The role of exchange–correlation functional on the description of multiferroic properties using density functional theory: the ATiO3 (A = Mn, Fe, Ni) case study

Renan Augusto Pontes Ribeiroa, Sergio Ricardo de Lazaro*a and Carlo Gattib
aDepartment of Chemistry, State University of Ponta Grossa, Av. Gen. Carlos Cavalcanti, 4748, 84030-900, Ponta Grossa, Brazil. E-mail: srlazaro@uepg.br; Fax: +55 4242203042
bCNR-ISTM Istituto di Scienze e Tecnologie Molecolari, Via Golgi 18, 20133 Milano, Italy

Received 26th August 2016 , Accepted 18th October 2016

First published on 19th October 2016


Abstract

In this study, ab initio density functional theory calculations were performed on ATiO3 (A = Mn, Fe, Ni) materials for multiferroic applications. Structural, magnetic, electronic and topological analysis were investigated as regard the A-site cation effect. Concerning the prediction of new multiferroic candidates in silico, the role of different exchange-correlation functionals was reviewed. From such properties was verified that PBE0 functional performs the best for ATiO3 compounds overcoming the deficiency of standard LDA and PBESol functionals and get the better of B3LYP, PBE0+D and B3LYP+D. Regarding the A-site cation effect, calculated structural parameters follow the ionic radii trend moving from Mn to Ni, resulting in more ordered AO6 clusters; while, the TiO6 distortion increases. Using magneto-structural and magneto-electronic parameters, the antiferromagnetic exchange coupling constant was rationalized from GKA rules. The increase of the antiferromagnetic order along the transition metal series was explained from the 3d orbital occupancy and A–O bond interactions. DOS and band structure profiles clarify the semiconductor behavior as regard the 3d crystal field splitting for both A-site and Ti cations. Further, electronic results suggest the existence of intermetallic connection (A–O–Ti–O–A); whereas, topological analysis in the framework of quantum theory of atom in molecules was carried out to evaluate the A–O and Ti–O bond interaction through the analyses of ∇2ρ sign at bond critical point and of the charge-transfer parameter obtained from net charges.


1 Introduction

Recently, the multiferroics materials have assumed a remarkable importance as research topic on the design of new technological devices. The most complete definition for this class involves two or more forms of ferroic orders (i.e. ferroelectricity, ferroelasticity, ferromagnetism, ferrotoroidicity) combined into a single crystalline structure. However, the main studies in this field are strictly related to candidates that combine a magnetic order (ferromagnetism, antiferromagnetism) with ferroelectricity, being denominated magnetoelectric multiferroics.1–3 The renewed interest on multiferroic materials was motivated from theoretical explanations about the physical reasons behind the scarcity of such compounds, as well to clarify the unusual phenomena and successfully predict new candidates.4–8

The theoretical understanding of the multiferroic materials depends of the precise description of structural, electronic and magnetic properties. In the last decades, the first-principles calculations, mainly Density Functional Theory (DFT), have played a fundamental role on solid-state investigations which notice the reliable description for different kind of properties.5,9–14 However, the multiferroic materials class corresponds to a challenging topic on first-principles calculations due to the strong correlation, which results from the unpaired electrons. In such materials, the partial filling of d or f orbitals induces a high localization summed to the strong interaction causing failures on the electronic structure representation, mainly for standard LDA or GGA exchange–correlation functionals.15 Such incoherency is associated with the Self-Interaction Error (SIE) that provides a spurious interaction of an electron with itself resulting in underestimated band gaps, erroneous localization for d–f orbitals, and the tendency toward magnetism.16–19 Countless approaches to correct the SIE limitation were proposed, being the GW approximation, self-interaction correction (SIC) and the DFT plus Hubbard potential (U) the most widely reported. In GW and SIC methods, the SIE is corrected through a formal procedure to calculate the self-energy, which is included in the splitting of occupied and unoccupied bands.20,21 In the other hand, DFT+U introduce a local correction to DFT based on a mean-field (HF) treatment of the Hubbard model characterized by two parameters, the on-site coloumb potential (U) and the exchange interaction (J) providing a correct interpretation for Jahn–Teller distortions and exchange interactions.15,22–25 Despite the notorious picture for the electronic structure of some transition metal oxides,15,25 the performance of such method is related to the system-dependent U and J parameters making the choice often arbitrary; besides, the wrong solutions for under-pressure systems or phase-transitions investigations.26

Hybrid functionals are an alternative approach to significantly improve the accuracy behind local (LDA) and semi-local (GGA) methods, corresponding to a mixing between the exact nonlocal Fock exchange and LDA/GGA exchange–correlation.27 Regarding the theoretical investigation of multiferroic materials, the hybrid Hamiltonian overcome the LDA and GGA approaches through a significant improvement of the band-gap description for semiconductors, as well through the correct localization of unpaired electrons in open-shell systems.15,28–32 Recently, these functionals have been widely applied to different multiferroic materials. Goffinet and co-workers report a systematic comparison between different exchange–correlation formalism to predict the structural, magnetic, electronic and dynamical properties of BiFeO3, showing the powerful prediction of hybrid functionals for multiferroic candidates.33 In addition, Stroppa and Picozzi confirmed this argument through a hybrid DFT/HF investigation of proper (BiFeO3) and improper (HoMnO3) multiferroic materials.34 Further, McDonnell and co-authors showed the agreement between hybrid DFT and experimental findings for optical properties.35 Subsequent studies in this field reveal that hybrid functionals calculations, mainly based on HSE formalism, are useful to predict new multiferroic candidates such as PbNiO3,36 YMnO3,37 BaMnO3,38 PbVO3,39 Bi3FeCrO3,40 among others.41,42 On the other hand, Heifets and collaborators investigated the thermodynamic stability of BiFeO3 by means of B3PW hybrid functional and noted a underestimated formation energy at 0 K suggesting that the further investigation about the performance of available hybrid functionals is needed.43

From this point of view, the present work proposes a systematic theoretical investigation comparing local (LDA), semilocal (GGA-PBESol), pure and energy dispersion augmented hybrid functionals (PBE0, PBE0+D, B3LYP and B3LYP+D) to evaluate the structural, magnetic and electronic properties of ATiO3 (A = Mn, Fe, Ni) multiferroic materials with LiNbO3-type structure. In the past few years, the scientific interest on such materials has increased due to the possibility of controlling the magnetism from antisymmetric Dzyaloshinskii–Moriya interaction. This interest was motivated from a theoretical structural–chemical investigation proposed by Fennie, where symmetry arguments and DFT+U calculations were combined to discuss the strong polarization-magnetism coupling in ATiO3 (A = Mn, Fe, Ni) ilmenite based materials when stabilized in R3c phase.44 Such predictions were validated by experimental reports about the synthesis of the high-pressure form of FeTiO3,45 MnTiO3,46,47 and NiTiO3 candidates.48 Despite the current theoretical studies developed for these materials,49–51 both exchange–correlation and A-site cation modification effect on structural, magnetic and electronic properties remains unclear.

