Oier
Arcelus
a,
Sergey
Nikolaev
b,
Javier
Carrasco
*a and
Igor
Solovyev
*bc
aCIC Energigune, Albert Einstein 48, 01510 Miñano, Aĺava, Spain. E-mail: jcarrasco@cicenergigune.com
bInternational Center for Materials Nanoarchitectonics, National Institute for Materials Science, 1-1 Namiki, Tsukuba, Ibaraki 305-0044, Japan. E-mail: SOLOVYEV.Igor@nims.go.jp; Fax: +81 29 860 4974; Tel: +81 29 860 4968
cDepartment of Theoretical Physics and Applied Mathematics, Ural Federal University, Mira str. 19, 620002 Ekaterinburg, Russia
First published on 27th April 2018
Magnetic properties of maricite (m) and triphlyte (t) polymorphs of NaFePO4 are investigated by combining ab initio density functional theory with a model Hamiltonian approach, where a realistic Hubbard-type model for magnetic Fe 3d states in NaFePO4 is constructed entirely from first-principles calculations. For these purposes, we perform a comparative study based on the pseudopotential and linear muffin-tin orbital methods while tackling the problem of parasitic non-sphericity of the exchange–correlation potential. Upon calculating the model parameters, magnetic properties are studied by applying the mean-field Hartree–Fock approximation and the theory of superexchange interactions to extract the corresponding interatomic exchange parameters. Despite some differences, the two methods provide a consistent description of the magnetic properties of NaFePO4. On the one hand, our calculations reproduce the correct magnetic ordering for t-NaFePO4 allowing for magnetoelectric effect, and the theoretical values of Néel and Curie–Weiss temperatures are in fair agreement with reported experimental data. Furthermore, we investigate the effect of chemical pressure on magnetic properties by substituting Na with Li and, in turn, we explain how a noncollinear magnetic alignment induced by an external magnetic field leads to magnetoelectric effect in NaFePO4 and other transition-metal phosphates. However, the origin of a magnetic superstructure with q = (1/2, 0, 1/2) observed experimentally in m-NaFePO4 remains puzzling. Instead, we predict that competing exchange interactions can lead to the formation of magnetic superstructures along the shortest orthorhombic c axis of m-NaFePO4, similar to multiferroic manganites.
Apart from their good electrochemical properties, other facets of potential interest in polyanionic compounds are however less explored. One prominent example is magnetism.3 Indeed, 3d-metal-based polyanionic compounds show unusual magnetic properties that can give rise to many appealing phenomena such as magnetoelectricity5–12 and even multiferroicity.13 The origin of such a rich magnetic behavior is a consequence of the combination of M–O–M super-exchange and M–O–X–O–M super-super-exchange interactions in these materials, which can induce the emergence of complex magnetic structures.
In this work, we focus on the magnetic properties of sodium iron phosphate (NaFePO4), the sodium counterpart of lithium iron phosphate (LiFePO4). LiFePO4 is one of the most studied cathode materials for today's Li-ion batteries.14 Yet the high abundance, environment-friendly nature, and low cost of sodium have made the research in Na-based electrode materials a topic of high interest,15–22 with an ongoing flurry of activity in NaFePO4 (see, e.g., Fang et al.23). NaFePO4 crystallizes in two different polymorphic forms: triphlyte (t) and maricite (m). m-NaFePO4 is the thermodynamically stable phase and shows many similarities with orthorhombically distorted transition-metal (TM) perovskite oxides,24,25 whereas t-NaFePO4 is isostructural to LiFePO4. Avdeev et al. studied both polymorphs by means of neutron powder diffraction (NPD) experiments and magnetic susceptibility measurements.26 The magnetic properties of NaFePO4 are indeed very interesting and rather complex. According to NPD measurements,26 t-NaFePO4 forms the same simple antiferromagnetic (AFM) order as LiFePO4,27 which is expected to reveal magnetoelectric effect.8 On the other hand, m-NaFePO4 tends to form a magnetic superstructure with the propagation vector q = (1/2, 0, 1/2),26 which implies the existence of competing magnetic interactions in the system. The situation is particularly interesting in the light that there are many examples of orthorhombic manganites, where such competition also results in the formation of magnetic superstructures that break inversion symmetry and give rise to the multiferroic behavior.28–30 The magnetic transition temperature, TN, is also very different, 50 K in t-NaFePO4 and only 13 K in m-NaFePO4, while the Curie–Weiss temperature (θ) is comparable in both phases and is about −80 K. The first theoretical attempt to study the magnetic properties of NaFePO4 was undertaken by Kim et al. using brute force total-energy calculations based on density functional theory (DFT) with a semi-phenomenological on-site Coulomb repulsion U.31 They concluded that while the magnetic structure of t-NaFePO4 can be understood by the spin exchange alone, the one of m-NaFePO4 apparently involves additional mechanisms, such as magnetic anisotropy.
