., 2020, 32, 915–952 CrossRef CAS.
  41. D. K. Chakrabarty, An Introduction to Physical Chemistry, Alpha Science International, Limited, 2001 Search PubMed.
  42. J. P. Perdew, A. Ruzsinszky, G. I. Csonka, O. A. Vydrov, G. E. Scuseria, L. A. Constantin, X. Zhou and K. Burke, Phys. Rev. Lett., 2008, 100, 136406 CrossRef PubMed.
  43. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  44. J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh and C. Fiolhais, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 46, 6671–6687 CrossRef CAS PubMed.
  45. A. P. Bartók and J. R. Yates, J. Chem. Phys., 2019, 150, 161101 CrossRef PubMed.
  46. M. Born and K. Huang, Am. J. Phys., 1955, 23, 474 CrossRef.
  47. J. L. Fleche, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 65, 245116 CrossRef.
  48. A. Togo and I. Tanaka, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 184104 CrossRef.
  49. M. I. Aroyo, J. M. Perez-Mato, C. Capillas, E. Kroumova, S. Ivantchev, G. Madariaga, A. Kirov and H. Wondratschek, Z. Kristallogr. - Cryst. Mater., 2006, 221, 15–27 CrossRef CAS.
  50. J. T. Hertz, Q. Huang, T. McQueen, T. Klimczuk, J. W. G. Bos, L. Viciu and R. J. Cava, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 77, 075119 CrossRef.
  51. D. G. Kellerman, J. E. Medvedeva, V. S. Gorshkov, A. I. Kurbakov, V. G. Zubkov, A. P. Tyutyunnik and V. A. Trunov, Solid State Sci., 2007, 9, 196–204 CrossRef CAS.
  52. G. M. Kale, S. S. Pandit and K. T. Jacob, Trans. JIM, 1988, 29, 125–132 CrossRef CAS.
  53. W. L. Roth, J. Phys. Chem. Solids, 1964, 25, 1–10 CrossRef CAS.
  54. B. J. Boyle, E. G. King and K. C. Conway, J. Am. Chem. Soc., 1954, 76, 3835–3837 CrossRef CAS.
  55. I. Barin, O. Knacke and O. Kubaschewski, Thermochemical properties of inorganic substances, Springer, Berlin, Heidelberg, 1st edn, 1991 Search PubMed.
  56. W. L. Roth, Phys. Rev., 1958, 110, 1333–1341 CrossRef CAS.
  57. K. T. Jacob, A. Kumar, G. Rajitha and Y. Waseda, High Temp. Mater. Processes, 2011, 30, 459–472 CAS.
  58. R. Shivaramaiah, S. Tallapragada, G. P. Nagabhushana and A. Navrotsky, J. Solid State Chem., 2019, 280, 121011 CrossRef CAS.
  59. C. Darie, P. Bordet, S. de Brion, M. Holzapfel, O. Isnard, A. Lecchi, J. E. Lorenzo and E. Suard, Eur. Phys. J. B, 2005, 43, 159–162 CrossRef CAS.
  60. O. Kubaschewski, C. B. Alcock and P. J. Spencer, Materials Thermochemistry, Pergamon, New York, 6th edn, 1993 Search PubMed.
  61. G. Shirane, S. J. Pickart and Y. Ishikawa, J. Phys. Soc. Jpn., 1959, 14, 1352–1360 CrossRef CAS.
  62. G. L. Humphrey, J. Am. Chem. Soc., 1951, 73, 1587–1590 CrossRef CAS.
  63. P. H. Carr and S. Foner, J. Appl. Phys., 1960, 31, S344–S345 CrossRef.
  64. G. K. Johnson, R. T. Grow and W. N. Hubbard, J. Chem. Thermodyn., 1975, 7, 781–786 CrossRef CAS.
  65. K. Chang and B. Hallstedt, Calphad, 2011, 35, 160–164 CrossRef CAS.
  66. D. Hobbs, J. Hafner and D. Spišák, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 68, 014407 CrossRef.
  67. M. J. Redman and E. G. Steward, Nature, 1962, 193, 867 CrossRef CAS.
  68. Y. C. Sui, Y. Zhao, J. Zhang, S. Jaswal, X. Z. Li and D. J. Sellmyer, IEEE Trans. Magn., 2007, 43, 3115–3117 CAS.
  69. A. Ghosh, S. Jana, M. K. Niranjan, F. Tran, D. Wimberger, P. Blaha, L. A. Constantin and P. Samal, J. Phys. Chem. C, 2022, 126, 14650–14660 CrossRef CAS.