2 Computational methodology

ATiO3 (A = Mn, Fe, Ni) materials crystallize in the acentric LiNbO3-type structure (R3c) under high pressure and temperature conditions through a cation reordering phase transition from ilmenite polymorph.48,52,53 The rhombohedral primitive unit cell for this polymorph contains two ATiO3 unit formulas (10-atoms) with A (A = Mn, Fe, Ni), Ti and O atoms located at (0; 0; u), (0; 0; 0) and (x, y, z), respectively, as showed in Fig. 1. Theoretical models proposed in this study are based on the experimental structural parameters obtained by X-ray measurements.48,53,54 The A, Ti, and O atoms were described by atom centered all-electron gaussian basis set such as 86-411d41G, 86-411(d31)G and 8-411G, respectively.55–57
image file: c6ra21465g-f1.tif
Fig. 1 Primitive (a) and conventional (b) crystallographic unit cell (R3c) for ATiO3 (A = Mn, Fe, Ni) materials. The black, blue, and red balls represent A, Ti, and O atoms, respectively. The blue and black polyhedral represent the [TiO6] and [AO6] clusters, respectively.

The ATiO3 (A = Mn, Fe, Ni) structures were fully-relaxed (lattice parameters and atomic positions) in the framework of Density Functional Theory (DFT) employing local (LDA),58 semilocal (PBESol),59 and hybrid (B3LYP,60,61 PBE0 (ref. 62)) exchange–correlation functionals. A general drawback of all local, semilocal or hybrid exchange–correlation functionals is the inability to describe long-range electron correlations (dispersive interactions) as van der Waals forces. These interactions correspond to a balance between electrostatic and exchange-repulsion forces, playing a fundamental role on the theoretical description of ferroelectric materials, such as ATiO3 (A = Mn, Fe, Ni) in LiNbO3-type structure. In this study, we include the dispersive effects on hybrid B3LYP and PBE0 exchange-correlation functionals through a London-type pairwise empirical correction proposed by Grimme.63 In this approach, the total energy (eqn (1)) and the empirical correction dispersion (eqn (2)) were described by the following equations:

 
EXC+D = EXC + Edisp (1)
 
image file: c6ra21465g-t1.tif(2)

Here, the summation involves all atom pairs and lattice vectors (g) excluding the i = j contribution for g = 0; s6 is a scaling factor that depends only of exchange–correlation functional used (1.05 for B3LYP, 0.6 for PBE0), c6ij is the dispersion coefficient for the atom pair ij, Rij,g is the interatomic distance between atoms i in the reference cell and j in the neighboring cells at distance |g| and fdmp is the damping function used to avoid near-singularities for small interatomic distances consisting of summed atomic van der Waals radii (i.e. Rvdw = Rvdwi + Rvdwj) and the steepness set to 20.63,64

All calculations were carried out using the periodic ab initio code CRYSTAL09.65 The convergence criteria for mono and bielectronic integrals were both set to 10−8 Hartree, while the RMS gradient, RMS displacement, maximum gradient and maximum displacement were set to 3 × 10−5, 1.2 × 10−4, 4.5 × 10−5 and 1.8 × 10−4 a.u., respectively. Regarding the density matrix diagonalization, the reciprocal space net was described by a shrinking factor set to 8, corresponding to 65 k points in accordance with the Monkhorst–Pack method.66 The accuracy in evaluating the Coulomb and exchange series was controlled by five thresholds, for which adopted values are 10−8, 10−8, 10−8, 10−8 and 10−16.

Experimental reports suggest a G-type antiferromagnetic ground-state for all ATiO3 (A = Mn, Fe, Ni) materials consisting of spins ferromagnetically ordered within (111) planes, while the adjacent planes are antiferromagnetically coupled.45,46,48 To determine the performance of the exchange–correlation functional in the magnetic ground-state description for strongly correlated materials, we consider two collinear magnetic configurations using the primitive (10-atoms) unit cell: (i) ferromagnetic (FEM), where the spins for all neighbors are parallel ordered; (ii) antiferromagnetic (AFM) for which the spins on the nearest neighbors are antiparallel ordered to each other. The exchange coupling constant was computed from the energy difference between the magnetic configurations using the Ising model:

 
Ĥising = ∑JijŜizŜjz (3)

Here Jij is the magnetic exchange coupling constant between A2+ neighbors, while Ŝz correspond to the operator for the z-component of spin momentum.

In order to understand the electronic structure of ATiO3 (A = Mn, Fe, Ni) multiferroic materials, theoretical results for density of states (DOS), band structure profiles and topological analysis were investigated from the optimized wavefunction. Topological analysis of the electron density ρ(r) and atom basin properties were obtained with TOPOND98 package67 within the framework of the quantum theory of atom in molecules (QTAIM) developed by Bader.68 This code provides a comprehensive evaluation of electron density (ρ) and of its Laplacian, as well as atomic basin properties and 2D–3D plots of periodic systems. In this study, we focus on the determination of so-called bond critical points (BCP), which are powerful tools to classify the chemical bonds. The critical points (CP) in the electron density and its Laplacian are the points where of the ∇ρ(r) vanishes. The CP can be classified in terms of their type (r, s), where r correspond to the number of non-zero eigenvalues (λ1–3) of the Hessian matrix, while s describe the difference between positive and negative eigenvalues. The BCP correspond to (3, −1) in terms of the (r, s) notation, indicating two negative curvatures associated with the interatomic surface orthogonal to the BCP, and one positive curvature related to the bond path at the BCP. Further, another section of TOPOND deals with atomic basin boundaries enabling the evaluation of several atomic properties such as Bader's net charge, atomic volume and others. A more detailed description of TOPOND98, QTAIM and topological analysis can be found in the ref. 69–73 and the references there in.