In this work, we aim at understanding the magnetic properties of NaFePO4 by means of effective model Hamiltonians constructed from first-principles electronic structure calculations in the Wannier basis for the set of magnetically active target Fe 3d bands. This approach treats the Coulomb interaction problem more rigorously than brute force total-energy calculations. Moreover, the model can provide deep insight into the microscopic origin of the magnetic interactions in polyanionic compounds and help to analyze the results of brute force calculations. Furthermore, we present a comparative study considering two different numerical implementations of the Wannier functions technique; one based on the pseudopotential method and the other on the minimal basis set linear muffin-tin orbital method.
The paper is organized as follows. In Section 2, we outline the details of the electronic structure calculations, as well as the methods used to construct and solve model Hamiltonians. Then, in Section 3, we present our main results for interatomic exchange interactions and relative stability of different magnetic states. We also consider the effect of chemical pressure on the magnetic properties and magnetoelectric effect in triphlytes. Finally, the results are recapitulated and summarized in Section 4.
![]() | ||
Fig. 1 ab and ac projections of the crystal structure of t- (a and b) and m-NaFePO4 (c and d). Na, P, Fe and O atoms are shown with yellow, gray, brown and red spheres, respectively. |
The calculated GGA density of states (DOS) for t- and m-NaFePO4 is shown in Fig. 2. Typically, the character of insulating TM oxides changes along the Ti–Cu33 series. Early oxides of Ti–Cr are normally regarded as Mott insulators, while late oxides exhibit a strong charge transfer character because the TM 3d states strongly hybridize with the O 2p states and, therefore, O atoms also contribute to the low-energy properties. While the properties of Mott insulators can be described by a conventional Hubbard-type model, charge transfer insulators require further model refinements in order to include the O 2p states explicitly. Yet, orthophosphates are a notorious exception to this general situation because PO43− species have a clear molecular character which leads to additional band splittings into bonding states (mainly formed by the O 2p orbitals) and antibonding states (formed by the P 3p orbitals). As a consequence, the bonding O 2p states are additionally shifted to the low-energy region, which increases their separation from the Fe 3d states. Thus, the Fe 3d states form a narrow band located near the Fermi level, which is very well separated from both the O 2p bands (from below) and the P 3p bands (from above), as shown in Fig. 2. This makes the description of these materials particularly suitable for applying a Hubbard model. In the following, we will construct such a model by employing the Wannier functions method.34,35 Then, we will solve this model in the mean-field Hartree–Fock (HF) approximation and calculate all relevant parameters of interatomic exchange interactions. Since the degeneracy of the ground state of NaFePO4 is lifted by the lattice distortion, the ground-state wavefunction can be described by a single Slater determinant, which justifies the use of the HF approximation.