  70. W. Jauch, M. Reehuis, H. J. Bleif, F. Kubanek and P. Pattison, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 64, 052102 CrossRef.
  71. R. Hoppe, G. Brachtel and M. Jansen, Z. Anorg. Allg. Chem., 1975, 417, 1–10 CrossRef CAS.
  72. G. Dittrich and R. Hoppe, Z. Anorg. Allg. Chem., 1969, 368, 262–270 CrossRef CAS.
  73. C. Pouillerie, E. Suard and C. Delmas, J. Solid State Chem., 2001, 158, 187–197 CrossRef CAS.
  74. J. Cao, H. Zou, C. Guo, Z. Chen and S. Pu, Solid State Ionics, 2009, 180, 1209–1214 CrossRef CAS.
  75. A. Rougier, C. Delmas and A. V. Chadwick, Solid State Commun., 1995, 94, 123–127 CrossRef CAS.
  76. J.-H. Chung, Th Proffen, S. Shamoto, A. M. Ghorayeb, L. Croguennec, W. Tian, B. C. Sales, R. Jin, D. Mandrus and T. Egami, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 71, 064410 CrossRef.
  77. H. Das, A. Urban, W. Huang and G. Ceder, Chem. Mater., 2017, 29, 7840–7851 CrossRef CAS.
  78. C. A. Marianetti, D. Morgan and G. Ceder, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 63, 224304 CrossRef.
  79. H. Chen, C. L. Freeman and J. H. Harding, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 085108 CrossRef.
  80. Z. Chen, H. Zou, X. Zhu, J. Zou and J. Cao, J. Solid State Chem., 2011, 184, 1784–1790 CrossRef CAS.
  81. S. Sicolo, M. Mock, M. Bianchini and K. Albe, Chem. Mater., 2020, 32, 10096–10103 CrossRef CAS.
  82. M. Manago, G. Motoyama, S. Nishigori, K. Fujiwara, K. Kinjo, S. Kitagawa, K. Ishida, K. Akiba, S. Araki, T. C. Kobayashi and H. Harima, J. Phys. Soc. Jpn., 2022, 91, 113701 CrossRef.
  83. A. Pulkkinen, B. Barbiellini, J. Nokelainen, V. Sokolovskiy, D. Baigutlin, O. Miroshkina, M. Zagrebin, V. Buchelnikov, C. Lane, R. S. Markiewicz, A. Bansil, J. Sun, K. Pussi and E. Lähderanta, Phys. Rev. B, 2020, 101, 075115 CrossRef CAS.
  84. S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. I. J. Probert, K. Refson and M. C. Payne, Z. Kristallogr. - Cryst. Mater., 2005, 220, 567–570 CrossRef CAS.
  85. Z. Q. Li and Y. Kawazoe, Phys. Rev. Lett., 1997, 78, 4063–4066 CrossRef.
  86. J. H. Lloyd-Williams and B. Monserrat, Phys. Rev. B: Condens. Matter Mater. Phys., 2015, 92, 184301 CrossRef.
  87. M. C. Payne, M. P. Teter, D. C. Allan, T. A. Arias and J. D. Joannopoulos, Rev. Mod. Phys., 1992, 64, 1045–1097 CrossRef CAS.