3 Results and discussions

3.1 Structural properties

In order to assess the validity of different exchange–correlation functionals to compute structural properties, the lattice parameters and unit cell volume obtained from DFT benchmark for ATiO3 (A = Mn, Fe, Ni) materials were investigated, as summarized in Table S1. Despite the fact that non-local PBESol functional slightly overcomes local (LDA) formalism for all ATiO3 material, it was observed that standard exchange–correlation functional underestimate the experimental results, while pure and dispersion augmented hybrid formalism are the closest. Surprisingly, B3LYP exhibits the worst performance among the hybrid functionals, resulting in overestimated structural parameters for all ATiO3 materials. Similar results were reported for other ferroelectric titanates, where Bilc and co-authors argue that the origin of B3LYP failures, mainly for lattice parameters and unit cell volume, is the Becke's GGA exchange part.31 On the other hand, PBE0 performs the best among the investigated exchange–correlation functionals. Further, such results are closer to the experimental measurements even when compared to the literature DFT+U results for ATiO3 (A = Mn, Fe, Ni) materials.44,49,50

From the results reported in Table 1, it can be seen that the inclusion of dispersive effects on hybrid B3LYP and PBE0 functionals induces a large unit cell contraction mainly for the c lattice parameter, resulting in a worse performance than non-augmented hybrid formalism. This error can be attributed to the nature of empirical Rvdw and c6ij parameters that does not take in account the chemical environment resulting from the crystalline ordering in solid state materials, where the coordination number controls the van der Waals radii. One way to overcome this problem consists of a empirical search of system-dependent Rvdw and c6ij parameters as function of desired properties, as performed by Albuquerque for TiO2 anatase.74

Table 1 Theoretical results for local magnetic moment (SμB), superexchange coupling constant (JA–A – in K) and A–O–A bond angle (Θ) of the ATiO3 (A = Mn, Fe, Ni) materials as function of exchange–correlation functionals
  MnTiO3 FeTiO3 NiTiO3
S J ΘMn–O–Mn S J ΘFe–O–Fe S J ΘNi–O–Ni
B3LYP 4.755 −8.74 118.44 3.762 −13.87 122.35 1.719 −9.30 126.56
B3LYP+D 4.748 −6.62 115.80 3.754 −10.45 120.90 1.716 −5.13 126.13
PBE0 4.788 −6.89 117.19 3.791 −11.34 121.82 1.751 −10.06 126.49
PBE0+D 4.785 −5.90 115.72 3.787 −10.08 121.05 1.749 −6.99 126.26
PBESol 4.597 −28.62 119.27 3.672 −7.83 125.82 1.528 21.30 128.21
LDA 4.524 −43.53 120.09 3.607 −12.49 126.43 1.480 6.16 128.85


Let us now briefly discuss the performance of different exchange–correlation functionals on the main bond distances in R3c multiferroic titanates. The LiNbO3-type polymorph of ATiO3 (A = Mn, Fe, Ni) can be described by two sub-units centered on A- and Ti-site cations, as presented in Fig. 1. Unlike the common ABO3 materials, both sub-units in R3c structure are centered on six-fold coordinated metals (AO6 and TiO6) linked up to oxygen atoms by short (3) and long (3) paths, enabling an A–O–Ti–O–A intermetallic connection via face and edge-sharing. Further, both AO6 and TiO6 clusters share the corners with other octahedrals through a A–O–A or Ti–O–Ti path. To balance the electrostatic repulsion between the metals, both A- and Ti cations are displaced from the octahedral central position, resulting in structural distortion that controls the ferroelectricity. The Table S2 summarizes the M–O (M = Mn, Fe, Ni, Ti) bond distances and octahedral distortion (Δ = 1/6Σ[(ddav)/dav]2) obtained from different exchange–correlation functionals.

Bond lengths calculated for both AO6 and TiO6 clusters show the general behavior observed for the lattice parameters and unit cell volume: (i) local (LDA) and semilocal (PBESol) functionals underestimates the experimental values; (ii) whereas the pure and dispersion augmented hybrid functionals exhibit a reasonable agreement. However, both B3LYP+D and PBE0+D functionals perform slightly worse than non-augmented hybrid functionals clarifying the dependence of Rvdw and c6ij parameters as function of chemical environment. Despite the same coordination number for A and Ti cations, the experimental results shown that both clusters exhibit different distortions, which are originated from distinct mechanisms. Inasmuch as the values proposed by Grimme63 for the 3d transition metals are equal in DFT-D2 method, the calculated structural properties show an incorrect picture as regard the crystalline arrangement. For instance, the octahedral distortion calculated with B3LYP+D and PBE0+D functionals suggests a reduced role of TiO6 clusters in the structural deformation, while overestimates the A-site contribution.

The latter investigation in this section involves the A-site cation effect on the structural properties of ATiO3 (A = Mn, Fe, Ni) materials in LiNbO3-type structure. For this purpose, the lattice parameters, unit cell volume and bond lengths obtained with hybrid PBE0 exchange–correlation functional were selected. Theoretical results presented in Table S1 reveal that the unit cell volume of ATiO3 materials decrease in order MnTiO3 > FeTiO3 > NiTiO3 following the A2+ ionic radii series. Such contraction is accompanied by a shortening of the A–O bond distances resulting in more ordered AO6 clusters. This tendency can be explained by the 3d orbital occupancy moving from the Mn to Ni, where the progressive filling of t2g levels induces an orbital contraction summed to the large crystal field splitting for high-spin configurations in order Mn2+ (t32g − e2g) < Fe2+ (t42g − e2g) < Ni2+ (t62g − e2g). Such results are similar to the general trend observed for Radtke and co-authors as regard the ATiO3 (A = Mn, Fe, Ni) materials in the ilmenite polymorph.75 On the other hand, the distortion on TiO6 clusters increases moving from Mn to Ni suggesting the fundamental role of second-order Jahn–Teller distortion (SOJT) on the stabilization of R3c polar structure for ATiO3 (A = Mn, Fe, Ni) materials in agreement with the polarization results proposed by Fennie.44

3.2 Magnetic properties

As previously discussed, the magnetic ordering of ATiO3 (A = Mn, Fe, Ni) materials in LiNbO3-type structure can be described by a G-type arrangement consisting of alternated (111) planes antiferromagneticly ordered. In such geometry each magnetic A2+ cation is surrounded by six neighbors through a A–O–A superexchange coupling constant, which is controlled by the 3d orbital occupancy and A–O–A bond angle following the Goodenough–Kanamori–Anderson (GKA) rules, as presented in Fig. 2. In order to investigate the performance of different exchange–correlation functionals on the description of magnetic ground-state for ATiO3 (A = Mn, Fe, Ni) materials, the exchange coupling constant along the A–O–A path (JA–A) was computed from the energy differences between FEM and AFM configurations without spin–orbit contribution, as summarized in Table 1.
image file: c6ra21465g-f2.tif
Fig. 2 Hexagonal unit cell (R3c) for ATiO3 (A = Mn, Fe, Ni) materials and G-type antiferromagnetic spin ordering representation (orange vectors). The black polyhedra represent seven (AO6) clusters denominated as (AO6)7 magnetic complex with A–O–A superexchange coupling constant JA–A.