![]() | ||
Fig. 2 Total and partial PBE DOS for t- and m-NaFePO4. The shaded blue area shows contributions from the Fe 3d states. The Fermi level is at zero energy. |
![]() | (1) |
![]() | ||
Fig. 3 (a) PBE (turquoise solid lines) and (b) Wannier interpolated (blue dashed lines) band structures for the low-energy Fe 3d states as obtained for t- and m-NaFePO4, respectively. The corresponding real-space Wannier functions centered at Fe sites are shown for t-NaFePO4 (c) and m-NaFePO4 (d). Notations of the high symmetry points of the Brillouin zone are taken from ref. 36. |
Then, the transfer integrals and the crystal-field splitting, which are incorporated in [tijab], are evaluated as the matrix elements of the GGA Hamiltonian in the Wannier basis, as will be discussed below in detail. The screened Coulomb interactions are evaluated by using constrained random-phase approximation (cRPA).37 in eqn (1) are the matrix elements of spin–orbit (SO) coupling, which are also derived from the GGA Hamiltonian. In addition to the QE approach, we performed similar calculations by using the linear muffin-tin orbital (LMTO) method.38,39 The details of the model Hamiltonian construction in this case can be found in ref. 35.
In order to evaluate the one-electron part of the model Hamiltonian, = [tijab] in eqn (1), we considered different computational schemes.
Scheme 1 (s1): The one-electron part is evaluated ‘as is’ and identified with the matrix elements of the GGA Hamiltonian calculated in the Wannier basis. In this case, already takes into account non-sphericity of the Coulomb and exchange–correlation potentials. Therefore, in order to avoid double counting, we have to remove them when dealing with Coulomb and exchange interactions in the Hubbard model. For these purposes, we employed the spherical parametrization of the screened Coulomb interactions obtained from cRPA, Û = [Uabcd], with the following values for the on-site Coulomb repulsion U = F0, intra-atomic exchange interaction J = (F2 + F4)/14, and the ‘nonsphericity’ B = (9F2 − 5F4)/441 (F0, F2, and F4 being the screened radial Slater's integrals), which fully specify Û in the atomic spherical environment. Then, in the HF calculations, we set Uaabb = U for any a and b, and Uabba = J for a ≠ b. This guarantees that the Coulomb part of the HF potential is spherical and the exchange part, besides intra-atomic Hund's rule, is responsible for the appearance of a discontinuity term, which is proportional to U–J and missing in ordinary GGA. Other effects are included in
and described at GGA level. Thus, the idea is similar to DFT+U.40 The logic here is the following: the electron configuration d6 of the Fe2+ ions in NaFePO4 is essentially non-spherical. According to the fundamental Hohenberg–Kohn theorem,41 this non-sphericity should be properly described by DFT, and GGA, as one of its possible approximations, should capture this effect in a certain form. The discontinuity of the exchange potential in the HF method does not affect the total energies of isolated ions and leads only to rearrangement of single-particle levels between occupied and empty states.42 However, in solids, where these levels become connected by transfer integrals, the discontinuity already contributes to the total energy and other ground-state properties, as is clearly seen, for instance, in the theory of superexchange (SE) interactions.43
Scheme 2 (s2): We additionally subtract from the matrix elements of the GGA exchange–correlation potential in the Wannier basis, 〈wi|Vxc|wj〉. The basic idea behind is to fully replace the exchange–correlation interactions in GGA by the ones obtained for the Hubbard model. Thus, in the HF calculations, we use the full matrix Û = [Uabcd], obtained in cRPA without additional approximations for the exchange part of the HF potential. At the same time, non-sphericity of the Coulomb potential is treated at GGA level and already included in
. Therefore, for the Coulomb potential in the HF method, we keep using the spherical form of Û, which is described by a single parameter, Uaabb = U. Moreover, we consider two approximations. In the first approximation, we subtract from
only site-diagonal matrix elements of the exchange–correlation potential, 〈wi|Vxc|wi〉. The logic is that since the Hubbard model (1) includes the on-site interactions, only the on-site interactions should be subtracted from GGA, assuming that other (intersite) interactions are described reasonably well at the GGA level. In the second approximation, we subtract both site-diagonal and intersite matrix elements, 〈wi|Vxc|wj〉.