  88. Materials Studio, BIOVIA, Dassault Systèmes (version 20.1.0.2728) 2020.
  89. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
  90. A. H. Romero, D. C. Allan, B. Amadon, G. Antonius, T. Applencourt, L. Baguet, J. Bieder, F. Bottin, J. Bouchet, E. Bousquet, F. Bruneval, G. Brunin, D. Caliste, M. Côté, J. Denier, C. Dreyer, P. Ghosez, M. Giantomassi, Y. Gillet, O. Gingras, D. R. Hamann, G. Hautier, F. Jollet, G. Jomard, A. Martin, H. P. C. Miranda, F. Naccarato, G. Petretto, N. A. Pike, V. Planes, S. Prokhorenko, T. Rangel, F. Ricci, G.-M. Rignanese, M. Royo, M. Stengel, M. Torrent, M. J. van Setten, B. Van Troeye, M. J. Verstraete, J. Wiktor, J. W. Zwanziger and X. Gonze, J. Chem. Phys., 2020, 152, 124102 CrossRef CAS PubMed.
  91. M. Torrent, F. Jollet, F. Bottin, G. Zérah and X. Gonze, Comput. Mater. Sci., 2008, 42, 337–351 CrossRef CAS.
  92. X. Gonze, B. Amadon, G. Antonius, F. Arnardi, L. Baguet, J.-M. Beuken, J. Bieder, F. Bottin, J. Bouchet, E. Bousquet, N. Brouwer, F. Bruneval, G. Brunin, T. Cavignac, J.-B. Charraud, W. Chen, M. Côté, S. Cottenier, J. Denier, G. Geneste, P. Ghosez, M. Giantomassi, Y. Gillet, O. Gingras, D. R. Hamann, G. Hautier, X. He, N. Helbig, N. Holzwarth, Y. Jia, F. Jollet, W. Lafargue-Dit-Hauret, K. Lejaeghere, M. A. L. Marques, A. Martin, C. Martins, H. P. C. Miranda, F. Naccarato, K. Persson, G. Petretto, V. Planes, Y. Pouillon, S. Prokhorenko, F. Ricci, G.-M. Rignanese, A. H. Romero, M. M. Schmitt, M. Torrent, M. J. van Setten, B. Van Troeye, M. J. Verstraete, G. Zérah and J. W. Zwanziger, Comput. Phys. Commun., 2020, 248, 107042 CrossRef CAS.
  93. F. Bottin, S. Leroux, A. Knyazev and G. Zérah, Comput. Mater. Sci., 2008, 42, 329–336 CrossRef CAS.
  94. C. Lee and X. Gonze, Phys. Rev. B: Condens. Matter Mater. Phys., 1995, 51, 8610–8613 CrossRef CAS PubMed.
  95. M. Chen, B. Hallstedt and L. J. Gauckler, J. Phase Equilib., 2003, 24, 212–227 CrossRef CAS.
  96. P. J. Linstrom and W. G. Mallard, J. Chem. Eng. Data, 2001, 46, 1059–1063 CrossRef CAS.
  97. G. J. Ackland, M. C. Warren and S. J. Clark, J. Phys.: Condens. Matter, 1997, 9, 7861 CrossRef CAS.
  98. C.-C. Lee, C.-E. Hsu and H.-C. Hsueh, J. Phys.: Condens. Matter, 2020, 33, 055902 CrossRef PubMed.
  99. J. L. Shapiro, B. F. Woodfield, R. Stevens, J. Boerio-Goates and M. L. Wilson, J. Chem. Thermodyn., 1999, 31, 725–739 CrossRef CAS.
  100. Y.-Q. Hu, L. Xiong, X.-Q. Liu, H.-Y. Zhao, G.-T. Liu, L.-G. Bai, W.-R. Cui, Y. Gong and X.-D. Li, Chinese Phys. B, 2019, 28, 016402 CrossRef CAS.
  101. V. Stevanović, S. Lany, X. Zhang and A. Zunger, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 115104 CrossRef.
  102. A. B. Shick, W. E. Pickett and A. I. Liechtenstein, J. Electron Spectrosc. Relat. Phenom., 2001, 114–116, 753–758 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cp01771k

This journal is © the Owner Societies 2023
Click here to see how this site uses Cookies. View our privacy policy here.