Regarding the magnetic ground state all exchange–correlation functionals successfully predict the antiferromagnetic exchange coupling constant for ATiO3 (A = Mn, Fe, Ni) materials, except for NiTiO3 where LDA and PBESol predict a ferromagnetic ordering in disagreement with the experimental predictions. Further, it was observed that LDA and PBESol provide smallest values for the local magnetic moment (S) comparing with hybrid formalism. If we apply a fully ionic picture, the A-site metal is a divalent cation resulting in Mn2+ (3d5), Fe2+ (3d6) and Ni2+ (3d8) valence orbital configuration, which would result in a magnetic moment of 5, 4 and 2 μB, respectively. Table 1 shows that pure and dispersion augmented hybrid functionals results exhibit a better agreement with expected values than local and semilocal formalism, which can be explained by the spurious self-interaction error presented by local and semilocal formalism, resulting in inaccurate pictures toward magnetism of strongly correlated materials.16–19 Furthermore, the inclusion of dispersion effect on hybrid functionals induces smallest values for JA–A suggesting a magneto-structural relation. Thus, DFT+D formalism overestimates the distortion degree for AO6 clusters resulting in longer A–O–A paths (Table S2) with high ΘA–O–A (Table 1) bond angle. Such distortion reflects on different covalent bond character for the different hybrid formalism, as presented in Table S3. It was observed that the inclusion of dispersive effects on hybrid functionals induces a higher ionic character for A–O interactions than pure formalism, in excellent agreement with A–O–A bond angle (Table 1).

The superexchange JA–A coupling constant follows the GKA rules through a virtual electron transfer between magnetic A2+ (A = Mn, Fe, Ni) cation using the molecular orbital originated from the mixing between transition metal 3d and oxygen 2p orbitals. Therefore, the covalent/ionic character of A–O chemical bonds play a fundamental role on the magnetic ground-state stabilization, since the GKA theory predict that the increase of covalent character for A–O bonds extends out the 3d metal wavefunction over the oxygen atoms facilitating the electron transfer responsible for the antiferromagnetic ordering.76 In this way, the high AO6 octahedral distortion obtained with dispersive augmented hybrid functionals induces an increased ionic character for A–O bonds (Table S3) resulting in more positive values for JA–A.

Looking at the chemical dependence of JA–A along the transition metal series (A = Mn, Fe, Ni) it was observed that the antiferromagnetic ordering strength increases with 3d orbital occupancy in agreement with theoretical results obtained for AO oxides, as well reported by Fennie using DFT+U.44,77 This tendency can be explained by the A–O covalent character that increases moving from Mn to Ni (Table S3), meaning an extended 3d–2p overlap stabilizing the antiferromagnetic ordering.

3.3 Electronic properties

In order to investigate the electronic structure of ATiO3 (A = Mn, Fe, Ni) materials, theoretical results for band structure profiles and density of states (DOS) were analyzed. Firstly, it was analyzed the effect of different exchange–correlation functionals on the band-gap description of strongly correlated ATiO3 materials, as presented in Fig. S1. It was noted that conventional exchange–correlation functionals (LDA and PBESol – Fig. S1m–r) underestimate the semiconductor behavior of ATiO3 materials, suggesting band-gap values around 0–1.5 eV. This failure can be clearly noted for the band structure of FeTiO3 obtained with LDA and PBESol functionals (Fig. S1n and q), where a metallic behavior was found. These results are originated from the self-interaction error presented in such functionals, as well observed for other strongly correlated oxides.19,30,49,78 On the other hand, pure and dispersive augmented hybrid functionals successfully predict the semiconductor behavior for all ATiO3 (A = Mn, Fe, Ni) materials without meaningful differences between B3LYP+D(PBE0+D) and B3LYP(PBE0) results, as presented in Table 2. Further, comparing our results with reported DFT+U one for MnTiO3 and NiTiO3, we noted that Hubbard corrected DFT results are very similar to the standard exchange–correlation functionals and smaller than hybrid formalism. It is important to clarify that experimental measurements of electronic band-gap for ATiO3 (A = Mn, Fe, Ni) materials have not been reported to date; however, the existence of ferroelectric properties in such materials suggest a semiconductor–insulator behavior following the electronic results reported for the ilmenite polymorph, where all ATiO3 candidates are semiconductors with band-gap around 2.0–3.5 eV.79
Table 2 Theoretical results for electronic band-gap (in eV) of ATiO3 (A = Mn, Fe, Ni) materials at different exchange–correlation functionals
  Mn Fe Ni
a Ref. 50.b Ref. 51.c Ref. 49.
B3LYP 3.89 (LΓ) 2.79 (TΓ) 4.42 (LΓ)
B3LYP+D 3.95 (ΓΓ) 2.77 (TΓ) 4.43 (LΓ)
PBE0 4.48 (ΓΓ) 3.30 (TΓ) 5.12 (LΓ)
PBE0+D 4.51 (ΓΓ) 3.30 (TΓ) 5.13 (LΓ)
PBESol 1.37 (LΓ) Metallic 1.78 (FB–Γ)
LDA 1.12 (LΓ) Metallic 1.57 (FB–Γ)
DFT+U 1.79a 2.35c
0.85b


Concerning the A-site cation effect on the electronic band-gap of the titanate materials, it was observed that the smallest excitation energy (Table 2) follows the order FeTiO3 < MnTiO3 < NiTiO3. Such effect can be explained from the 3d site orbital occupation and the crystal field splitting for t2g and eg levels. Moving from Mn to Ni, the electrons are included in t2g (t32g, t42g and t62g), while eg remains constant (e2g) increasing the crystal field splitting. In this way, the valence band region (VBR) is expected to be sensibly different along the transition metal series, in agreement with theoretical results presented in Fig. S1. These results show that the VBR for NiTiO3 has the higher degeneration character, while MnTiO3 and FeTiO3 show an internal gap around 1.0 and 1.5 eV, respectively. This increased internal band-gap for FeTiO3 is originated from the energy difference between double-occupied t22g and degenerated single-occupied t2g orbitals. Further, the t22g orbitals moves back in energy, the single-occupied t2g states are displaced to higher energy resulting in a reduced band-gap comparing to nickel and manganese titanates. Regarding the electronic excitation nature, theoretical results obtained with hybrid formalism show a direct band-gap for MnTiO3 between ΓΓ points, while FeTiO3 and NiTiO3 have indirect band-gaps at TΓ(LΓ) points.