Scheme 3 (s3): This scheme is based on the LMTO method.35 The additional atomic-spheres approximation (ASA) in the LMTO method leads to some limitations for treating the electronic structure of NaFePO4. In our calculations, we tried to choose the atomic spheres in LMTO so as to reproduce the electronic structure of the more accurate QE calculations. Nevertheless, the good aspect of the LMTO method is that it provides another possibility for elimination of non-spherical on-site Coulomb and exchange–correlation interactions in GGA: in ASA, all these interactions are spherically averaged inside the atomic spheres. Therefore, we do not need to worry about double counting of this non-shpericity, so all non-sphericity arises from the screened Coulomb and exchange interactions in the Hubbard model. In ASA, one should only include the additional correction due to non-sphericity of intersite Madelung potential.35 As somewhat technical aspect, we had to use additional approximations in the process of cRPA calculations in the LMTO basis.35 Namely, instead of constructing proper Wannier functions, we used pseudo-atomic LMTO basis to calculate Û. Therefore, since the atomic orbitals are more localized than the Wannier functions, the effective Coulomb interaction is a bit larger compared to full-scale QE cRPA calculations.
The corresponding crystal-field splitting obtained from the diagonalization of the site-diagonal part of is displayed in Fig. 4. A key parameter here is the energy separation of the crystal-field orbital 1, populated by the minority-spin electron, from the rest of the orbitals. For instance, this parameter controls the strength of the spin–orbit coupling effects and typically competes with the crystal-field splitting. In LMTO (s3) and QE (s1, without any corrections), the splitting between the levels 1 and 2 is of the order of 100 meV. However, subtraction of 〈wi|Vxc|wj〉 in s2 additionally splits off the orbital 1. The shape of the orbital 1 is also important. Its effect will be considered latter, separately for m- and t-NaFePO4.
Having constructed and solved the effective model (1) by means of the mean-field HF approximation, one can calculate the one-electron (retarded) Green functions, and evaluate parameters of isotropic exchange interactions by using the local force theorem and considering perturbation theory to second order in infinitesimal spin rotations.44 The corresponding expression is given by:
![]() | (2) |
![]() | (3) |
Fig. 5 shows the parameters of the effective Coulomb and exchange interactions obtained in cRPA. As expected, the effective Coulomb repulsion calculated with the LMTO method is larger than the one obtained with QE. For instance, the parametrization in terms of U, J, and B (see Section 2) yields U = 3.7 and 2.9 eV for LMTO and QE, respectively. This is understandable, because the effective U in LMTO is evaluated within the basis of pseudo-atomic 3d orbitals, which are more localized than the proper Wannier orbitals used in the QE method. In all other respects, the behavior of matrix elements of Û is rather similar between the two methods, and the parameters J and B are close for both QE and LMTO methods: J = 0.9 eV and B = 0.1.
![]() | ||
Fig. 5 Matrix elements of Coulomb (a and c) and exchange (b and d) interactions for m-NaFePO4 as obtained in the framework of cRPA in QE (a and b) and LMTO (c and d) calculations. |
The behavior of transfer integrals can be illustrated by considering SE interactions, which are directly related to ij. The corresponding expression is given by:47
![]() | (4) |
![]() | ||
Fig. 7 (a) Distance dependence of SE interactions as obtained in LMTO (s3) and QE calculations with (s2) and without (s1) subtraction of site-diagonal matrix elements of the GGA exchange–correlation potential. (b) Results obtained in the QE calculations after subtraction of both site-diagonal and off-diagonal matrix elements of the GGA exchange–correlation potential (s2). The FM and AFM contributions are shown by open red and cyan symbols, respectively, and the total contributions are shown by filled symbols. The symbols denote the type of interactions (see Fig. 6 for the notation of atomic positions). |
Method | J 12 | J 33′ | J 13 | J 34′ | J 14 | J 14′ | θ |
---|---|---|---|---|---|---|---|
s1, SE | 0.02 | −0.05 | −0.39 | −1.30 | −1.61 | −0.12 | −80.4 |
s1, LF | −1.20 | −0.08 | −0.38 | −1.27 | −1.55 | −0.13 | −91.1 |
s2, SE | −3.19 | −0.11 | −0.11 | −0.07 | −1.32 | −0.13 | −76.5 |
s2, LF | −2.19 | −0.12 | −0.30 | −0.72 | −1.54 | −0.17 | −87.1 |
s3, SE | −3.54 | −0.07 | −0.17 | −0.66 | −0.95 | −0.12 | −86.5 |
s3, LF | −3.07 | −0.06 | −0.18 | −0.67 | −0.97 | −0.12 | −87.1 |
Finally, we note that the superexchange interactions obtained with the s2 scheme, after subtraction of both site-diagonal and off-diagonal elements of the GGA exchange–correlation potential, are strongly overestimated (Fig. 7b): the off-diagonal elements 〈wi|Vxc|wj〉 appear to be large and change the values of the transfer integrals. Thus, the construction of the model parameters in the scheme s2 is clearly unphysical: if one tries to correct GGA by adding only on-site electron–electron interactions, it is reasonable to subtract only the local (or site-diagonal) part of 〈wi|Vxc|wj〉.
Using the obtained parameters of exchange interactions, one can easily evaluate the Curie–Weiss temperature as . Depending on the computational scheme (s1–s3) and method to calculate the exchange parameters (SE versus the local force theorem), θ varies from −77 to −91 K (Table 1), being in fair agreement with the experimental value of −83 K reported by Avdeev et al.26
The next important questions is how consistent the obtained parameters are with the experimental magnetic structure of m-NaFePO4. Since the m-phase has the same symmetry as perovskite TM oxides, we can follow the same strategy as for the perovskites and, as the first step, compare the energies of FM, A-type AFM (the FM ac layers are antiferromagnetically coupled along b), C-type AFM (FM chains propagating along b and antiferromagnetically coupled in the ac plane), and G-type AFM (AFM coupling between nearest neighbors in all three directions) states. We have found that the A and C states are close in energy, and considerably lower than the FM and G states. The energy difference between C and A states varies from 8 meV per NaFePO4 in the s3 scheme till 6 meV per NaFePO4 in the s1 scheme. The A state is stabilized due to the AFM coupling J12. The C state arises due to a joint effect of several long-range interactions (e.g., J13 and J14 in Table 1) that compete with J12 and tend to make the coupling in the b chains ferromagnetic. The formation of the FM chains in the second case is consistent with the experimental data.26 However, NPD measurements reveal a more complex magnetic structure, which is described by the propagation vector q = (1/2, 0, 1/2) and corresponds to the AFM coupling between chains separated by the translations along a and c.26 In order to explore such possibility, we investigate the stability of the A and C states with respect to the incommensurate spin-wave excitations. Namely, we search for the eigenvalues ωα of the 4 × 4 matrix , where the dimensionality corresponds to the number of magnetic sublattices in the unit cell,
ij = ±Jij for the ferromagnetically (+) and antiferromagnetically (−) coupled bonds,
′(q) is the Fourier image of
ij between sublattices
and
′, and