In order to understand the electronic structure of the ATiO3 candidates, total and atom-resolved density of states were investigated, as presented in Fig. 3. From now, all results are based on PBE0 formalism, which best performs the structural, magnetic and band-gap results for strongly correlated materials proposed in this work. Toward the electronic state contribution to both valence (VB) and conduction band (CB) regions, it was noted that transition metal titanates posses the same pattern of distribution. The general order can be described by a VB predominantly composed by oxygen states highly hybridized with A-site cation valence orbitals (A = Mn, Fe, Ni), while the CB region was mostly based of empty valence orbitals from titanium atoms. However, it was observed on the A-site an increasing cation contribution to the formation of CB moving from Mn to Ni due to sequential filling of 3d orbitals, suggesting a possible electron transfer between 3d orbitals. Further, these results clearly indicate that 3d orbitals for both A-site cation and titanium atoms are divided in low-(t2g) and high-lying (eg) energy levels under the trigonal crystal field. However, the splitting energy for Ti(3d) orbitals is higher that of the A(3d) orbitals for MnTiO3 and NiTiO3 cases (Fig. 3a and b), which can be explained by the more severe distortional degree observed for TiO6 cluster comparing to (Mn/Ni)O6 clusters (see Table S2), in excellent agreement with theoretical and experimental results.79,80 Unlike the (Mn/Ni)TiO3 materials, FeTiO3 (Fig. 3c) exhibits a larger crystal field splitting for Fe(3d) orbitals than Ti(3d), mainly due to the 3d orbital occupancy. In trigonal systems, like R3c crystalline structure, the distortion on MO6 octahedral induces the splitting of t2g levels into non-degenerated a1g and double-degenerated eπg states, while the eg becomes eσg. Thus, the 3d6 orbital occupancy of Fe2+ cation induces the electron pairing in non-degenerated a1g orbital, while eπg states remains partially occupied, which allows an internal splitting gap between such states summed to the eπg and eσg contribution, resulting in higher crystal field splitting energy. Furthermore, such electronic configuration changes the upper VB region from highly hybridized 3dxy/yz–2py/z states in (Mn/Ni)TiO3 (Fig. 3a and b) materials to non-bonding 3dz2 states for FeTiO3 (Fig. 3c), in agreement with the reduced band-gap (Table 2) and stronger antiferromagnetic coupling (Table 1).


image file: c6ra21465g-f3.tif
Fig. 3 Total and atom-resolved density of states (DOS) for MnTiO3 (a), FeTiO3 (b) and NiTiO3 (c) materials at PBE0. In all cases the Fermi level was set to zero.

Regarding the bonding overlap, Fig. 3 allows to point out that A–O bond interaction is originated from anion-metal overlap between 2–0 eV, 3–0 eV, 2–0 eV energy range of upper VB in MnTiO3, FeTiO3 and NiTiO3, respectively. On the other hand, Ti–O bonding interactions are concentrated along the CB region at 4–6 eV, 3–5 eV, 5–7 eV energy range moving from Mn to Ni. From the combination between bonding overlap it is possible to suggest the existence of A–O–Ti–O–A intermetallic connection, as reported for the ilmenite structure.81 However, in this case the structural-induced charge ordering occurs in both ab plane and c-axis direction.

3.4 Topological analysis

The topological analysis of the electron density (ρ) is a powerful tool to investigate the electronic properties associated with the M–O bond paths in crystalline materials. In this study, it was focused on the properties of ρ(r) at A–O (A = Mn, Fe, Ni) and Ti–O BCPs, as presented in Table 3. A simple discussion for bond interactions within the framework of the QTAIM may be customarily performed in terms of the dichotomous classification based on the sign of ∇2ρ, which is equal to 4 × (2G(r) + V(r)), where G is the positive definite electron kinetic energy density and V is the electron potential energy density (which is always a negative quantity). Thus, the sign of ∇2ρ reflects the dominant energy contribution on the local expression of the virial theorem. The Laplacian of ρ is given by the sum of the three principal curvatures of the electron density ρ, where λ1,2 are the negative curvatures and define, through their associated eigenvectors, a plane perpendicular to the bond path at BCP, while λ3 is positive and its associated eigenvector is tangent to the bond path. Resuming, ∇2ρ expresses the competition between charge accumulation toward the bond path (λ1,2) for shared-shell (covalent and polar) bonds, and the contraction of electronic charge within the atomic basins (λ3) for closed-shell (ionic, H-bonds and vdW) interactions.73 However, from Sc to Ge the outermost N-shell of charge depletion and concentration is missing in the atomic Laplacian description and the ∇2ρ is always positive for 4th-row atoms (including Ti, Mn, Fe, Ni), making the dichotomous classification inadequate.82
Table 3 Electron charge density (ρ × 100), Laplacian (∇2ρ × 100), Hessian eigenvalues (λn × 100), ellipticity, and Bader's charge obtained by topological analysis in the framework of quantum theory of atoms in molecules for ATiO3 (A = Mn, Fe, Ni) multiferroic materials. All results are in atomic units. Rx correspond to the atomic dBCP for M–O (xy) interactions
  Topological properties Bader's Charge
dM–O Rx Rx/dM–O ρ 2ρ λ1 λ2 λ3 ε A Ti O
MnTiO3 Mn–O 2.116 1.047 0.495 6.2 30.0 −7.7 −7.6 45.3 0.009 1.559 2.402 −1.322
Ti–O 1.873 0.942 0.503 7.2 31.2 −10.3 −10.2 51.7 0.010
FeTiO3 Fe–O 2.087 1.031 0.495 6.5 32.3 −8.5 −8.2 49.1 0.031 1.493 2.400 −1.297
Ti–O 1.867 0.939 0.503 6.9 29.6 −9.7 −9.7 48.9 0.007
NiTiO3 Ni–O 2.042 0.988 0.517 6.6 38.4 −7.3 −6.8 52.5 0.078 1.376 2.411 −1.260
Ti–O 1.856 0.934 0.503 6.8 28.7 −9.6 −9.5 47.8 0.010


Concerning the BCP distance, obtained Rx/dM–O relations for Mn–O and Fe–O bond interactions indicate that the critical point is almost equidistant to the involved atoms, whereas for Ni–O is dislocated toward oxygen, suggesting a negative charge population for nickel atoms according to the increased bonded radius. On the other hand, reported Rx/dM–O values for Ti–O are invariant along the transition metal series, clarifying that the A-site cation modification induces a local charge disorder in the ATiO3 materials. Furthermore, it was observed that the BCP electron density (ρ) increases at A–O bonds moving from Mn to Ni, while reduces for Ti–O, indicating that oxygen atoms prefer to localize their valence electrons in the AO6 cluster at the expense of TiO6, according with the increase of the A-site cation electronegativity.