. If all ωα(q) are positive, the magnetic state is stable. If some of them are negative, the state is unstable for given q. The results are shown in Fig. 8, using parameters obtained in the SE theory for the s1 scheme (other schemes provide qualitatively a similar picture). One can see that for the calculated value of U the A state is stable (all ω's are positive), while for the C phase some of the ω's are negative, even for q = 0. Thus, the magnetic ground state is expected to be of the A type. This is clearly inconsistent with the experimental data. To some extent, the problem can be resolved by decreasing the value of U, which increases the weight of the FM contributions [mainly the first term in eqn (4)] to J12 and eventually makes this coupling ferromagnetic. For instance, the SE interaction JSE12 in the s1 scheme gradually increases from 0.01 meV to 4.58 meV when the Coulomb repulsion U changes from 2.94 eV to 1.47 eV. In reality, the underestimation of the FM contributions to the exchange coupling may be related to the lack of direct interatomic interactions in the considered Hubbard model, eqn (1),48 which is mimicked by the change of the value of U in the expression for the SE interactions eqn (4). Anyway, for U* ≈ 0.5U (U being the calculated value) the A state becomes unstable for q = 0. Instead, the C state is stabilized. Furthermore, the minimal ωα(q*) is obtained for q* ≈ (0, 0, 0.36), which is close to the Z-point in Fig. 8. Thus, one could expect the formation of the incommensurate magnetic ground state. For U* ≈ 0.5U, both A and C states are also unstable for the experimental q = (1/2, 0, 1/2) (the U-point in Fig. 8). However, according to the spin-wave analysis, the energy of this state is expected to be higher than that of the incommensurate state corresponding to q*. It is true that the energy difference between q* and U points of the Brillouin zone is small (about 1 meV). However, it is also true that the spin-wave dispersion along the Z–U direction is practically flat, meaning that there are no magnetic interactions along the a axis, which would stabilize the experimental q = (1/2, 0, 1/2) superstructure. This seems to be reasonable taking into account that a is the largest, while c is the shortest lattice parameter in m-NaFePO4 (a/b = 1.310 and a/c = 1.782). Therefore, it is fair to expect the existence of strong magnetic interactions which would lead to the formation of magnetic superstructures along the c axis, but not along the a axis. Finally, we note that if the magnetic ground state of m-NaFePO4 was indeed incommensurate with q* ≈ (0, 0, 0.36), this phase would be multiferroic, similar to the perovskite manganites with the same crystal and magnetic structures.28–30 It would be interesting to explore this possibility experimentally.
![]() | ||
Fig. 8 Spin-wave dispersion in the A- (a and c) and C-type (b and d) AFM state of m-NaFePO4. (a and b) Obtained by using parameters of the SE interactions and scaling the value of the Coulomb repulsion from U = 2.94 eV to 0.5U = 1.47 eV. (c) and (d) were obtained by using parameters reported in Table 2 of ref. 31. The notations of the high symmetry points of the Brillouin zone are taken from ref. 36. |
Regarding this apparent disagreement with the NPD data,26 we would like to note that the parameters of interatomic exchange interactions for m-NaFePO4 were also derived in ref. 31 by mapping the total energies, obtained for different magnetic structures in the DFT+U approach, onto the spin model of eqn (3). This procedure is different from the one used in this work. On the one hand, it takes into account the contributions of the oxygen band, which may be important. On the other hand, it suffers from ambiguities related to the choice of parameters and details of implementation of the DFT+U approach. Nevertheless, the main conclusion is similar to ours: if we use the exchange parameters reported in ref. 31 and evaluate ωα, we readily obtain that the magnetic ground state is also A and there is no sign of instability towards a magnetic superstructure with q = (1/2, 0, 1/2). The relativistic SO interaction can lead to the canting of spins, similar to other perovskite TM oxides, crystallizing in the centrosymmetric Pnma structure.35,45 However, the SO interaction alone cannot lead to the formation of superstructures. Thus, the microscopic origin of the q = (1/2, 0, 1/2) order, reported experimentally, is still puzzling.
Without SO interaction, there are four possible collinear arrangements of the Fe spins in the primitive cell: ↑↑↑↑ (F), ↑↓↓↑ (A1), ↑↓↑↓ (A2), and ↑↑↓↓ (A3), where the arrows indicate the directions of spins at sites 1–4 (see Fig. 1). Amongst them, F and A3 contain spacial inversion Î as it is, while in A1 and A2 Î is combined with time reversal . Thus, the latter two magnetic phases allow for magnetoelectric effect, which will be discussed in more detail in Section 3.4. The magnetic ground state is A1, in agreement with experiments.26 In the HF approximation, the energy difference between F, A2, and A3 (on the one hand) and A1 (on the other hand) is 16.5 (26.0), 12.6 (22.7), and 2.1 (2.0) meV per NaFePO4 in s1 (s3), respectively.