Then, the integral atomic basin properties in the framework of QTAIM were calculated, as presented in Table 3. Theoretical results show that the A-site cation net charge decreases moving from Mn to Ni, in accordance with the increased BCP electron density (ρTable 3), suggesting a higher covalent character for A–O bond interactions. Meanwhile, the Ti charge remains almost constant along the series, enabling to point out that the A-site cation replacement has a primary effect on oxygen atoms that withdraw an almost constant number of electrons from Ti. Such statement allows to interpret the effect of A-site cation on the charge balance through a local disorder that modifies, mainly, the A–O bond character, which is responsible for several properties of the ATiO3 candidates, e.g. the antiferromagnetic order enhancement along the transition metal series from the extended 3d–2p overlap.

Further, we classify the A–O and Ti–O bond interaction using the integral atomic basin properties (charge) obtained from Bader's theory. For this purpose, we use an independent coordinate through a charge-transfer scale proposed by Mori-Sanchez and co-workers, as follows:83

 
image file: c6ra21465g-t2.tif(4)

N represents the number of involved atoms, qb and qo correspond to the net charge from Bader's analysis and nominal oxidation state in fully ionic picture (A2+, Ti4+ and O2−), respectively. Mori-Sanchez and co-workers developed and applied such tool to classify a large number of crystalline materials as ionic, covalent or metallic.83 In this study, we use the charge-transfer scale to classify both A–O and Ti–O bond interactions, comparing the general trends with theoretical analysis of ρ (Table 3). The obtained c parameter for A–O bond interaction are 0.68 (Mn), 0.66 (Fe) and 0.64 (Ni), suggesting an increased covalent character along the series in agreement with the reported values of the electron density (Table 3). For Ti–O bonds the c parameters were calculated as 0.65 (Mn), 0.64 (Fe) and 0.63 (Ni), indicating an enhancement of the covalent character along the series, also for A–O bonds. Such effect can be explained by the charge balance spread along the A–O–Ti–O–A intermetallic connection, once the Ti charge is stable along the series and the A-site cation replacement reflects in less negative oxygen anions. The calculated charge-transfer constant (c) for the entire crystal are in 0.64–0.67 range, proving the existence of polar component along the M–O interactions, as expected for ferroelectric materials.

4 Conclusions

Density functional theory calculations employing local (LDA), semilocal (PBESol), pure (B3LYP and PBE0) and dispersive augmented (B3LYP+D and PBE0+D) hybrid functional were carried out to investigate the structural, magnetic and electronic properties of ATiO3 (A = Mn, Fe, Ni) multiferroic materials at LiNbO3-type structure.

Regarding the role of exchange–correlation functionals on the description of multiferroic properties, local (LDA) and semilocal (PBESol) formalism fails to describe the structure, as well the antiferromagnetic ordering and semiconductor behavior of ATiO3 materials. Indeed, hybrid DFT/HF functionals are better for all investigated properties of the strongly correlated ATiO3 materials. However, the inclusion of dispersive effects at hybrid Hamiltonians induces higher deviations when compared to the pure one. Thus, PBE0 best performs the investigated properties along the employed exchange–correlated functionals. Further, comparing such results with reported DFT+U is possible to argue that hybrid functionals overcome the Hubbard corrected DFT formalism, widely used for multiferroic investigations.

Calculated lattice parameters shown that unit cell volume for ATiO3 (A = Mn, Fe, Ni) materials is controlled by the A-site cation ionic radius. Following such behavior, the AO6 cluster becomes more ordered along the transition metal series in agreement with the 3d orbital contraction resulting from the increased orbital occupancy. However, the distortion degree of TiO6 increases, suggesting the enhancement of SOJT effect – a fundamental stabilization criterion of polar structures as R3c polymorphs.

The magnetic properties were evaluated by the A2+ next neighbors exchange coupling constant. All investigated transition metal titanates are antiferromagnetic following the GKA rules through a virtual electrons transfer between 3d orbitals of the A2+ cation and oxygen 2p states. Further, we verify the dependence of JAA as regard the covalent character of A–O–A bond interactions, which increases moving from Mn to Ni resulting in more antiferromagnetic materials.

Electronic properties section clarifies the semiconductor behavior of ATiO3 (A = Mn, Fe, Ni) multiferroic materials with band-gap around 3.3–5.2 eV. Furthermore, the nature of band-gap excitation was rationalized as direct for MnTiO3, while (Fe,Ni)TiO3 materials posses indirect excitation nature. Regarding the A-site cation effect on band-gap, the calculated values increase along the transition metal series in accordance with the 3d orbital occupancy and crystal field splitting (Δo). Such effects were deeply investigated from DOS analyses showing that both A-site and Ti cations 3d orbitals are divided in a1g, eπg and eσg levels, being the ΔTi-3d higher than ΔA-3d due to the octahedral distortion degree. Bonding overlap was also observed in DOS results, suggesting the existence of A–O–Ti–O–A intermetallic connection, as widely reported for the ilmenite polymorph. Therefore, we can argue that hybrid HF/DFT formalism are the most suitable choice to investigate the strongly correlated multiferroic candidate, especially to design new candidates in silico.

Topological analysis of ρ and Bader atomic charges reveal the main effect behind the A-site cation replacement along the ATiO3 candidates, which corresponds to a charge concentration in the AO6 clusters at the expense of TiO6, following the electronegativity trend along the transition metal series. Such effect was discussed as a local charge disorder, which has a primary effect on A–O bond interactions, being responsible to control several properties of LiNbO3-type materials.