The interatomic exchange interactions obtained by using different techniques are summarized in Table 2. Like in m-NaFePO4, we find a consistent description for Jij, even despite a substantial change of the orbital ordering between the s1 and s3 schemes (Fig. 9).
Method | J 13 | J 11′ | J 12 | J 1′4 | J 14 | J 1′1′′ | θ | T N |
---|---|---|---|---|---|---|---|---|
s1, SE | −3.62 | 0.14 | −1.66 | −0.57 | 0.09 | −2.16 | −133 | 34 |
s1, LF | −3.39 | 0.07 | −1.49 | −0.48 | 0.05 | −1.93 | −124 | 34 |
s3, SE | −4.20 | −0.14 | −0.94 | 0.03 | −0.14 | −1.32 | −127 | 50 |
s3, LF | −5.92 | −0.21 | −1.30 | −0.08 | −0.22 | −1.73 | −177 | 68 |
The nonvanishing interactions spread up to the 6th coordination sphere around each Fe site. The ↑↓↓↑ AFM order is stabilized mainly by J13 and J12. Moreover, we note the existence of the relatively strong AFM interaction J1′1′′, which tends to form a long-periodic magnetic structure along b.
Thus, our theoretical calculations correctly reproduce the ↑↓↓↑ magnetic ground state of t-NaFePO4. The theoretical Néel temperature, obtained by using Tyablikov's random-phase approximation (RPA) and parameters listed in Table 2 is about 34–68 K, which is also in fair agreement with the experimental value of 50 K reported in ref. 26. However, the theoretical Curie–Weiss temperature seems to be systematically overestimated by a factor of 1.5–2.
Compound | J 13 | J 11′ | J 12 | J 1′4 | J 14 | J 1′1′′ | θ | T N |
---|---|---|---|---|---|---|---|---|
NaFePO4 | −4.20 | −0.14 | −0.94 | 0.03 | −0.14 | −1.32 | −127 | 50 |
LiFePO4 | −7.83 | −0.29 | −2.62 | −0.59 | −0.37 | −2.65 | −261 | 93 |
LiMnPO4 | −4.41 | −0.19 | −1.06 | −0.14 | −0.16 | −0.64 | −121 | 55 |
Pij = εji![]() | (5) |
The symmetry properties of ij within two coordination spheres around one of the Mn sites is shown in Fig. 10. Then, according to the HF calculations with SO coupling, the easy magnetization direction in LiMnPO4 is a (= x). The canting of spins away from the a axis caused by the DM interactions is small. Indeed, the strongest isotropic exchange interaction is J13 = −3.94 meV, while the DM vector for the same bond is dij = (0.03, −0.02, 0.05) meV, i.e. about two orders of magnitude smaller. The corresponding Néel and Curie–Weiss temperatures can be estimated as 55 K and −121 K, respectively (Table 3). They are in fair agreement with the experimental data, TN = 33–45 K10,57 and θ = −87 K.57 The experimental parameters derived from inelastic neutron-scattering measurements were reported in ref. 9. The main interactions [when multiplied by −2S2 according to our definition of the spin model, eqn (3)] are J13 = −6.00 meV, J11′ = −0.95 meV, J12 = −0.45 meV, (J1′4 + J14)/2 = −0.78 meV and J1′1′′ = −2.5 meV. Again, we find reasonable agreement with our theoretical results (Table 3). Strictly speaking, instead of the ME ↑↓↓↑ state the experimental magnetic structure of LiMnPO4 reported in ref. 9 is ↑↑↓↓, which gives rise to weak ferromagnetism.57,58 However, according to our HF calculations, these two states are nearly degenerate with the energy difference of about 1 meV per LiMnPO4.
![]() | ||
Fig. 10 Fragment of the crystal structure of LiMnPO4 illustrating symmetry properties of the pseudovectors ![]() |
Next, we consider deformations of a nearly collinear ↑↓↓↑ spin structure by applying a magnetic field along the b (= y) or c (= z) axis (see Fig. 11). In the first case (H‖b), the finite component of [ei × ej] is c. Then, by combining the phases of zij and εji in all bonds (see Fig. 10), it is easy to see that the total polarization at each Mn site should also be parallel to b (see Fig. 11). Using the model parameters, the change in the polarization
induced by the FM canting of spins along y can be estimated as pyi = 1.08 μC m−2, which is the same for all Mn atoms. The direct HF calculations for our Hubbard model [eqn (1)] with an applied magnetic field yield pyi = 0.61 μC m−2. Then, the matrix element of the ME tensor αyy can be found as αyy = pyi∂ey/∂Hy, where ∂ey/∂Hy is estimated by using parameters of the considered spin model [eqn (3)]. This yields αyy ≈ 0.02 ps m−1.
In the case of H‖c, the finite component of [ei × ej] is b. Then, by combining yij and εji, one can find that pi = (±0.24, 0, 0.25) μC m−2. Thus, after summation over the unit cell the x component is canceled out (see Fig. 11) and the total polarization is parallel to z. The direct HF calculations with an applied magnetic field give a close value of pzi = 0.33 μC m−2. Then, the corresponding component of the ME tensor can be estimated as αzz ≈ 0.01 ps m−1.
Regarding the type of AFM ordering and theoretical values of Néel and Curie–Weiss temperatures, we have obtained very good agreement with the experimental data for t-NaFePO4. In this case, the experimental ↑↓↓↑ structure is mainly stabilized by the AFM interactions in the Fe–Fe bonds 1–2 and 1–3 (and equivalent to them bonds). Nevertheless, like in the previous study,31 we could not reproduce the experimental q = (1/2, 0, 1/2) superstructure in m-NaFePO4. The experimental frustration index |θ|/TN in m-NaFePO4 is significantly larger than in t-NaFePO4 (about 6.4 and 1.7, respectively), suggesting that interatomic magnetic interactions in m-NaFePO4 should be more complex and involve competing AFM interactions connecting Fe sites beyond nearest neighbours. However, the theoretical interactions around each Fe site are limited by practically the same number of coordination spheres in m- and t-NaFePO4. Apparently, some long-range AFM interactions are missing in the electronic structure calculations for m-NaFePO4. The reason is not quite clear: while the magnetic superstructure along the shortest orthorhombic c axis can be understood by adjusting the value of the on-site Coulomb repulsion U, the one formed along the longest a axis is puzzling. In any case, smaller value of U makes the theoretical A-type AFM ground state unstable. This could be the key point for understanding the origin of the experimental q = (1/2, 0, 1/2) order. If the Coulomb repulsion is small, the conventional SE theory breaks down and one has to consider additional interactions, which appear in higher orders with respect to ij/U. These interactions operate via intermediate transition-metal sites and are typically referred to as the super-SE interactions. Particularly, they play a very important role in multiferroic manganites and are believed to be responsible for the formation of long-periodic magnetic structures.59 At present, it is not clear whether the super-SE interactions can be relevant to the magnetic properties of m-NaFePO4. In order to explore such possibility, it is vital to understand: (i) the physical mechanism, which would reduce the value of the effective Coulomb repulsion U in m-NaFePO4, and (ii) the origin of the orbital ordering, which would assist the super-SE mechanism in stabilizing the AFM coupling along the a axis. The combination of these two factors, the relatively small value of U and the particular orbital ordering, was found to be crucial for the behavior of long-range magnetic interactions in manganites.59 We hope that our analysis will stimulate further theoretical and experimental studies to clarify the origin of the q = (1/2, 0, 1/2) order in m-NaFePO4.
Finally, we have investigated the effect of chemical pressure and explained the properties of magnetoelectric effect in the triphlyte phase of TM phosphates.
This journal is © the Owner Societies 2018 |