In this way, all computed parameters are useful to understand the nature of intriguing properties of multiferroic materials class. Moreover, the systematic exchange–correlation functional investigation validates the mightiness of DFT plus hybrid functionals to successfully predict new candidates in this field.

Acknowledgements

The authors acknowledge UEPG, Capes and Araucaria Foundation for financial support.

References

  1. C. Ederer and N. A. Spaldin, Curr. Opin. Solid State Mater. Sci., 2005, 9, 128–139 CrossRef CAS.
  2. N. A. Hill, Annu. Rev. Mater. Res., 2002, 32, 1–37 CrossRef CAS.
  3. S. Picozzi and A. Stroppa, Eur. Phys. J. B, 2012, 85, 240 CrossRef.
  4. N. A. Hill, J. Phys. Chem. B, 2000, 104, 6694–6709 CrossRef CAS.
  5. J. Varignon, N. C. Bristowe, E. Bousquet and P. Ghosez, C. R. Phys., 2015, 16, 153–167 CrossRef CAS.
  6. T. Birol, N. A. Benedek, H. Das, A. L. Wysocki, A. T. Mulder, B. M. Abbett, E. H. Smith, S. Ghosh and C. J. Fennie, Curr. Opin. Solid State Mater. Sci., 2012, 16, 227–242 CrossRef CAS.
  7. N. A. Spaldin and W. E. Pickett, J. Solid State Chem., 2003, 176, 615–632 CrossRef CAS.
  8. N. A. Spaldin and M. Fiebig, Science, 2005, 309, 391–392 CrossRef CAS PubMed.
  9. A. Stroppa, M. Marsman, G. Kresse and S. Picozzi, New J. Phys., 2010, 12, 093026 CrossRef.
  10. M. Goffinet, P. Hermet, D. I. Bilc and P. Ghosez, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 014403 CrossRef.
  11. X. Rocquefelte, K. Schwarz and P. Blaha, Sci. Rep., 2012, 2, 759 Search PubMed.
  12. A. Stroppa and S. Picozzi, Phys. Chem. Chem. Phys., 2010, 12, 5405–5416 RSC.
  13. A. Jain, G. Hautier, C. J. Moore, S. P. Ong, C. C. Fischer, T. Mueller, K. A. Persson and G. Ceder, Comput. Mater. Sci., 2011, 50, 2295–2310 CrossRef CAS.
  14. J. Hafner, C. Wolverton and G. Ceder, MRS Bull., 2006, 31, 659–668 CrossRef.
  15. H. Jiang, Int. J. Quantum Chem., 2015, 115, 722–730 CrossRef CAS.
  16. I. Solovyev, N. Hamada and K. Terakura, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 53, 7158–7170 CrossRef CAS.
  17. P. Mori-Sanchez, A. J. Cohen and W. Yang, J. Chem. Phys., 2006, 125, 201102 CrossRef PubMed.
  18. C. Franchini, R. Podloucky, J. Paier, M. Marsman and G. Kresse, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 195128 CrossRef.
  19. F. Tran, P. Blaha, K. Schwarz and P. Novák, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 155108 CrossRef.
  20. A. Svane and O. Gunnarsson, Phys. Rev. Lett., 1990, 65, 1148–1151 CrossRef CAS PubMed.
  21. L. Hedin, Phys. Rev., 1965, 139, A796–A823 CrossRef.
  22. A. I. Liechtenstein, V. I. Anisimov and J. Zaanen, Phys. Rev. B: Condens. Matter Mater. Phys., 1995, 52, R5467–R5470 CrossRef CAS.
  23. S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J. Humphreys and A. P. Sutton, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 57, 1505–1509 CrossRef CAS.
  24. V. I. Anisimov, J. Zaanen and O. K. Andersen, Phys. Rev. B: Condens. Matter Mater. Phys., 1991, 44, 943–954 CrossRef CAS.
  25. H. J. Kulik, J. Chem. Phys., 2015, 142, 240901 CrossRef PubMed.
  26. T. Eom, H.-K. Lim, I. W. A. Goddard and H. Kim, J. Phys. Chem. C, 2015, 119, 556–562 CAS.
  27. A. D. Becke, J. Chem. Phys., 2014, 140, 18A301 CrossRef PubMed.
  28. F. Iori, M. Gatti and A. Rubio, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 115129 CrossRef.
  29. P. Rivero, I. d. P. R. Moreira, G. E. Scuseria and F. Illas, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 245129 CrossRef.
  30. C. Franchini, R. Podloucky, J. Paier, M. Marsman and G. Kresse, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 195128 CrossRef.
  31. D. I. Bilc, R. Orlando, R. Shaltaf, G.-M. Rignanese, J. Íñiguez and P. Ghosez, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 77, 165107 CrossRef.
  32. C. Franchini, J. Phys.: Condens. Matter, 2014, 26, 253202 CrossRef PubMed.
  33. M. Goffinet, P. Hermet, D. I. Bilc and P. Ghosez, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 014403 CrossRef.
  34. A. Stroppa and S. Picozzi, Phys. Chem. Chem. Phys., 2010, 12, 5405–5416 RSC.
  35. K. A. McDonnell, N. Wadnerkar, N. J. English, M. Rahman and D. Dowling, Chem. Phys. Lett., 2013, 572, 78–84 CrossRef CAS.
  36. X. F. Hao, A. Stroppa, P. Barone, A. Filippetti, C. Franchini and S. Picozzi, New J. Phys., 2014, 16, 015030 CrossRef.
  37. A. Prikockyte, D. Bilc, P. Hermet, C. Dubourdieu and P. Ghosez, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 214301 CrossRef.
  38. J. Varignon and P. Ghosez, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 140403 CrossRef.
  39. X. Ming, F. Hu, F. Du, Y.-J. Wei and G. Chen, Eur. Phys. J. B, 2015, 88, 212 CrossRef.
  40. M. Goffinet, J. Iniguez and P. Ghosez, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 86, 024415 CrossRef.
  41. S. Picozzi and A. Stroppa, Eur. Phys. J. B, 2012, 85, 240 CrossRef.
  42. J. Hong, A. Stroppa, J. Íñiguez, S. Picozzi and D. Vanderbilt, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 054417 CrossRef.
  43. E. Heifets, E. A. Kotomin, A. A. Bagaturyants and J. Maier, J. Phys. Chem. Lett., 2015, 6, 2847–2851 CrossRef CAS PubMed.
  44. C. J. Fennie, Phys. Rev. Lett., 2008, 100, 167203 CrossRef PubMed.
  45. T. Varga, A. Kumar, E. Vlahos, S. Denev, M. Park, S. Hong, T. Sanehira, Y. Wang, C. J. Fennie, S. K. Streiffer, X. Ke, P. Schiffer, V. Gopalan and J. F. Mitchell, Phys. Rev. Lett., 2009, 103, 047601 CrossRef CAS PubMed.
  46. A. Aimi, T. Katsumata, D. Mori, D. Fu, M. Itoh, T. Kyamen, K. Ichi Hiraki, T. Takahashi and Y. Inaguma, Inorg. Chem., 2011, 50, 6392–6398 CrossRef CAS PubMed.
  47. A. M. Arévalo-López and J. P. Attfield, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 88, 104416 CrossRef.
  48. T. Varga, T. C. Droubay, M. E. Bowden, R. J. Colby, S. Manandhar, V. Shutthanandan, D. Hu, B. C. Kabius, E. Apra, W. A. Shelton and S. A. Chambers, J. Vac. Sci. Technol., B: Nanotechnol. Microelectron.: Mater., Process., Meas., Phenom., 2013, 31, 030603 Search PubMed.
  49. X. Hao, Y. Xu, C. Franchini and F. Gao, Phys. Status Solidi B, 2015, 252, 626–634 CrossRef CAS.
  50. C. Xin, Y. Wang, Y. Sui, Y. Wang, X. Wang, K. Zhao, Z. Liu, B. Li and X. Liu, J. Alloys Compd., 2014, 613, 401–406 CrossRef CAS.
  51. X. Deng, W. Lu, H. Wang, H. Huang and J. Dai, J. Mater. Res., 2012, 27, 1421–1429 CrossRef CAS.
  52. B. A. Wechsler and C. T. Prewitt, Am. Mineral., 1984, 69, 176–185 CAS.
  53. J. Ko and C. T. Prewitt, Phys. Chem. Miner., 1988, 15, 355–362 CrossRef CAS.
  54. K. Leinenweber, W. Utsumi, Y. Tsuchida, T. Yagi and K. Kurita, Phys. Chem. Miner., 1991, 18, 244–250 CrossRef CAS.
  55. I. d. P. R. Moreira, R. Dovesi, C. Roetti, V. R. Saunders and R. Orlando, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 62, 7816–7823 CrossRef CAS.
  56. F. Cora, Mol. Phys., 2005, 103, 2483–2496 CrossRef CAS.
  57. M. D. Towler, N. L. Allan, N. M. Harrison, V. R. Saunders, W. C. Mackrodt and E. Aprà, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 5041–5054 CrossRef CAS.
  58. S. H. Vosko, L. Wilk and M. Nusair, Can. J. Phys., 1980, 58, 1200–1211 CrossRef CAS.
  59. 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.
  60. A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS.
  61. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS.
  62. C. Adamo and V. Barone, J. Chem. Phys., 1999, 110, 6158–6170 CrossRef CAS.
  63. S. Grimme, J. Comput. Chem., 2006, 27, 1787–1799 CrossRef CAS PubMed.
  64. R. Dovesi, V. R. Saunders, C. Roetti, R. Orlando, C. M. Zicovich-Wilson, F. Pascale, B. Civalleri, K. Doll, N. M. Harrison, I. J. Bush, P. D'Arco and M. Llunell, CRYSTAL09 User's Manual, 2009 Search PubMed.
  65. R. Dovesi, R. Orlando, B. Civalleri, C. Roetti, V. R. Saunders and C. M. Zicovich-Wilson, Z. Kristallogr.–Cryst. Mater., 2005, 220, 571–573 CAS.
  66. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Condens. Matter Mater. Phys., 1976, 13, 5188–5192 CrossRef.
  67. C. Gatti, TOPOND-98: An Electron Density Topological Program for Systems Periodic in N (N = 0-3) Dimensions. User’s Manual, 1999 Search PubMed.
  68. R. F. W. Bader, Atoms in Molecules: A Quantum Theory (International Series of Monographs on Chemistry), Oxford University Press, 1994 Search PubMed.
  69. L. Bertini, F. Cargnoni and C. Gatti, Theor. Chem. Acc., 2007, 117, 847–884 CrossRef CAS.
  70. C. Gatti, Phys. Scr., 2013, 87, 048102 CrossRef.
  71. C. Gatti, V. R. Saunders and C. Roetti, J. Chem. Phys., 1994, 101, 10686–10696 CrossRef CAS.
  72. C. Matta and R. Boyd, The Quantum Theory of Atoms in Molecules: From Solid State to DNA and Drug Design, Wiley, 2007 Search PubMed.
  73. C. Gatti, Z. Kristallogr.–Cryst. Mater., 2005, 220, 399–457 CAS.
  74. A. R. Albuquerque, M. L. Garzim, I. M. G. dos Santos, V. Longo, E. Longo and J. R. Sambrano, J. Phys. Chem. A, 2012, 116, 11731–11735 CrossRef CAS PubMed.
  75. G. Radtke, S. Lazar and G. A. Botton, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 155117 CrossRef.
  76. J. Goodenough, Magnetism and the chemical bond, Interscience Publishers, 1963 Search PubMed.
  77. G. Fischer, M. Däne, A. Ernst, P. Bruno, M. Lüders, Z. Szotek, W. Temmerman and W. Hergert, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80, 014408 CrossRef.
  78. J. He and C. Franchini, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 86, 235117 CrossRef.
  79. A. Agui and M. Mizumaki, J. Electron Spectrosc. Relat. Phenom., 2011, 184, 463–467 CrossRef CAS.
  80. S. W. Chen, P. A. Lin, H. T. Jeng, S. W. Fu, J. M. Lee, J. F. Lee, C. W. Pao, H. Ishii, K. D. Tsuei, N. Hiraoka, D. P. Chen, S. X. Dou, X. L. Wang, K. T. Lu and J. M. Chen, Appl. Phys. Lett., 2014, 104, 082104 CrossRef.
  81. R. A. P. Ribeiro and S. R. de Lazaro, RSC Adv., 2014, 4, 59839–59846 RSC.
  82. C. Gatti and D. Lasi, Faraday Discuss., 2007, 135, 55–78 RSC.
  83. P. Mori-Sanchez, A. M. Pendas and V. Luana, J. Am. Chem. Soc., 2002, 124, 14721–14723 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra21465g

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.